Go to the Nature Institute’s home page

Organisms and Their Evolution
A book by
Stephen L. Talbott


Organisms and Their Evolution
Agency and Meaning in the Drama of Life

Download an informal, preliminary PDF of the entire book

This is a complete draft of an online book freely available, with proper attribution, for noncommercial, personal use (including classroom use). The above title is the final title for the book, which previously bore the temporary title, “Evolution As It Was Meant To Be — and the Living Narratives That Tell Its Story”. This material is part of the Biology Worthy of Life project of The Nature Institute. Copyright 2023 by Stephen L. Talbott. All rights reserved. Date of the current draft: April 7, 2023. Html files for individual chapters (accessible from this page) may be more recently revised than the whole-book pdf.

The book will remain for the foreseeable future a “living document”, which is to say that it will be subject to change. Some of those changes may result from your comments, if you are so kind as to send those comments to Stephen L. Talbott (stevet@netfuture.org).

Table of Contents

Preface

Part 1: The Life of Organisms

Chapter 1: The Keys to This Book

Read about this chapter (+/-)

This introductory chapter offers a brief summary of the central themes of the book: 1) every animal comprises more than a set of physical and chemical interactions, but possesses an agency through which it weaves a life story; 2) more generally, every animal’s life narrative is an outward expression of interior meaning; 3) The meaningful, narrative character of life demands a holistic style of understanding and explanation; and 4) biologists suffer from a kind of blindsight through which they are unable to incorporate much of what they actually know about organisms into their scientific explanations.

(show less)

Chapter 2: The Organism’s Story

Read the chapter introduction (+/-)

Organisms are purposive (“teleological”) beings. Nothing could be more obvious. The fact of the matter is so indisputable that even those who don’t believe it really do believe it. Philosopher of biology Robert Arp speaks for biology as a whole when he writes,

Thinkers cannot seem to get around [evolutionary biologist Robert] Trivers’ claim that “even the humblest creature, say, a virus, appears organized to do something; it acts as if it is trying to achieve some purpose”, or [political philosopher Larry] Arnhart’s observation that … “Reproduction, growth, feeding, healing, courtship, parental care for the young — these and many other activities of organisms are goal-directed”.

And yet, despite his acknowledgment that we “cannot get around” this truth, Arp again speaks for almost the entire discipline of biology when he tries, with some delicacy, to take it all back: “with respect to organisms, it is useful to think as if these entities have traits and processes that function in goal-directed ways” (his emphasis). This as if is a long-running cliché, designed to warn us that the organism’s purposive behavior is somehow deceptive — not quite what it seems. The goal-directedness is, in the conventional terminology, merely apparent or illusory. Certainly it must not be seen as having any relation at all to human purposive activity — an odd insistence given how eager so many biologists are to make sure we never forget that the human being is “just another animal”.

Others have commented on this strange reluctance to acknowledge fully the purposiveness that is there for all to see. The philosopher of science, Karl Popper, said that “The fear of using teleological terms reminds me of the Victorian fear of speaking about sex”. Popper may have had in mind a famous remark by his friend and twentieth-century British evolutionary theorist, J. B. S. Haldane, who once quipped that “Teleology is like a mistress to a biologist; he cannot live without her but he’s unwilling to be seen with her in public”.

We find — and will later explore further — this same unwilling yet inescapable conviction of purposiveness at the foundations of evolutionary theory. The theory, we are often told, is supposed to explain away the organism’s purposes — “naturalize” them, as those who claim to speak for nature like to say. But at the same time the theory is itself said to be grounded solidly in the fact that organisms, unlike rocks, thunderstorms, and solar systems, struggle to survive and reproduce. If they did not spend their entire lives striving toward an end, or telos, in this way, natural selection of the fittest organisms (those best qualified to survive and reproduce) could not occur. So it is not at all clear how selection is supposed to explain the origin of such end-directed behavior.

This double stance — believing and not believing, acknowledging and explaining away — constitutes, you could almost say, the warp and woof of biology itself. Look for “purpose” in the index of any biological textbook, and you will almost certainly be disappointed. That term, along with others such as “meaning” and “value”, is effectively banned. There is something like a taboo against it. Yet, in striking self-contradiction, those textbooks are themselves structured according to the purposive activities, or tasks, of organisms. Biologists are always working to narrate goal-directed achievements. How is DNA replicated? How do cells divide? How does metabolism supply energy for living activity? How are circadian rhythms established and maintained? How do animals arrive at the evolutionary strategies or games or arms races through which they try to eat and avoid being eaten?

Such questions are endless, and their defining role is reflected on every page of every textbook on development, physiology or evolution. A research question is biological, as opposed to physical or chemical, only when it is posed in one way or another by the organism’s purposive, future-oriented activity. The puzzle is that the answers biologists are willing to offer, on the other hand, are rooted with equal consistency in the assumption that organisms have no purposes. The reigning conviction is that explanations of physical and chemical means effectively remove any need to deal scientifically with the ends that alone could have prompted our search for means in the first place.

My larger argument in this book will be that this conviction about the adequacy of physical and chemical descriptions is misbegotten, with devastating effects upon many fields of biological understanding, and particularly evolutionary theory. It hardly needs emphasizing that if organisms really are purposive beings — if the fact of purposive activity is not an illusion — then a biological science so repulsed by the idea of purpose that its practitioners must avert their eyes at the very mention of it … well, it appears that these practitioners must feel threatened at a place they consider foundational. And with some justification, for admit to what they actually know about organisms would be to turn upside down and inside out much of the science to which they have committed their lives.

“Purpose” — an idea that will need careful qualification in different biological contexts — gives us but one of several intimately related avenues of approach to what is distinctive about the life of organisms. In the remainder of this chapter I will briefly sketch a few of these avenues.

(show less)

Chapter 3: What Brings Our Genome Alive?

Read the chapter introduction (+/-)

Throughout most of the twentieth century, genes were viewed as the “agents” responsible for an organism’s development, activity, and evolution. Their agency was said to lie in their ability to “regulate”, “organize”, “coordinate”, and “control” physiological processes. DNA, the bearer of these genes, became the “Book of Life” — the essential maker of organisms and driver of evolution. And this view remains stubbornly entrenched today, despite many changes in our understanding. A leading behavioral geneticist has recently written a book entitled, Blueprint: How DNA Makes Us Who We Are.

Nevertheless, the idea that genes are the decisive “first causes” of life — and, more generally, that molecules at the “bottom” ultimately explain everything that happens at larger scales — has come in for a great deal of criticism in recent years. This criticism, as we will see, is fully justified. But the issues can be subtle, as is suggested by an apparent paradox. Philosopher of biology Lenny Moss, who wrote the valuable book, What Genes Can’t Do, has remarked:

“Where molecular biology once taught us that life is more about the interplay of molecules than we might have previously imagined, molecular biology is now beginning to reveal the extent to which macromolecules [such as DNA], with their surprisingly flexible and adaptive complex behavior, turn out to be more life-like than we had previously imagined.”

In a similar vein, I myself wrote a decade ago:

Having plunged headlong toward the micro and molecular in their drive to reduce the living to the inanimate, biologists now find unapologetic life staring back at them from every chromatogram, every electron micrograph, every gene expression profile. Things do not become simpler, less organic, less animate. The explanatory task at the bottom is essentially the same as what we faced higher up.

But if all this is true, what are we to make of Harvard geneticist Richard Lewontin’s declaration, itself hardly disputable, that “DNA is a dead molecule, among the most nonreactive, chemically inert molecules in the living world. That is why it can be recovered in good enough shape to determine its sequence from mummies, from mastodons frozen tens of thousands of years ago, and even, under the right circumstances, from twenty-million-year-old fossil plants … DNA has no power to reproduce itself. Rather it is produced out of elementary materials by a complex cellular machinery of proteins. While it is often said that DNA produces proteins, in fact proteins (enzymes) produce DNA … Not only is DNA incapable of making copies of itself, aided or unaided, but it is incapable of ‘making’ anything else.”

Many astute observers have echoed Lewontin’s remarks, and I have never seen anyone question them, including those who remain enamored of the “Book of Life”. So which is it? When we peer at DNA, do we see a dead molecule or the secret of life? As it happens, there is a simple answer: if we are looking at a molecule conceived in the usual way as a bit of mindless, inherently inert stuff, then, according to our own conceptions, we see only dead stuff. But if we observe the molecule as a system of forces and energies capable of participating and being caught up in the creative life of the cell and organism, then we can hardly help recognizing — and perhaps even reverencing — the living performance unfolding before our eyes.

Saying this is one thing; making it both meaningful and profound is quite another — and that is one task of the present book. So let us begin.

(show less)

Chapter 4: The Sensitive, Dynamic Cell

Read the chapter introduction (+/-)

Throughout a good part of the twentieth century, cell biologists battled over the question, “Which exerts greater control over the life of the cell — the cell nucleus or the cytoplasm?” From mid-century onward, however, the badge of imperial authority was, by enthusiastic consensus, awarded to the nucleus, and especially to the genes and DNA within it. “Genes make proteins, and proteins make us” — this has been the governing motto, despite both halves of the statement being false (which will become ever clearer as we proceed).

The question for our own day is, “Why would anyone think — as the majority of biologists still do — that any part of a cell must possess executive control over all the other parts?” We have already caught our first glimpse of the performances in the nucleus (see Chapter 3), and these hardly testify to domination by a single, controlling agent. Now we will broaden our outlook by making a first approach to the rest of the cell — the cytoplasm, along with its organelles and enclosing membrane.

It would be well to remind ourselves before we proceed, however, that, whatever else it may be, an organism is a physical being. Its doings are always in one way or another physical doings. This may seem a strange point to need emphasizing at a time when science is wedded to materialism. And yet, for the better part of the past century problems relating to the material coordination of biological activity were largely ignored while biologists stared, transfixed, into the cell nucleus. If they concentrated hard enough, they could begin to hear the siren call of a de-materialized, one-dimensional, informational view of life.

The idea of a genetic code and program proved compelling, even though the program was never found and the supposedly fixed code was continually rewritten by the cell in every phase of its activity. So long as one lay under the spell woven by notions of causally effective information and code, problems of material causation somehow disappeared from view, or seemed unimportant. And so, freed from “mere” material constraint, programmatic Information became rather like the Designer of the intelligent design advocates.

Surely, even if they are not the decisive causes usually imagined, genes do connect in some manner with the features they were thought one-sidedly to explain. But this just as surely means they must connect physically and meaningfully, via movements and transformations of substance testifying to an underlying narrative (Chapter 2) — not merely logically, through the genetic encoding of an imagined program. And what we saw in Chapter 3 about the significant movements and gesturings of chromosomes is only the beginning of the story.

(show less)

Chapter 5: Our Bodies Are Formed Streams

Read the chapter introduction (+/-)

In this materialist era, we like our reality hard and our truths weighty and rock solid. We may accept that there are states of matter less substantial than rocks, but in our imaginations we turn even fluids and gases into collections of tiny particles more or less closely bound together. Similarly, in our reconstructions of physiological processes, material structures come first, and only then can movement, flow, and meaningful activity somehow occur.

How, after all, can there be movement without things to do the moving? (It’s easy to forget that energy, fields, and forces are not things!) Ask someone to describe the circulatory system, and you will very likely hear a great deal about the heart, arteries, veins, capillaries, red blood cells, and all the rest, but little or nothing about the endless subtleties of circulatory movement through which, for example, the structured heart first comes into being.

Yet there is no escaping the fact that we begin our lives in a thoroughly fluid and plastic condition. Only with time do relatively solid and stable structures precipitate out as tentatively formed “islands” within the streaming rivers of cells that shape the life of the early embryo. As adults, we are still about seventy percent water.

One might think quite differently based on the scientific rhetoric to which we are daily exposed. This could easily lead us to believe that the real essence and solid foundation of our lives was from the beginning rigidly established inside those very first cells. There we find DNA macromolecules that, in a ceaseless flood of images, are presented to us as crystalline forms in the shape of a spiraling ladder — a ladder whose countless rungs constitute the fateful stairway of our lives. So, too, with the all-important proteins and protein complexes of our bodies: we have been told for decades that they fold precisely into wondrously efficient molecular machines whose all-important functions are predestined by the DNA sequence.

The trouble is, biological researches of the last few decades have not merely hinted at an altogether different story; they have (albeit sometimes to deaf ears) been trumpeting it aloud as a theme with a thousand variations. Even the supposedly “solid” structures and molecular complexes in our cells — including the ones we have imagined as strict determinants of our lives — are caught up in functionally significant movement that the structures themselves can hardly have originated. (See Chapter 3, “What Brings Our Genome Alive?”, and Chapter 4, “The Sensitive, Dynamic Cell”.)

Nowhere are we looking either at a static sculpture or at controlling molecules responsible for the sculpting. In an article in Nature following the completion of the Human Genome Project, Helen Pearson interviewed many geneticists in order to assemble the emerging picture of DNA. One research group, she reported, has shown that the molecule is made “to gyrate like a demonic dancer”. Others point out how chromosomes “form fleeting liaisons with proteins, jiggle around impatiently and shoot out exploratory arms”. Phrases such as “endless acrobatics”, “subcellular waltz”, and DNA that “twirls in time and space” are strewn through the article. “The word ‘static’ is disappearing from our vocabulary”, remarks cell biologist and geneticist Tom Misteli, a Distinguished Investigator at the National Cancer Institute in Bethesda, Maryland.

Everywhere we look, shifting form and movement show themselves to be the “substance” of biological activity. The physiological narratives of our lives play out in gestural dramas that explain the origin and significance of structures rather than being explained by those structures.

Hannah Landecker, a professor of both genetics and sociology at UCLA, having looked at the impact of recent, highly sophisticated cellular imaging techniques on our understanding, has written: “The depicted cell seems a kind of endlessly dynamic molecular sea, where even those ‘structures’ elaborated by a century of biochemical analysis are constantly being broken down and resynthesized.” And she adds: “It is not so much that the structures begin to move, but movements — for example in the assembly and self-organization of the cytoskeleton — begin to constitute structure”.

And in a paper that appeared as I was writing this chapter, a team of biochemists from Duke and Stanford Universities point out how inadequate is our knowledge of the action of biomolecules when all we have is a frozen structure of the sort commonly reported in the literature. “In reality”, they say, “all macromolecules dynamically alternate between conformational states [that is, between three-dimensional folded shapes] to carry out their biological functions”:

Decades ago, it was realized that the structures of biomolecules are better described as “screaming and kicking”, constantly undergoing motions on timescales spanning twelve orders of magnitude, from picoseconds [trillionths of a second] to seconds.

Why, after all, should we ever have expected our physiology to be less a matter of gesturings than is our life as a whole?

(show less)

Chapter 6: Context: Dare We Call It Holism?

Read the chapter introduction (+/-)

The centrality of living wholes within biology seems beyond argument. These have not been “put together” or built by an external agency. They are never the results of a physical activity that starts with non-wholes. Biology gives us nothing but beings that have never existed except as wholes possesssing the formative powers that enable them to pass through further stages of physical development.

The one-celled zygote is already a functioning whole. It does not gain further cells through the addition of “building blocks” assembled by an engineer or designer, but rather through an internal power of reorganization and subdivision in which the entire organism participates. All the parts are orchestrated in a unified performance that yields new cells, and particular kinds of cells, just where they are needed. The orchestrating power of the whole can hardly be determined by the particular parts it is bringing into being and orchestrating.

Where the physicist may prefer unambiguous, isolated, and well-defined “point” causes, the biologist never has such causes to theorize about. A biological whole is never absolute, and never perfectly definable as distinct from its environment. Further, its actions are always multivalent, like the meaning of a sentence in a profound and complex text. Its activities interpenetrate one another, like the events of a story.

The wonderfully insightful twentieth-century botanist, Agnes Arber, captured well the polar tension between organic wholeness, on one hand, and contextual embeddedness, on the other:

The biological explanation of a phenomenon is the discovery of its own intrinsic place in a nexus of relations, extending indefinitely in all directions. To explain it is to see it simultaneously in its full individuality (as a whole in itself), and in its subordinate position (as one element in a larger whole.

Every ecological setting, every organism within that setting, every organ within the organism, and every cell within the organ is a whole providing a context for its own interrelated parts, and at the same time is itself contextually embedded within larger wholes. “Context”, “whole”, and “part” can never be rigid, absolute terms in biology. They are bound up with interweaving spheres of activity.

We need to gain some practice in thinking, not with the single, distinct point-causes of the physicist (or at least the classically minded physicist), but rather with the actual narrative qualities of biological activity. The perplexing issues surrounding attempts at holistic thought may thereby lend themselves more easily to our efforts at understanding.

(show less)

Chapter 7: Epigenetics: A Brief Introduction

Read the chapter introduction (+/-)

You and I harbor trillions of “sub-creatures” in our bodies. I am not referring to the microorganisms in our guts, but rather the cells we consider our own — the constituents of our muscles and brains, our livers and bones, our lenses and retinas. Each of these cells, embedded in its supportive environment, sustains a dauntingly complex and unique way of life. If we had first discovered such cells floating singly in a pool of water and had observed them through a microscope, we would have judged them to be distantly related organisms. Phenotypically (that is, in visible form and function) one cell type can differ from another as much as an amoeba differs from a paramecium.

All the cells in the human body have descended from a single cell (zygote) with a single genome. And just as hundreds of different cell types have arisen from that one zygote, so, too, have the multicellular, intricately organized entities we know as lung, heart, eye, kidney, and pancreas, along with all our other organs. Supremely interdependent as these are, each is nevertheless a functioning organic world of altogether distinctive character.

For the past century these facts of development have been thought to present a (largely ignored) problem for the gene-centered view of life. The developmental biologist Frank Lillie, who had directed the prestigious Marine Biological Laboratory at Woods Hole, Massachusetts, and would go on to become president of the National Academy of Sciences, remarked in 1927 on the contrast between “genes which remain the same throughout the life history” of an organism, and a developmental process that “never stands still from germ to old age”. In his view, “those who desire to make genetics the basis of physiology of development will have to explain how an unchanging complex can direct the course of an ordered developmental stream.”

This ordered developmental stream, of course, includes generation of the hundreds of different cell types in our bodies. It is hard to understand how a single genomic “blueprint” — or any other way of construing a fixed genetic sequence — could by itself provide the definitive causal basis for these hundreds of radically distinct ways of living. If the blueprint is compatible with all of them, do we have compelling grounds for thinking that it fundamentally determines any one type of cell, or organ, let alone all of them together? One might reasonably expect that other factors direct the developmental process toward particular outcomes of such different sorts.

A more balanced understanding arises when we watch how every cell displays its character through its life as a whole. That character, in all its qualitative richness, somehow seems decisive. DNA is caught up in a seamless and integral way of being. When we grasp this integral nature, we quickly realize that the idea of DNA as the crucial causal determinant of the whole is an impossible one. As a specific kind of liver cell passes through its developmental lineage, it must sustain its entire organization in a coherent and well-directed manner from one cell generation to the next — including, for example, the cytoskeletal and cell membrane organization described in Chapter 4. It must also bring about and orchestrate the elaborate performances of its chromosomes we saw in Chapter 3 — performances that are unique to this type of cell and that chromosomes themselves have no way to set in motion.

Every individual part, including DNA, is shaped by, and gives expression to, the character of a larger whole. As functional participants in diverse physiological processes, our genes do not in fact “remain the same throughout life”. They, like all parts of a cell or organism, gain their identity and meaning only within the context of innumerable, interpenetrating, living narratives (Chapter 2).

(show less)

Chapter 8: The Mystery of an Unexpected Coherence

Read the chapter introduction (+/-)

We heard in “The Organism’s Story” that living activity has a certain future-oriented (“purposive” or “intentional” or end-directed) character that is missed by causal explanations of the usual physical and chemical sort. This is true whether the end being sought is the perfection of adult form through development, or the taking of a prey animal for food.

An animal’s end-directed activity may, of course, be very far from what we humans know as conscious aiming at a goal. But all such activity nevertheless displays certain common features distinguishing it from inanimate proceedings: it tends to be persistent, so that it is resumed again and again after being blocked; it likewise tends to be adaptable, changing strategy in the face of altered circumstances; and the entire activity ceases once the end is achieved.

This flexible directedness — this interwoven play of diverse ends and means within an overall living unity — is what gives the organism’s life its peculiar sort of multi-threaded, narrative coherence. Life becomes a story. Events occur, not merely from physical necessity, but because they hold significance for an organism whose life is a distinctive pattern of significances.

The idea of narrative coherence, like the related idea of a governing context (Chapter 6), is a mystery for all attempts at purely physical explanation. This is why even the explicit acknowledgment of an organism’s striving for life — central as it may be for evolutionary theory — is discouraged whenever biologists are describing organisms themselves. It sounds too much as if one were invoking inner, or soul, qualities rather than material causes — acknowledging a being rather than a thing. And it is true that our physical laws, however combined, nowhere touch the idea of striving.

Biologists much prefer to identify single, definitive causes. The cell nucleus with its genome has long been viewed as the seat of such causation. But, as we saw in our discussion of epigenetics, the single-minded pursuit of genetic causes has forcibly redirected our attention to epigenetics, where we have discovered that genes are circumscribed and given their meaning by the narrative life of the entire cell and organism.

In what follows below we will consider this narrative coherence in a more detailed way — first, in relation to one of the many activities of the cell that can be considered under the heading of “epigenetics”. Then we will look more briefly at a startling phenomenon that, already on its face, renders absurd the idea of central genetic control. In both cases we will be focused on molecular-level activity, which is precisely where we have been most strictly trained to expect the absence of any coherence other than that of “blind mechanism”.

(show less)

Chapter 9: A Mess of Causes

Read the chapter introduction (+/-)

The difficulties in talking about causes in biology have been recognized for at least two centuries. It’s just that the issues were largely set aside in the era of molecular biology due to the expectation that our rapidly growing powers of minute analysis would bring full causal understanding. Biology would soon be rid of its troublesome language of life in favor of well-behaved molecular mechanisms. And yet today, after several decades of stunning progress in molecular research, the struggle to fit our understanding of living activity into the comfortable garb of familiar causal explanation looks more hopeless than ever.

On one hand, most biologists seem unaware that there is a problem here — or, at least, they are unwilling to betray their awareness in professional circles. On the other hand, their scientific descriptions could hardly signal more dramatically the failure of the usual causal explanations. We seem to be looking here at another illustration of blindsight (Chapter 2).

In a previous chapter we considered epigenetics, which is commonly taken to be about the way epigenetic “marks” on chromosomes alter gene expression. But no sooner did epigenetics gain biologists’ attention than researchers began puzzling over the question, “Do epigenetic marks alter gene expression, or do changes in gene expression alter the marks?” And (as we will see in Box 9.1), the question is still with us. According to Luca Magnani, a cancer researcher at Imperial College London,

It’s an absolutely legitimate question and we need to address it. The answer is either going to kill the field [of epigenetics], or make it very important.

“Either kill the field or make it very important”. The comment expresses absolute confidence that we can discover unambiguous causation, which will in turn settle the matter: either epigenetic changes cause gene activity (in which case they are very important), or they are mere effects of that activity, with little significance. It must be one way or the other. The general idea is that, if something is to contribute to scientific understanding, it must be the indisputable cause of an indisputable effect. And yet, as we will now see, this stubborn insistence on causal clarity continually prods biologists to offer embarassingly incoherent explanations.

(show less)

Chapter 10: What Is the Problem of Form?

Read part of the chapter introduction (+/-)

It is well known that amphibians such as frogs and salamanders have a remarkable ability to regenerate severed limbs. What may not be so commonly realized is that, if you graft the tail bud of a salamander onto the flank of a frog tadpole at the place where a limb would normally form — and also near the time when metamorphosis of the tadpole into a frog will occur — the grafted organ first grows into a salamander-like tail, and then, in some cases, more or less completely transforms into a limb, albeit a dysfunctional one. Among other changes, the tip of the tail turns into a set of fingers.

The experiment can remind us, in passing, how biologists commonly try to learn about life by severely disrupting it. But the current thing to note is that, in this particular experiment, the transformation of the tail into an approximate limb cannot be the result of local causes, since the local environment of the fingers-to-be is a tail, not a limb. The power of transformation is, in a puzzling manner, holistic. The part is caught up within the whole and moves toward its new identity based, not merely on local determinants, but also on the form and character of a whole that is not yet physically all there.

In a rather different vein, Harvard biologist Richard Lewontin once described how you can excise the developing limb bud from an amphibian embryo, shake the cells loose from each other, allow them to reaggregate into a random lump, and then replace the lump in the embryo. A normal leg develops. Somehow the currently unrealized form of the limb as a whole is the ruling factor, redefining the parts according to the larger, developing pattern. Lewontin went on to remark:

Unlike a machine whose totality is created by the juxtaposition of bits and pieces with different functions and properties, the bits and pieces of a developing organism seem to come into existence as a consequence of their spatial position at critical moments in the embryo’s development.

But how can this be? How can spatial position within a not yet fully realized form physically determine not only the future and proper sculpting of that form, but also the identity of its parts?

(show less)

Chapter 11: Why We Cannot Explain the Form of Organisms

Read part of the chapter introduction (+/-)

Questions of form have seemed oddly resistant to the biologist’s quest for explanation. Darwin himself seemed to sense the difficulty in that famous instance where he recoiled from contemplating the subtle perfections in the form of the eye: “To suppose that the eye with all its inimitable contrivances for adjusting the focus to different distances, for admitting different amounts of light, and for the correction of spherical and chromatic aberration, could have been formed by natural selection, seems, I freely confess, absurd in the highest degree”.

Of course, as Darwin quickly added, his theory convinced him that he was merely suffering from a lack of imagination. All that was really needed were the creative powers of natural selection acting through eons upon an endless supply of small, helpful changes. But his underlying malaise was not so easily vanquished: “It is curious”, he wrote to the American botanist Asa Gray in the year following publication of the Origin, “that I remember well [the] time when the thought of the eye made me cold all over, but I have got over this stage of the complaint, and now small trifling particulars of structure often make me very uncomfortable. The sight of a feather in a peacock’s tail, whenever I gaze at it, makes me sick!”

We can assume that Darwin got over that stage of the complaint as well. But, thankfully, the biologist is still now and then allowed, if not a complaint, at least an honest expression of wonder.

(show less)

Chapter 12: Is a Qualitative Biology Possible?

Read the chapter introduction (+/-)

The philosopher, Ronald Brady, once wrote about his undergraduate experience this way:

When I began college as a chemistry major my enthusiasm for science was somewhat dampened by meeting a professor of chemistry who pointed out the difference between my own goals and those he, as an experienced professional, would call mature. My passion, he noted, was entirely focused on direct experience — my sense of chemical change was invested in sensible qualities: in smells, colors, the effervescence of liquids, the appearance of precipitates, the light and violence of flame, etc. But, he countered, this was probably closer to medieval alchemy than to chemistry. The latter is really a matter of molecular and atomic events of which we can have only a theoretical grasp, and the sensible experience on which my excitement centered was secondary ... I was reminded of him when I spoke to a morphologist at Berkeley about my interest in Goethe’s attempt to approach science by keeping to direct experience. The morphologist responded: “You are interested in this approach because you are a Nature appreciator, while I am a productive scientist.” It is always nice to see where one stands.

Ever since the Scientific Revolution, physical scientists have held to the conviction that, whereas nature speaks decisively in the language of mathematics, the qualities of nature are not actually qualities of nature, but rather additions provided “from outside” by human subjectivity. And where physical scientists have led, biologists have done their best to follow.

If, as is commonly thought, qualities reside outside the bounds of any rigorous science, including biological science, then the very idea of a qualitative biology is self-contradictory. There can be no such science. Since this entire book is founded on the contrary assumption — an assumption explicitly defended in Chapters 13 and 24 — it feels obligatory to provide some examples of what a qualitative biology might look like.

In what follows I offer three such examples of widely differing sorts. The first involves the study of a single animal, the second a study of leaf sequences on certain plants, and the third a study of systemic morphological, behavioral, and other patterns recognizable in evolved groups of organisms, yet inexplicable in terms of present evolutionary theory.

(show less)

Chapter 13: All Science Must Be Rooted in Experience

Read the chapter introduction (+/-)

In previous chapters we have seen how organisms, as centered agents, present us with rich, narrative contexts — mortal performances that proceed, with characteristic expressiveness and intention, through the stages of a life drama unique to a particular species. And yet, as we have also seen, a powerful urge drives biologists to ignore, as far as they can, every living feature of those performances.

They ignore, for example, what it must really mean when they say that animals “strive” to maintain their life, or that a wound “heals” itself, or that an organism “adapts” to its environment, or that it “perceives” a threat and “responds” to it. (Physical objects in general — stones, clouds, and dust storms — do not strive, heal, adapt, perceive, or respond.) But it is all too easy for any scientist to side-step such meanings and analyze the organism’s story into lifeless sequences of precisely lawful molecular interactions. And since there appear to be no gaps in the molecular-level picture, the resulting explanations seem complete. Only the organism is missing.

In other words, seamless as they may be in their own impoverished terms, such explanations are not in fact complete. They miss the simply observed fact that molecular-level interactions in an organism are always caught up in, and governed by, the higher-level pattern of a life story. We always find ourselves watching the meaningful coordination of causal processes in an extended narrative — an end-directed coordination that cannot be explained by the processes being coordinated. This is why explanations that never move beyond physics and chemistry stop short of biology.

Non-living explanations do, however, have one advantage: they conveniently avoid all those troublesome words I use throughout this book in discussing organic contexts and life stories — words such as intention and purposiveness, idea and thought, agency and end-directedness, interests and meaning. Most biologists prefer to have nothing to do with such terms.

One problem with those words is that they evoke features of our own inner lives — our human experience. It is, of course, healthy to avoid an anthropomorphic projection of human experience upon other organisms, where it does not belong. But we, too, are organisms, and we have good reason to ask: Where does living human awareness belong in our biological science? If we ignore the character of our own life and experience, can we fully understand a world that has been capable of producing us? Where can we gain our scientific ideas, if they are not empirical — if they are not expressions of our most rigorously considered experience? And can we reasonably assume that our own experience has nothing at all in common with that of our evolutionary forebears?

(show less)

Technical Supplement to Part 1

Chapter 14: How Our Genes Come to Expression (It Takes an Epigenetic Village)

Read part of the chapter introduction (+/-)

If your understanding of genetics comes from your newspaper’s science section, or a popular science magazine, or any other source intended for the general public, then you will not have been given the remotest glimpse of what actually goes on with the genes in our bodies. In fact, geneticists themselves have been known to lament how limited their knowledge of gene-related activity is, simply because the demands of professional specialization scarcely allow a wide field of view.

But it turns out that a wide field of view is the one critical prerequisite for any adequate understanding of genes. Only a broad survey can illustrate how every gene, like a significant word in a text, receives its full meaning only through the interweaving and converging influences issuing from all the elements of its context.

My aim here is to offer such a wider, “epigenetic” view — and to do so in the briefest space possible. If I succeed, you will begin to sense a biological landscape that reconfigures many long-standing assumptions, not only about genetics itself, but also about the character of living processes.

(show less)

Chapter 15: Puzzles of the Microworld

Read from the chapter introduction (+/-)

Einstein, so it is said, was led to his theory of special relativity due in part to his having imagined what it would be like to “ride on a light beam”. Might we possibly discover equally strange things if we tried to imagine what it would be like to dwell within an individual living cell?

Unlike Einstein with his task, ours would be much simpler. It would not require bold new understandings in physics, but simply a willingness to imagine the changing play of already recognized physical laws at different dimensions. And, fortunately, we have at least one scientific paper, written thirty years ago, that has already done much of the work of imagining the startlingly different conditions of life at the scale of the cell.

That 1990 paper was written by Guenter Albrecht-Buehler of the Northwestern University Medical School in Chicago. He began his professional life as a physicist before moving into cell biology. However, unlike what you might expect of a physicist, one of his larger concerns was rooted in the conviction that we cannot build up an understanding of organisms by starting from the molecular level. His paper, titled “In Defense of ‘Nonmolecular’ Cell Biology”, has not, in my judgment, received the attention it deserves. The present article represents my effort to summarize only that part of the paper dealing with the wildly unexpected consequences of differences of scale, and then to offer a few additional comments of my own.

(show less)

Part 2: Extending the Organism’s Story:
Toward an Evolutionary Narrative

Chapter 16: Let’s Not Begin With Natural Selection

Read the chapter introduction (+/-)

Evolutionary theorists tend to become frustrated when many of the rest of us fail to “get” the revolutionary and convincing simplicity of natural selection, that primary engine of adaptive evolution also known as “the survival of the fittest”. For example, Niles Eldredge, a paleontologist and, for several decades, a curator at the Museum of Natural History, has wondered, “Why do physicists, who have the reputation of being among the best and the brightest, have such a hard time with the simple notion of natural selection? For simple it is”. He then quotes Charles Darwin:

As many more individuals of each species are born than can possibly survive; and as, consequently, there is a frequently recurring Struggle for Existence, it follows that any being, if it vary however slightly in any manner profitable to itself, under the complex and sometimes varying conditions of life, will have a better chance of surviving, and thus be naturally selected.

“The concept”, Eldredge writes, “is definitely simple enough. This description of natural selection may be a bit longer than the elegantly brief F=MA [force equals mass times acceleration — Newton’s second law of motion]. Conceptually, however, it is hardly more complicated.”

(show less)

Chapter 17: Evolution Writ Small

Read the chapter introduction (+/-)

One might think that the natural place to look for an understanding adequate to the evolutionary history of life would be the powers of self-transformation we observe in the evolving organisms themselves. But it can be dangerous to look in a clear-eyed manner at the creative potentials of living beings. One risks having to acknowledge the evident wisdom and agency so vividly on display. In an era of institutionalized materialism, any suggestion that these inner powers are vital to the entire evolutionary story can only produce the sort of discomfort associated with a taboo.

On the other hand, Stephen Jay Gould ran afoul of no taboo when he effectively ascribed this same wisdom and agency to natural selection. Countering the questions we heard voiced in Chapter 16 about what sort of creative principle could explain the “arrival of the fittest”, he asked (referring to several giants of twentieth-century evolutionary biology), “Why was natural selection compared to a composer by Dobzhansky; to a poet by Simpson; to a sculptor by Mayr; and to, of all people, Mr. Shakespeare by Julian Huxley?”

The answer, Gould said, is that the allusions to poetry, musical composition, and sculpture helpfully underscore the “creativity of natural selection”:

The essence of Darwinism lies in its claim that natural selection creates the fit. Variation is ubiquitous and random in direction. It supplies the raw material only. Natural selection directs the course of evolutionary change. It preserves favorable variants and builds fitness gradually.

On its face, Gould’s argument was a puzzling one. His answer to the question how creative variation arises amounted to saying nothing more than “It is everywhere” (“variation is ubiquitous”) — which, one might have thought, only added urgency to the need for an explanation. He seemed to suggest that, because organisms are so expert and prolific at producing new possibilities of life, the evolutionist can simply take their powers of achievement for granted. Because organisms so abundantly provide raw materials for creative work, we are somehow free to declare natural selection the agent performing this work. It need only preserve all those wonderfully effective new traits.

How easy it is, apparently, to forget that the so-called “raw materials” being preserved are never merely raw materials! At the first appearance of any beneficial change, the creative work has already been accomplished. We find ourselves looking, not at random raw materials, but at a viable feature harmoniously incorporated into all the tightly interwoven complexity of a living being. The only power we know to be capable of such incorporation is that of the organism telling its own story, a story always reflecting the qualitative, dynamic character and developmental potentials of a particular species.

This harmonious incorporation of new features, founded upon whole-cell inheritance and manifested in whole-organism processes of development, is the only place where we see creative evolutionary change originating. The spreading of an already-existing change through a population is not where we see evolutionary novelties arising.

So Gould’s response shows us that one of the evolutionist’s strategies for coping with taboo agency is immediately to turn the question, “How does creative change arise?” into the different question, “How does creative change, once arisen, spread through a population?” The switch of topics is not hidden, but occurs in plain sight. Only a habit of blindsight relative to the organism’s agency seems able to explain the prevailing inattention to such an obvious evasion of a real biological question.

None of this means we need to doubt whatever is true in the idea of natural selection. Selective mortality certainly occurs in one sense or another. Not every organism lives out a full life. But the mere elimination of problematic traits (or defective organisms) through mortality is not the same thing as profoundly transforming the integral unity that every organism is.

The point is not terribly subtle. There is simply nothing in the idea of natural selection that points to the creative capacities necessary for producing new adaptive features — for producing, say, a four-chambered heart (with all its organism-wide implications) from a three-chambered one. There is only the living being whose agency and activity natural selection necessarily assumes and which, as a result of blindsight, evolutionists have unconsciously transferred to a mystical “mechanism” of selection somehow operated by the inanimate world.

So here is our main question for this chapter: What do organisms show us, directly, compellingly, and uncontroversially, about their own powers of organic transformation? Much of the first half of this book contributes to an answer, especially at the physiological and molecular levels of observation. But in the present, evolutionary context, it will be well to look at the organism from a new angle.

(show less)

Chapter 18: Teleology and Evolution

Read the chapter introduction (+/-)

Every organism is continually dying in order to live. Breaking-down activities are prerequisites for building up. Complex molecules are synthesized, only to be degraded later, with their constituents recycled or excreted. In multicellular organisms such as vertebrates, many cells must die so that others may divide, proliferate, and differentiate. Many cancers reflect a failure to counterbalance proliferation with properly directed tearing-down processes.

You and I have distinct fingers and toes thanks to massive cell death during development. The early embryo’s paddle-like hands give way to the more mature form as cells die and the spaces between our digits are “hollowed out”. In general, our various organs are sculpted through cell death as well as cell growth and proliferation. During development the body produces far more neurons than the adult will possess, and an estimated ninety-five percent of the cell population of the immature thymus gland dies off by the time the mature gland is formed.

Despite all this life and death, I doubt whether anyone would be tempted to describe an embryo’s cells as “red in tooth and claw”. Nor do I think anyone would appeal to “survival of the fittest” or natural selection as a fundamental principle governing what goes on during normal development. The life and death of cells appears to be governed, rather, by the form of the whole in whose development the cells are participating.

But this has been a truth hard for biologists to assimilate, since it has no explanation in the usual causal sense. One way to register the problem is to ask yourself what you would think if I suggested that organisms in an evolving population thrive or die off in a manner governed by the evolutionary outcome toward which they are headed — that the pattern of thriving and dying off becomes what it is, in some sense, because of that outcome. It is not a thought any evolutionist is likely to tolerate.

But perhaps the occasional intrepid researcher will be moved to inquire: “Why not?” After all, we can also ask about the cells populating our bodies: do they thrive or die off in a manner governed, in some sense, by the forthcoming adult form? And here the answer appears to be a self-evident “yes”.

Perhaps, when we have come to accept what we see so clearly in individual development, we will find ourselves asking the “impossible” question about evolutionary trajectories: Does natural selection really drive evolution, or is it rather that the evolving form of a species or population drives what we think of as natural selection? Are some members of an evolving species — just as with the cells of an embryo’s hands — bearers of the future, while other members, no longer being fit for the developing form of the species, die out?

What makes this idea seem outrageous is the requirement that inheritances, matings, interactions with predators, and various other factors in a population should somehow be coordinated and constrained along a coherent path of directed change. Unthinkable? But the problem remains: Why — when we see a no less dramatic, life-and-death, future-oriented coordination and constraint occurring within the populations of cells in your and my developing bodies — do we not regard our own development as equally unthinkable?

Few would imagine that our own well-directed development from embryo to adult is owing to an external guiding power or to a conscious “aiming” or planning. Nor need we think that the “developmental path” of evolution is owing to guidance such as an external breeder might supply. Rather, the idea would be that the evolutionary narrative, like the developmental one, arises from the agency and developmental powers of cells, organisms, and communities of organisms, as they express their own character and realize their potentials in the presence of the prevailing environmental challenges and opportunities.

So the question is this: do we have any less reason to expect a coordinating agency at play in a population of organisms pursuing an evolutionary trajectory than we do to expect a coordinating agency at play in a population of cells pursuing a developmental trajectory?

Our answer will depend on our willingness to take seriously a plain fact of our experience — a fact stressed throughout the first half of this book: agency and intention, wisdom and meaning, are given expression by organisms in a way that belies our expectations for collected bits of inanimate matter.

(show less)

Chapter 19: Development Writ Large

Read the chapter introduction (+/-)

We have found throughout the preceding chapters that all biological activity, including at the molecular level, is thoroughly and irreducibly directive. Some biologists explicitly acknowledge the fact, and all biologists implicitly recognize it in their choice of descriptive language (see Chapter 2, “The Organism’s Story”).

This leads naturally to the central conclusion of this book — a conclusion I will develop in this chapter — which is that we already know more than enough to say that evolution is a purposive, or directive, or telos-realizing, process. Many readers, I suspect, will have difficulty coming to terms with this conclusion. But, as we will see, it is simply a matter of admitting to ourselves what we in fact know quite well. After all, an understanding of the directiveness of living activity, however repressed, is the only thing that lends to existing theory any appearance of plausibility.

The essence of this “unacknowledged knowledge” lies in the striking truth that living activities are quite unlike inanimate processes. Whether conscious or unconscious, they are, simply as a matter of observable fact, effective preparations for the future. In this sense they are purposive, and the purposes are carried out with an incomprehensible wisdom and facility. A cell replicating its DNA, proceeding through division, and intricately coordinating its ever-changing patterns of gene expression; higher animals mating and providing for their young; a zygote undergoing all the “miraculous” transformations of embryonic development — these activities are, in terms of the prevailing principles of biological explanation, all but out of reach. Processes we conventionally accept as “causal” do not explain a developing organism’s living narrative — its ceaseless adjustment and coordination of causal activity so as to move directively toward a characteristic future that is not yet there.

But such telos-realizing narratives are also so “boringly” familiar that we cannot help taking them for granted. We assume their decisive role in every biological context we look at, and cannot “un-know” them even when we are theorizing from a position that ignores or denies them. And so we have the two sides of biology today: an uneasy, theoretical disregard of what seems ungraspable or dangerously mysterious, and a carefree, unexamined taking-for-granted of the powers at work in those all-too-familiar mysteries.

My aim in this chapter — an aim grounded in all the preceding chapters — is to facilitate the changed angle of vision that can enable the reader to grant full recognition to what is already known. I want to jog evolutionary thinking out of its customary pathways.

(show less)

Chapter 20: Inheritance and the Whole Organism

Read the chapter introduction (+/-)

In 1923 Wilhelm Johannsen, the Danish plant physiologist and pioneering geneticist who had earlier given biologists the word “gene”, expressed concern about the way genes were being conceived as neat, cleanly separable causal units. He made the following curious remark, which remains today as intriguing as ever, despite its never having prompted much serious discussion within the field of genetics as a whole:

Personally I believe in a great central ‘something’ as yet not divisible into separate factors. The pomace-flies in Morgan’s splendid experiments continue to be pomace flies even if they lose all “good” genes necessary for a normal fly-life, or if they be possessed with all the “bad” genes, detrimental to the welfare of this little friend of the geneticists (Johannsen 1923, p. 137).

The pomace-fly, of course, was the fruit fly (Drosophila melanogaster) that Thomas Hunt Morgan, in his Princeton University laboratory, was famously converting into a “model organism” for genetic studies. Thanks to procedures for mutating genes, controlling the mating of the flies, and tracing the inheritance of traits, this was the fateful period during which “genetic” was becoming synonymous with “heritable”. The fact that whole cells — and not merely genes — pass between generations was progressively losing its significance in the minds of biologists interested in inheritance and evolution.

Johannsen saw that this new genetic work was based on the assumed existence of separate and independent causes of traits, and therefore left untouched what might easily be seen as the central problem of inheritance: the faithful reproduction of kind, or type — that is, the maintenance of the integral unity that harmonizes all the particular traits and parts of an organism and gives that organism its characteristic way of being. While mutated genes might result in (typically pathological) differences in certain narrowly conceived traits, this sort of change never reached through to the fundamental identity of the organism. Whatever the introduced variations (mutations), the pomace-flies always remained pomace-flies.

Johannsen’s problem arises because we can hardly help recognizing the distinctive unity of a living being — a unity we have difficulty equating to any particular parts. Rather, the organism seems in some way responsible for its parts. We never see an organism being constructed or assembled from already-existing parts. In its development it works to bring them about — to differentiate them out of a prior unity. Every organism is the power to do this work, and the power is not derivable from its results. If some of its parts become deformed, the organism works out of its unity to compensate for the deformities, doing so according to the way of being of its own kind.

But what sort of genetically investigated differences was Johannsen dismissing as disconnected from the problem of the whole? In his brilliant, and still decisively relevant 1930 book, The Interpretation of Development and Heredity, the British marine biologist E. S. Russell took up Johannsen’s concern. “When we say that a child shows a hereditary likeness to its father”, Russell wrote, “we mean that it resembles its father more closely than it does the average of the population. The likeness is observable in respect of those individual characteristics that distinguish the father from the rest of the race” (emphasis added). Much the same can be said of the child’s resemblance to its mother.

It’s also possible that there will be no particular resemblance to either parent. “But yet in all three cases the child would show the characteristics of its species and its race — it would be a human child, distinguishable as belonging to the same racial type as its parents”. As Russell then noted, this general resemblance in type, whereby all members of a species share an entire manner of development and way of being, can hardly be compared to the inheritance of this or that inessential variation wherein a parent happens to differ from most other members of the species. But such inessential variations have been a main focus of geneticists’ investigations for the past century.

The distinction between a fundamental, shared nature, on one hand, and individual peculiarities that occur within that shared nature, on the other, has practical implications for genetic research:

The broad general resemblances of type give no hold for experimental or statistical treatment, and have accordingly on the whole been ignored. But it is this general hereditary resemblance which constitutes the main problem. [The gene theory] deals only with differences between closely allied forms, and with the modes of inheritance of these differences; it leaves the main problem quite untouched as to why, for example, from a pair of Drosophila only Drosophila arise. It takes for granted the inheritance of Johannsen’s “great central something” — the general hereditary equipment of the species (Russell 1930, pp. 269-70).

We could also add here that the species’ capacity to produce variations in offspring was drastically understated by the methods the researchers employed — a problem that continues to this day. These laboratory methods, by keeping conditions as uniform as possible, enabled geneticists to isolate more or less reliably reproducible “causes”. This strengthened their conviction that biological causation could be approached on the model of the physical sciences. In other words, the design of the experiments was such as to reinforce pre-existing beliefs. After all, without reliable, unambiguous, isolated causes and effects, how could one come up with a publishable paper?

What this overlooks is that every organism is a thoroughly holistic (contextual) being whose entire business might be seen as the continual redefining of its own part-relations, or causal interactions, in response to different environments. Keeping those environments constant in the laboratory was a way of repressing the full expression of the organism’s unity as it might have manifested under varying circumstances — a unity that could not be summed up in terms of a set of discrete and fixed causes and effects.

A vast (and almost overwhelming) amount of research today has had to be aimed at elucidating the context-dependent activity of organisms that was overlooked earlier. “Context-dependent” has become a byword of contemporary genetic and molecular biological investigations. And yet the pathology of the earlier work, compulsively driven as it was by the effort to isolate unambiguous causes fit for a successful publication, continues to distort these newer investigations. (On this, see Chapter 9.)

(show less)

Chapter 21: Inheritance, Genetics, and the Particulate View of Life

Read the chapter introduction (+/-)

This chapter about the gene-centered (“genocentric”) view of evolving organisms shouldn’t have needed to be written. Today genocentrism has been challenged from so many different sides, by so many leading biologists, and in such an ever more insistent manner, that it might easily seem a waste of time to raise the usual issues afresh here. So I won’t.

But there are caveats. One is that, despite the criticism, the idea of the masterful, controlling gene remains as strongly entrenched as ever in the minds of most biologists. This is especially true of evolutionary theorists, for whom the word “genetic” has long been synonymous with “heritable”. In other words, for purposes of evolutionary theory genes substitute for the entire, one-celled organism (fertilized ovum or newly divided cell) that passes between generations. Which means that, as participants in an inheritance-based evolutionary lineage, organisms themselves effectively cease to exist for the theorist.

The century-long habit of genocentrism is seemingly resistant to all criticism. As three Duke University biologists summarized the matter in 2017, “Everyone understands” that the idea of a definitive gene for this or that feature of an organism “is a distortion of the biological facts, yet, as a profession, we have yet to rid ourselves of this crutch” (Gawne, McKenna and Nijhout 2018).

Much of the criticism of genocentrism has arisen from the field of evolutionary-developmental biology (“evo-devo”). Yet even here, according to a leader in that discipline, “increasing gene centrism characterizes the field today”:

This reductionist attitude continues to be upheld, even though overwhelming evidence points to the fact that it is not gene expression and regulation that singularly define body structures but the systemic processes of interaction between genes, cells, and tissues as well as the physics and physiologies of the involved entities and their interactions with numerous factors of the environment (Müller 2019).

A second caveat, even more discouraging, is that the critics themselves leave the door wide open for the persistance of genetic reductionism. This is because few if any well-positioned, reputable biologists are willing, at the risk of reputation and career, to speak out against the reigning materialist dogma of their profession. What isn’t generally recognized is that this dogma invites, as an inevitable counter-movement, devout respect for the all-determining, informational gene — a machine-like gene intelligently designed and engineered from outside by the “creative forces” of evolution. In this way the theorist employs the gene as a stand-in for everything that lives in the organism — an organism that has become invisible, yet whose life still nags at evolutionists in a blindsighted sort of way.

(show less)

Chapter 22: A Curiously Absolute Demand for Stable Variation

Read the chapter introduction (+/-)

There could hardly be a more frequently stated requirement for natural selection than this: any beneficial genetic or other variation occurring in an organism, if it is to be evolutionarily relevant, must be stable, heritable, and long-lasting down through the generations. If a given variation is likely to pass away after a generation or two, or if it quickly suffers further change, then the normally long and slow process of selection will not have time to spread the variation (“fix it”) throughout a population. Patrick Bateson was giving voice to a universal consensus when he wrote, “For the Darwinian evolutionary mechanism to work, something must be inherited with fidelity”.

But we might want to ask: Is it certain material products of an organism’s activity that must be stably inherited? Or is it rather the capacity for activity — an activity through which not only do particular products arise, but also a life is sustained and the character of a species (Chapter 20) is consistently expressed? Think, for example, of the apparently chaotic “catastrophe” that overtakes a goliath beetle larva prior to its ultimate, glorious transformation into an adult form (Box 17.1). Here, one set of material products all but disappears during a relatively brief but extremely intense period of change upon change, only to be replaced by an altogether different set. What is the organizing power, and what are the organizing ideas, through which this all-encompassing transformation occurs? And how are the organizing ideas and power passed from one generation to the next?

When we talk only about the inheritance of discrete products of activity, we have already shown a willingness to ignore the more fundamental problem of the origin of viable new traits, which require much more than some new bits of matter. Even if we are thinking only of the development of the color patch (speculum) on a duck’s wing feathers (Figure 11.2), we still need to embrace in thought a huge range of interactions that are possible only as an expression of an integrated and living whole.

The obstacle to a proper reckoning with change and inheritance lies in the focus on isolated products that are seen primarily in relation to a specific molecule (DNA) and its genes — genes whose mathematically calculable spread through a population is then thought of as equivalent to the spread of traits, which in turn is taken to be evolution. It is the demand for this sort of sterile calculability that leads to a one-sided emphasis on stable variation (gene mutations) rather than on the potent activity of self-transformation that organisms put on such obvious display.

Such, I think we will see, is the heart of the matter. But we may get a fuller grip on the issues by starting with the popularly effective case Richard Dawkins makes for the particulate, “gene’s-eye view” of evolution.

(show less)

Chapter 23: The Evolution of Consciousness

Read a brief comment about this chapter (+/-)

The history of language gives us a reflection of the evolution of human consciousness. It points to a time before we became aware of our own existence as distinct Selves observing a world “out there” and separate from ourselves. How accurate (or inaccurate) our own experience of a subject-object dualism is can be properly assessed only against the background of the evolution of consciousness, so far as it can be traced. Oddly, though, contemporary evolutionary theory has not had much to say about the evolution of consciousness. This ought to concern us, because the evolved character of our own consciousness unavoidably frames and limits our theory of evolution. The ancients read the “face” of what we think of as the outer world rather as we today can still read the face of another human being — that is, they read material phenomena as expressions of inner being. Can we learn anything by setting their experience over against our own alienation from the outer world? This may be the same question as, “Can we overcome the dualistic heritage of the past several centuries?”

(show less)

Part 3: The Underpinnings

Chapter 24: How the World Lends Itself to Our Knowing

Read the “Where are we now?” section from the end of the chapter (+/-)

We have, throughout this book, been bumping up against questions of epistemology: How are we situated in the world as knowers? How does our knowing emerge from our experience, and what is the relation between the resulting knowledge and the world’s reality? The questions began already in the Preface, where I suggested that a good part of our thinking about the molecular realm is really just an illegitimate projection of our qualitative experience of the world onto the blank screen of an unknown world of particles. This particle-world is falsely imagined to be non-qualitative and mind-independent and to exist somehow “behind” our experience. I explicitly reinforced this concern in Chapter 13 by offering examples of the projection, and throughout the book I have appealed to an interior (meaningful, ideal, agential, and purposive) dimension of living activity that is hard to square with current notions about what constitutes acceptable biological explanation.

There is not much in their training to encourage biologists to entertain questions on this fundamental level. And there is a great deal in the powerful taboo against non-materialist thought that discourages such questions. Nevertheless, I have tried to show in this chapter that even a cursory look at the role of qualities and thought in our engagement with the world (including our scientific engagement with the world) decisively undermines the entire materialist framework of current biology. This book, you might say, has been an exercise in raising questions that are simply invisible — because forbidden — under the present scientific regime.

I realize that some of the questions I raise may seem almost surreal to those raised and disciplined within the current environment. (For example, “With whom are we in conversation as we engage with the world around us?”) But a science that is altogether closed off from unfamiliar questions — especially when there is a strong case to be made for asking those questions — is not a healthy science. And surely we should never make it a requirement to have a full answer to a question before asking it.

(show less)

Chapter 25: Some Principles of Biological Understanding

Read the introduction to this chapter (+/-)

Failure to recognize the reality of the world’s interior dimension (Chapter 24, “How the World Lends Itself to Our Knowing”) is the central fact underlying virtually all the limitations, illusions, and distortions of today’s science. The twelve principles set forth in this chapter represent my effort to suggest some of the biological consequences of our acknowledging the interior dimension of the world.

Everything that distinguishes biology from the physical sciences derives in one way or another from the inner life of organisms. This remains true, I think, despite our difficulty in even beginning to imagine or characterize that inner life, and despite the fact that it must vary almost beyond all possibility of recognition between a one-celled organism and a human being.

Still, even those complex features commonly treated as definitive of life, such as the capacities for reproduction and self-maintenance (which I do not deal with here) are obvious manifestations of a well-directed wisdom, all the way down to the molecular level. This word “wisdom” needs to be understood, not as an occasional and foolish eruption of empty sentiment, but rather as a pointer to effective, end-directed, and meaningful life processes that surely must be distinguished from, yet just as surely must be evolutionarily continuous with, conscious human intention and reason. What I mean by this will, I hope, become clear in the discussion of the various principles given below.

Here are the principles we will look at:

Principle 1: The world manifests itself, by nature, as a content of experience.

Principle 2: Organisms are focal centers of agency.

Principle 3: Biology, as a science of organized wholes, cannot be understood in terms of mechanized parts.

Principle 4: Every organism is, first of all, a becoming, not a material structure.

Principle 5: Biology requires portraits of specific character, not applications of universal law.

Principle 6: Understanding the human self requires us to distinguish our own inner activity from its products.

Principle 7: Organisms in general lack human-like selfhood.

Principle 8: There exists a wide spectrum of consciousness in organisms.

Principle 9: An organism has its own sort of Interior Dimension.

Principle 10: A healthy science acknowledges the mystery implied by its own ignorance.

Principle 11: The mystery of time is central to the life of organisms.

Principle 12: Humans are a key to evolution.

(show less)




This document: https://bwo.life/bk/index.htm

Steve Talbott :: Organisms and Their Evolution — Table of Contents