ÔLife, Environment and PeopleÕ (BB30108) - Encouraging creative and critical biological and scientific enquiry into issues concerning human relationships with the living world

 

Course Notes

 

INTENTION

 

My intention in the course is to provide an opportunity for us to reflect and learn together about how to apply our scientific and biological knowledge effectively and creatively in a social and environmental context. This joint reflection and learning will include an enquiry into methods of scientific enquiry, perception and communication in order to identify possible limitations in current thinking and prospects for the development of approaches that can enhance and deepen our understanding of our human relationships with the living world.

 

An important aim will be to enhance our awareness of the uncertainties implicit in the dynamic processes that underlie the functioning and ecological and evolutionary responsiveness of living systems, with a view to developing patterns of working with these systems that can enhance quality of life. I will draw attention to the need to seek complementary relationship rather than opposition, and contrast this with purely objective approaches aimed at imposing control. I will suggest that working with living systems requires sensitivity to the receptivity and responsiveness of indeterminate (dynamically bounded) forms that is often absent when working upon Human-made constructions. The co-evolution of life forms in tune with their environmental context cannot be equated with the prescriptive precision of the assembly line. Life is not constructed from a set of pre-existing Ôbuilding blocksÕ: it is much richer, deeper and more intimately involved with its ever-changing living space than that metaphor implies.

 

My underlying personal commitment is to a future in which we human beings can enjoy more empathic relationships with one another, other life forms and our surroundings. For that future to emerge I feel - in common with a variety of other thinkers in the arts, humanities and sciences - that it is necessary to look in a fresh, more contextually aware way at the fundamental nature of life and evolutionary (irreversible) processes. I began to try to develop this fresh look in the following book:

 

A.D.M. Rayner (1997) Degrees of Freedom – Living in Dynamic Boundaries (Imperial College Press)

 

You are recommended to read this book as an adjunct to the course. Although it has its faults and is not as up-to-date, clear or fully developed as IÕd like it to be, it does provide an overview of many of the basic themes covered by the course and the issues that arise from them. I am currently writing a new book about Ôthe vital place of space in lifeÕ, and in the meantime have compiled a multi-authored  Ôe bookÕ entitled ÔInclusionality: The Science, Art and Spirituality of Place, Space and EvolutionÕ. The latter can be downloaded (or you can purchase a CD directly from me for £10) via my homepage at http://people.bath.ac.uk/bssadmr, where numerous other writings, artworks and links can also be found.

 

 

Another recommended book, which raises similar and related issues whilst coming from a more conventional biological, environmental and sociological standpoint, and contains numerous interesting citations and quotations is:

 

D. Suzuki with Amanda McConnell (1999) The Sacred Balance – Rediscovering Our Place in Nature (Bantam Books)

 

 

For an excellent, wide-ranging and very readable account of the history and possible future development of the environmental movement, see:

 

Spowers, R. (2002) Rising Tides. Edinburgh: Canongate Books.

 

And for an insightful and accessible critique of modern and postmodern ideas concerning the dynamic reality of our living space, see:

 

Spretnak, C. (1999). The Resurgence of the Real: Body, Nature and Place in a Hypermodern World. New York: Routledge.

 

 

Other reference sources that you may find helpful are listed at the end of these notes.

 

 

Since the intention of this course is to reflect and learn together, you are not necessarily expected to agree with, understand or re-iterate everything I say in order to do well! My perspectives, like anyone elseÕs, inevitably reflect my unique personal situation and experience. Indeed, understanding how it is impossible for us as individuals to obtain a comprehensive Ôall-roundÕ view, and how this limitation may be overcome by sharing our unique perspectives with one another, is one of the themes I will be considering. I therefore prefer to think of myself as your experienced educational ÔfacilitatorÕ, rather than authoritative training instructor. Correspondingly, what I will be looking for, hoping for, is that you will use the course as a way of opening up your own horizons, finding ways to think critically and creatively about environmental and social concerns and how your own knowledge and experience can contribute to greater understanding. The ÔRound TableÕ discussion sessions on topical issues at the end of the lecture series are intended to help you develop these skills.


STRUCTURE OF THE COURSE

The course schedule is given below. You will see that there are nine double ÔLecturer-led SessionsÕ and three double ÔRound Table SessionsÕ. For timetabling purposes it has been necessary to split each double session into two sessions separated by a gap. This is not what I want ideally, but the gap will at least offer an opportunity to pause for reflection. It has also been necessary to separate Biology Students from Natural Sciences, Psychology and Management students, which is anathema to me, but will keep the group size small enough for all to participate readily.

 

In each of the Lecturer-led sessions I will initiate a conversation about a scientific/biological theme, based upon, but not necessarily exactly corresponding with the notes provided in the following pages that outline some of my own perspectives and relevant knowledge. Please donÕt be daunted if you donÕt understand the notes on first reading: understanding is something that I hope will evolve during our conversations. As I have already indicated, I wonÕt be expecting or wanting you to ÔreproduceÕ the information and ideas contained in these notes for assessment purposes (see below). This especially applies to the opening discussion of ÔrationalityÕ and ÔinclusionalityÕ, the full significance of which will probably only really begin to clarify once the course has been completed. So, please be patient.

 

In each of the Student-led ÔRound-TableÕ sessions, a third of the class will form a group who will choose a suitable environment-related issue or question for discussion. This discussion group will organise a Ômini-conferenceÕ in which each group member will represent a particular aspect of the topic. The remainder of the class will provide an ÔaudienceÕ who will be free to question the discussion group and comment on the issues raised. Suitable topics for discussion might include: ÔsustainabilityÕ; ÔbiotechnologyÕ; ÔglobalizationÕ; ÔconsumerismÕ; Ôthe nature of timeÕ; Ôclimate changeÕ; ÔconflictÕ; ÔconservationÕ; Ômodes of governanceÕ; ÔpollutionÕ; Ôscience and religionÕ etc, etc

 

Lecturer-led Session 1 ÒRationalistic and ÔInclusionalÕ Views of NatureÓ (14, 18 Feb [B]; 15, 18 Feb [NS, PS, MN])

Lecturer-led Session 2 ÒWater as the Dynamic Context of LifeÓ  (21, 25 Feb [B]; 22, 25 Feb [NS, PS, MN])

 

Lecturer-led Session 3 ÒDivided and United States I – Genes and CellsÓ  (28 Feb, 4 Mar [B]; 29 Feb, 4 Mar [NS, PS, MN])

Lecturer-led Session 4 ÒDivided and United States II – Bodies, Societies and CommunitiesÓ(7, 11 Mar [B];8 , 11 Mar [NS, PS,MN])

Lecturer-led Session 5 ÒDeath and AlternativesÓ  (14, 18 Mar [B]; 15, 18 Mar [NS, PS, MN])

 

Lecturer-led Session 6  ÒEvolutionary Transformation I – Creativity, Opportunity Space and Life HistoryÓ (11, 15 Apr [B]; 12, 15 Apr [NS, PS, MN])

Lecturer-led Session 7  ÒEvolutionary Transformation II – Elemental Themes and Mathematical AssumptionsÓ (18, 22 Apr [B]; 19, 22 Apr [NS, PS, MN])

Lecturer-led Session 8  ÒPeople and Environment I – Challenging UnpredictabilityÓ (25, 29 Apr [B]; 26, 29 Apr [NS, PS, MN]

Lecturer-led Session 9 ÒPeople and Environment II–Managing Life and EnvironmentÓ (2?, 6 May [B]; 3, 6 May [NS, PS, MN])

 

 

Round Table Session 1   (9, 13 May [B]; 10, 13 May [NS, PS, MN])

 

Round Table Session 2   (16, 20 May [B]; 17, 20 May [NS, PS, MN])

 

Round Table Session 3   (23, 27 May [B]; 24, 27 May [NS, PS, MN])


 

ASSESSMENT OF THE COURSE

 ÒIn order to do well, however, you will need to justify your position using relevant information, citations and examples, and show a critical awareness of other possible viewsÓ

 

In a course as intentionally co-explorative as this one, it is important that the means of assessing your learning should take account of the diversity of human perception and learning styles, and correspondingly different but valid and complementary forms of scientific expression and enquiry. Together with an independent internal assessor, I will need to assess your learning both as Ôuniquely situated individualsÕ and as participants in a Ôcollective processÕ in which there isnÕt a pre-determined Ôpath to successÕ. So, I will seek to encourage you to develop your personal potential rather than measure and rank your performance against a prescriptive ÔnormÕ. I hope you will therefore experience the assessment as an enjoyable and challenging adjunct to your education, which acknowledges the quality of critical and creative enquiry you express, rather than a narrow Ôonce-and-for-all-timeÕ judgement that disregards your personal context and potential. In other words, the aim is to recognize and reward high quality work of all kinds, with due regard to your situation and experience. 

 

Following consultation with colleagues, it has been decided this year for the first time to use a variety of solely COURSE ASSESSMENT methods. This will give you the fullest possible opportunity to express your learning from the course in a balanced, scholarly way without imposing unrealistic time constraints (as in an exam). 

 

COURSE ASSESSMENT ONE will concern your learning from the ÔLecturer-led sessionsÕ. It will receive a total of 40 % of the final marks for the course. It will involve providing an answer to the following question: How, in your view, may the application of scientific and biological knowledge and concepts in a social and environmental context be influenced by our human perceptions of space and boundaries? Your answer should contain no more than 1500 words of text, excluding figure legends and references. It may be accompanied by one or two illustrative figures with legends. Reference sources should be cited in a bibliography at the end. My use of Ôin your viewÕ enables me not to be too prescriptive with regard to what material you include (which can be from any sources, including other courses) in your answer, and to encourage you to develop and express your own thinking based on your own knowledge and experience. In order to do well, however, you will need to justify your position using relevant information, citations and examples, and show a critical awareness of other possible views. In other words you should avoid making any statements or assertions without backing these up with evidence and indicating possible sources of contention and uncertainty. I suggest that you prepare to answer this question by keeping a personal record of the conversations during the Lecturer-led sessions and thinking about how they relate to what you have heard and read about elsewhere. These conversations will hopefully reveal the diversity of scientific viewpoints that can arise about social and environmental issues, due to distinctive forms of cognition and associated logical premises.

 

COURSE ASSESSMENT TWO will concern your learning from the Student-led ÔRound TablesÕ (which is likely also to incorporate learning from the ÔLecturer-ledÕ sessions), and will receive 40 % of the final marks for the course. It will involve answering the following question: On the basis of the Round-Table Session in which you participated, consider an environment-related issue or question of your choice from as wide a variety of scientific, biological and other relevant perspectives as possible.  This question is framed so that firstly, you donÕt simply reproduce your individual contribution to a Round Table session (although you might well include this).  Secondly, I want you to ensure that your answer makes some attempt to apply your scientific and biological knowledge (awareness of information) and understanding (comprehension) rather than coming solely from a Ônon-scientificÕ (e.g. purely political, theological, artistic) perspective. I recognise and will take into account the fact that the balance of ÔscientificÕ, ÔbiologicalÕ and ÔotherÕ perspectives may vary between students attending different degree courses. As in course assessment 1, your answer should contain no more than 1500 words of text, excluding figure legends and references. It may be accompanied by one or two illustrative figures with legends. Reference sources should be cited in a bibliography at the end. You will gather that participation in a high quality round-table session is a way in which you can collectively help one another to do well in this assessment. Again, you should avoid making any statements or assertions without backing these up with evidence and indicating possible sources of contention and uncertainty.

 

 

COURSE ASSESSMENT THREE is intended to give you the freedom to express your learning about any aspect of the course that you like, based on a submitted piece of personal project work (a 1500 word essay or equivalent), and will receive 20 % of the final marks.  The actual form of your submission is up to you. Feel free to use conventional scientific expression if that suits your purpose, or, if you would like to try something ÔdifferentÕ, artwork, music, performance or poetry. I am happy to encourage diverse forms of exploration and expression, and use these myself, to open up new possibilities for scientific enquiry, understanding and communication.

 

 

 

EVALUATION CRITERIA

 

In view of the distinctive aspirations of this course, your work will be assessed using somewhat different criteria from those used in other courses. These criteria are as follows:

 

 

Reflective Quality:- does the work accurately and thoughtfully reflect themes emerging during the course? Are the scientific ideas that are conveyed and/or challenged fairly represented, in a way that demonstrates sound critical judgement/ understanding/scholarship in your own learning?

 

Creativity:- does the work display imaginative  thought and (where applicable) practical resourcefulness in relation to the theme/subject matter addressed?

 

Communicative Quality:- does the work communicate a clear message and/or evoke imagination and thought?

 

Quality of Execution:- is there evidence of skilful work?

 

Endeavour:- is there evidence of care and effort?

 

Was your contribution to class/round table discussion especially commendable? [NB This criterion is included because we may wish to reward special cases]

 

 


INTRODUCTION: THE PROBLEM OF RATIONALISTIC EXCLUSIVITY AND THE NEED FOR AN ÔINCLUSIONALÕ VIEW OF ENVIRONMENTAL RELATIONSHIPS

 

Preview: ÒInclusionality is an awareness that space, far from passively surrounding and isolating discrete massy objects, is a vital, dynamic inclusion within, around and permeating natural form across all scales of organization, allowing diverse possibilities for movement and communication. Correspondingly, boundaries are not fixed limits - smooth, space-excluding, Euclidean lines or planes - but rather are pivotal places comprising complex, dynamic arrays of voids and relief that both emerge from and pattern the co-creative togetherness of inner and outer domains, as in the banks of a riverÓ.

 

Summary: Modern understanding of dynamic processes of all kinds, from subatomic to universal scales and encompassing the evolution of living systems, continues to be restricted by the rationalistic treatment of boundaries as discrete limits and space as distance between material objects. Such treatment is founded mathematically in the geometry of Euclid and arithmetic of discrete numerical units, which formed the basis for Newtonian mechanics and the development of objective, quantitative science aimed at prediction and control. It lacks, however, an evidence base in being founded on the illusion that matter ultimately consists of solid, massy particles surrounded by (and hence excluding) non-interactive space. This illusion leads to the dualistic Ôparadoxes of completenessÕ that underlie the interpretation of change as the consequence of imposing purely external force upon discrete (isolated) and hence independent bodies, seen in life forms as the location of individual selves.

 

                        A radically more creative, evidence-based and ultimately less environmentally, socially and psychologically damaging perspective emerges when it is acknowledged that space inextricably permeates within, through and around - and not just outside - all physical forms, from sub-atomic to galactic in scale. This space opens up the possibility for movement and communication at the same time that it introduces uncertainty into what would otherwise be a static and impenetrable purely material world and universe. It is the source of fluidity that pools the swirling contents of the universe gravitationally together, like the solvent in a solution of solutes. Without it there wouldnÕt be any room for change.

 

There are no paradoxes of completeness in this perspective because there are no discrete bodies - no isolated wholes. Far from being complete, fixed Euclidean surfaces that simply divide insides from outsides, boundaries both form and are formed by the togetherness of inner and outer realms as complex, variably resistive, dynamic, space-incorporating transitions. As such, they both distinguish and reciprocally couple the local, inner (ÔindividualÕ) and non-local, outer (ÔcollectiveÕ) aspects of uniquely situated flow-form features of energy-space - relational places rather than discrete entities - nested over all scales and identifiable in life forms as Ôcomplex selvesÕ.

----------------------------------------------------------------------------------------------------------------------------------------

 

 

 

How do you view your ÔenvironmentÕ? Do you think of it as Ôwhatever is left over when you have singled yourself out of itÕ – whatever is ÔoutsideÕ of yourself? Or are you more personally involved in it? Is it something that you must adapt to and control if you are to survive and thrive? Or is it some kind of dynamic Ôliving spaceÕ that you incorporate into your life, which somehow becomes you as you become it? How far does this living space extend? Is it local or universal? How do you regard ÔothersÕ within this living space? Are they independent agents whose performance directly impinges on your own well being, that you collide and compete with and exploit to serve your own needs? Or are they receptive co-habitants capable both of transforming and of being transformed by the living space you share with them? Are they separate from you or a part of you? Are you a part of or apart from your environment? Is Ôyour environmentÕ the same as or different from ÔothersÕÕ environmentÕ? Which is of primary significance in shaping lives – the living space or its contents? Can ÔcontentsÕ precede their spatial context? Can contents meaningfully be separated from their context, such that their resultant isolation does not radically alter, if not utterly negate their behavioural possibilities?

 

Questions, questionsÉ.where do they come fromÉwhat do they meanÉ.why should we care?

 

Most fundamentally, I think that these questions reflect the way we perceive and respond to sources of certainty and uncertainty in our lives. The way that we answer them, both scientifically and culturally, therefore fundamentally affects our human relationships with one another and the living world that we inhabit.

 

For thousands of years, faced with the variability of our surroundings and needing to secure the means of our own survival, we human beings have tended to make an enemy out of uncertainty. The principal adversarial tactic that we have brought to bear on this enemy has been to try to exclude or confine it by imposing closure upon it. That is, we have tried mentally and/or physically to box uncertainty inside or outside absolutely fixed and sealed boundaries. This tactic has been reinforced psychologically by the predominant use of our object-defining binocular vision as a means of distinguishing between ÔselfÕ and ÔotherÕ, ÔfriendÕ and ÔfoeÕ, which gives rise to the cognitive illusion of an absolute demarcation between ÔsomethingÕ and ÔnothingÕ. The resultant focus on solid, particulate ÔmatterÕ as all that matters has in turn been at the core of systems of rationalistic philosophical, mathematical and scientific inquiry, founded on the discrete logical premise of the Ôlaw of the excluded middleÕ – the notion that everything is either A or not A.

 

By these means we have consciously or unconsciously sought to impose control over the wildness that we perceive both in nature and, if we allow it free access and expression, within ourselves. Essentially we attempt to ignore or even eradicate variability in a quest to standardize nature and human nature and establish and apply the laws and rules through which their behaviour can be rendered definable and predictable. And where we canÕt practicably ignore variability, we attribute it to Ôstatistical chanceÕ - the ÔrandomnessÕ of ÔindependentÕ (discrete/isolated) events – which we account for in terms of probability distributions and associated Ôrisk analysesÕ.

 

The fundamental problem with this approach is that it is liable to lead us into false senses of both security and insecurity, through which we are liable to bring about real damage to others and ourselves as well as fail to fulfil our creative human potential. This is because it fails to take account of overwhelming evidence - implicit in modern scientific theories of relativity, quantum mechanics and complex, non-linear systems - that it is impossible fully to exclude or confine the fundamental source of uncertainty anywhere, at any scale, in the ÔknownÕ Universe.  Indeed, without this source of uncertainty, life, the Universe and everywhere, including human beings would lose their dynamic, evolutionary, flow-form quality and become an impenetrable solid mass.

 

When we fail, literally, to see this fundamental source of uncertainty, and hence disregard it, we tend to inquire in a very particular way, which has become the hallmark of orthodox objective scientific methodology based on what I have called impositional logic. This kind of inquiry is made in the full expectation - providing we gather sufficient, accurate data - of fulfilling our prescriptive objectives and hence delivering outcomes in the form of definitive conclusions. Our research hence becomes extraordinarily goal-oriented, so much so that we feel obliged to overlook other possibilities as they crop up en-route. The destination becomes more important than the journey and the end may increasingly seem to justify the means. But in the process, our inquiries become prejudicial, prone to find what we are looking for and ignore the rest, so reinforcing the loop of closure from solid perception to definitive expectation, to objective inquiry to solid perception. Then when it all goes wrong, and outcomes arenÕt as expected, we search desperately for someone or something to blame, whilst overlooking the assumptions and tools of our inquiry.

 

So what is this fundamental source of uncertainty that we try to close down upon and treat as ÔnothingÕ? Well, actually, it is nothing material, but far from being an Ôabsence of material presenceÕ that we can discount, it is rather a Ôpresence of material absenceÕ that is vital to the dynamic world and Universe we inhabit. I am speaking here of the physical space that we can think of as the ubiquitous zero viscosity bathing fluid and fluidizing solvent, which, inseparably together with electromagnetic energy gives dynamically bounded form to Universal features everywhere, nested over all scales from subatomic to galactic. The awareness of this inextricable Universal presence of material absence as a dynamic inclusion within around and through all has been termed ÔinclusionalityÕ. It gives rise to a form of reasoned inquiry that is very different from and more evidence-based than that founded on the law of the excluded middle. The hallmark of this form of inquiry is its openness to possibility and resultant mindful, honest, response to the opportunities for serendipitous discovery and dangers of misadventure in an uncertain world.

 

Restrictive Practices: A Chamber of Horror Stories

As I have already intimated, orthodox methods of objective inquiry come close to Inquisitorial interrogation where the kinds of answers extracted may be misleading, to say the least. The inquiry invariably begins with the selection and abstraction of a sample, which is placed within some actual or theoretical limiting boundary or reference frame and then studied in isolation from its natural context. A part of nature is excised and brought under scrutiny within the imposed framework of the sampling grid, laboratory, containing vessel, experimental apparatus or mathematical construct.

 

The underlying hope of this kind of inquiry is that the small picture it provides of the part realistically represents the big picture of the whole from which the part was abstracted. But it is rather like trying to represent a river by scrutinizing the contents of a cup dipped into it! No form of inquiry based on the deliberate ignorance of spatial context can comprehend the behaviour and properties of a complex, dynamic system. True, comparison of the properties of water contained in a cup with those of the river may yield valuable insights into the dynamic possibilities of the latter and how these are affected by isolation within a fixed boundary. But to extrapolate from what can be defined within a fixed container to the uncertainties of the open field makes nonsense. Whatever certainty we may gain about the properties and behaviour of our isolated sample or system, comes at the expense of profound uncertainty about the applicability of our conclusions to the wider dynamic context.

 

Unfortunately, such extrapolation continues to be the mainstay of all kinds of technological product testing under highly restrictive conditions prior to release onto the open market. All kinds of tales of the unexpected may then unfold, from the decimation by a hitherto unknown fungal disease of a maize crop carrying a Ômale sterilityÕ factor, to thalidomide-induced deformities and the BSE epidemic.

 

What undermines public trust here is not so much the unexpected outcome in itself, for this aligns with our personal human experience of uncertainty, but the pretence of certainty on the part of the producers and the scientific and governmental communities. The product testing appears more and more like an exercise in self-deception designed to provide reassurance to suit vested interests, than an open-minded exploration of possibilities.

 

Inclusional Inquiry: Openness to Possibility

As I will discuss in more detail later, through including and coming to terms with the fundamental source of uncertainty, inclusionality leads to a new view of evolutionary (irreversible) processes of all kinds in terms of contextual transformation - the continual reconfiguration of spatial possibility in relation with informational, electromagnetic, inner-outer linings (dynamic boundaries). Such processes cannot be prescriptively defined and predicted in the long run, but they can be understood and their varied developmental potentialities anticipated, in much the same way that we can understand and anticipate, but not exactly forecast weather patterns. 

 

Inclusional inquiry is itself just such an evolutionary process, analogous to a coin rolling on edge or a riverÕs flow both shaping and being shaped by landscape via its banks and valley sides, that reciprocally balances, or attunes, inner with outer potentials via their intermediary interfaces in a resonant relationship.  Moreover it sees such dynamic balancing as vital to the health, diversity and creative and responsive potential of living systems as inner-outer couples and complex flow-forms. By the same token, disruption of this balance, through imposition of two-way or one-way closure, is seen as de-vitalizing, as when the coin settles on heads or tails.

 

Correspondingly, the fundamental intention of inclusional inquiry is not to oppose and provide a competitive ÔalternativeÕ, but rather to reframe and augment existing practice so that it can be made more congruent with our growing human knowledge and experience of dynamic processes. As befits an inquiry that is therefore necessarily open-ended and experience-based, its methods and approaches cannot usefully be prescribed in a laboratory manual or recipe book, which sets out an exact sequence of instructions targeted towards a specific end goal or product. Although it may well incorporate objective, quantitative methodologies and practices, there is an important place also for subjective improvisation and the development of craft and artistry – the kind of qualitative approaches that distinguish a ÔchefÕ from a ÔcookÕ and ÔeducationÕ from ÔtrainingÕ.  Subjectivity is important here because complex issues cannot adequately be understood from a single, pre-determined perspective but need to be viewed from Ôall-roundÕ so that their different contributory aspects can be appreciated and mindfully held together.  ÔIndividualÕ researchers can help to fulfil this need through the development of their contextual awareness and imagination, whilst making and interpreting their uniquely situated observations and sharing these in a complementary rather than competitive way with others in a truly collective, participatory enterprise.

 

So, my double-edged question to you is this. Do we want to open up or close down - can we see ourselves working imaginatively and openly together in a way that respects, values and anticipates diversity, or do we prefer to stick with the context-free concrete certainties of a one size fits all abstract mentality? Yes, I know, this question is loaded. But to me it all comes down to a matter of trust in human experience, mindful experimentation and careful monitoring, neither denying nor exaggerating risks or opportunities, but holding them sensitively in the balance, always prepared to accommodate whatever surprises emerge along the way. IsnÕt that what learning is all about?

 

 

COMBINING ARTISTIC AND SCIENTIFIC PERSPECTIVES OF IMPLICIT SPACE AND EXPLICIT INFORMATION

 

Perhaps, most fundamentally, what I am suggesting is that in order to attune more empathically with one another, other life forms and our surroundings, there is a need for a radical re-orientation in the way we think about the nature of contextual space. To achieve this will involve more than tinkering with our scientific methodologies and ways of seeing. It will involve a profound, and for many people disturbing upheaval of the intellectual landscape our rationalistic vision has conditioned us to believe in. In the words of novelist, Lindsay Clarke (pers. com.), it will involve nothing less than a Ôtransformation of consciousness in our timeÕ, a Ôshift away from the fissive mythology of positivism back towards a lively sense of the sacredÕ, a sense of Place. This is where I think our future environmental and social understanding can benefit especially from a re-integration of scientific and artistic perspectives.

 

For much of my own adult life, I tried to keep my work as a professional biological scientist separate from my personal pleasure in expressing my feelings artistically. Eventually, however, I began to think that something was missing from my practice of science, and that this absence was restricting both my understanding and enjoyment of what I was doing. I wondered whether it might be possible to restore this missing something by bringing my artistic sense of kinship with the living space within and around my self into communion with my scientific knowledge. I also wondered whether science in general might benefit from the inclusion of an artistic perspective, both in its approach to understanding nature and communicating that understanding to others. I felt this question was relevant because it was becoming very clear to me in the light of occurrences like BSE and debates over climate change and genetically modified organisms etc, that in the public mind, science had a serious communication problem. It came across as insensitive, unintelligible and increasingly unreliable as a basis for making any kinds of decisions about how to live in and interact with the Ôworld about usÕ. Moreover, it seemed to bring our materialistic worldview into a collision course with our emotional and spiritual yearnings for a sense of belonging and meaning.

 

And so, a few years ago, I took the opportunity to begin to explore this possibility personally, by preparing and presenting a painting entitled ÔFountains of the ForestÕ as part of my Presidential Address to the British Mycological Society, in 1998.

 

At about the same time, an interview with a well-known scientist in response to growing interest in re-connecting Art and Science, was broadcast on BBC Radio 4. To my dismay, the scientist pronounced that Art and Science were completely different human endeavours and should therefore keep a respectful distance from one another. I was taken aback, because the scientist seemed to be arguing that difference was a reason for staying apart, whereas I thought it was a reason for partnership, an opportunity to realize the new possibilities implicit in complementary viewpoints, as in a mutually beneficial symbiosis.

 

                        So, the difference between me and the well-known scientist seemed to lie in our attitude to difference. He wished to exclude it, for fear of the contamination, take-over and dysfunction it could bring about; I wished to include it for the new opportunities it might bring. He wanted Art and Science to agree to differ – each to adopt their own distinctive view of the world and not intrude upon one another, especially not Art into Science. I wanted them to differ to agree – to discover through their diverse perspectives a common but many-stranded reality, all views of which were necessarily partial but for that very reason also unique contributions to the overall picture, as in a hologram.

 

So, what, then, is the difference between Art and Science? We all assume thereÕs a difference, donÕt we? And so many of us have been required to make an either/or choice between these paths during our so-called ÔeducationÕ. However, this seeming difference between what C.P. Snow notoriously described in 1959 as Two Cultures, each using and abusing the other but rarely recognizing their common origin in philosophical inquiry, appears only to be a relatively recent phenomenon. Throughout human history, we have creatively combined acute, careful observation, description and technique with our ability to imagine possibilities, in our quest to understand and make our way in the world about us. ÔScienceÕ itself was only distinguished from ÔNatural PhilosophyÕ as a separate human endeavour in the nineteenth century. And the iconic figure of Leonardo da Vinci continues to be held up as the epitome of the creative potential that arises through the combination of artistic and scientific perspectives.

 

                        So, what happened to split our artistic and scientific Genius? Increasingly, postmodern philosophers are prone to link this split to the time that the French philosopher RenŽ Descartes spent locked in an oven to keep himself warm during a bleak winter. This insular experience is said to have led Descartes to proclaim ÔCogito ergo sumÕ and divide ÔmindÕ from ÔmatterÕ - the ÔCartesian SplitÕ, upon which the Scientific Revolution and Enlightenment movements are said to be based. Rather than attribute so much responsibility to one lonely person, however, I suggest that that this division was the ultimate rendering of an idea that had been a very long while in the making. I see it as the product of the Ôcognitive illusionÕ I have alluded to above, to which we are all subject in the way we perceive and consequently interpret the Ôworld about usÕ.

 

Most fundamentally, I think that art can help us out of this cognitive illusion and the trap of rationalistic impositional logic and 'either-or' thinking, in a way that can re-connect and so revitalize our reason with our emotions and our sense of ÔselfÕ with our Ôliving spaceÕ. It can do so by drawing our attention to the dynamic, reciprocal relation between inner and outer space, figure and (back)ground.

 

Our science can then literally inform our understanding of this relation through knowledge of the distribution of energy-matter that inseparably lines and so distinguishes these spaces without ever fully sealing them in or making them discrete - for that would be to destroy their vitality and render them static. It is, after all, the microscopical and telescopical investigative tools of science that have actually provided the evidence that space permeates everywhere, across all scales from sub-atomic to galactic. So, when we look more closely into the apparent solidity of a tree or any other natural form, we find that it is actually full of holes. Boundaries that at a distance appear as smooth limits or barriers between insides and outsides are actually complex relational surfaces adjoining domains with different arrays of voids and physical relief. Yet it is this very finding that analytical science, working literally in isolation has been unable to assimilate into its own theory and practice and reconcile with the impositional logic upon which this practice and mathematical underpinning has, historically, been founded.  Analytical science, working in isolation, has therefore been unable to derive meaning from its own findings and is riddled with contradictions, ambiguities and paradoxes, including those associated with the notion of Ôsurvival of the fittestÕ. And so it is that scientific materialism has placed itself at loggerheads with artistic and emotional awareness. 

 

By combining explicit scientific knowledge with the implicit spatial awareness of art, an inclusional logic emerges, which recognises that at no physical scale in a dynamic system can any thing be isolated as a discrete, independent object that can only be moved by a purely external force. Rather, everywhere is coupled with everywhere.

 

 

 

WATER – THE DYNAMIC CONTEXT OF LIFE, AND ITS INCLUSIONAL SIGNIFICANCE

 

When considering the potential for life to exist on other planets, the first thing that space scientists look for is water. This is no coincidence. Life forms, as we know them here on Earth, can never come into being or be active, though they can survive in a dormant state, without water. This is because water is the receptive medium into and through which life forms gather and distribute the energy that puts them in motion via processes of photosynthesis, chemosynthesis, digestion, respiration, transport and translocation. Water provides the link between generations, through and in which genetic information can flow and be exchanged and expressed in endlessly diverse forms. As such, water is and always has been the indeterminate dynamic context in which life forms thrive, diversify and respond to and influence their surroundings and neighbours—an ÔartistsÕ mediumÕ whose physical properties both constrain and contribute to lifeÕs heterogeneity and versatility.

 

A start can be made towards understanding the dynamic role of water in life by asking what possibilities for innovation and relationship exist in just a single droplet of water. Inclusionally, a droplet of water is a pool of energy-space, a dynamic context whose surface-tense boundary is the informational interface between its inside and outside.

 

The surface area of the droplet can be altered by assimilating or discharging energy sources across its boundary. Assimilative processes result in expansion. At low input rates, this expansion is isotropic (equal in all directions), thereby minimizing the resultant increase in free surface. At higher rates, Ôsymmetry-breakingÕ occurs, so that the droplet polarizes into a rivulet or subdivides into branches that, in being both separate and connected to one another to some extent, have only a degree of freedom. At even higher rates, the droplet may dissociate into smaller droplets and ultimately molecules. Viewed at a snapshot in time, these entities may appear to be discrete individual units, but this ignores the historical trajectories that link them to a common origin. Such trajectories are only apparent when viewed dynamically, whence their indeterminate capacity for expansion and change reveals discreteness to be an illusion of isolated observations.

 

Assimilative processes causing progressive subdivision of an initially coherent state can be termed Ôself-differentiationÕ. These processes generate the exponentially increasing amounts of exposed free surface (i.e. they result in boundary externalization) characteristic of individual and population growth. The emergence of such surface has, confusingly, been referred to as Ôself-organizationÕ, or Ôorder out of chaosÕ due to misinterpreting the location of system boundaries and so supposing that the surface arises from a random rather than coherent state. In fact, chaos, in the form of the deterministic proliferation of free surface in a forced system, is a state of increased order¾increased informational boundary, even though the connectivity of this boundary becomes progressively more tenuous (see later).

 

As the surface generated by self-differentiation takes shape, its options for change become constrained by what has already been produced. Moreover, since this surface cannot be fully sealed, it inevitably dissipates as well as gathers sources of free energy and so is only sustainable as long as supplies donÕt run out. If self-differentiation were to continue without the replenishment of external energy sources, it could therefore only end irreversibly in a boundless, fully incoherent state with all the systemÕs energy converted into entropy. This is the true meaning of the Ôunsustainable growthÕ that Ômore equals betterÕ capitalist economics and neo-Darwinian notions of evolutionary fitness promulgate. Processes, complementary to self-differentiation, which can be called Ôself-integrationÕ, may, however, prevent this fate.

 

Self-integration counteracts the dissipative effects of self-differentiation through the coalescence, sealing in and/or redistribution of boundaries (i.e. boundary internalization), so conserving energy within the system and enabling it to rejuvenate. In the case of water, vapour may condense into droplets, droplets may coalesce into pools and pools freeze into a myriad of ice forms, with a release of stored energy accompanying each reduction in free surface.

 

Self-integration can either lead to a reiteration of an initial coherent state, and concomitant loss of boundary generated by self-differentiation, or to synergistic amplification of this state and resultant spiralling out to larger scales and new horizons. Synergism can be due to boundary fusion and networking, whereby more power can be distributed through the system by replacing resistances in series with resistances in parallel. It can also result from Ôautocatalytic flowÕ, whereby passage of current through a resistive field lowers the resistance to subsequent passage.

 

                       Such are the creative possibilities for differentiation and integration of form even in a droplet of pure water. Now, allow materials to be incorporated or dissolved within the dropletÕs contents, changing their viscosity, matric, electrical and osmotic potential, or added to the surface of the droplet to form an insulating coating or envelope. Harnessed in this way, the dynamic potential for elaboration of diverse water forms becomes even greater. These formsÕ permeability, deformability, and continuity and consequent receptivity, responsiveness and conductivity can thence be varied according to whether their circumstances are appropriate for gathering in, exploring for, conserving or recycling energy sources. As they gather sufficient energy to begin to flow, they will, over time, both create and follow paths of least resistance in their surroundings, as in river systems. By taking substance out from their catchment, much as a hypha of a wood decay fungus might dissolve and absorb wood substance in the course of its growth, rivers effectively make their own inductive space. The same is true of all organic life forms and perhaps all universal features that emerge and dissipate through the reciprocal dynamic relationship between inner and outer inductive spaces or ÔholesÕ (as opposed to ÔwholesÕ). They are capable both of bringing about and responding to environmental change.

 

In those water forms that we have come to regard as organisms, materials added to and enveloping water constrain and enable the expression of diversity over scales ranging from the boundaries of molecular to social and ecosystem domains. These materials may be organic or inorganic. They may originate outside the organismÕs boundaries; they may be synthesized within, by gene action, or they may be produced by interaction at boundaries between internal and external reagents. They include the carbohydrates, fats, proteins, nucleic acids and other metabolites found in living cells. They include the oxidatively cross-linked hides, bark layers, cuticles and cell walls that protect and contain the living contents of innumerable forms of plant, animal and fungal life as they move or grow to form branching trajectories through space and time. They include the calcium-enriched shells and coatings of invertebrates and algae. They also include the earthy highways, byways, dams and buildings created by animals ranging from termites and earthworms to moles, beavers and human beings as they open up and seal off paths of least resistance in their surroundings to provide shelter and avenues of communication.

 

In modern times, the dominance of analytical approaches to the management of life processes has led to an increasing focus on internal genetic ÔinformationÕ as the principal means by which the form and functioning (ÔphenotypeÕ) of organisms is determined, subject only to the moderating influence of external environmental variables. Consequently, bioengineers and biotechnologists have sought means of altering this information to suit human requirements, raising many concerns about the ethics and effects of such Ôgenetic engineeringÕ on human health and the environment. Such concerns can only be heightened when, as a result of scientific exclusivity, genetic engineering is practised in a state of ignorance of dynamic context, whence there can be no wise appraisal of, or response to, its possible repercussions.

 

Viewed inclusionally, in the continuous dynamic context of harnessed water, phenotype is not, however, as genetic determinism would have it, a direct genetic function of environmental variables. Rather, genes are variables whose influence, along with other factors, on boundary properties affects the pattern in which water is arrayed, and re-arrayed, through space and time.

 

 

DIVIDED AND UNITED STATES OF BEING

 

The processes whereby what was once coherent becomes driven along separate paths by self-differentiation only to be brought back together and re-empowered by self-integration are evident at many scales and in many contexts of life. Correspondingly, my aim in this section of the course is to enhance awareness of how ÔecologyÕ – most fundamentally, the study of pattern, process and relationship in environmental context – is relevant from molecular to social and psychological scales of biological organization. By the same token, I shall attempt to draw attention to the ways in which the distinctive features of all these scales of biological organization are relevant to understanding environmental relationships.

 

The ability to visualize common processes of differentiation and integration across all organizational scales strongly depends, however, on the viewpoint of the observer, whether this is from inside or outside the observed, or both. Only when the viewpoint is truly inclusional, i.e. incorporating the dynamic interface between inner and outer space, can the observer truly be in empathic rapport with the observed. Only then is it possible to understand both the creative scope and limitations of ourselves and others, which arise through the relation between future and past that shapes the indeterminate (dynamically bounded) story (history, trajectory) of our lives. For human beings this can be a difficult viewpoint to stay with as we oscillate between subjectivity and objectivity, regarding life in terms of self or other with determinate (absolutely bounded) beginnings and endings.

 

                        Finding the appropriate viewpoint, the appropriate dynamic boundary between inside and outside, from which to perceive the fundamental indeterminacy that characterizes all living things requires varying degrees of imagination. Perhaps it is easiest with those life forms, including many fungi and plants, which grow rather than move bodily from place to place. Here, branching structures, with a large free surface across which to assimilate sources of energy, emerge from a germinating spore or seed. As with the river that both shapes and is shaped by the landscape it flows through, the boundaries of these structures map their own life history. Even though their movement may be barely discernible over short intervals of time, creating the illusion of stillness, the mindful observer can therefore readily trace the dynamic trajectories through which their present pattern arose out of their past. Commonly, shifts from self-differentiation to self-integration are clearly visible as branches emerge, diverge, die out and combine forces. 

 

                        For life forms that move bodily from place to place, including animals like ourselves, more imagination is required. It is all too easy to view the bodies of these forms, as they appear in snapshots of space and time, as discrete ÔindividualsÕ.  Yet, when the movements and interactions of these bodies are traced through their contextual landscape, or when the developmental processes that convert egg to adult are viewed, the riverine pathways of self-differentiation and integration soon become evident. Such pathways also occur in the tracery of all kinds of genealogies, in scientific cycles between Ôparadigm shiftsÕ, in psychological journeys from dependence to independence to interdependence as the self is born and rediscovered and in the philosophical journey from classical to postmodern and beyond.

 

 

 

 

 

Genetic Differentiation and Integration

 

Of all things, genes have become the focus for the enshrinement of discretist views of life. Herein are the supposed foundational, atomistic Ôbuilding blocksÕ that can be independently sorted out and assembled via the agency of natural selection to yield the phenomenal diversity of the living world. Herein are the ÔinformationalÕ instructions for life, couched in an alphabetic code of just four letters. As human beings have deciphered this code, working out how it can be translated into proteins that provide structure and catalyse chemical reactions in living cells, so the aspiration to develop a God-like power to possess and control nature, and even human nature, has been reinforced. Driven by this aspiration, it may seem that manipulating life will become no more difficult conceptually than manipulating the letters of the alphabet to form words and sentences of language, or bytes of digital code to create a computer program. BUT there is more to creating meaning than assembling code. There is no simple, single formula for deriving meaning from a set sequence of code, any more than there is such a formula for understanding the sequence of letters in this paragraph. Meaning depends on context, and if context changes, so does meaning.

 

So it is that genes, as with any other source of information, only make sense in context. This context is a changeable watery envelope of potentially complex interrelationships between genes, other genes and the other stuff of life (see Ôcell differentiation and integrationÕ, below). Outside this context, genes have no means of expression, no vitality of their own, as is evident from the inert state of viruses beyond the confines of their host cells.

 

Moreover, although the nucleic acid sequence of a geneÕs message must have definite start and end points if it is to be translated into functional protein rather than nonsense, the boundaries of genes are by no means fixed and isolated for all time. To begin with, there is evidence that particular sequences of DNA are capable of giving rise to more than one product due to overlapping translatable information or varied processing. Also, in eukaryotic organisms (plants, animals and fungi) genes often contain sequences called introns whose information has to be spliced out before translation into protein.

 

A crucial general issue concerns the fact that genes are not packaged as discrete entities but are linked together to varying degrees on sets of chromosomes. The number and size of chromosomes varies greatly between different organisms. The significance of this lies in the fact that linked genes cannot be re-assorted into varied combinations with one another (i.e. ÔrecombinedÕ) as readily as can genes on separate chromosomes. Although there are processes (see below) which can cause recombination even of  (and within) genes on the same chromosome, the closer together¾more closely linked¾genes are, the less likely they are to dissociate.

 

For several reasons, very close linkage can be both a cause and a consequence of interdependency between genes, resulting in the formation of  ÔsupergenesÕ or Ôco-adapted gene complexesÕ. Closely linked genes can either have a common origin, arising by self-differentiation from an ancestral gene that has duplicated and diversified, or a disparate origin, through self-integration of distinct DNA sequences. Interdependence due to common origin can involve the formation of complex proteins containing different sub-units whose interaction is vital to function, as in the red blood cell pigment, haemoglobin. Interdependence of genes having a disparate origin occurs when the genes code for proteins with complementary functions, a classical example being those involved in the assimilation of lactose in the bacterium Escherichia coli. Such interdependence may generally be augmented by common regulatory controls that ensure that the genes are co-ordinately expressed.

 

Recombination of genes, a process that manifestly involves an act of separation followed by a reunion, can be brought about via a variety of mechanisms. The most familiar of these mechanisms is sexual reproduction, which is a characteristic feature of eukaryotic organisms.

 

Sexual reproduction encompasses three important events: cell-fusion (ÔplasmogamyÕ), nuclear fusion (ÔkaryogamyÕ) and nuclear division (ÔmeiosisÕ). The first two of these events involve the bringing together of single (ÔhaploidÕ) sets of genes from each of two parents and their integration to produce a double (ÔdiploidÕ) set. The third event separates the double set back into single sets containing potentially new combinations of chromosomes (through Ôindependent segregationÕ) and new combinations of genes on the same chromosome (through Ôcrossing overÕ of DNA sequences as partnered, ÔhomologousÕ, chromosomes separate from one another). The progeny of sexual reproduction therefore tend to differ genetically both from one another and from their parents. Depending on whether the interval between plasmogamy and meiosis is short or long, parents and offspring may respectively contain either a single set of genes (as in many ÔlowerÕ fungi and plants) or a double set (as in many ÔhigherÕ fungi, plants and animals, including human beings).

 

Since it does not literally produce Ômore of the sameÕ, sexual reproduction is not truly re-production at all, but rather a dynamic interplay between ÔselfÕ and ÔotherÕ, differentiation and integration, within an interbreeding population (in effect a gene ÔpoolÕor ÔnetworkÕ) that provides scope for innovative exchanges and partnerships. It therefore makes much more general sense to talk of sexual ÔrecreationÕ or ÔrejuvenationÕ than reproduction. As will be discussed further below, the idea that sex is primarily Ôfor reproductionÕ has underlain many spurious evolutionary arguments and inappropriate comparisons with asexual mechanisms that really do result in the duplication and proliferation, i.e. ÔcloningÕ, of the same genetic material.

 

Although sex is usually regarded as characteristic of eukaryotic organisms, genetic recombination mechanisms also exist in bacteria. Some of these mechanisms are reminiscent of sex; others indicate how, with or without human intervention, genes can transcend the boundaries even of very different organisms by means of Ôhorizontal transferÕ.

 

The three known mechanisms of genetic recombination in bacteria are known as ÔconjugationÕ, ÔtransductionÕ and ÔtransformationÕ. Conjugation involves a sex-like process in which one cell, often described as ÔmaleÕ acts as donor of genetic information to a recipient ÔfemaleÕ cell, to which it attaches by means of structures called sex ÔpiliÕ. Transduction involves a virus as the agent of transfer. When the nucleic acid of such a virus enters a host cell, it can combine with some of the host nucleic acid, which can then be transferred when the resulting virus particles infect other host cells. Transformation involves the uptake of foreign DNA by a bacterial cell from its immediate environment. Along with transduction, it is an important process in Ôgenetic engineeringÕ whereby foreign DNA can be incorporated into cells and then ÔclonedÕ.

 

When transformation is used in genetic engineering, it often involves non-chromosomal pieces of DNA called ÔplasmidsÕ which can multiply independently. One plasmid, the Ti (Tumour-inducing) plasmid in the bacterium Agrobacterium tumefaciens demonstrates how, even without human intervention, genes can be transferred between fundamentally different kinds of organisms. A region, known as T-DNA, is transferred from the Ti plasmid into the cells of a host plant, where it integrates with the host DNA. The T-DNA contains information which induces the host cells to proliferate and produce unusual amino acids called ÔopinesÕ that serve as specific nutrients for the bacteria. Here, then, the bacterium could be said to genetically engineer its host in such a way as to provide itself with a unique natural habitat.

 

The behaviour of the T-DNA in the Ti plasmid is characteristic of a general class of entities called Ômobile genetic elementsÕ which can relocate between and so rearrange different parts of a DNA molecule. Two important kinds of mobile elements are known as Ôinsertion sequencesÕ, which do not contain genes, and ÔtransposonsÕ, which do. In bacteria, transposons often carry information that confers resistance to antibiotics.

 

Some kinds of DNA rearrangement are vital to successful patterns of cell differentiation in multicellular organisms (see below). One of the most complex examples, and one that beautifully illustrates the creativity of integration-differentiation interplay at the genetic level, occurs in the immune systems of vertebrates (animals with backbones). This system has much in common with sexually interbreeding populations in that it can be thought of as an open-ended genetic network in which separate, genetically differentiated entities are distributed within a dynamic context. In the immune system, however, the context is defined by the body boundary of a multicellular animal, and the entities are cells.

 

Just one gene encoding a polypeptide about 100 amino acid units long may have been the ancestor of the immune system. This gene has duplicated and diversified over evolutionary time to give rise to a superfamily of genes and protein products of such versatility that it is capable of detecting close to an infinite range of ÔforeignÕ molecules. Even artificial molecules will elicit an immune response, implying that the system can operate without any previous exposure to the molecules in question. This versatility stems from the freedom of members of immunity gene superfamilies to integrate with one another in a multiplicity of combinations. These different combinations produce a multiplicity of protein products with different boundary configurations.

 

 

Cell Differentiation and Integration

 

If genes, their component nucleotides and the amino acids and polypeptides they give rise to represent the modern notion of Ôbuilding blocks of lifeÕ, this discretist metaphor was given a firm foundation by the discovery that many life forms are composed of ÔcellsÕ. Indeed, anyone who has cursorily examined a section through plant or animal tissue under a microscope is inclined to have received a first impression of what looks remarkably like brickwork. Like many first impressions, however, this one is potentially very misleading. Cells are by no means isolated entities, but are dynamically bounded and capable of forming complex interrelationships.

 

Cells are the primary containers of the watery context of life in and through which genetic material is expressed, transferred and exchanged. This watery context or ÔcytoplasmÕ is the medium in which the energy conversions necessary to sustain a dynamic existence take place. It is retained and internally partitioned by boundaries of variable complexity.

 

In the ÔprokaryoticÕ cells of bacteria, there is a cell wall (except in organisms known as ÔmycoplasmasÕ) and plasma membrane consisting of a lipid bilayer around the cytoplasm (see below). There are no internal partitions, however, between the genetic material¾which consists of simple closed loops and lengths of DNA¾and cytoplasm, or between different cytoplasmic components. There is therefore relatively little scope for differentiation of specialized functions.

 

The ÔeukaryoticÕ cells of plants, animals and fungi are, by contrast, markedly internally partitioned into cytoplasmic regions and components that serve specialized functions. It is widely thought that this complex internal structure is the evolutionary result of ÔendosymbiosisÕ – the Ôself-integrationÕ of prokaryotic cells. The chromosomes are complex packages of DNA and protein (Ônucleoprotein) contained in a true nucleus that is surrounded by a double layer of membranes that communicates with cytoplasm and ultimately the outside through pores and internal labyrinths within the membrane network known as the Ôendoplasmic reticulumÕ. There are membrane-bound organelles such as ÔmitochondriaÕ, ÔchloroplastsÕ, ÔGolgi bodiesÕ, lysosomes and peroxisomes. There may be one or more solution-filled vacuoles. Within the Ôcell sapÕ (ÔcytosolÕ) may be ÔcytoskeletalÕ proteins, ÔmicrofilamentsÕ, ÔmicrotubulesÕ, and ÔmicrotrabeculaeÕ which through their association-dissociation interplay provide for support and movement within the cell. In other words, eukaryotic cells are parcels of parcels, externally bounded and internally partitioned by membranes.

 

It is vital to appreciate that cell membranes constrain but do not altogether prevent the flow of materials into, out of and within cells. As is vital to a dynamic system, they are partial boundaries, which serve to maintain a balance between the opening up and sealing off of communication channels as circumstances vary.

 

The principal ingredients of cell membranes are ÔphospholipidÕ molecules. These elongated molecules are attracted to water (ÔhydrophilicÕ) at one end and repelled by water (ÔhydrophobicÕ) elsewhere. When placed in contact with water, they therefore tend to line up, with the hydrophilic ÔheadsÕ facing outwards and hydrophilic tails facing inwards, so forming a self-sealing double layer (ÔbilayerÕ). This layer is Ôsemi-permeableÕ: it allows passage of water, gases such as oxygen and carbon dioxide, and small, relatively hydrophobic molecules such as ethanol, but is impermeable to most water-soluble molecules.

 

Also present in cell membranes are a variety of proteins that provide channels through which particular kinds of substances, including ions, can be transported. Where this transport is ÔactiveÕ, i.e. consumes chemical energy, it serves as a pump which maintains an electrical potential difference between the inside and outside of cells. This potential difference is essential to staying alive: it has been estimated that as much as one third of an animalÕs energy expenditure serves this need. It is also responsible for the ability of cells to maintain a higher solute concentration in their interior than outside. This higher solute concentration results in a more negative water potential (the free energy of water in a system relative to that of a reference pool of pure water, with a potential of zero) inside than outside the cell, causing water to tend to enter by means of ÔosmosisÕ.

 

Transfer between the inside and outside of cells can also be achieved through re-arrangements of the plasma membrane that result in enclosure (enabling importation) and exclosure (enabling exportation). These re-arrangements depend on the ability of cell membranes to break and reseal (differentiate and integrate).

 

Cell walls provide an additional retaining boundary, outside the plasma membrane, by which the shape and behaviour of cells can be moderated. The relative deformability or rigidity of this boundary affects the extent to which cell boundaries will give way or Ôballoon outÕ under pressure resulting from uptake or throughput of resources. This property can be altered through the production and association or dissociation of microfibrillar components such as chitin and cellulose, augmented by the association and dissociation of cytoskeletal elements in the underlying cytoplasm. In fungi, the extensibility of the boundary can be confined to a particular, newly generated, region of the cell wall, resulting in a highly polarized pattern of cell growth to form an apically extending tube known as a hypha.

 

In addition to their strength properties, the permeability of cell walls, i.e. their resistance to passage of materials, is crucial, especially in circumstances where there are possibilities of losing resources or taking in toxins. Here, there are advantages in being able to seal the walls with hydrophobic materials. These materials include suberin (the main component of cork) and lignin (a component of woody tissues) in plants, and various polyphenolic and proteins (ÔhydrophobinsÕ) in fungi.

 

Some organisms exist throughout most of their lives as single cells – as the earliest life forms are generally assumed to have done. These organisms should not, however, necessarily be regarded as ÔsimpleÕ. The degree of sophistication that has been achieved, through internal partitioning, by some eukaryotic single-celled organisms, known as ÔprotistsÕ is remarkable. The ciliate, Paramecium, for example, has within its cell boundary structures analogous in function to the mouth, throat, anus and kidneys of a mammal. As a general theme for all life, such a high degree of individual sophistication is generally associated with a relatively independent life style but can impede the ability to enhance scope through co-operative interactions with others.

 

No matter how successful unicellular organisms might be, like all other life forms they have to be able to proliferate – make more of themselves – if they are not to die out. There are a number of ways in which such proliferation can be achieved. Firstly, there can be a simple expansion of the cell boundary, either isotropically (in all directions) or in a particular direction. For both physical and functional reasons, such expansion cannot be sustained indefinitely, however, and sooner or later results in subdivision either into separate cells or into branches and consequent self-differentiation.

 

Subdivision into branches is a characteristic feature of fungi and some other organisms with filamentous cell organization. Branches have only partial freedom in that they remain interconnected to one another through their points of origin. In fungi, they give rise to an indefinitely expanding, collectively organized structure known as a mycelium.

 

Fungal mycelia are physically organized as indeterminate, versatile systems of interconnected hyphae. These systems can span heterogeneous environments of potentially great complexity, ranging over space and time scales from micrometres to kilometres and seconds to millennia and in which energy is often in very variable supply.

 

The versatility of fungal mycelia becomes evident as soon as a spore takes up water and nutrients, so expanding isotropically at first and then breaking symmetry with the emergence of one or more indeterminately expanding, protoplasm-filled germ tubes. Alternatively, a determinate unicellular pattern may be maintained for greater or lesser periods, as in yeasts.

 

Once polarity has been established, the hyphal tubes may become internally partitioned by valve-like ingrowths known as septa, and branch in either a tributary-like or a distributary-like pattern. The branches either diverge or converge and fuse (anastomose). Whereas some parts of the system are in close contact with the nutrient source, others become sealed off or emerge beyond the immediate sites of assimilation. The branches may remain diffuse or they may aggregate to form protective, reproductive or migratory structures. Whilst some parts of the system continue to expand, others degenerate.

 

The biological utility of such a changeable dynamic structure becomes clear whenever fungi are observed growing in heterogeneous environments. For example, in moist woodland soil, networks of mycelial cables interconnect the roots of neighbouring plants as well as decaying wood or leaves. The processes leading to the formation of such networks can be revealed experimentally by growing the relevant fungi in sets of chambers. These chambers are isolated from one another with respect to diffusion through the growth medium, but interconnected by passageways that allow particular portions of the mycelium to grow between and across separate domains. Here it is possible to see how, purely by changing its boundary properties in response to local circumstances and without any central administration, a mycelium can generate a persistent network which is reinforced along avenues of successful exploration.

 

The special properties of mycelial networks arise from the fact that they connect boundary resistances in parallel rather than in series, as in branched systems, so increasing conductivity and reducing the tendency to self-differentiate/break symmetry. Networking enables stable ÒestablishmentsÓ to form at the same time as allowing multiple redistributional options through a potentially huge number of sub-circuits. It also makes possible large amplifications of organizational scale as a result of the increase in waterpower that can be delivered to a local site of expansion or emergence¾the system may literally be capable of ÒmushroomingÓ. For the latter to occur, however, energy input has to exceed throughput capacity. This may be difficult for a system that automatically Òself-limitsÓ by minimizing the proliferation of assimilative boundary, but can be achieved through degenerative mechanisms. These mechanisms allow dis-integration of part of the network and consequent redistribution to outgrowth sites, as in fairy rings.

 

Where subdivision is into separate cells, and these cells are free either to drift or propel themselves apart, then they will retain a relatively independent existence from one another. There are, however, two main reasons why cells may not be free to separate from one another. They may be restricted by the boundaries or surfaces within or upon which they grow. They may be attracted together and/or adhere when they come into contact. The process of subdivision therefore automatically implies the possibility of cells assembling together, ultimately to form complex, interdependent organizations.

 

 

Tissue Differentiation and Integration

 

Even in assemblies of initially identical cells, processes occur that inevitably lead to the differentiation of a heterogeneous structure. For example, those cells on the margin of the aggregation will form a boundary layer that insulates all other cells in the interior from the external environment.

 

These processes result in the formation of colonies of cells that range from simple associations of similar, relatively autonomous ÔunitsÕ, such as a yeast colony, to coherent organizations with a clear Ôdivision of labourÕ between interdependent cells with distinctive functional roles, such as a sponge. In the latter case there may be further scope for association of simple colonies into compound or aggregate colonies – literally colonies of colonies of cells! Yet, complex as these structures may be, they are limited in the scale and sophistication of their operation by the absence of a further self-integrative process. This process takes them beyond being an assortment of cellular entities to being entities in themselves, with power fully transferred from the local member to the global membership, so bringing scope for much more effective co-ordination.

 

When cells of different kinds within a life form develop into distinctive arrays or ÔtissuesÕ, then that life form becomes a ÔmulticellularÕ organism. Some multicellular organisms consist of inner and outer layers of tissue, that become elaborated outwardly (i.e. self-differentiate) into a profusion of forms with maximal free surface, from the polyps and medusae of jellyfish and their allies, to the root and shoot systems of plants. Others become elaborated internally, within a self-contained boundary with minimal free surface (i.e. via self-integration), into a variety of distinct but interconnected subdomains or ÔorgansÕ, as in many complex animals.

 

Crucial to the functional integrity of multicellular organisms is some means of union that transcends the boundaries between neighbouring cells, tissues and organs. Correspondingly, channels are opened by means of passages (e.g. Ôgap junctionsÕ and ÔplasmodesmataÕ) between adjacent cells, and by electrical conduits (nerves) and vascular pipelines between tissues and organs. Without such union, neighbours can only compete. Such competition results in net transfer of resources to those capable of exerting the most active demand (i.e. to those that already ÔhathÕ, as in metabolically active metastatic cancer cells which do not open gap junctions with their neighbours) and so to the inevitable breakdown of mutual relationships. On the other hand, the presence of passages can render the system vulnerable to the spread of damaging agencies through its interior, and to extreme demand by competitive components or parasites (Ôpower drainsÕ) that take but donÕt give through their semi-permeable boundaries. Damage-limitation mechanisms that seal off channels in the face of violation are therefore as essential to surviving as are boundary-opening mechanisms to thriving.

 

The processes of differentiation and integration that give rise to the interconnected cells, tissues and organs of a multicellular organism commonly begin with the fertilization of an egg cell to produce a diploid ÔzygoteÕ. The zygote then proliferates to form an embryo, which undergoes a variety of developmental processes culminating in the emergence of an adult.

 

At first, most embryos consist of little more than a group of more or less similar cells. For tissues to form, some kind of re-organisation has to occur, so that the cells become heterogeneously distributed into distinctive layers and subdomains where they follow different developmental pathways. The way that this re-organization occurs contrasts markedly between most higher plants and animals, reflecting the difference between those forms that grow from place to place and those that move bodily from place to place.

 

Basically, in higher plants, the embryo becomes polarized into an elongated structure to the tips of which new cells are added either by proliferation from a single apical cell, or a group of cells known as a ÔmeristemÕ. Further tips may then arise by means of branching. All the cells and tissues of what is known as Ôthe primary plant bodyÕ arise from these tips. In woody plants, secondary lateral meristems known as ÔcambiaÕ then thicken the trunks and branches of roots and stems by giving rise to additional layers of wood (to the inside) and bark (to the outside). The localization of cell division within apical meristems also occurs in some colonial invertebrates (see below) and is a basic feature of indeterminate multicellular structures, analogous to the extending tips and branches of hyphae and other cellular filaments.

 

By contrast, in the majority of animal embryos the production of new cells is not localized, but occurs within all the developing organs and tissues. Here, from the viewpoint of an external observer, development appears to be highly prescriptive, occurring in a set sequence and directed towards a specific, determinate, functional end point, the sexually mature adult. Viewed from within the developing body boundary, however, indeterminate processes analogous to those seen in plants and fungi are evident, and even minute variations in the interactive dynamics of these processes have the potential to result in radically different overall outcomes. In other words, it is possible to generate radically different kinds of organisms by varying their developmental context and without radically changing their genetic information content. Hence, for example, chimpanzees and human beings can have almost identical DNA but very different phenotypes.

 

Following fertilization of an animal egg, rounds of mitosis accompanied or followed by cytoplasmic cleavage cause cell and/or nuclear numbers to increase in a series of doublings. As division continues, an internal space or ÔblastocoelÕ develops, preparatory to a remarkable self-integrative phase of boundary-infolding, known as ÔgastrulationÕ, which culminates in the formation of inner, outer and intermediate tissue layers: endoderm, ectoderm and mesoderm. Cells within these layers then undergo self-differentiation, and ultimately become specialized for distinctive roles in skin, nerve, gut, muscle, connective tissue, bone, blood vessels, liver, kidneys etc.

 

The processes that follow gastrulation are generally considered to be administered by a genetic programme that controls the activation and inactivation of distinctive sets of genes. For this programme to give rise to an appropriate sequence of changes, it is important for the developing embryo to be buffered, as far as possible, from the influence of a variable external environment. The developmental context is therefore internally self-regulated, within the confines of an enclosing boundary that minimizes exposure to the outside, at least until the moment of birth.

 

Progression through the developmental programme both equips the emerging adult for engagement with its real-world contextual boundaries and narrows down its options for change. The condition of universal possibilities (ÔtotipotencyÕ) from which development begins in the zygote leads, through self-differentiation, into increasingly narrow, bifurcating paths of specialism, entry to each of which is conditional upon those paths that have already been followed. In the midst of actual gain, there is therefore also inevitable loss of potential, a loss of what might have been.

 

This process of simultaneous proliferation and narrowing down of options from the initial coherent state of the zygote is known as determination. It has been likened to a ball rolling through a bagatelle-like landscape dissected by a bifurcating series of valleys. As the ball rolls and follows one fork or another it makes irrevocable ÔdecisionsÕ that fix its fate.

 

Here there is a fundamental difference between determinately developing animals and indeterminately developing plants. With few exceptions, determined animal cells cannot change their developmental course. So long as they remain alive, even fully differentiated plant cells (those that are in their final functional form) can, however, regenerate into whole organisms. In artificial culture, such regeneration involves Ôde-differentiationÕ into an uncoordinated mass, known as ÔcallusÕ. In nature it commonly involves passage through an intermediary conservational or ÔstorageÕ phase (see below).

 

In general, determination and subsequent differentiation are achieved ÔepigeneticallyÕ, i.e. via changes in the expression rather than the content of genetic material. Such changes imply that sets of genes can somehow be activated or inactivated according to requirements.

 

There is much evidence that these changes are effected by means of ÔcontrolÕ genes, i.e. genes which govern the expression of other genes. These genes include the ÔhomeoticÕ genes, first discovered in the fruit fly where mutations in them can have such curious effects as inducing a leg to develop in place of an antenna. These genes all contain a ÔconsensusÕ sequence, specifying similar sets of 60 amino acids characteristic of regualtory proteins known as Ôtranscription factorsÕ. These transcription factors are capable of binding directly to DNA and thereby promoting or inhibiting gene expression.

 

In the fruit fly, the homeotic genes function at the end of a hierarchical cascade of gene expression that specifies distinctive developmental domains. The first genes to function, are maternal genes which are expressed in cells surrounding the egg and give rise to head-to-bottom and front-to-back contextual gradients of protein morphogens known as dorsal and bicoid. These gradients regulate the expression of Ôgap genesÕ, which regulate two tiers of Ôpair-ruleÕ genes, which regulate the Ôsegment polarityÕ genes, which define the front and back boundaries of individual body segments. Notice here that this cascade begins in mother cells, rather than in the egg itself: there is therefore no abrupt demarcation of control between the parent and the offspring.

 

As has already been implied, the separation of distinctive life-maintaining functions into local domains or tissues with specialized attributes allows each to function with maximum efficiency and minimum interference. This very separation and specialism of parts leads, however, to their increasing interdependence – an inability to function in isolation. It is therefore essential to have in place some kind of re-integrational infrastructure or set of conduits that allows transmission of resources and information between specialisms. These conduits either conduct liquid, as in the ÔvascularÕ systems of plants and animals, or electricity, as in the nervous systems of animals. The pattern of development of nervous and vascular infrastructures is fundamentally indeterminate, resembling that of a fungal mycelium as they connect up their sites of supply and discharge.

 

Nervous systems contain two types of cells: elongated ÔneuronsÕ that transmit electrical impulses and variously shaped ÔgliaÕ that provide packaging around the neurons. The neurons are often bundled together into cable-like structures known as nerves.

 

Neurons commonly have four distinct regions: a cell body, a tributary-like gathering-system of ÔdendritesÕ, a distributive channel or ÔaxonÕ and a junction between the cell body and axon, the Ôaxon hillockÕ. During development, axons can both branch in a delta-like pattern and elongate at their tips. In so doing, they maintain and proliferate connections at specialized junctions known as ÔsynapsesÕ, both with tissues (especially muscles and glands) and with other neurons.

 

The extent and rate of spread of electrical charge along a neuron depends on two properties, the permeability of the cell boundary to ions and the conductivity of the cell interior. Wide, well-insulated neurons therefore conduct nerve impulses further, faster and more efficiently than narrow, uninsulated ones. Correspondingly, the distributive components of neurons, i.e. axons are both wide and well-insulated. In vertebrates, the insulation is provided by specialized glial cells, known as ÔSchwann cellsÕ that wrap their plasma membranes around individual axons to form a many-layered coating known as ÔmyelinÕ.

 

The transmission of a nerve impulse across a synapse depends on the action of chemicals called ÔneurotransmittersÕ. Acetylcholine is an example of a neurotransmitter that enhances transmission, whilst gamma amino butyric acid is an example of a neurotransmitter that impedes transmission. Depending on the identity of the neurotransmitter, sysnapses may be ÔexcitatoryÕ or ÔinhibitoryÕ, respectively propagating or resisting a nerve impulse. A single neuron may receive impulses from up to thousands of other neurons that synapse with it. Whether such a neuron fires an impulse depends on the overall balance between inhibitory and excitatory signals that it receives.

 

                        Higher plants typically consist of two complementary, potentially competitive but ultimately interdependent systems – roots and shoots, interconnected by two sets of conduits or vascular tissues, known as ÔxylemÕ and ÔphloemÕ. Water and mineral nutrients absorbed by roots are distributed through xylem, whereas the photosynthesized products of shoots are distributed through phloem. In herbaceous plants and in young shoots and leaves, xylem and phloem are associated with one another in cable-like Ôvascular bundlesÕ or veins. In perennial plants, the xylem is normally contained in a central cylinder of wood, whereas phloem is a component of bark. The external boundary of bark consists of an insulating layer of dead cells impregnated with a hydrophobic corky substance known as ÔsuberinÕ.

 

                        The vascular systems of animals likewise consist of an interdependent combination of gathering and distributive conduits, namely veins, lymph ducts and arteries. Veins and arteries are surrounded by relatively impermeable layers of muscle and connective tissue which are particularly thick in arteries. At their gathering and distributing end-points, however, these major blood vessels characteristically branch into progressively finer sets of thin-walled, permeable channels known as ÔcapillariesÕ. The capillary systems enable oxygen, carbon dioxide, nutrients and waste products to be transferred between tissues and blood stream, and are organized into patterns that resemble the intricate venation of a leaf or branching mycelial network of a fungus.

 

                        Tissues or groups of cells that are not tapped into a capillary supply cannot proliferate. The expansion of capillary systems so that they maintain and/or establish connections with the tissues that they supply is therefore important in embryonic development. At later stages of life, however, the proliferation of capillary systems can hasten the onset of processes resulting in disease and death, as in the ÔvascularizationÕ of solid tumours.

 

                        As well as being routes for passage of resources and waste products, vascular systems provide channels for transmission of chemical signals known as ÔhormonesÕ. In animals, hormones are usually produced in specialized tissues or organs and can either be hydrophobic or hydrophilic. The binding of hydrophilic hormones to receptor molecules on cell surfaces usually leads to a rapid change in cell activity, as with the Ôfight or flightÕ hormone, adrenaline. Hydrophobic hormones, by contrast, usually lead to changes in gene expression, as with the steroid ÔsexÕ hormones. Plant hormones, of which six basic kinds are known (auxin, gibberellic acid, ethylene, cytokinins, abscisic acid and brassicolides) all affect some aspect of growth or development through changes in cell boundary properties and internal metabolism. Their complex, counteractive interplay results in varied patterns of shoot and root emergence, extension, branching and loss.

 

 

Social Differentiation and Integration

 

Many of the organizational principles upon which social structures are based reflect themes, and elaborate upon mechanisms that occur in cells and tissues. As organisms proliferate in unrestrictive, energy-rich regimes, they dissociate into highly subdivided, dissipative, competitive assemblages that are unsustainable in the absence of continual replenishment of resources. As conditions become locally or generally restrictive, however, due to external impediments or as a consequence of resource-depletion by the organisms themselves, so less dissipative, more coherent organizations become favoured. From this stage onwards, the outcome of counteraction between associative and dissociative trends both determines and is determined by patterns of resource supply from the environment and leads to increasingly complex and heterogeneous organizational patterns.

 

                        Multicellular organisms associate into colonies in much the same way that has already been described for single-celled organisms. Simple gatherings of similar organisms, such as herds, flocks and shoals of animals or clumps, tufts and stands of plants occur whenever the organisms are contained by an external boundary or are attracted to one another in some way or fail to detach fully when they proliferate. The boundaries of these simple gatherings are sufficiently fluid to enable them to generate an immense variety of patterns by both creating and following paths of least resistance. Hence a herd of wildebeest can migrate from dry lands to wet lands along a delta-like array of well-worn paths, colonies of mosses can follow and accentuate cracks in walls and pavements, and flocks of birds or shoals of fish shimmer and fenestrate through air and sea currents.

 

                        Simple assemblages give rise to social structures through differentiation and division of labour amongst specialisms integrated into coherent organizations by means of various kinds of communicative infrastructures. Two examples of such social structures, the one essentially indeterminate and plant-like, the other more self-contained, are found in certain Cnidaria (jellyfish etc), the hydroids and the siphonophores.

 

                        Hydroids consist of individual ÔpolypsÕ whose gut cavities are all connected to one another, usually by a tubular system of erect branching ÔstemsÕ and creeping stolons or ÔhydrorhizaÕ. The hydrorhiza extend outwards and give rise to further erect stems, so increasing the size of the colony. Whilst the polyps in any one colony are all genetically identical to one another, they occur in a variety of forms. Many are feeding polyps, equipped with tentacles that trap prey. Others are reproductive, giving rise to free-swimming medusae that drift away and engage in sexual union. Yet others may be equipped with stinging cells that protect the colony or paralyse prey.

 

                        Siphonophores are complex assemblies not only of different kinds of polyp, but also different kinds of medusae, all interconnected to one another. As well as having sexual functions, the medusae may be modified into swimming bells, protective flaps or a gas-filled float. A well-known example is the Portuguese-man-of-war, Physalia.

 

                        Another kind of social organization, in which organisms are behaviourally rather than materially interconnected, is found in some kinds of insects. Here, entities serving reproductive, feeding and protective roles are allocated to distinctive ÔcastesÕ.

 

These entities are closely related to one another, a feature that is also broadly true of those social groupings formed by vertebrates in which there is clearest evidence of functional specialization and division of labour: prides of lions, troops of baboons, packs of hyaenas, societies of meerkats etc. The same would originally have held for human societies formed from extended family groupings. With increased mobility and development of communications infrastructures, however, human societies have become ever more culturally and genetically diverse, a fact which brings some of the most exciting and liberating opportunities but is underlain by some of the deepest tensions of our time.

 

                        Of fundamental concern here is the exclusivity or inclusivity of the dynamic boundary that exists between ÔselfÕ and ÔotherÕ. In an arena where there is a fixed supply of resources and entities are unable to integrate, then those entities will inevitably compete – the more so, the more similar are their requirements, as with members of the same species. Such competition brings the risk to ÔselfÕ of becoming excluded and replaced by ÔotherÕ. This risk can be overcome by self-integrating with ÔotherÕ and so pooling resources. Where ÔotherÕ is genetically the same as ÔselfÕ, such integration does not compromize genetic survival. Where ÔotherÕ is genetically different, however, there is a real danger of a mismatch, leading to internal conflict between opposing self-interests, culminating in takeover or degeneration. Hence it is understandable that throughout the living realm can be found Ôrejection mechanismsÕ that maintain separation between genetically different members of the same species. These mechanisms underlie all kinds of seemingly ÔantagonisticÕ or ÔselfishÕ responses that accompany and reinforce self-differentiation between neighbours, limiting their scope for co-operation but ensuring maintenance of their individual identity and consequent diversity, so long as circumstances do not change. There comes a time in the lives of many organisms, however, when their contextual boundaries do change and these mechanisms must be overridden if a renewable future is to be ensured. This is when sexual union occurs, heralded by all manner of chemical exchanges and behavioural courtship displays that induce return to a primal state in which ÔotherÕ is accepted as ÔselfÕ.

 

                        Both competitive and co-operative interactions also occur between members of different species, where they play fundamentally important roles in the structure and development of ecosystems and symbioses. Competition between species can be very intense wherever there is at least some overlap in requirements for resources. It is generally thought, however, that in stable ecosystems such competition is minimized by partitioning or differentiation into more narrowly defined or specialized activities. For example, different species of wading birds co-existing on a mud flat have different shapes and sizes of bills appropriate for different sources of food.

 

                        Differences between species can allow them not only to co-exist without mutual interference, but also to form associative networks and partnerships that divide labour between specialisms. In some cases these associations arise through mutual effects on external environmental conditions, as in many examples of ecological successions where there is a build up of biological and environmental complexity over time. Other examples arise as a result of the direct use of one life form by another as a resource. Such associations range from temporary use of ÔotherÕ as a food item or means of transport to persistent, intimate partnerships (mutualistic symbioses). Even temporary associations can result in intricate sets of interrelationships such as food chains and food webs. Mutualistic symbioses have become a dominant force in the generation and maintenance of biological diversity.

 

                        In terrestrial environments, the majority of higher plants would be unable to thrive without forming mutualistic partnerships, known as ÔmycorrhizasÕ, with fungi that invade and act as absorptive accessories to their roots. The fungi extend out, as mycelium, into soil and thereby provide their plant partner with improved access to mineral nutrients and water in exchange for organic compounds produced by photosynthesis. The mycelium can also interconnect different plants – even of different species. By providing communication channels between the plants, mycorrhizal mycelia are thought to enable adult plants to nurture seedlings, to reduce competition and to enhance efficient usage and distribution of soil nutrients.

 

                        Apart from forming mycorrhizas, the roots of some plants form associations with bacteria, blue-green bacteria and actinomycetes, that are capable of converting atmospheric nitrogen into ammonia by a process known as Ônitrogen fixationÕ. These associations have considerable importance in the generation and maintenance of soil fertility.

 

                        Where higher plants are unable to establish in terrestrial habitats, then surfaces that would otherwise be bare become covered by another kind of symbiotic union, lichens. Lichens consist of a photosynthetic filling of green algal or cyanobacterial cells sandwiched between layers of fungal mycelium. Being tolerant of extremes of temperature and water availability, they grow very slowly, contributing over many years to processes of rock erosion and soil formation, and are a source of a unique variety of chemical compounds. They are, however, extremely sensitive to atmospheric pollution and so easily lost as a result of human intervention in natural ecosystems.

 

                        Not only terrestrial plants, but also many animals depend on mutualistic symbioses. The guts of many animals contain assemblages of microorganisms and protists that both benefit from and can aid digestive processes. Some of these associations are indeed essential to digestion, and the activities of different members of assemblages complement one another. Such complementation occurs between the fungi and bacteria that inhabit the rumen of ruminant mammals, and the bacteria and protists that inhabit the guts of lower termites. Some animals even cultivate partners that can aid digestion: amongst insects these include the wood wasps, ambrosia beetles, higher termites and attine ants.

 

                        Mutualistic symbioses are also of great importance in marine communities. The reef-building corals, for example, depend on the presence of photosynthetic protists, ÔzooxanthellaeÕ within their tissues and so cannot exist below depths where an adequate supply of light can penetrate. The corals benefit through the provision of photosynthetic products and enhanced production of calcium carbonate (limestone) for skeletal support. The zooxanthellae obtain nitrogen and phosphorus from the food caught by the polyps as well as gaining shelter within the animal tissues.

 

                        Mutualistic symbioses have the potential to become so intimate, with the partners so interdependent that they become literally inseparable, so that what originated as partnership between selves becomes, in effect, one and the same self. Such inseparability may be accompanied by a transfer of genetic control between the partners. As has already been mentioned, it is now widely thought that eukaryotic cells arose in this way, and that DNA-containing organelles like mitochondria and chloroplasts are derived from bacterial and blue-green bacterial ancestors. Over time, these organelles may have lost their autonomy through transfer of genes to the nucleus.

 

                        It should always be remembered, however, that partnerships brought about by self-integration of boundaries are potentially unstable if those boundaries are re-asserted, and that many mutualistic symbioses may have passed through a parasitic phase in which there was benefit only to one partner. A number of degenerative diseases and male-sterility phenomena, for example, are believed to result from mitochondrial inheritance.

 

                        Parasitism is itself, of course, an extremely widespread phenomenon amongst life forms. It is usually viewed negatively from a human perspective, as a cause of disease and death. In the overall scheme of planetary life, however, it may play an invaluable role in keeping populations in check and in providing a means for redistribution of resources. Indeed, what may be regarded as Ôself-parasitismÕ plays an essential role in the life histories of many life forms.

 

 

 

DEGENERACY AND INNOVATION – HOW DEATH BECOMES LIFE

 

In our own lives, many of us recognize that to be able to move on and find new horizons on the shifting stage of our existence, we have to relinquish something of ourselves that we may cherish yet which binds us to our past. We have to find some way of putting that past outside of ourselves, drawing whatever resources from it that we can, and leaving it behind, however painful that may be. Otherwise we become stuck, entangled in a world that is all memory and no fresh experience. We need to be able to forget.

 

So it is with all life. New boundaries emerge out of the mortal remains of old boundaries. The destruction of the old and its incorporation into the new are as vital an aspect of self-integrative processes as are the fusion and sealing of boundaries in the establishment of partnerships and networks. Indeed, without such redistribution, partnerships and networks can become just that – retentive, inflexible establishments that resist change.

 

 

Metamorphosis

 

Amongst animals, perhaps the most striking manifestation of emergence into a new pattern of life out of the remains of an old pattern involves the conversion from a larva to an adult, e.g. the transformation of a tadpole into a frog or a caterpillar into a butterfly. Such transformations broadly correspond with the conversion of assimilative or growing phases into distributive sexual phases, and involve very obvious stages of boundary redistribution.

 

                        In the case of a tadpole, the tail and gills which are appropriate for a life in water degenerate and become replaced by the legs and lungs that enable frogs to make their way on land. The degeneration and resorption of the tail is a redistributional process that involves the developmentally programmed death (known as ÔapoptosisÕ) of cells.

 

                        Degenerative processes are even more apparent during insect metamorphosis, where virtually the entire muscle system of a larva is absent from adults. This transition also involves conversion from soft-bodied forms with relatively deformable external boundaries to hard-bodied forms with a rigidified, armour-like ÔexoskeletonÕ. The soft-bodied forms are able to enlarge partly because of the expandability of their skin or ÔcuticleÕ and partly because once the cuticle has reached the limits to which it can be stretched, it is separated off and discarded. Often there are several such moults (ÔecdysesÕ) between separate larval stages (ÔinstarsÕ). The timing of these moults is associated with a counteractive interplay between hormones related to the hardening off of old and generation of new cuticle in tune with environmental conditions. When the final instar reaches its size limit, the cuticle is hardened by a tanning process involving the action of phenol-oxidizing enzymes, to form a pupa. This pupa is a self-integrational phase which arrests further expansion and seals in the resources accumulated by the feeding larva. Emergence from the pupa then entails the degeneration of larval tissues, abandonment of the pupa casing and activation of embryonic cells that have lain dormant during proliferation of larval tissues from the egg.

 

                        The degenerative processes associated with metamorphosis have always been recognized as a vital and therefore prescriptively programmed part of the life of animals with distinct life cycle stages. In other walks of animal life, death has usually been regarded as an inescapable and fundamentally undesirable consequence of ÔimperfectionÕ, the result of infection, damage, accidents, predation, toxins etc. More recently, however, it has been recognized that programmed cell death limits the proliferation of cells that would otherwise develop unsustainably, at the expense of the self-integrity of the organisms, as in cancers.

 

                        Amongst plants, processes analogous to animal metamorphosis occur during what are called Ôvegetative reproductionÕ and the Ôalternation of generationsÕ. Vegetative reproduction entails mitotic division to produce groups of cells or ÔplantletsÕ that can mature into adult plants. These plantlets can arise by means of resource redistribution from leaves or explorative structures known as ÔstolonsÕ, or from various kinds of Ôstorage organsÕ (rhizomes, tubers, corms, bulbs and bulbils). The alternation of generations involves the production of distinct plant bodies, consisting of haploid ÔgametophytesÕ and diploid ÔsporophytesÕ as a consequence of meiotic division. In mosses and liverworts (ÔbryophytesÕ) the gametophyte is the dominant, independently growing life cycle stage. It eventually produces male and female sex organs (ÔantheridiaÕ and ÔarchegoniaÕ respectively). Fertilization of the egg cell in the archegonium results in the outgrowth of the sporophyte, virtually as a parasite upon the gametophyte. Meiosis in the sporophyte produces spores which germinate into gametophytes. In the ferns and their allies (ÔpteridophytesÕ) the gametophytes are more transient, consisting either of small, independently growing ÔprothalliÕ, or spores. The sporophyte which emerges from the fertilized archegonia soon becomes an independent plant in its own right. The trend for reduction of the gametophyte generation is continued in the higher plants.

 

 

Autodegeneracy

 

Whereas cell death in animals can be recognized as a means whereby a new form of life can be expressed, as in metamorphosis, or growth can be held in check, the unconfined spread of degeneration through the determinate animal body can mean only one thing: mortality. Such autodegeneracy is therefore regarded as a disorder.

 

                        In organisms with indeterminate body boundaries, however, autodegeneracy plays a valuable role as a means of redistribution of resources from redundant to actively expanding domains. The importance of this role in fungal mycelia has already been mentioned, where it enables efficient allocation of resources between nutrient-rich and nutrient-poor sites and the freeing up of Ôfairy ringsÕ so that they can expand indefinitely rather than become ÔgridlockedÕ.

 

                        In many plants, the action of phenol-oxidizing enzymes can insulate cell boundaries with lignin. Combined with the degeneration of protoplasm, this leads to the production of that familiar, predominantly dead, but functionally vital redistributive tissue known as wood. Plant stems and roots therefore degenerate from inside-out, with the formation of wood keeping pace with outward expansion due to secondary thickening (see above). In mature trees, the redistributive process is continued through the agency of fungi which harness the destructive power of oxygen to decay the core wood, causing ÔheartrotÕ. Heartrot results in the hollowing of tree trunks, so providing a huge variety of habitats for animals as well as allowing the tree to recycle itself by proliferating roots within its own internal ÔcompostÕ of decomposing remains.

 

                        Degeneration also often occurs amongst the central branches of plants such as heather, so that an annulus analogous to that of fairy rings is formed. Similarly, structures such as stolons often degenerate as the plantlets that they once interconnected become established in their own right.

 

                        Plants also degenerate on the outside, where leaves and branches are drained of their resources, senesce, die and ultimately become detached as they get left behind the expanding boundary of the system. The final act before detachment usually involves the formation of a corky sealing-off or Ôabscission zoneÕ.

 

                        The enormous scale on which these losses are sustained is easily overlooked by eyes focused solely in the short term. They may be envisaged, however, when viewing a mature tree, by imagining the huge number of branches that must have been produced during the course of its lifetime. Alternatively, if any branching system is closely examined from its tips back along its length, huge numbers of detachment scars where branches have died and been shed will be found. Only a small proportion of branches ultimately persist and become reinforced into main thoroughfares. If it were not for continuous self-pruning, all trees would be dense thickets.

 

                        Similar patterns of autodegeneracy are evident in many other examples of indeterminate systems, for example in the paths worn by migrating animals, in human communication systems, settlements and civilizations, in phylogenetic trees of evolutionary ancestry, in the evolution of ideas and knowledge, in fading memoriesÉ.

 

 

Decomposition

 

As just alluded to when mentioning ÔheartrotÕ, death is but the beginning of the processes through which boundaries that have outlived their time are degraded into forms through which their stored energy can be returned to rejuvenate life. These processes are collectively known as decomposition, and without them planetary life would soon grind to a halt for lack of carbon dioxide to sustain photosynthesis. Life would become a museum relic of its former self, for none to see or celebrate, unlessÉUnless set alight by fire. Ah, yes! Fire! For what is decomposition, really, but a slow-burning fire? A fire, damped down by water of its own making, through the combination of hydrogen fuel with oxygen, releasing chemical energy in the process. A fire slow enough to provide for new life.

 

 

DIVERSE OPTIONS – ALTERNATIVE LIFE STYLES

 

Metamorphoses of the kind described above allow organisms to divide labour between distinct life stages, but lack versatility because these stages occur in a specific sequence, e.g. larva -> pupa -> adult -> egg -> larva. There is therefore no scope for changing form to correspond with locally unpredictable changes of circumstances. If a particular stage does not encounter an environment in which it can proliferate, subsequent stages cannot be produced, no matter how suitable conditions might be for them. This has great practical significance in controlling ÔpestÕ organisms, for example, because it is only necessary to target one life cycle stage to eliminate the remainder. Mosquito adults can be targeted via their larvae.

 

                        More versatility can be achieved by producing multiple developmental options. Here, each option can be specialized for a particular role, but there is no set sequence by which one option arises from another.

 

 

Alternative Phenotypes

 

Where multiple developmental options are expressed as distinctive, determinate body forms, they can be called Ôalternative phenotypesÕ. There are many examples. There are protists that develop big or small mouths depending on the size of their prey. There are parasitic wasps that do or do not possess wings and bushy antennae depending on which host their larvae grow up in. There are butterflies that have different colours and body patterns depending on the time of year at which they emerge from the pupa. There are the castes and morphs of social insects that differ in the way their body boundaries expand depending on how the larvae are fed.

 

 

Mode-Transitions

 

In indeterminate body forms, varied developmental options can be expressed in different parts of the same, interconnected system, as Ômode transitionsÕ. Commonly, for example, different leaves on the same plant can have radically different forms depending on when and where they are formed, a condition known as ÔheterophyllyÕ. For example, leaves developed in strong sunlight (Ôsun leavesÕ) tend to be thicker but narrower and held in a less horizontal orientation than those produced in shade (Ôshade leavesÕ). Many aquatic plants produce highly dissected or strap-shaped leaves underwater, but leaves with broad blades upon or emerging out of water. In ivy (Hedera helix), the leaves on flowering stems are unlobed, whereas those on non-flowering stems are lobed. The lobed leaves have tributary-like patterns of venation, with wide-angled, erratic branches of many different widths, whereas the unlobed leaves have distributary-like patterns of branching.

 

                        Structures that can be recognized to originate developmentally as leaves can also undergo a variety of transfigurations that suit them for different functional roles. The different parts of flowers, the sepals, petals, stamens and carpels are all modified leaves. Leaves produced at the base of flowering stems are known as ÔbractsÕ and can sometimes be brightly coloured like petals. Leaves and parts of leaves can be modified into the coiling tendrils of climbing plants, the protective scales around buds, the storage leaves of bulbs and the spines that deter herbivores. Some of the most extraordinary leaf-modifications occur in plants that extend their supplies of nitrogen by capturing and digesting insects and crustacea – butterworts, bladderworts, sundews, venus fly traps and pitcher plants.

 

                        The indeterminate axes of plants and fungi, from which determinate offshoots such as leaves arise, can themselves exhibit an array of alternative forms or states. Transitions between these states can occur gradually or abruptly and may generally be regulated by changes in boundary permeability, deformability and internal partitioning. In fungi, proliferation may be in the form of cells (e.g. yeasts) or mycelium. The mycelium may be densely branched but slowly extending or sparsely branched but rapidly extending. The hyphae may be septate or not septate, anastomosed or not anastomosed, assimilative or not assimilative, diffuse or aggregated. In plants, slow-dense and fast-sparse branching patterns also occur and have respectively been regarded as representing ÔphalanxÕ and ÔguerrillaÕ formations. Plant root systems are often divided into relatively highly branched, absorptive Ôshort rootsÕ of limited duration and less branched, indefinitely extending, conductive Ôlong rootsÕ. Equivalent alternations occur between stoloniferous and rooting stages of plants like strawberries and the nomadic and settled phases of animal societies.

 

 

 

EVOLUTIONARY SYNTHESIS: LIFE CYCLES AND SPIRALS AND THE ORIGINS OF CREATIVITY

 

The Hole in the Mole – an ÔinclusionalÕ poem, by Alan Rayner

 

I AM the hole

That lives in a mole

That induces the mole

To dig the hole

That moves the mole

Through the earth

That forms a hill

That becomes a mountain

That reaches to sky

That connects with stars

And brings the rain

That the mountain collects

Into streams and rivers

That moisten the earth

That grows the grass

That freshens the air

That condenses to rain

That carries the water

That brings the mole

To Life

 

 

Having reviewed the many ways in which life forms generate and sustain diversity through self-differentiation and self-integration, it is now appropriate to explore the deeper evolutionary question of why they do this, and how inclusional approaches to this question move on from rationalistic approaches. I will begin by focusing on the dynamic origin of the creative potential that makes it possible for life forms both to bring about and respond to change. I will then consider the contrasting ways in which this possibility can actually be realized at dynamic boundaries, depending on elemental circumstances. Finally, I will discuss how the relational approach to these issues is affected by mathematical assumptions and methodologies.

 

 

 

 

 

Evolutionary Ebb and Flow: Problems and Opportunities

 

Nowhere is the dislocation of discrete contents from their dynamic spatial context more obvious, or more profound in its influence on the way we regard our relationships with one another and other life forms, than in the evolutionary biological notions of Ônatural selectionÕ and Ôsurvival of the fittestÕ. The implicit Ôfixed framingÕ in these notions, following on from Malthusian principles of limits to population growth, is evident in the way that Ônatural selectionÕ is commonly portrayed as a ÔpressureÕ. This pressure intensifies as population growth squeezes out available resource/space so that ultimately only those entities with particularly favoured characteristics can endure.

 

                        There are deep inconsistencies embedded in these notions, arising from the associated dislocation of changes in organisms (and their genes) from changes in their environment. This dislocation results in the loss of co-creative power and coherence from the dynamic system, and their delegation to some external agency, rather as with a cine film that requires a projector and an observer lacking resolving power to create the illusion of movement captured in its freeze frames. Moreover, the resultant placement of action and reaction in linear sequence raises endless, unanswerable questions of precedence and origin: Ôwhich came firstÕ - Ônucleic acidÕ or ÔproteinÕ, ÔnatureÕ or ÔnurtureÕ, ÔchickenÕ or ÔeggÕ etc, and where did these agencies come from, and how? And it renders the evolution of complex form from disparate ÔindependentÕ components astronomically unlikely – celebrated examples being the vertebrate eye and, even more fundamentally, the living cell with all its closely co-ordinated relationships between fine structure and metabolic processes.

 

On the one hand, the environment is treated as a ÔgivenÕ - a passive fixture imposed upon its living contents. On the other hand these contents are treated as passive, pre-formed, discrete units, lacking relationship with others, which can thereby only respond in a prescriptive way to the environmental circumstances on which they are imposed. Although changes in organisms and changes in environment are both recognized as essential to evolution, the actual mechanism(s) underlying their simultaneous and complementary relationship is obscured, so that this relationship appears instead to be sequential and adversarial.

 

Attention then focuses on how, as the putatively primary evolutionary mechanism, adaptive and purely genetic changes in these contents are enforced through competition in a confined space, rather than how the context, which actually includes and simultaneously both shapes and is shaped by these contents, transforms. Far from creating the observed diversity of living form, the effect of adaptation and competition in a fixed space would actually be the inexorable drive towards hegemonic monoculture, through the removal of variation implicit in the notions of Ôcompetitive exclusionÕ and Ôadaptive peaksÕ. Such hegemony conflicts not only with the observed diversity in natural biological communities, but also with the widespread occurrence within and between closely related populations of the process of sexual ÔreproductionÕ (a contradiction in terms, since the word, ÔreproductionÕ, implies Ômore of the sameÕ whereas sexual recombination produces variety). This process has always been a conundrum because it reduces the ability to make more of the same genetic self (i.e. truly to ÔreproduceÕ), which is the putative basis for evolutionary ÔfitnessÕ under Ôshort term selection pressureÕ.  Meanwhile, far from enhancing ÔfitnessÕ in the form of  ÔefficiencyÕ, the operation of systems at their most intensely competitive under conditions of Ôresource-limitationÕ would greatly increase the wastage that is actually prevented under such conditions in natural systems by pooling and reduced consumption. Natural selection, as it is most widely and popularly represented, is a profoundly counter-evolutionary mechanism, which, if it existed, would greatly reduce the energy-efficiency and impede the innovation that it is supposed to promote. Furthermore, the notion of producing increasing order and complexity through natural selection is not only self-contradictory, but also appears to contradict another derivation from impositional logic, the second law of thermodynamics, which views the irreversibility of natural processes in terms of the inexorable increase of ÔentropyÕ.

 

Inclusional logic, by contrast, radically changes our understanding of irreversible (evolutionary) change, according to principles that are common to all kinds of physical, chemical and biological systems, and that restore co-creative power and coherence to the dynamic relation between content and context. Rather than beginning, through the imposition of a fixed reference frame, with an assumption of stasis that then has to be ÔforcedÕ into action from ÔoutsideÕ, the very nature of nature is understood to be dynamic. And with this understanding, our concepts of causality and uncertainty also change. Rather than regarding change as externally enforced and measurable as a progression through space referenced to intervals of absolute time, all change is understood to involve the transformation of space and consequent simultaneous alteration in both content and context and their reciprocal relationship. And this simultaneous, reciprocal alteration, where content and context co-creatively shape one another can be thought of as attunement or resonance, rather than adaptation.

 

So, unlike the impositional logical perception that when only one thing moves, everything else remains fixed, in inclusional logic when one ÔthingÕ – a place somewhere – moves, the shape of possibility space everywhere transforms. And this contextual transformation is experienced uniquely at every location as a shift in the inductive pull of a potential energy field, extraordinarily rich with ever-changing evolutionary opportunity. This field is invisible and intangible to the external observer, but provides the locale for the emergence of complex form through synergistic processes that have been referred to, albeit from the perspective of impositional logic, as Ôself-organizationÕ.

 

Since such transformation necessarily involves a change in content-context, it is by its very nature irreversible and unrepeatable - unable to return directly or indirectly to exactly the same place that it emanated from. Far from being reproductive, producing more of exactly the same, natural processes are continually recreative and autocatalytic- opening up and building upon new possibilities. As was said so long ago by Heraclitus, Ôyou can never step in the same river twiceÕ. Content and context, stream and catchment, continually re-shape one another in an ever-transforming flow of place. This place is dynamically framed by itself as a resonant coupling of inner with outer energy-space, as was effectively recognized, albeit in a conventional mathematical framing, by the communication theory of Dennis Gabor (1946). Long neglected scientifically, but now being rediscovered, this theory provided the basis for GaborÕs Nobel Prize-winning invention of holography, key to which was the notion of a Ôcomplex signalÕ as a reciprocal combination of real and imaginary components, rather than an independent pulse of information.

 

 

Impositional and Inclusional Logic in ÔSimpleÕ and ÔComplexÕ Depictions of ÔSelfÕ, ÔDeathÕ and ÔCommunityÕ

Taken to extremes, the primacy given to individual survival in natural selection theory can result in the conclusion that Ôthere is no such thing as society/communityÕ, because the requisite co-operation in such a collective organization would compromize individual Ôself-interestÕ. Both diversity and co-operation are deeply problematic concepts according to this view, and so, if they are to be desired or tolerated at all in human societies, can only be sustained by legal and educational enforcement. As Dawkins ironically put it, ÔLet us try and teach generosity and altruism, because we are born selfishÕ!

 

                        However, such conclusions about the nature and occurrence of Ôself-interestÕ, ÔselfishnessÕ, ÔaltruismÕ and ÔsurvivalÕ inevitably depend very fundamentally on how the notion of ÔselfÕ is actually perceived. Here can be found perhaps the most far-reaching difference between impositional, fixed framing and inclusional, dynamic framing of evolutionary processes, with regard to how we relate to one another and our living space.

 

                        Using impositional logic, the notion of Ôindividual selfÕ as an independent body annihilated by death is simple and unambiguous, and the conclusion that evolution thereby entails inherently ÔselfishÕ processes focused on the survival of genes that prescriptively define this ÔselfÕ is inescapable. But with this conclusion come the paradoxical inconsistencies and lack of coherence described in the previous section.

 

                        Using inclusional logic, however, the isolation of the simple, fixed notion of self becomes subsumed by the togetherness of complex, dynamic forms (in effect Ôflow formsÕ) comprising inner, outer and intermediary spatial domains, all of which are vital to their distinct, but not discrete, identities. Rather than being unitary or binary, ecocentric or egocentric, such Ôcomplex selvesÕ represent ternary couplings of inner with outer, of the kind alluded to by ShakunleÕs Ôfluid logic numbersÕ (see below). Their behaviour is therefore ultimately intractable to impositional logic, as was implicitly acknowledged by Newton ÔhimselfÕ in his analysis of the Ôthree body problemÕ. Moreover, this behaviour can neither be regarded as intrinsically ÔselfishÕ nor ÔaltruisticÕ, because neither the disregard of the outer (ÔcollectiveÕ/ ÔweÕ) nor inner (ÔindividualÕ/ÕIÕ) aspect is evolutionarily sustainable in such a co-creative system.

 

                        The concepts of Ôcomplex selfÕ and Ônested holeynessÕ were anticipated by Koestler in his descriptions of ÔholonsÕ - as ÔJanus-facedÕ entities combining individual and collective aspects, and ÔholarchiesÕ - as nested arrays of holons, in his ÔOpen Hierarchical Systems TheoryÕ. Even more pertinent was the description of a ÔRussian dollÕ kind of nesting, by Caldwell et al (1997), who recognized that the resultant conflation of ÔinformationÕ deriving both from content and context was inconsistent with the notion of an external Ônatural selectorÕ. This recognition is made all the more potent when the necessary incompleteness, and consequent transformability (indeterminacy) of space-incorporating boundaries is introduced. We then can make the full transition from a view of ÔselfÕ as an object, to an appreciation of self as a place. Not only is every ÔplaceÕ necessarily both a grouping of smaller ÔplacesÕ and grouped with others in some larger ÔplaceÕ, but the incompleteness of boundaries ensures that there is communicative spatial relationship and the possibility for transformation across all scales.

 

Only through the development of an explicitly ternary logic, via the introduction of a dynamically balancing, intermediary agency, can the paradoxes resulting from the severance of inner from outer be avoided. In this ternary, Ôdynamic framingÕ, complete sealing of boundaries would disrupt and stifle flow, whereas total dissolution of boundaries would end in featurelessness. So both the pursuit of absolute individual autonomy (independence and immortality) through the completion of external boundaries, and of absolute collective unity (dependence and self-abandonment) through the obviation of internal boundaries are evolutionarily untenable. By contrast, a holey (i.e. space-including and hence permeable or porous) intermediary boundary provides the possibility for energy transfer between dynamically coupled inner and outer inductive domains. Closing in (decreasing holeyness) of boundaries results in ÔinformationÕ, the constructive shaping of local ÔfeaturesÕ and increased resistance to energy transfer both from outer to inner (inspiration/ in-welling) and from inner to outer (expiration/out-welling). Opening out (increasing holeyness) of boundaries results in ÔexformationÕ, and consequent decreased resistance to energy transfer.

 

The complementary interdependence of generative and degenerative processes via dynamic boundaries between inner and outer is therefore inescapable. Space, though we may perceive it rationalistically as  ÔimperfectionÕ, cannot be excluded from a vital, evolutionary system, try as we might in the pursuit of ÔperfectionÕ in the form of individual or collective completeness (wholeness). Such ÔperfectionÕ would imply eternal stasis. Rather, in the excitable, dynamic world and universe that is drawn towards balanced relationship, outside yields to and feeds the growth of inside, which yields in turn to outside in natural renewable cycles and spirals. These natural inspirations and expirations are disrupted, and even reversed, by the severance of one from the other.

 

                        In this inclusional view, there is therefore nothing problematic about co-operation and diversity, nor, for that matter, about outwardly ÔaggressiveÕ behaviour that sustains diversity through the assertion of local identity. Rather, what we have, as many ecologists implicitly or explicitly recognize in natural ÔecosystemsÕ and development of their increasingly complex and interdependent inhabitant ÔcommunitiesÕ through autocatalytic stages (i.e. ÔseresÕ) of succession, is a dynamically creative Ôtogetherness in diversityÕ or Ôcomplementarity of labourÕ.  Here, the collective and individual, Ôthe forest and the treeÕ, both necessarily incomplete, continually reconfigure one another as they explore and manifest their common-space realm of possibilities.

 

Nonetheless, we continue to find it virtually impossible to apply this understanding to evolutionary co-creativity and communication, due to our continuing adherence to impositional logic. Even when we proclaim ÔinterconnectednessÕ, we are prone mentally to envisage connections as solid transactional ÔstringsÕ or ÔtiesÕ across space that are inserted between initially discrete entities, rather than as conduits or ÔpipelinesÕ of included space that grow relationally into place.  This transactional Ôjoining up of dotsÕ is evident in the metaphor of ÔwebÕ and in modern Ônetwork theoryÕ whereby each of the connected entities is regarded as a ÔnodeÕ or ÔhubÕ whose influence corresponds with the number of connections that radiate out from its self-centre. Such thinking is being applied increasingly not only to human organizations but also to natural ecosystems, where the most influential hubs represent what have been called Ôkeystone speciesÕ, with the inference that less connected entities are more readily dispensed with. Although such constructs of entities plus connections are portrayed as highly effective communication systems, examination of their structure reveals them to be highly resistant to flow and transformation as well as fundamentally unlike actual biological networks like blood systems, nerve systems and fungal mycelia. The latter consist of variably permeable and deformable tubes capable of highly versatile and re-distributive responses to their local circumstances. The formation of lateral connections or ÔanastomosesÕ, which connect the tubes Ôin parallelÕ, greatly increases the conductivity of these systems, as does their lack of hubs, which would actually serve as ÔbottlenecksÕ.

 

 

 

 

Creativity and Breakthrough: Neoteny, Macroevolution and Paradigm Shifts

 

 Suchantke has quoted A. Takhtajan as stating that Ôin comparison to most Dicotyledons, the typical Monocotyledons (including all primitive forms) are characterised by a certain ÒinfantilisationÓ in the vegetative realm. The sort of simplifications they display look like the products of a truncated ontogenesis. Their cambium is reduced to an axial activity and the main root does not develop. The leaves either remain completely undifferentiated or divide indistinctly into stalk and blade, and resemble in their venation the incompletely developed leaf-organs of Dicotyledons (stipules, bracts, bud-scales, sepals etc). Taken together these facts have led me to suppose that neotenyÉhas played a decisive role in the origin of the MonocotyledonsÉNeotenous development is, to my way of thinking, the key to understanding their morphological peculiarities.Õ Suchantke then adds that Ôin contrast to anagenetic development, which favours the formation of bodily structures, the complete flow of morphic potential into the physical body, in which it becomes completely bound and literally ÔsmotheredÕ, juvenilisation signifies a return to the origin, in the sense of a state in which all possibilities are still (or once again) open. It is certainly not a regression to some past state, but through it the organism is able to begin something new on a different level, or is primed for such developmentÉthe process of juvenilisation is a prelude to a genuine leap in evolution, i.e. the emergence of something new and completely unprecedented.Õ

 

This account of the evolutionary origins of monocotyledons (grasses, lilies etc) is but one example which fully accords with the idea that the capacity for self-integrative return in life spirals is the source of creativity that leads to radical change under the inductive pull of evolutionary opportunity. There are many other examples. All animals with backbones (vertebrates) are believed to be descendants from the pelagic larval stages of sea squirts. Homo sapiens itself may owe its extraordinary scope for discovery and innovation to being a neotenous ape. An immense variety of domestic animals have found niches in the company of people through their juvenile qualities of trust and playfulness. Cultivated plants, protected from competition in the nursery and so enabled to shed their physical and biochemical combat gear have likewise assumed forms far removed from their wild progenitors under the inductive pull of human influence. Peacocks have elaborated fantastic tails under the induction of female space (an alternative view of so-called Ôsexual selectionÕ). The child-like qualities of questioning authority and playing with ideas used by scientists and artists who have been foci for major shifts in thought or expression are unmistakable.

 

 

Life History Strategies and Horses for Courses – the Short and the Long of It All

 

The transformation of creative potential into the diverse modes of actual life depends on how life forms respond to and bring about variations in their dynamic context. Of fundamental importance here are patterns of availability of life-supporting resources. Are these resources widespread or local, renewable or finite, plentiful or sparse, temporarily available or persistent, easy or difficult to access and assimilate? Depending on the answers to these questions, the balance between self-differentiation and self-integration during life histories, and corresponding deformability, permeability and continuity of boundaries can shift markedly. This shifting balance results in distinctive Ôlife historyÕ or Ôecological strategiesÕ. These strategies differ in whether the pitch of the life spirals that they are associated with is short or long, and whether the diameter of the spirals stays within or extends beyond fixed limits.

 

                        Where resources are temporary, plentiful and easily accessible and assimilable, the emphasis is on short-term exploitation through rapid self-differentiation. Life spirals have a short pitch and small diameter. Boundaries are permeable, deformable and dissipative. Innovation and synergism are minimal. Primacy in finding resources before competitors do is paramount. There is the boom-bust capitalist economics of unbridled proliferation, regardless of waste or effect on the living space. There is no conservation of the past and no preparation for the future other than through reproduction and dispersal to new sites of exploitation. This is the pattern, closest to the neo-Darwinian paradigm, of ruderal and ephemeral life forms that prosper briefly in sites of recent ecological disturbance, including many that human beings call ÔweedsÕ and ÔpestsÕ. It is also, all too commonly, the unsustainable pattern of human depredation of our environment.

 

                        Where resources are relatively accessible and assimilable, and renewable or persistent but not continually plentiful, longer-term patterns of territoriality and synergism develop. These patterns are marked by a shift in emphasis towards self-integrative processes of boundary-sealing, boundary-fusion and boundary-redistribution. Past structures provide foundations for future occupation. Dissipation is lessened and innovative versatility enhanced in spatially and temporally changeable living spaces. The pitch of life spirals lengthens and their diameter enlarges. There are the pluralistic economics of sustainability. These are the patterns of combative and/or collaborative life forms, dominant at the ÔclimaxÕ or zenith of ecological successions and evolutionary Ages where spatial scale and heterogeneity are maximal. They are also the dominant patterns in human societies, paradigms and civilizations that have attained their most established form prior to degeneration, revolution and creative re-emergence.

 

                        Where resources are inaccessible due to adverse environmental circumstances, or are difficult to assimilate, the stage is set for specialist life forms, few in number and often slow to proliferate or rejuvenate, but strong in resilience and uniqueness. There may be protracted periods of dormancy. Self-integrative processes are maximized and dissipation minimized. There are the economics of stasis. The pitch of life spirals may be long and the diameter wide. This is the pattern of life forms notable for their rarity and endurance. It is the pattern of the recluse and indigenous in human societies.

 

                        It is important to appreciate that the descriptions of the three basic strategies just given are intended only to provide a rule of thumb for understanding the evolutionary and ecological origins of diverse patterns of life. They should be regarded as relative and context-dependent rather than categorical. They often overlap, and any particular life form may exhibit more than one strategy over the course of its life spiral(s), as in metamorphoses, alternative phenotypes and mode transitions.

 

 

The Elemental Interplay of Contextual Transformation

 

As has been emphasized, the dynamic interplay between self-differentiation and self-integration depends fundamentally on the availability of external energy supplies (ÒresourcesÓ). High availability allows high input rates and consequent self-differentiation of dissipative free surface (boundary-externalization). Restricted availability leads to self-integration into coherent organizations capable of exploring for, conserving and recycling resources, and so connecting across times/spaces of shortage between times/spaces of plenty.

 

Studies with fungi have helped to highlight the importance of the relationship between sources of fuel, oxygen and water in the boundary chemistry that mediates this interplay. Fungi, like the majority of earthly organisms, have become addicted to oxygen as a potent but potentially dangerous source of chemical energy through its role as an acceptor of electrons. These electrons are passed from the earthy fuel of carbohydrates, fats and proteins through the consuming fire of respiration. Oxygen – of which water, through the constructive light-fire of photosynthesis, is itself a source – accepts them one at a time, so generating extremely reactive chemical species, including free radicals, that are only rendered innocuous once the electrons have been fully incorporated into water. Any imbalance between the supply and demand for electrons allows the reactive species to accumulate, disrupting the chemical integrity of living cells and causing degeneration. This fate can be prevented in watery environments by quenching and excreting reactive species by means of antioxidant compounds that can then play additional roles, e.g. as antibiotics. In terrestrial environments, though, it is necessary to prevent excess oxygen entering cells from the gaseous phase, where it diffuses 10,000 times more rapidly than through water, and this is done by actually using the oxygen to produce polymeric, oxygen-impermeable coatings. These coatings change the boundary permeability and consequent pattern-generating potential of life forms that possess them.

 

Here, then can be seen the fundamental way in which the contextual responses of life forms to the threat and promise of oxygen has shaped their evolutionary course. These responses cause boundaries to open, seal, fuse and degenerate, so enabling energy sources to be gathered, distributed, conserved and recycled as circumstances dictate, just as I have described in preceding sections of this script. 

 

 

The Elementary Mathematics of Contextual Transformation

 

The contextual dynamics of living systems are, then, fundamentally and irrevocably counteractive as a consequence of the role of their boundaries both in the transfer and in the containment of resources. In the sense that ÔpositiveÕ can be regarded as ÔsupplyÕ or ÔpressureÕ and ÔnegativeÕ as ÔdemandÕ or Ôlower pressureÕ (ÔvacuumÕ), there is both interdependence and counteraction between positive and negative, mediated at dynamic boundaries between insides and outsides. Consequently, there is always a tendency for the system to be drawn towards its Ôground zeroÕ condition. The positive is drawn in, and so restrained from expanding to plus infinity, by the negative. The negative draws in the positive and so is prevented from contracting to minus infinity.

 

Correspondingly, transfer of energy from external supplies across the boundary that gives externally perceptible form to a living system, down actively maintained gradients of water and electrical potential, generates a positive internal pressure. This internal pressure is counteracted by the resistance to expansion of the boundary and external environment and lessened by dissipative loss from the boundary. At the same time as increasing internal pressure, input of energy across the boundary reduces the external pressure, making the boundary more prone to expand, but also more prone to dissipative loss. When there is an exact balance between positive and negative, there is neither expansion nor contraction of the systemÕs boundary. The maintenance of this balance at any value of the internal pressure greater than zero, requires, however, that there be a continual input of energy sources from outside to compensate for losses – much as would be needed to keep a leaky balloon inflated. Any diminution in the rate of input will therefore result in contraction of the system – a withdrawal of boundary as in self-integration. On the other hand any enhancement of the rate of input will result in an autocatalytic expansion of dissipative free surface, as in self-differentiation, up until such time that this rate of input becomes unsustainable.

 

How can this counteractive situation be translated into mathematical terms? The problem here is that the number system upon which so much mathematics has been built assumes, consciously or unconsciously, that numbers are discrete units, independent and separate from one another like clothes pegs strung out along a washing line, even when infinitely large or infinitesimally small. There is no inbuilt tendency for larger (Ômore positiveÕ) positive numbers to become smaller (Ômore negativeÕ) positive numbers unless sustained by energy input. Nor is there any tendency for larger (Ômore negativeÕ) negative numbers to become smaller (Ômore positiveÕ) negative numbers unless sustained by energy output. In other words there is no implicit neutralizing flow from high pressure to low pressure – for numbers to flow into and out of each other according to supply and demand – and so no capacity for reciprocal transformation. Paradox is deeply embedded in the system, as becomes evident whenever finite is compared with infinite or infinitesimal, or when linear measure is forced into circular measure to produce ÔirrationalÕ numbers like p.

 

One way of approaching this issue, whilst staying literally within the confines of a discrete number system, is through the use of non-linear equations (note that the difference between two sides of an exactly balanced equation is indeed zero). Here processes of becoming larger as the consequence of addition are accompanied by a proportionally increasing subtraction, such that more positive values Ôfeed backÕ into less positive values.

 

A well-known example of a non-linear equation is the Ôlogistic difference equationÕ. This equation relates the actual number of entities (x) as a proportion of the maximum possible number (1) in a current population to the number of entities in the next generation (xnext) in terms of the net rate of reproduction (r) per head of population as follows:

 

xnext = rx – rx2

 

where x varies between zero and 1.

 

                        Here, the potential for increase in x, due to the reproductive drive, r, resulting from resource acquisition is countered by the negative feedback term, rx2. When this equation is iterated (i.e. when the output xnext value is used repeatedly to input the next x value) from some low initial positive value (if initiated from exact zero it will remain zero for eternity), the rx2 term increasingly constrains the increase in x. Ultimately, there is no net increase when x is equal to 1 – 1/, representing the Ôequilibrium population sizeÕ or Ôcarrying capacityÕ of the population.

 

                        For values of r between 1 and 3, the equilibrium population size ranges from zero to 2/3, and iteration of the equation from low values results in an initial increase in x. This increase either leads directly to attainment of the equilibrium value if r<2, or, if r>2 to a series of progressively smaller fluctuations (i.e. Ôdamped oscillations) above and below the equilibrium value. For values of r<1, x becomes zero. For values of r>3, however, the population is driven over a threshold where it becomes unstable, unable to attain a single equilibrium state, unless arriving by some infinitesimally small chance at exactly the requisite value of 1-1/r, and instead subdividing or ÔbifurcatingÕ into a series of alternative states. Here, as r is increased, x values come to oscillate around first two, then four, then eight É2n values in a so-called Ôperiod doublingÕ cascade. At r = 3.57, deterministic ÔchaosÕ first becomes evident, as x values vary unrepeatedly and at r = 4, all x values between 0 and 1 become possible.

 

                        Some fundamentally important conclusions can be drawn from these properties. Firstly, in counteractive systems that are driven hard enough, long term behaviour becomes fully unpredictable (or, rather, predictably unpredictable) due to their Òsensitivity to small changes in initial conditionsÓ that become amplified by feedback. Secondly, those systems that proliferate most freely, i.e. with the highest r values, are the ones most prone to instability – to Ôboom or bustÕ (cf. earlier discussion of r-selected life history strategies). Thirdly, what appear to be statistically uncorrelated variables are not necessarily independent.

 

                        An understanding of how this complicated behaviour arises can be obtained by plotting the solutions generated by iterating non-linear equations on a map whose co-ordinates define the state of the system (i.e. its displacement and rate of change) in Ôphase spaceÕ. By joining the positions produced on this map by successive iterations, a kind of Ôfate pathÕ or ÔtrajectoryÕ is derived, which describes how the characteristics of the system change as energy is fed through it. This trajectory can show a variety of behaviours depending on the condition, known as an ÔattractorÕ, towards which it is drawn. The simplest attractors are Ôfixed pointÕ attractors – equilibrium states that once arrived at cease to change. This is the condition for the logistic equation with r<3. In cases where damped oscillations occur, the trajectory spirals around before reaching the fixed point. Where the trajectory cycles repetitively around two or more values, the attractor is known as a Ôlimit cycleÕ. Where the trajectory wanders non-repetitively, but nonetheless never exceeds certain bounds, the attractor is known as a ÔstrangeÕ, ÔchaoticÕ or ÔfractalÕ attractor.

 

                        The concept of a ÔfractalÕ attractor relates to a kind of geometry of irregular structures that cannot usefully be described in classical Euclidean terms of smooth curves and surfaces arrayed in integral dimensions of 1, 2 and 3. When these structures are examined more and more minutely, more and more irregularities come into view. Their lengths, areas and volumes are therefore infinite when viewed at infinitesimal scales, even though they can be circumscribed within finite planes, spaces or space-times (note that this is an example of the paradoxes caused by comparing finite and infinite using discrete number systems). Moreover, they commonly exhibit the property of Ôself-similarityÕ whence, if a small part of the structure is magnified it looks similar to a larger part of the structure.

 

                        The problem of quantifying fractal structures is intractable if approached conventionally, using standard units of length, area, volume etc. It can be solved, however, by relinquishing the notion, arising from discretist thinking, that dimensions can have only integral values of 0, 1, 2, 3, 4 etc, and allowing them also to have fractional (hence ÔfractalÕ) values (but note that fractions are also discrete – see below). The fractal dimension of a structure can be calculated from the equation:

 

M = krD

 

where M is the material ÔcontentÕ of a portion of the structure, r is the radius of the field in which this portion of content is contained, and D is the dimension. D can readily be found from the relationship between the logarithms of M and r for different fields of view. If the structure is homogeneous, then D will have an integral value. If it is heterogeneous, D will be fractional.

 

                        A very useful feature of the fractal dimension is, then, that it indicates how uniformly a structure permeates space in terms of the relation between content and form. For example, on a flat surface a branching structure with predominantly radiately aligned axes has a fractal dimension close to 1, corresponding with wide coverage and very uneven density, whereas a structure with D close to 2 has relatively more tangential axes and hence a more even density.

 

                        Correspondingly, fractal dimension is a measure of the degree of self-differentiation of boundary free surface (the degree of ÔfreedomÕ or ÔsubdivisionÕ) of a system. It relates directly to the way a coherent (homogeneous) initial state breaks symmetry in response to energy input sufficient to exceed its carrying capacity, and begins to subdivide or branch.

 

In purely mathematical formulations of the concept, however, fractal subdivision only occurs within the pre-imposed set limits of the overall, containing boundary of the system or attractor. In applications to real systems, the ultimate unsustainability of fractal structures, their consequent dependence on self-integrational processes and the ways in which these processes enable boundary limits to be exceeded or changed have generally been overlooked.

 

The development and application of non-linear mathematical models over the last few decades has therefore remained limited: the hallmark Ôsensitive dependence on initial conditionsÕ may be an artefact of  Ôhuman dependence on fixing initial conditionsÕ (i.e. imposing closure) and hence an indirect rather than direct insight into openly inductive (indeterminate) holes. This is not to say that it has not revolutionized thinking by laying to rest notions of long-term predictability and precision in dynamic systems. But the subject-object, form-content, inside-outside, positive-negative, figure-ground, nature-nurture, information-environment, self-other, genotype-phenotype, all-none, present-absent dichotomies remain firmly in place, with all the attendant paradoxes and inconsistencies.

 

Perhaps the time is coming not only to relinquish the cherished certainty of the linear, but also to pass beyond the protective veil of the discretely closed off into a world where, indeed, Ôall is flowÕ. More direct insight into this world could be achieved by moving from binary to ternary logic. Here, every feature and number is viewed both in terms of its own Ôself-identityÕ, and its Ôpossibility spaceÕ of what it contains and what it is contained by, hence connecting it relationally to all space-time. Numerically, this can be achieved by assigning a number a ÔtripletÕ code. This code consists of the number itself (its Ôpresent or Ôzero conditionÕ) together with the number immediately smaller (its ÔpastÕ condition of what it has been and could become again) and immediately larger (its ÔfutureÕ condition of what it could become) than itself. For example, the number conventionally represented as Ô2Õ could be represented instead as Ô123Õ. Rendered geometrically, these relational, Ônatural logicÕ numbers depict infinity in recurved, spiral form (Shakunle, 1994) that may well prove deeply to reflect the inductive-assertive flow dynamics of evolutionary processes. For the moment, however, this ÔnewÕ mathematics is in its infancy, so weÕll have to wait and Ôwatch this spaceÕ.

 

 

PEOPLE AND ENVIRONMENT: FROM CONFRONTATION TO RAPPORT

 

And so, having witnessed the interdependence between life forms and their dynamic context across all scales, we come at last to the reality of practical environmental management. What do we do? What can be done? What should be done?

 

In a sense we are all environmental managers, both individually and collectively, through the mutually transforming relationships we establish with our living space. Some of us may end up as professional environmental managers, trying to advise others how to conduct their business in ways that sustain and enhance quality of life rather than damage it. All professions contain some element of environmental management founded in a shared responsibility to look after our living space. One thing is certain: if we neglect our living space we ultimately neglect ourselves.

 

How, then, can we establish a mutually supportive relationship with our living space – one that can endure and develop rather than decline? How, in particular, can we use our understanding of living systems to enhance this relationship and guide our responses to environmental issues as they arise?

 

In order to address these questions, it is necessary to have some idea of how the principles of working with dynamic, responsive systems might compare and contrast with those of working upon what we have regarded as ÔinanimateÕ systems. Here the varied ways that I have been describing in which living systems, contain, take in and distribute or redistribute energy sources become paramount.

 

 

Challenging Unpredictability – Cast Iron or Bamboo?

 

Ask any conventionally inclined analytical scientist what they find most problematic about living systems, and they will be sure to point their finger at the mainstream theme of this script. Living systems are inherently complex, variable and, in the long term, unpredictable. You just canÕt rely on them. You just donÕt know where you are with them. Their performance just isnÕt always reproducible. Just ask an honest biologist about what will happen if you alter living conditions in some way and (s)he wonÕt give you a straight answer. The best you can do is just get bogged down in statistics that donÕt prove anything.

 

                        But perhaps the very properties of living systems that from one point of view appear problematic represent wonderful opportunities in another perspective. Do we really want to ignore these opportunities? The modern upsurge of interest in bioengineering and biotechnology shows that we donÕt. But our continuing efforts to control and constrain biological systems, including ourselves, in order to make them behave less erratically, is testimony to our reluctance to accept, value and work with them as they are or could naturally become rather than as we might desire them to be. We continue to try to exclude rather than include their inherent possibilities, to make them conform to prescriptive specification rather than allow them to expand both our horizons and their horizons into a less certain but potentially far richer world.

 

                        What I aim to do next, therefore, is to highlight those very characteristics of living systems that distinguish them from the ÔinanimateÕ and make them so challenging to work with, but which can open new vistas of opportunity if only we can be receptive to them. In order to do this, I will compare just one example of a biological system, bamboo, with some familiar manufactured products.

 

 

Static Rigidity or Dynamic Flexibility?

 

                        You know precisely, wellÉfairly precisely, where you stand with a set of building blocks or steel girders – how they can be incorporated into an assembled structure, how much space they will take up, how much stress or strain can be imposed on them before they break. The same could not be said for bamboo, whose strength and other properties will vary along its heterogeneous lengths and depend hugely on the conditions under which it has grown, not to mention the genotypes of the plants from which it originates. A precision-built bamboo structure is a contradiction in terms. No two such structures can be expected to look or perform quite the same. Yet, try to bend the blocks or girders into place, or alter them in some way to fulfil some new or unforeseen function, and their limitations will quickly become apparent. By contrast, if you put your mind to it, you can construct an entire living space, complete with walls, floors, roofs, plumbing system, window blinds, furniture, houseplants, artworks and musical instruments by bending and working bamboo. The loss of precision, predictability and controllability involved in working with the biological system is offset by its resilience and versatile responsiveness, together with the aesthetic appeal that comes from its being Ôcloser to natureÕ and thereby perhaps closer to our own nature.

 

The question that needs to be asked, therefore, is just how important are precision and conformity to us, really? To be sure, the quest for precision and conformity has brought us a long way, enabling us to navigate, keep time, drive cars, fly aeroplanes, assemble computers, send space craft to the planets and so on. But how far has this quest limited us to the straight and narrow, denying us the opportunity to wander off-route to discover who knows what? How much has it caused us to lay waste to the resources of the Earth, bringing about irreversible declines? How much has it made us busy ourselves generating dissipative structures that bring no long-term prosperity, at the cost of our human needs for belonging and relationship? How much is it necessary? Can we identify contexts in which it is and is not appropriate? Does it make us warm and at ease, or cold and alienated?

 

 

Manufactured Product or Productive Resource?

 

To make a brick or a girder, considerable energy is needed to extract raw materials from the environment, purify and heat them and get them into the requisite shape. In other words there is an inevitable human and environmental cost entailed in converting natural resources into manufactured products of uniform quality. But what if, as is the case with bamboo, the resource is to a greater or lesser extent already the product because it has grown rather than been forced into shape? A huge amount of conversion cost is thereby avoided, the energy having been provided renewably, by courtesy, ultimately, of sunlight.

 

Here perhaps lie the most exciting possibilities of living systems – that they can be grown into such a diverse array of structural and functional forms. Rather than us having to give them their shape, they produce it themselves through the transformation of renewable resources. Imagine the cost of making a horse or a flower – assuming of course that we ever could! But we donÕt have to. We can help them make themselves. All we have to do is relate to them – through understanding, to discover how to nurture and guide their creativity in ways that meet both their needs and ours.

 

 

Disposal Problem or Convertible Opportunity?

 

What happens when a product comes to the end of its useful life, through decay or obsolescence? Does it then become a disposal problem? Or can it be converted to another use? Perhaps because as mortal beings we so much fear the end of things, we often donÕt give much thought to these questions, laying them aside until the fateful day when they finally demand attention. By which time it is all too often too late. Manufactured things therefore tend to end up by cluttering up and increasingly, e.g. through their toxicity and radioactivity, even threatening our lives. This may be interesting for archaeologists attempting to unbury our history, but gets in the way of our present and future. By contrast, biological products are, by their very nature, readily combustible, decomposable and convertible to other uses. If the worst comes to the worst they can always be burned, preferably as a source of fuel. But there is also plenty of scope to bioconvert them into other products. For example they may, for a low energy cost, be composted, to produce fertilizer, or used to grow edible mushrooms, or converted into animal feed.

 

 

Management of Living Processes

 

Biological management has three main thrusts. Firstly it can involve intervening directly within the boundaries of a biological system to improve, control or remedy its operation. Secondly it can involve applying a biological system to some other process that we wish to manage. Finally, it can involve what has been called ÔbiomimeticsÕ – mimicking biological systems as a means of process management.

 

In this section I will consider the first of these thrusts. As in the previous and subsequent sections I will focus on three main kinds of themes. Firstly I will consider whether the approach is one that imposes upon or brings out the potential of the system. Secondly I will examine whether it uses artificial contrivances or relies upon inherent pattern-generating capabilities. Thirdly I will reflect on the extent to which it seeks immediate solutions to problems without regard to future repercussions.

 

 

                        Constraint or Facilitation?

 

                        Why harness a horse? Do we wish to impose control over the animal, to place its potential waywardness in check and make it do as we desire? Or are we seeking a way to gain access to its horsepower, a means of communication that opens up the scope for many and varied partnerships? Our responses to these questions will hugely influence the design of any harness we might manufacture. They are worth thinking about because they define the attitudes we bring to any kind of management that seeks to draw power from or remediate a living system. Ultimately putting on some kind of harness is the way that we influence the boundary properties of the system. But does this harness constrain or facilitate? Does it confine movement or does it allow freedom of movement? Does it make possible new kinds of movement? Does it impose or release pressure?

 

 

                        Artificial or Natural?

 

                        To begin with, is the harness just referred to artificial or natural – and, indeed, where are the dynamic boundaries between artificial and natural? Perhaps a good way of thinking about these issues is by reference to ourselves. Down the ages, there have been many ways in which we have sought to enhance what we can do by embellishing our basic bodies with varied forms of clothing, tools and housing. In so doing, we have greatly extended our phenotypic range. Moreover, some of us continue to entertain longings for immortality through reconstructing ourselves from a set of bionic replacement parts that dispense with the vulnerability of our flesh and blood. We might have artificial limbs, artificial hearts, artificial guts, artificial circulation fluids É digitized brains. But would we lose some vital aspect of ourselves in the process? Could there come a time when Human Being becomes pure Machine, alienated from our natural context and inhabiting a world populated by biomachines of our own making? Personally I doubt whether such a time or such a world could ever be possible because of the intrinsic limitations, already alluded to, of non-biological materials and processes. Time and again bioengineers attempting to design an artificial heart, or suchlike, experience the problems of assembling devices that no matter how precise or intricate fail to work in the long term because of their inability to keep in tune with a changeable context. In fact, imprecision is a vital ingredient in the attunement of living systems with their context, and it is now widely recognized, for example, that an irregular and complex heartbeat is healthy, whereas a regular, predictable one is deadly. The best substitute for a living mechanism or process may ultimately be another living mechanism or process of the same kind. It may be better in the long term to grow than to make replacement parts.

 

                        Once again, the fundamental issue here is the kind of attitude that underlies the thinking that we bring to bear on the problem. This time the question of attitude concerns the light in which we view living substance. Do we see the latter as something that needs to be replaced with something more dependable? Or do we idealize it as something with mystic powers that must be good in the long run and must remain pure, uncontaminated by the human quest for knowledge and control, if it is not to turn against us? Or do we try, in all humility, to understand it both from inside-out and from outside-in, finding ways to relate to and augment its possibilities by fusing its boundaries with the human-made.

 

Human beings are, after all, products of nature and so any things we make are also, in a sense, products of nature, even though we might distinguish them as artificial or artefactual. Would we call a snailÕs shell, a beaverÕs dam or a birdÕs nest ÔartificialÕ? No. Why treat what we might make as any different? In the end it is not the question of the distinction between natural and artificial that is at issue, but rather the relationship between what is within a living system and what the system makes of the world by transforming its surroundings. Is this relationship mutually beneficial, or abusive, such that one gains at the expense of the other or both lose out? Do human beings become enslaved, liberated or rendered useless by their own constructions? Do other life forms become empowered or disempowered through their interactions with human beings?

 

                       

                        Short term or Long Term?

 

The idea of empowerment through fusion of the self with the self-made or indeed non-self-made is implicit in Donna HarrawayÕs concept of the ÔcyborgÕ – that synthesis of the human and the machine that we have all become due to the now virtually seamless relationship between our selves and our accessories. It is also implicit in the very idea of interdependence between the insides and outsides of dynamically bounded systems and hence the creative evolutionary process itself. So to attempt to ignore or prevent it is both unrealistic and to forestall our future. On the other hand, to think that its outcome can be fully circumscribed in advance, or that this outcome will necessarily prove to be beneficial is foolhardy in the extreme.

 

In an inherently unpredictable context, short-term gain may very possibly turn out to be long term pain, and vice versa. For example, making the car an extension of our selves may well take us places, but it may also damage our environmental context and bring in its train all kinds of compulsive drives that disturb our peace and unsettle our relationships. Faced with such uncertainty, the best we can do is to follow what has become known as Ôthe precautionary principleÕ and keep a weather eye open. We should neither assume that all will be fine nor indeed that all will be devastation, but rather tread carefully, constantly alert to possibilities and prepared to question the outcome of our endeavours – whether we really are getting what we want or need. Do we, for example, really need to live longer and longer, thereby denying scope for rejuvenation? Do we want the things we make to last forever? What will we do with them when they have reached the end of their useful life? Is built-in obsolescence a sensible way of maintaining employment? Do we really need more food production to fill an ever-burgeoning number of mouths that increase in direct response to supply? Or, rather, do we need better quality and distribution of food to sustain the population we already have, whilst preventing those disparities that divide us into obese and malnourished? Do we need more roads to carry more vehicles over longer distances, or more effective local distribution programmes? Is it good to become locked in to the virtual reality of computer networks whilst losing sight of the real world in which we live? Are our relationships between ÔselfÕ and ÔotherÕ turning out as we might wish, or are they leading into unforeseen restrictions and misadventure?

 

Here, I have little personal doubt that the greatest threat to human and other quality of life comes not from attempting to manage our environment, which is quite ÔnaturalÕ in its own way, but in the arrogance of Ôassuming controlÕ and thereby failing to ask the questions that need to be asked. Sadly, this is the arrogance that has become increasingly characteristic of a kind of science and technology that alienates itself from its context by not allowing for relationship and concerns itself, like an ephemeral life form, only with the short-term exploitation of plenty. This is the arrogance that preens itself as ÔobjectiveÕ and Ôvalue-freeÕ and ÔpragmaticÕ, whilst casting aspersions on any attempts to be more inclusional or long-sighted. This is the arrogance that assumes it will be fine to breed and plant monocultures, apply herbicides and pesticides, remove habitats, alter growth parameters, feed sheepÕs brains to cattle etc on an unprecedented scale, only to be found out by disease, malnutrition and environmental destruction. It is this arrogance that has finally, if belatedly, aroused public concern, if not hostility, most recently expressed in the debate about the development of genetically modified organisms in which DNA is transferred, ÔunnaturallyÕ, across species boundaries. The public is right to be concerned, if not about the technique itself then about the context in which it is being applied in a state of wilful ignorance. But to allow the alienating approach of some scientists to be a reason to alienate science, to assume that the entire scientific endeavour is tainted and should therefore be thrown out, would be to discard the baby with the bath water. There is much that is good and creative in the baby if it is nurtured in a state of questioning awareness.

 

 

Management using Biological Processes

 

Biological systems can be the agency as well as the agent of management. Here, what can be done with, rather than to, the biological system is contingent upon the kinds of special properties discussed earlier in the light of the particular example of bamboo, and how these properties are harnessed, as discussed in the previous section.

 

 

                        Biomaterials

 

                        Depending on circumstances and type, biomaterials can have the advantages (or, from another perspective, disadvantages) of flexibility, heterogeneity, convertability, resilience, digestibility, degradability, renewability, aesthetic appeal and low environmental and economic cost of production. They are not uniformly reproducible or permanent. They are not suitable, therefore, for industries in which compatibility of components depends upon an exact match with prescriptive specifications that do not change. On the other hand biomaterials may be the appropriate choice for responsive, Ôfuzzy logicÕ designs where precision is not called for and may indeed be ineffective in the longer term, leading to inevitable deterioration of performance. In fact it might be appropriate to question how much longer precision engineering, with its attendant high production and maintenance costs and lack of margin for error, can continue to hold sway as understanding of and demand for dynamically responsive systems grows.

 

 

 

                        Bioproduction

 

                        As I have said, the great thing about biological systems is that given adequate nurture, they grow. All we have to do is ensure that they get what they need and they will elaborate a wondrous array of physical and chemical forms. All that creative potential, all that sophisticated wizardry of molecular, cellular and community structure is at our fingertips, without us having to make or assemble any of it! All we have to do is learn how to apply this creative potential to our own needs. Ah, there comes the rub! We have to understand the relationship between their needs and ours and between what they can do and what we can do. To begin with we need to know our selves and their selves from inside-out and from outside-in. Without such knowledge, without such understanding, our relationship is liable to be superficial, unproductive and abusive. Indeed, that is how our current relationship may stand – a long way short of fulfilling its potential.

 

                        The pharmaceuticals industry is a case in point. Following upon the long tradition of herbal remedies for ailments, the discovery of penicillin triggered an enhanced appreciation of the biosynthetic power of organisms and how this power might be harnessed for mass production. Natural product discovery became the order of the day, and much effort was invested in devising the best methods for large-scale cultivation of producer microorganisms in particular, culminating in the design of complex, submerged liquid ÔfermentersÕ. The latter are, in effect, large, stirred tanks containing growth medium in which conditions of aeration, nutrient supply, mineral ion content etc are precisely monitored and regulated in order to optimize production.

 

                        At first all seemed to be very well, with the success of the natural product discovery and production systems contributing in no small measure to the expansion of some pharmaceuticals companies into the multinational organizations that they are today. New products and new producer organisms were regularly discovered and cultivated.

 

                        Nowadays, however, the future for biological production of pharmaceuticals is seen by many as much more bleak and threatened by the quicker, more ÔpreciseÕ, more ÔcontrollableÕ methods of Ôrecombinant chemistryÕ and purely chemical manufacture. Organisms, if they are valued at all, are used more as ÔleadsÕ in the discovery of biologically active compounds than as agencies for production of these compounds. Faced with the vagaries of biological production, required to be competitive in the short term, disinclined to innovate or replace old plant with new plant, lacking a deep understanding of why, when and where organisms produce compounds and what to do about it, the industry becomes conservative. It falls back on what it thinks it already knows about.

 

                        This situation may partly have arisen because the methods for discovery and production that were at first so successful are not suitable for the vast majority of potential producer organisms. In fact these methods of Ôhigh throughput screeningÕ, whereby large numbers of candidates are tested over a short time scale, and submerged liquid fermentation, which favours rapid proliferation as dispersible units rather than interconnected systems, favours organisms with ephemeral traits. Little time or space is allowed for a candidate organism to develop and display its full range and repertoire. As in human societies dominated by short-term economics, Ôlate developersÕ are rejected before theyÕve had a chance. A huge potential like that below the exposed tip of an iceberg languishes untapped, beyond conscious apprehension. The importance of self-integrative processes and of dynamic contextual boundaries that both create and respond to heterogeneous conditions via a complex, free-radical chemistry dependent on the balance between oxygen and fuel supply, is overlooked. Until, perhaps, the Titanic ship of human short-termism meets its destiny and is forced either to think again or sink again!

 

                        But there are more problems for the pharmaceuticals industry than those of understanding the potential and needs of producer organisms. These additional problems relate to our understanding of our selves, of our own needs, and what unexpected repercussions and Ôside-effectsÕ might arise from incautious use of biologically active compounds. Bitter experience has made us wise after the event, forcing us to recognize that the seemingly incisive Ômagic bulletÕ of the chemically purified Ôwonder drugÕ might not be as precisely targeted within the complex, interconnected systems of our bodies as we might have expected. Moreover, the target can fight back through drug-resistance – in fact we encourage it to do so through the drug over-use that creates the space, the new context, for the innovative microorganism, virus or cancer cell to move in, liberated from competition with its neighbours. In effect the agent of disease brings about its own evolution by eliciting a human response which changes the context. This kind of repercussion, or co-evolutionary feedback, is in fact relevant to any human attempt to control a living, responsive system by biological or other means, and so needs to be borne very clearly in mind. The way to counter it is through cautious integration of a multiplicity of complementary approaches. Consciously or unconsciously, this has been, and may yet increasingly once again become the way of many empirically based remedies.

 

 

                        Biodegradation

 

                        As well as being productive, biological systems also have ways of being destructive, ultimately breaking down even the most elaborate physical and chemical structures into small molecules. This destructive power is often regarded as a problem when it affects materials of practical value to people. These materials include the food we eat, the fabrics we wear, the structures we house ourselves in, the glass we see through, the machinery that we equip ourselves with and the fuel and lubricants that power and service that machinery. They also include the cosmetics that we make ourselves up with and the medicines we treat ourselves with. In fact, given appropriate conditions of moisture, temperature, aeration and nutrient supply, just about anything we use can be rendered useless by other life forms, and the economic losses resulting from such ÔbiodeteriorationÕ are enormous. The best way of minimizing this deterioration is by prevention, through understanding the needs of the causal organisms and not allowing these needs to be met: for example if we donÕt want timber to decay, we keep it dry or non-aerated.

 

                        This very same destructive power of living systems that can cause such losses when allowed to occur in an inappropriate context is, however, vital to the sustainability and rejuvenation of natural ecosystems and to our own efforts to dispose of, remediate or recycle waste or hazardous materials. Such beneficial application is termed ÔbiodegradationÕ, and, having only recently developed much environmental concern, we are no doubt at the bottom of a very steep learning curve as to how to make best use of it beyond keeping a compost heap in our back yards. As ever, the aim should be to understand as much as possible about the context in which the needs and potential of the degradative systems can be met. Then it may be possible to develop new or improved approaches to fertilizing soils, reducing pollution damage, revitalizing water, producing foods and medicines etc.

 

 

Management Based Upon Biological Processes

 

As well as being agents and agencies for the direct application of management principles, biological systems can be valuable models upon which to base management strategies. Look carefully enough and it is generally possible to find a biological precedent for just about any human discovery or invention. Examples of such precedents range from the sonar equipment of a bat or dolphin to the magnetic compass of a migrating bird or bacterium, the microscopic hearing aid of a parasitic fly and the genetic manipulation of its host by a crown gall bacterium. It doesnÕt take much wit to appreciate the likelihood that there could be many more innovative ways of doing things out there to be copied from the living world. That is, there could be if only we knew how to look for them and recognise them when we see them rather than only with the benefit of hindsight. Perhaps a good place to begin is in a state of awareness of the problem-solving, opportunity-finding capacity of living systems, and thence to look to those systems for insight whenever we encounter a problem or opportunity.

 

                        But first, a word of caution may be necessary. It is widely considered, as a powerful argument in favour of biomimetics and as a by-product of neo-Darwinian adaptational thinking, that the solutions to problems found by living systems are ÔoptimalÕ, i.e. the best possible product of cost-benefit analysis. If that were so, however, life would have stopped evolving significantly long ago. But it hasnÕt. Life continues to change and to be changed by its dynamic context, and its future is contingent upon its past and present condition. It works within the constraints and facilities of the watery medium in which it is expressed. It is impeded and refined by the short-term competitive pressure that is intolerant of any departure from currently successful modes of operation. Only when this pressure is released, often via some kind of co-evolutionary effect, can a new innovative phase commence. So, in looking to life for insights into how to do whatÕs ÔbestÕ, itÕs important to realize that this ÔbestÕ may only be ÔbestÕ in the context of a specific set of boundary properties that may change. If this context-dependence is not understood, there is a danger that our search may be limited to specific, Ôright or wrongÕ applications closed off from the possibilities embedded in the indeterminacy of living systems. Indeed it may be that it is this indeterminacy and resultant capacity to bring about and cater for change that might be most opportune for us to emulate.

 

 

 

 

                        Design for Responsiveness and Resilience

 

                        By emulating the capacity of life forms to vary their boundary properties of deformability, permeability and continuity according to circumstances it may be possible to increase our ability to design versatile, resilient systems that are not rendered dysfunctional or outmoded by changes in conditions.

 

 

                        Design for Innovation, Renovation and Efficiency

 

                        By incorporating self-integrative processes, it may be possible to produce creative designs with capacities for learning, recall and efficient switchover from dissipative, assimilative structures to energy-conserving distributive and redistributive structures.

 

 

                        Design for Decommissioning

 

                        By emulating the ways in which living systems degenerate and reconfigure we can design structures that donÕt store up disposal problems for the future

 

                        Life-cycle analysis

 

                        By taking account of all the energetic demands of a design throughout the dynamic trajectory from its inception to its decommissioning, rather than at a snapshot in time, a more inclusional picture of its environmental impact can emerge. This kind of analysis is already practised, though it is limited conceptually through not being done in the context of dynamically bounded systems.

 

 

 

Additional Reference Sources

 

History of Human-Environment Relationships

 

Ponting, C. (1991). A Green History of the World.

 

Diamond, J. (1997). Guns, Germs and Steel.

 

J. Bate (2001). The Song of the Earth, London: Picador

 

 

Biodiversity

 

Capra, F. (1996). The Web of Life.

 

Perlman, D.L. & Adelson, G. (1997). Biodiversity: Exploring Values and Priorities in Conservation.

 

Wilson, E.O. (1992). The Diversity of Life.

 

 

Environmental Philosophy

 

Kumar, S. (2002) You Are Therefore I Am – A Declaration of Dependence. Green Books.

 

Capra, F. (2002) The Hidden Connections

 

Bortoft, H. (1996). The Wholeness of Nature: GoetheÕs Way of Science.

 

Sessions, G. (1995). Deep Ecology for the 21st Century.

 

Leopold, A. (1966) A Sand County Almanac.

 

Orr, D. (1994). Earth in Mind: On Education, Environment and the Human Prospect.

 

Pojman, L.P. (1994) Environmental Ethics.

 

Roszak, T. et al. (1995). Ecopsychology.

 

J. Blatter and H. Ingram (2001). Reflections on Water – New Approaches to Transboundary Conflicts and Co-operation, MIT Press.

 

D. Boyle (2000). The Tyranny of Numbers – Why Counting CanÕt Make Us Happy, London: Harper Collins.

 

D. Rothenberg and M. Ulvaeus (2001). Writing on Water – A Terra Nova Book, MIT Press.

 

Landry, C (2000) The Creative City. Comedia, Earthscan.

 

 

Websites

 

http://people.bath.ac.uk/bssadmr This is my personal homepage where you can find numerous artworks and recent essays

 

www.goodshare.org

www.i-sis.org

 

 

 

Alan Rayner

January 2005