Genes and Life: The Need for Qualitative Understanding


Genes and Life: The Need for Qualitative Understanding

Craig Holdrege

Which of our genes make us human?" An article with this title appeared last fall in Science (Gibbons, 1998). The article reports that there is hardly any difference between the DNA from humans and chimpanzees.

Approximately 98.5% of the DNA is the same. A photo of a chimp standing upright accompanies the article, with this caption: "Chimpanzees may adopt the occasional two-legged pose, but they differ dramatically from humans in anatomy and behavior." Given the similarity of humans and chimps at the DNA level, and the manifold differences at the biological and behavioral levels, we might conclude that DNA has little to do with the essential differences between human and chimp.

But the author of the article comes to a very different conclusion. She states:

This means that a very small portion of human DNA is responsible for the traits that make us human, and that a handful of genes somehow confer everything from an upright gait to the ability to recite poetry and compose music.

This statement reveals the conviction that DNA is the basic cause of all biological and psychological phenomena: organisms are bundles of myriad separate traits that are determined by genes. It follows from this postulate that all differences between organisms can only stem from differences in DNA. The 1.5% of "unique" human DNA—which may include only about fifty functional genes—must be the cause of everything human. Humanness is seen to be genetically modified "chimpness." Conversely, if scientists knew these "human" genes, they might be able to "convert a default-mode great ape into a human," as New York Times science writer Nicholas Wade put it (October 20, 1998).

Since its advent twenty-five years ago, genetic engineering of organisms has developed into a paradise for the quantitatively oriented, combinatorial intellect. DNA from any organism can be transferred into any other organism. There are no perceived boundaries; everything appears to have become interchangeable. True, the technical prowess still lags behind the vision of "anything goes," and in any given experiment there is only about a one percent success rate. But this fact is looked upon as a technical problem to be overcome on the path toward the creation of organisms as custom-made, fine-tuned genetic mechanisms.

At the gateway to the temple of modern science stands written: banish all qualities, pursue quantity. Galileo, Descartes, and Locke furthered this cause by stating that qualities do not actually exist. Qualities are deemed subjective epiphenomena of reality, which is thought to be quantitative. This conception became the underlying ideology of science: qualities are subjective and there is, therefore, no science of qualities; science pursues the quantitative, because that is the nature of reality.

But as the holistic thinker and neurologist Kurt Goldstein points out:

Biological knowledge is not advanced by simply adding more and more individual facts. The facts which are gradually included in the "whole" as parts can never be evaluated merely quantitatively, in such a way that the more parts we are able to determine the firmer our knowledge becomes. In biology every fact always has a qualitative significance. (Goldstein, 1971, p. 30)

Goethe and a Qualitative Approach

One example of the striving toward a science that encompasses the qualitative took place in a controversy at the end of the 18th century. By that time comparative anatomists had discovered that humans and mammals have essentially the same anatomical structure. Finally it came down literally to one bone of contention: the premaxilla. The premaxilla (which is actually a pair of bones, also called the premaxillary bones) forms the outermost (distal) part of the snout of the upper jaw in mammals (see figure 1). When anatomists investigated the human skull, they didn't find the premaxillary bones. Looking at the human skull from the front, the premaxillary bones "should" have been between the two upper jaw bones. Instead, the anatomists saw only the suture where the two upper jaw bones meet.



Figure 1. Skulls of the white-tailed deer, mountain lion, and human being. In the human being the premaxillary bones (p) are visible only when the skull is viewed from below (basal view). m: maxilla (upper jaw bone); n: nasal bone;p: premaxillary bones.

Several well-known anatomists declared in their treatises that the human being has no premaxillary bones, thereby distinguishing us from animals. Most of the scientists of those pre-Darwinian times were steeped in the Judeo-Christian tradition and carried the deeply-felt conviction that there were essential differences between the human being and animals. But their scientific studies were showing more and more similarities. The premaxilla was evidently felt to be the last bastion separating humanity from "the lower beasts."

The poet and scientist J.W. von Goethe rejected this view that the concept of humanness could hinge upon a single bone. With vehemence he set out to find the premaxilla, and he succeeded (Goethe, 1988, p. 111ff). He had to look more carefully than his contemporaries had done, since the premaxillary bones can usually only be seen when you look behind the incisors at the upper palate (where you put your tongue when you say "tea"). There Goethe saw, in some skulls, two little bones that lie between the upper jaw bones (see figure 1). These are the premaxillary bones. Goethe soon discovered that others had made the same observation long before him. But their studies had been ignored, forgotten or overlooked.

You may have difficulty appreciating Goethe's enthusiasm for the discovery of this bone—an enthusiasm he expresses in letters to his friends Herder and Knebel (quoted in Schad, 1985, my translations):

(To Herder:) I have found—neither gold nor silver, but it gives me the greatest joy—the premaxilla of the human being.... It is like the keystone to the human being; it's not missing; it's there! But how!(To Knebel:) The difference between the human being and animals is not to be found in any given particular....The agreement within the whole makes each creature what it is, and the human being is a human being just as much through the form and nature of the upper jaw as through the form and nature of the last two bones of the little toe.

The discovery of the premaxilla supported Goethe's view that what distinguishes the human being from animals does not lie "in any given particular." For Goethe the presence of the premaxilla did not make the human being into a "mere" animal, nor would its absence have guaranteed a distinction between human and animal. What was important was not that the bone was present, but how, "since the agreement within the whole makes each creature what it is." He saw that the premaxilla—if viewed with an eye directed toward understanding the whole organism—could reveal humanness, just like any other bone. This insight was the source of his deepest joy.

The Whole and the Part

Can we see how a single bone is related to the organism as a whole? Long-legged, long-necked hooved mammals like the deer also have elongated skulls and the premaxillary bones are long and slender (see figure 1). The same is true for the other bones of the front part of the skull: the nasal bone, upper jaw (maxilla) and the lower jaw (mandible) are all long and taper toward a point. Just as the front (distal) part of the skull is elongated so is the distal part of the limb: the foot and toe bones are especially long. The deer stands on the tip of its two hoofed toes and the heel is halfway between the ground and the torso. This part of the leg is mainly tendon and bone, while the shorter upper leg bones are surrounded by muscle. Similarly, the rear part of the skull and jaw are embedded in muscle, while the snout is bony.

By contrast, the mountain lion has a much more compact skull, which is mirrored in the shorter and broader premaxillary bones (see figure 1). The mountain lion—as is typical for the cat family as a whole—also has a more compact body than a deer. Its neck and legs are relatively short and muscular. It stands not on the tips of its four toes, but on the first toe joint. Just as the mountain lion latches onto and penetrates through the skin of its prey with its claws, so does it pierce and tear flesh with its sharp and pointed canine and cheek teeth. Clearly, head and limb fit together.

What about in the human being? Goethe realized that the way the individual bones—including the premaxillary bones—are formed in the human being is an expression of our upright posture (cf. Schad, 1985). In the human skull the facial and jaw bones do not protrude forward. Developmentally this fact is expressed in the rapid closing of the upper jaw bones that grow over the premaxillary bones, which themselves remain undeveloped. Being short and round allows the skull to balance on the vertical spine. In the four-legged mammal, the brain case remains small and sits behind the protruding face and jaw, reflecting the animal's overall horizontal orientation. By contrast, the human skull has a large brain case, which rises above the face and jaw, mirroring the vertical orientation of the whole human body. The fact that we have no snout is related to our ability to speak and also express our emotions and thoughts through facial expression. The large cranial vault is an expression of the inwardness out of which we can act.

In contrast to the deer or mountain lion, in the human being the femur—the body-near, proximal part of the leg—is the longest bone in the leg (and in the whole body) while the feet and toes are short; the heel rests on the ground. Mirroring this relation in the head, the brain case, as the part of the skull closest to the body axis, expands, while the facial cranium remains small. Now we can see how justified Goethe was in writing: "A human being is a human being just as much through the form and nature of the upper jaw as through the form and nature of the last two bones of the little toe."

All of our anatomical, as well as our mental and spiritual capacities can be seen in relation to the upright posture. Goethe's friend Herder recognized this connection clearly, and expressed it precisely and beautifully:

Because the human being has to learn all things, because it is our instinct and calling to learn everything like our upright gait, we learn to walk by falling and come often to truth only through error. The animal is carried forward securely in its four-legged gait; the more strongly expressed proportions of its senses and drives are its guides. The human being has the advantage of a king to look to far horizons, upright and with head held high. Of course, we also see much darkly and falsely. We forget our steps only to be reminded when stumbling on what a narrow basis the whole head and heart edifice of our concepts and judgments rests.... The human being is the first to be set free in creation. We stand upright. The balance of good and evil, of false and true hangs in us. We can search, we shall choose. Just as nature gave us an overviewing eye to guide our gait, so also do we have the power not only to place the weights, but—if I may put it this way—to be the weights on the balance. (Herder, 1982, p. 65)

Herder's description can give you an impression of how far it is possible to go in relating seemingly mundane biological facts to the inner nature of humanity. When we see how each aspect of a being tells us something about it as a whole, we are no longer thinking quantitatively. The part is not an isolated "thing." It is a revelation and realization of a larger context. As Hegel might have phrased it: the part of an organism goes beyond itself to become what it is essentially, namely, a member of a whole. When we begin to see the same tendencies, the same gestures in different parts, the whole begins to speak. What speaks is not a quantity, not a thing, but a quality—that of the specific wholeness of human or chimp. Nothing is interchangeable. The chimp is through and through chimp. The human is through and through human. In order to adequately grasp the inherent nature of a living being, we must think qualitatively. This is what organisms tell us.

This understanding is direly lacking in a genocentric view that hopes to discover the difference between human and chimp in fifty genes. It is not only lack of insight, but hype to give 1.5% of the genome credit for human biology—let alone credit for the ability to recite poetry and compose music. The secret—the "cause"—of humanness or chimpness is not to be discovered in our genes, although some day it may be possible to discover how genes themselves relate to the integral nature of a being.

Genetic Manipulation
The prowess of genetic engineers arises from their ability to ignore the organism as an integrated, interactive being, while gaining a knowledge of isolated substances and small-scale biochemical processes. When scientists think of an organism as a genetic mechanism, then there is every reason to try to exchange parts in order to "improve" the mechanism's function. The ideal of genetic engineering is to create and control the smoothly functioning, predictable bioreactor—the plant that is resistant to pests and herbicides, the farm animal that produces meat or pharmaceutical substances with maximal efficiency, the human being genetically vaccinated against an array of infectious diseases.

For example, if larger, faster-growing pigs are desired, why not implant growth hormone DNA from, say, cattle into pigs? Government, university, and industry scientists have been working on this task for more than a decade (Pursel, 1998). There are now transgenic pigs that produce large amounts of bovine growth hormone. These pigs grow faster, utilize feed more effectively, and have less carcass fat than their normal cousins. This is more than the researchers had hoped to achieve, since they didn't expect the pigs to produce leaner pork. But these are not the only changes that take place. The pigs are also very susceptible to stress, have a high incidence of gastric ulcers, dermatitis, arthritis, lameness, and renal disease, and the boars often lack libido. Because of these (and other) "side-effects," such transgenic pigs have not yet become marketable.

This example is not an isolated case. Genetic engineering experiments often produce unexpected, negative results. Once we learn to see organisms as the integrated beings they are, where every part and process is related to every other part and process, then we begin to expect such unexpected results. They are not "accidents," even if they cannot be foreseen. The transgenic organism reacts as a whole to the foreign substance (DNA). Sometimes the reactions are subtle, often they are crass and harmful. For example, in addition to the changes described above the transgenic pigs stop the secretion of their own porcine growth hormone, partially balancing out the effects of the genetic manipulation (Pursel et al., 1987).

When genetic knowledge and the technology are re-introduced into the arena of life, they enter a realm where context is everything. If there is one thing we can be sure about, it's that unexpected problems will arise. That's the most certain prognosis. This fact makes the blatant undervaluation of the way transgenic crops will affect other organisms and the environment as a whole especially insidious. Many crops (soybeans, corn, potatoes, cotton, and so on) are being genetically modified to become resistant to herbicides and pests. In 1998 about thirty percent of all soybeans planted in the USA were genetically engineered for either pesticide or herbicide resistance. Recently scientists genetically engineered a weed (Arabidopsis, the workhorse of plant genetics) to become herbicide resistant. Completely unexpectedly, many of the transgenic plants changed their reproductive behavior (Bergelson et al., 1998). Normally Arabidopsis is self-fertilizing, but after the genetic manipulation many of them started to cross-pollinate. The researchers have no idea why this happened and how it is related to their manipulation. But they do recognize that such an unexpected result should give us pause to stop and think about transgenic crops. Soybeans, for example, are normally self-fertilizing. If anything comparable happened to them, then the herbicide resistance could spread via pollen to plants that are not genetically engineered. Imagine the predicament of the organic farmer whose neighbor plants transgenic soybeans.

Of course no one knows whether soybeans will change in this way, but at least there should be concern. Most ecologists are urging more stringent practices and better oversight of transgenic crops. But the USDA has actuallyreduced the oversight of field trials and commercialization of transgenic crops, playing into the hands of the large biotech companies that produce the seeds as well as the herbicides.

The producers of transgenic organisms, and very often government agencies as well, would like to capture us in their vision of complete control, and to have us believe that laboratory experiments and clearly circumscribed field tests "prove" the efficacy and safety of their products. But neither genetic thinking, nor the techniques themselves are adapted to the complexity of life; they are maximally effective only under maximally controlled conditions. Life is not about isolation; it is about interpenetration and mutual dependence.

The tendency within science and genetic technology is not to change the approach, but to find ways to increase control when problems arise, counteracting the problems already created by using more of the same kind of methods. It might be possible, for example, to build stalls to support arthritic, lamed pigs that can hardly walk and carry their body weight. Or couldn't transgenic pollen be modified to self-destruct when it reaches the air? (But what if a few don't?) Precisely such "solutions" show the grotesque nature of viewing and treating organisms as objects to be manipulated and completely controlled. This approach can only lead to the creation and perpetuation of more unhealthy conditions.

The Need for Qualitative Understanding

The quantitative approach can continue to feed our desire, and increase our ability, to control aspects of life. But it is also starkly reflected in the mirror of unhealthy side-effects and the continual race to solve problems we have caused with solutions that cause more of the same kind of problems. Following this path, organisms will increasingly come to reflect our disregard for the integrity of living processes. Organisms are resilient, but can they withstand indefinitely the onslaught of manipulation based on combinatorial, mechanistic thought?

A shift to a qualitative approach in science can help lead us beyond this dilemma. This shift is radical and difficult, since it entails cutting through hype and sacrificing our will to power. But we can leave behind the desire to control, striving instead to acknowledge and understand the integrity of our fellow creatures. When we begin to see organisms as qualitative wholes in which even a seemingly insignificant bone carries the signature of the whole, then the organism starts to come alive for us. We establish a relation from being to being. The more our understanding becomes centered in the other, the more we can develop ways of action that take the integrity of that being into account. How else is responsible action possible?

As long as scientists view organisms as mechanisms in a world separate from them, there is no real question of responsibility. Ethical considerations in this case are after-the-fact and viewed as being outside the scientific process. For this reason they are also usually ineffective. Only when we incorporate the qualitative into the scientific process—when our way of viewing consciously includes the other being from the outset—can we begin to heal this split. Responsibility involves an inner relation to the world that is bracketed out of a quantitative approach.

Although genetic engineering is clearly a proof of the efficacy of modern science, its results show just as clearly the necessity to change our ways of viewing and treating life. It is as though transgenic plants and animals were shouting at us to stop viewing and treating them like inanimate objects that can be modified according to our wishes and desires, and to earnestly begin to acknowledge, understand, and deal with them as the beings they are.

Craig Holdrege is founder of The Nature Institute and author of Genetics and the Manipulation of Life: The Forgotten Factor of Context. A lengthened version of this essay will be presented at a conference on "Goethean Science in Holistic Perspective," Teachers College, Columbia University, May 20-22, 1999.

References

  • Bergelson , J. et al. ( 1998). Promiscuity in Transgenic Plants. Nature 395:25.
  • Gibbons, A. (1998). Which of Our Genes Makes Us Human? Science 281:1432-1434.
  • Goethe, J. (1988). Scientific Studies, edited by D. Miller. New York: Suhrkamp.
  • Goldstein, K. (1971). Human Nature. New York: Shocken Books.
  • Goldstein, K. (1995). The Organism. New York: Zone Books.
  • Herder, J. (1982). Ideen zur Philosophie der Geschichte der Menschheit, vierter Band. Berlin: Aufbauverlag.
  • Pursel, V. (1998). Modification of Production Traits, in Animal Breeding - Technology for the 21st Century, edited by A. Clark. Netherlands: Harwood Acad. Publ.
  • Pursel, V. et al. (1987). Progress on Gene Transfer in Farm Animals. Veterinary Immunology and Immunopathy 17:303-312.
  • Schad, W. (1985). Stauphaenomene am menschlichen Knochenbau, in Goetheanistische Naturwissenschaft, Bd. 4 Anthropologie, pp. 9-29, edited by W. Schad. Stuttgart: Verlag Freies Geistesleben.
  • Simmel, M. (1968). The Reach of the Mind (Essays in Memory of Kurt Goldstein). New York: Springer Verlag.
  • Original source: In Context (Spring, 1999, pp.11-15); copyright 1999 by The Nature Institute

Doing Goethean Science


Craig Holdrege
The Nature Institute
http://www.janushead.org/8-1/Holdrege.pdf

Practicing the Goethean approach to science involves heightened methodological awareness and sensitivity to the way we engage in the phenomenal world. We need to overcome our habit of viewing the world in terms of objects and leave behind the scientific propensity to explain via reification and reductive models. I describe science as a conversation with nature and how this perspective can inform a new scientific frame of mind. I then present the Goethean approach via a practical example (a study of a plant, skunk cabbage) and discuss some of the essential features of Goethean methodology and insight: the riddle; into the phenomenon; exact picture building; and seeing the whole.

Beginnings: Sensing Boundaries
I have vivid memories of Mr. Sinn’s 9th grade science class. We did experiments with glassware, tubes, and Bunsen burners—that was neat. But then Mr. Sinn taught us how to explain the results of our experiments. He described processes—he must have been talking about molecules—that we didn’t see. These became schemes with letters and numbers on the blackboard. We now were supposed to know what had really been going on. And I was lost. I didn’t get it. What did the blackboard diagram have to do with what we’d been observing? This was an unsettling experience that had significant consequences: I avoided science like the plague in high school.

It is also my first memory of the kind of experience that I have had repeatedly since then and that has been key in my pursuing Goethe’s approach to science. It is the experience of confronting what are called scientific explanations and feeling (in thought) a distinct sense of dissatisfaction. How can a phenomenon be explained by something that is supposed to underlie it and that is always less than the phenomenon itself? I have been amazed that what a large community of people feels to be an explanation leaves me with the question: what do I have now? What am I doing by leaving the phenomenon in order to explain it? Let me give a few examples.

In a college botany course I learned why plants that grow in shady places have broader and larger leaves than plants that grow in full sunlight. The reason given is that plants growing in shade don’t receive as much light to do photosynthesis. Therefore they grow larger surfaces with which they can capture more light and produce more organic matter via photosynthesis. Plants have developed this strategy to survive and reproduce in shady habitats. This is a typical functional explanation that makes perfect sense—until you think the matter through a bit further. The larger the surface area a plant creates, the more substance it needs to build up and sustain its larger body. Wouldn’t it be just as effective for the plant to stay very small with narrow leaves? In this way it wouldn’t have to do so much photosynthesis since it could stay small. Both explanations make sense. I have yet to find a functional explanation of a phenomenon for which one couldn’t find equally plausible alternatives. Such evolutionary explanations always fall short. They fall short because they are an attempt to get a grip on a complex biological phenomenon from only one narrow and limited perspective. I have shown this to be the case in celebrated textbook examples such as industrial melanism in the peppered moth and the long-necked giraffe (Holdrege 2003, 2004).

I can formulate the problem in another way. Every biology student learns that the fundamental question of biologists confronting a phenomenon is: what is the underlying mechanism? It may be a Darwinian survival strategy or a hormonal or genetic mechanism. In the search for such mechanisms two essential things happen. First, you isolate the phenomenon out of its context within the organism as a whole and, second, you seek to explain it in terms of a reduced set of quasi-mechanical processes. In the end what you come up with is a simplified picture of a phenomenon caused by an abstractly conceived underlying mechanism. (The neurologist Kurt Goldstein has elucidated this problematic side of science in his seminal work on holistic science, The Organism (Goldstein, 1939/1995).)

Take, for example, a “trait” that is explained by a “gene.” (For an in depth discussion see Holdrege, 1996.) When Mendel discovered hereditary patterns in pea plants he focused his attention on particular characteristics such as seed shape or flower color. He mentally abstracted these characteristics from the plant. He could only do this with “clear and distinct” characteristics, that is, ones that don’t vary much under changing conditions. He did not look at flower color as it changes from bud formation through flower maturation and wilting, nor did he worry about the slight variations in color that occur between different specimens of the same breeding line. Flower color grasped as a Mendelian trait entails bracketing out developmental changes and variation. The resultant trait is an isolated, distinct feature and it is quite straightforward to go from it to an underlying particulate factor—later called a gene—that is inherited and responsible for the appearance of the trait through the generations. The history of genetics shows the power of  this way of viewing and working with organisms.

The problem is that both the trait and the gene are products of abstraction, so that one is explaining an abstraction with an even greater abstraction. The red flower color of a strain of pea plants is much more than a genetic trait, and the biochemical component of inheritance is much more than the genetic code. This is becoming increasingly clear even within the field of genetics, so that some geneticists question the value of the concept of the gene altogether. Geneticist William Gelbart writes:
For biological research, the 20th century has arguably been the century of the gene. The central importance of the gene as a unity of inheritance and function has been crucial to our present understanding of many biological phenomena. Nonetheless, we may well have come to the point where the use of the term “gene” is of limited value and might in fact be a hindrance to our understanding of the genome. Although this may sound heretical, especially coming from a card-carrying geneticist, it reflects the fact that, unlike chromosomes, genes are not physical objects but are merely concepts that have acquired a great deal of historic baggage over the past decades. (Gelbart, 1998)
It’s interesting how reality tends to catch up with science at some point.

I’ve sketched here a couple of experiences I’ve had with scientific concepts. Over the past 25 years I’ve become keenly aware of what these concepts do not tell us and do not reveal. And because they are often used as if they told us something beyond their own narrowly circumscribed domain, they often mislead and cover up phenomena. Becoming aware of such boundaries is significant, because—to paraphrase Hegel—in our gaining an awareness of a boundary we have already begun to transcend it. We are at the gateway to a new kind of understanding, to the further evolution of science that Goethe inaugurated.

Delicate Empiricism: Science as a Conversation
The realization that the phenomena we confront are always richer than the abstractions we use to explain them is central to a Goethean approach. This realization is the expression of a two-fold awareness or sensitivity that Goethe points to with his expression “delicate empiricism” (Goethe, 1829, in Miller, 1995, p. 307). First, we experience a phenomenon (a mouse, a wooded swamp, a range of blue hills in the distance, or the clouds moving across the sky) as a kind of fullness that calls forth wonder, curiosity, questioning. We want to get to know it better, or as Goethe states it radically, “become utterly identical with it” (ibid.). This is empiricism, because we orient all our striving around the phenomena themselves. A phenomenon is what meets the eye but we also experience it is as something more, as a kind of surface that is pregnant with a depth we may be able to plumb. But we realize that we will not fathom these depths with models and theories, which more likely than not will lead us away from the phenomenon itself. This brings us to the second mode of sensitivity: we are acutely aware of the thoughts we bring to the phenomenon, how we interact with the world through thinking. We know that in conceiving thoughts we can both illuminate and color our experience. The more we are aware of the thoughts we bring, the more transparent and illuminating they can be. We must become delicate in the way we work with our concepts in our efforts to let the depths of the phenomena disclose themselves.

Goethe describes the process of gaining knowledge in the following way:
When in the exercise of his powers of observation man undertakes to confront the world of nature, he will at first experience a tremendous compulsion to bring what he finds there under his control. Before long, however, these objects will thrust themselves upon him with such force that he, in turn, must feel the obligation to acknowledge their power and pay homage to their effects. When this mutual interaction becomes evident he will make a discovery which, in a double sense, is limitless; among the objects he will find many different forms of existence and modes of change, a variety of relationships livingly interwoven; in himself, on the other hand, a potential for infinite growth through constant adaptation of his sensibilities and judgment to new ways of acquiring knowledge and responding with action. (Goethe, 1807; in Miller, 1995, p. 61)
In Goethe’s view science entails “mutual interaction” with the phenomena. Engaging in this process we discover the “limitless” nature of connections and relationships in the world but at the same time our potential to continually grow and adapt ourselves to new, more adequate ways of knowing. Doing Goethean science means treading a path of conscious development. The question accompanying every aspect of the work is, “How can I make myself into a better, more transparent instrument of knowing?” In traditional science, we are much more likely to ask, “How can I find ways of adapting the phenomena to my specific approach so that I can answer my question?”

I have found the metaphor of conversation increasingly helpful in illuminating the nature of a Goethean approach to science. The metaphor brings to consciousness that doing science is a back-and-forth between partners in an ongoing process. It accentuates a kind of inner attitude that lies at the heart of doing Goethean science, one very different from the frame of mind one normally associates with science (although it informs, but often not explicitly, the work of many good scientists). Here, expressed in fairly general terms, are some of the elements of science-as-conversation. (See also Talbott, 2004.)

1) When I enter into a conversation with nature my interest has been sparked by some experience, my attention has been caught. I’m presented with a riddle and begin asking questions, observing, and pondering. In this way I give the conversation an initial focus. If the interaction between me and nature has no focus it can easily become chit-chat and not a conversation.

2) But if the focus I bring is too narrow and too rigid (for example, a narrowly defined hypothesis), we don’t have a conversation, we have a drill (one-sided questioning). In any productive conversation the process itself is paramount. It’s not just about me answering my pre-formulated questions, but centrally about what happens along the way. There will be surprises, moments of silence, tension. The back and forth between me and nature is dynamic and I attend to this process as an integral part of the conversation.

3) Taking the conversation-as-process seriously means realizing that it is open-ended. I don’t know where we’re going to arrive. With this awareness present at every moment, the conversation is imbued with an atmosphere of openness. I could also describe this attitude as a kind of animated looking forward to things unexpected that may arise.

4) Nature is my partner in the conversation. If I truly mean this and don’t take the statement as a feel-good clichĂ©, then I’m acknowledging that nature is something in its own right. I may not, at the outset, be able to say more than that. But the recognition of the other as something in its own right is a pre-condition for any conversation. This recognition infuses respect into the conversation and gives it dignity. In saying this I don’t mean that geologists will no longer crack open rocks with their hammers or botanists will stop pressing plants. However, knowing that I am involved in a conversation makes me more circumspect and I become more sensitive in what I think and do. I may ask, for example, whether I may be going too far and transgressing boundaries. I’m not talking here about abstract, prescriptive directives—since the conversation is a process, I can’t know what will emerge out of it beforehand. But in any case, it is carried by an attitude of respect.

5) An essential feature of the conversation is that I listen to what nature has to say. Receptive attentiveness allows us to hear and see with fresh ears and eyes. It’s the quality of open interest in what the other has to say. But it would not be a conversation if I only listened. I respond and interject. I am actively giving form to the conversation through my questions, observations and the new concepts I bring in. A vibrant conversation needs the movement between receptive attentiveness and active contributing.

6) In the course of any real conversation the partners change and evolve—they are in a different place than they were at the outset. It is easy to see that I as a scientist change in this conversation. I have gained new experiences, taken new qualities into myself and gotten to know the world more deeply. But what about nature? In a simple sense, any time we interact with nature through an experiment, we change nature. Field ecologists have recently discovered that even touching and marking plants in the field can affect their growth (Cahill, et al. 2001). Goethe’s seminal essay “The Experiment as Mediator Between Object and Subject” (Goethe, 1792; in Miller, 1995, pp. 11-17) shows his keen awareness of science as a way of interacting with nature. Experiments don’t “prove,” they mediate a relationship. We are interwoven with nature and weaving new fabric when we do science.

There is another dimension to nature evolving in the conversation. Inasmuch as nature—the phenomenon I’m engaging with—has been recognized, worked with, and taken up into the human mind, it is appearing in a new form. Nature finds a new expression through the process of human knowing. This may seem to be a radical idea, but it is actually just a description of the process itself (see Steiner, 1894/1999). Unfortunately, most of us are held captive by the notion of the world “out there”, separate from us “in here.” The moment we wake up to the fact that we are part of the world and engaging in a conversation with her to get to know her (and ourselves) better, the captivity of a dualistic world view ends. We are freed to engage as participants in the world.

7) This realization helps us to see one more facet of science-as-conversation: I become aware that I am taking on a responsibility. I’m engaging in the world and whatever the outcome of the conversation, it will bear in part my stamp. I put to rest once and for all the comfortable specter of something called “value-free” science engaged in by some detached being called a scientist. Science is all about participation, and I can’t distance myself from the process and its results.

So much for an introductory overview. The idea of science as a conversation grows out of the doing. But once you’ve become conscious of it, it becomes a kind of scientific conscience—an inner guide—for all further work: Am I aware enough of the process? Is a back-and-forth occurring? Am I listening or pushing an agenda? When your work becomes infused with a circumspect attitude of questioning wedded to a strong desire to engage in the phenomena, you can see what Goethe wanted to express with the phrase, “delicate empiricism.” And you can also understand why he added that its practice belongs to a “highly evolved age,” since it is dependent on transformation within the human being. Goethe’s science involves the consciously evolving scientist.

In the following sections of this essay I will try to illuminate more fully the process of doing Goethean science by way of an example. I’ll show how science-as-conversation can unfold and also discuss additional important features of a Goethean approach.

Engaging the Conversation
When I moved to the Northeast twelve years ago I met new habitats, plants and animals. Early in March I was down in a wooded swamp and saw some strange looking plants—maroon and yellow, fist-sized buds that emerged directly out of the icy ground. They had a beautifully curved and pointed form, something like the hats of elves you see in children’s books. Nothing else in the wetland showed any sign of spring in this grey, frozen world. I was captivated—I began my journey to get to know the skunk cabbage.
So that’s how it begins. Something captivates your interest, and you move towards it. For me this meant returning to the skunk cabbage again and again—in all seasons and at different times of the year. I did this over a period of six years, in which time I also read everything I could get my hands on regarding skunk cabbage (which wasn’t a whole lot). What was my purpose? What was my goal? I know of no other answer than to say: I wanted to get to know the skunk cabbage. I felt it as a riddle that drew me towards it.
I didn’t have a particular hypothesis that I wanted to test. I didn’t want to “explain” the plant or its features in terms of competition or survival. Since I had been practicing the Goethean approach for many years, it wasn’t very hard to avoid the trap and narrowing effect of wanting to explain. But when I’d go out with other people, I’d often be asked question such as: Why does it flower so early? Why does it heat up? Why do its leaves grow so large? I could tell them what some scientists thought and perhaps point out alternative explanations. But I’d also say, and that is more important, that before we can see whether it’s even meaningful to ask such questions, we have to get to know the plant much more intimately. And just those “why” questions can hinder us from doing so.

So the conversation began. I began building up a picture of the plant’s development through the year. To do this I made lots of detailed observations. The plant itself is a unity that transforms over time. I had a vague sense of that unity, but I had to get to know it by bringing together discrete observations. As Henry Bortoft puts it, “the way to the whole is into and through the parts” (Bortoft, 1996, p. 12). In every part you discover new phenomena and new questions arise. One danger here is that you let yourself get pulled into an endless process of analysis, where the whole and, successively, each part dies into further analysis. The only antidote I know to this problem—which is a major problem for all of science—is to periodically disengage from analysis, step back and ask yourself: How does this all relate to the skunk cabbage? I began my journey wanting to get to know the plant better. So I have to continually try to place all the knowledge I gain through engaging in the parts (analysis) back into the context of the plant as a whole (synthesis).
I tried to go down to the wetland to observe every week or two. I sketched the plants to help me look more carefully, and I also took photographs. Sometimes I would have specific questions: How are the flowers really shaped? Are the skunk cabbages that grow at the wetland edge any different from those that grow in the wetter core area? But at other times I would purposely go out without a content focus, with the attitude “let’s see what comes today.” I know some of my most interesting observations—such as discovering bees on their first outing of the season were visiting skunk cabbage flowers—came when I was walking with an “unframed mind.” As Thoreau writes,

Be not preoccupied with looking. Go not to the object, let it come to you . . . What I need is not to look at all—but a true sauntering of the eye. (13 Sept. 1852, Journal 5: 343-44; in Dassow Walls, 1999)
Of course, we never engage in observing with a fully unframed or unfocused mind. We always have some kind of intentionality or attentiveness that orients us toward the world (or we fall asleep). But we can work on developing a kind of open, listening awareness—which is what Thoreau is pointing to with his expression “sauntering of the eye”—that is very different from going out with a specific question one wants to answer.
Through the weaving interplay of focused observation and open awareness you come to know the phenomena. It’s the most time-extensive part of doing science and most fruitful when you can move back and forth between the poles of focus and opening outward.
Exact Sensorial Imagination and Living Understanding
After I go out and observe, I make a point of actively re-membering the observations. With my mind’s eye I inwardly recreate the form of the leaves, I inwardly sense the colors and the smells, and so on. This process of conscious picture building is what Goethe called “exact sensorial imagination” (Goethe, 1824; in Miller, 1995, p. 46). It entails using the faculty of imagination to experience more vividly what I have observed. I try to be as precise as possible—and will often notice where I haven’t observed carefully enough, which I try to do the next time I’m out. When you do this kind of conscious picture building, you grow more and more connected to what you’re observing.

But there’s something else. The plant begins to reveal itself as a process. When we begin observing, we have many separate images, and we have to overcome separateness to begin seeing the plant as the living creature it is. The life of a plant plays itself out in the ongoing unfolding and decay of organs (leaves, stalks, flowers, etc.). We are presented with a drama of transformation that we can enter into. But we can’t enter into it through observation alone. We need to utilize our faculty of imagination to connect within ourselves what is already connected within the plant. As Goethe writes:

If I look at the created object, inquire into its creation, and follow this process back as far as I can, I will find a series of steps. Since these are not actually seen together before me, I must visualize them in my memory so that they form a certain ideal whole.
At first I will tend to think in terms of steps, but nature leaves no gaps, and thus, in the end, I will have to see this progression of uninterrupted activity as a whole. I can do so by dissolving the particular without destroying the impression. (Goethe, 1795, in Miller 1995, p. 75)
So to begin to grasp the flow of life and its specific qualities in skunk cabbage, you have to work to make your thinking fluid (process-oriented) and dynamic. In Goethe’s words, “If we want to approach a living perception (Anschauung) of nature, we must become as mobile and flexible as nature herself” (Goethe, 1807; translation by CH; in Miller p. 64).

I’d now like to give what you might call a report on my conversation with skunk cabbage. But it actually wants to be more than that. It’s an attempt to give a portrayal, to paint a picture in words that will let you see something of the unique qualities of this plant. I hope to give you a glimpse of another being, although I’m all too aware of my inability to adequately express what I have met.

Skunk Cabbage—A Portrayal
To find the first spring plant in flower in our region—the edge of the Taconic range southeast of Albany, New York—you have to get out before it feels much like spring at all. It’s March, the ground is still frozen, and frost comes nearly every night. Walking through the woods down a soft slope, you see the grey and brown tree trunks, a coloring mirrored in the ground litter of leaves from the previous year. There is no green. Not only the temperature but the whole mood of the woods is cool.
At the base of the slope there is wooded wetland—a flat expanse in Figure 1. A group of skunk cabbage spathes and leaf buds in March. which patches of ice spread around islands of bushes and small trees. In this still, quiescent world, little centers of emerging life are visible, the first sign of early spring—four-to-six-inch-high, hood-like leaves that enclose the flowers of skunk cabbage. (See figures 1 and 2; all drawings are by the author.)

Both color and shape are striking. Some leaves are deep wine-red or maroon, while in others this background coloring is mottled with dots or stripes of yellow or yellow green. The shape is hard to describe: it is like a spiral, sculpted hood drawn around itself, leaving only a narrow opening on one side. Not only the colors, but also the specific shapes are manifold; some are pointed and strongly twisted, others rounder and squat. As my eye sweeps over the twenty or thirty plants before me, my gaze is brought into a spiraling movement when it tries to rest upon any single specimen. The deep color is warm, the sculpted form alive.
Looking at skunk cabbage on one of the first warm, sunny March afternoons (it’s maybe 50° F) with the light shining through the leafless trees and shrubs and illuminating the wetland floor, I often sense for the first time that spring is on its way. On such days I’ve even seen the first bees of the year flying in and out of the skunk cabbage hoods.

The hood is, in botanical terms, a highly modified leaf called a spathe. The spathe wraps around itself to form a space that encloses a spherical head of flowers, called a spadix (see figure 2). The spathe functions as a bud that holds and protects the flower head when it emerges out of the ground. But it is a bud that never unfolds. When the flowers are full in bloom, they are still enwrapped by the spathe. You can see the flower head only by peeking inside the narrow opening in the spathe.

The roundish flower head (about 2 cm in diameter) has a spongy consistency like the spathe itself. It consists of numerous small, tightly packed individual flowers (see figure 3). They have no petals, which make up the showy part of the flower in most plants. Rather, they have four inconspicuous, fleshy, straw-colored sepals (which in many plants form the bud leaves enclosing the petals) that never really unfold.

The flowers “bloom” when the stamens grow up between and above the sepals and release their pale yellow pollen. Following this the style grows out of the middle of each flower to be pollinated by insects carrying pollen from other flower heads. All of this happens within the enclosing spathe. These first flowers of spring never leave their protective enclosure.

A couple of times I’ve been lucky enough to see spathes growing up through a thin layer of ice, the ice melted around the spathe in a circular form. This is an indication of skunk cabbage’s remarkable capacity to produce heat when flowering. If you catch the right time, you can put your finger into the cavity formed by the spathe and when you touch the flower head, your finger tip warms up noticeably. I have measured the temperature at the base of the flower head numerous times and have found it to be as warm as a 61° F when the surrounding air temperature was only 32° F. Biologist Roger Knutson found that skunk cabbage flowers produce warmth over a period of 12-14 days, remaining on average 20° C (36° F) above the outside air temperature, whether during the day or night. During this time they regulate their warmth, as a warm-blooded animal might!

Physiologically the warmth is created by the flower heads breaking down substances while using a good deal of oxygen. The rootstock and roots store large amounts of starch and are the likely source of nutrients for this break down. The more warmth produced, the more substances and oxygen consumed. Knutson found that the amount of oxygen consumed is similar to that of a small mammal of comparable size.
We must imagine that as the spathe grows out of the usually frozen ground, the flower head heats up and the warmth radiates outward. While in this heating phase, the flowers bloom, releasing pollen and being pollinated by insects. Not only can you see the first insects flying around between skunk cabbages, but you also find beetles and spiders crawling around within the warm enclosures of the spathes. You can even discover a spathe opening veiled with a spider net.
Shunk cabbage spathe; the front part has been cut off to show the flower head (spadix).

The flowers also release a noticeable odor at this time. On a calm day coming down to the wetland you can smell a lightly pungent, somewhat skunk-like odor. If you put your nose to the opening of a spathe or break off a small piece and crush it between your fingers, the scent is markedly stronger. Small flies and other insects are attracted to the flowers by the smell.

Due to the warmth production, a constant circulation of air in and out of the spathe occurs. From the flower head, warmth is generated and the air moves up and outward, while cooler air is drawn into the spathe. A vortex is formed with air streaming along the sculpted, curved surfaces of the spathe. In a habitat with numerous skunk cabbages, a microcosm of flowing warmth and odiferous air is created in which the first insects of spring fly.

This is the world of skunk cabbage over a number of weeks in March and sometimes into April: on the one hand, the enclosed, protected life just peering out of the still wintry earth and a flower that remains in a bud; on the other hand, the active, warmth-, movement-, and scent-emanating organism that creates a unique environment for the first stirrings of insect life. Skunk cabbage mirrors the quality of early spring—flowering at ground level in a bud that doesn’t open, while at the same time helping to create the environment for its own development.

When the spathe emerges out of the ground, there is often the tip of a large bud next to it, sticking an inch or two out of the ground (see figure 4). This bud contains all the leaves that will develop on the plant and is often already visible in the previous fall, having developed in the summer and overwintered. Only when the spathe slowly begins to wilt does this tightly-packed bud of leaves begin to grow. It grows longer than the spathe and is shaped like the tip of a spear. Then, when the days begin to get noticeably warmer at the end of April and into May, the bud unfolds rapidly. It’s clear that skunk cabbage now needs outer warmth to develop. The bright green leaves unfold in a beautiful spiraling pattern. Each leaf is rolled in upon itself and at the same time enwraps the next leaf. It’s the closest thing to an archetypal process of unfolding you can imagine.

Gradually a large, funnel-shaped rosette of long-stalked leaves forms. The largest leaves reach three, occasionally four feet in length. By mid-May this surge of growth peaks and the wetland is flooded with green patches of skunk cabbage. The leaves are oblong in shape and have a long leaf stalk. The leaf stalks are thick, but also easy to crush. They have no woody fibers and consist primarily of air and water inlaid with soft plant matter. This consistency extends, untypically, into the flowering part of the plant: both spathe and flower are watery and spongy. By contrast, think of the distinct difference you find in a wild rose between the hard, prickly, wooded stems carrying divided, fibrous leaves on the one hand and the refined, almost rarified petals on the other.

A crushed leaf also exudes a skunk-like odor, and ingested leaf juice calls forth a strong inflammatory reaction in the mouth and esophagus of human beings. Skunk cabbage not only produces its own warmth, it also stimulates warmth processes. Few creatures eat the leaves. I’ve seen leaf buds and also spathes that have been nibbled upon. In one instance the wetland was covered with a late March snow and tracks of wild turkeys led up to the buds, which apparently they had eaten from. Early Swedish settlers in Pennsylvania gave skunk cabbage the name “bear-weed,” since bears were known to feast on the buds and leaves.
In our area the leaves of the trees and bushes unfold in May and a homogenous dark green canopy has formed by mid-June. At this time the leaves of skunk cabbage begin to decay. They don’t dry up and fall onto the ground to become part of the leaf litter that is slowly decomposed by fungi over the next year. Skunk cabbage has its own characteristic way of decaying. The leaves get small holes in them, begin to hang down and parts turn black and somewhat slimy. Eventually the leaves sink to the ground and dissolve. This dissolution occurs rapidly, so that already by the end of July or early August the leaves are gone. You only find a few remnants of the bases of the leaf stalks. What dominated the appearance of the wetland in May has disappeared in August.

As strange as this way of decomposing at first seems, after studying the plant more intensively you begin to see how it fits with other characteristics. While growing, a plant is in its most fluid state. It then forms hard fibers, which, in biochemical terms, is a process of condensation and drying out. When the plant dies even more water is lost, and decay of the woody fibers sets in. Skunk cabbage stays in the watery phase; its substances don’t condense and dry out. Therefore the dying leaves appear to evaporate, since they are mostly water, and almost no dry matter is left on the ground to decay. Skunk cabbage unfolds rapidly and disappears rapidly.

The Whole in the Part
Through this sketch I want to give you at least a partial view into the life of skunk cabbage. (For a more a more complete portrayal, see Holdrege Figure 4. The development of skunk cabbage from early Spring to July, when its leaves begin to dissolve. We can see its unique characteristics, but we can also see more. We can see how the various aspects of a plant’s development, also in relation to its habitat, express certain unified tendencies.

When I see such relations, I sense that I’m finally beginning to actually meet and understand the plant, seeing through all the details to its unity and coherence. But at the same time, it’s a new kind of territory. The terrain is difficult. Where before I had seemingly solid objects—the different parts of the plant in their shape, size, consistency, etc.—now I’m dealing with the qualities that are expressed through these parts. And qualities aren’t things. It’s a real struggle to express these qualities so that someone else can see what you’re talking about.

Skunk cabbage expresses in many of its features a bud-like quality. Its flowers are housed in the large bud-like spathe, never extending out of this mantle. Skunk cabbage blooms in a bud at the time of year in which most flowers, later to unfold, are still tightly encased in their buds. Its flowers never reach the full light of day and the parts of a flower that normally unfold are highly reduced. While the petals are missing altogether, the small, fleshy sepals, all tightly packed into a sphere, open only enough to let the stamens and style slightly protrude. The flower head remains a big, fleshy bud within the bud-like spathe.
When the plant grows, leaf upon leaf unwraps out of the large bud. Since the stem of the plant never elongates but remains in the ground, the leaves never grow apart. Instead, they form a funnel-shaped rosette. The rosette is only fully open, that is, the leaves spread out in horizontal fashion, when the leaves are dying. Their life is in the unfolding bud; being unfolded signals decay. And skunk cabbage never stops laying down new buds, so that an established plant contains within it the spathe and leaf buds, not just for the next season, which is typical for perennials, but for a number of years to come.

We can go further and view these bud-like qualities in connection with skunk cabbage’s dependency upon a wet environment. When I asked students in a field ecology course how they could determine where the wetland begins, they would often answer, “skunk cabbage shows you.” Its roots need to be bathed in muddy soil throughout the year.

Skunk cabbage is not only dependent upon water, but also brings qualities of water—such as fluidity, movement, continuity, and the tendency to form surfaces—to expression. Early in spring, when stasis reigns in the wetland, skunk cabbage brings movement and life. The spathe grows out of the frozen ground and expresses in its form the congealed movement of spiraling surfaces. With the help of water, solid starch transforms into fluid sugar sap. Rising from the roots and rootstock, the fluid sugar is utilized in all growth processes. Moreover, large amounts of sugar are broken down to produce the warmth in the flower head. This transformation from solid starch to flowing sugar sap to radiating warmth is mediated by water and brings movement into the dormant landscape of early spring.

The radiating warmth in turn brings the air and insects into motion. When the leaves grow, you can almost see the water moving out of wet soil through the roots into the leaves, swelling and unfolding them. The leaves have a large, undulating surface that is like a conduit for water. They don’t have a thick, waxy cuticle that prevents transpiration. As a result, water is continually flowing out of the soil, into and through the plant, and into the air, increasing the humidity of the lower layer of air in the wetland.

When skunk cabbage leaves decompose, they don’t dry up and crumble; they dissolve. With few fibers, they consist mainly of water and air, as do the spathe and flowers, and disintegrate into these elements. Skunk cabbage embodies wateriness, growing and decaying in its watery world.

The Unity of the Organism
As the process of knowing unfolds—the conversation with the plant—you begin to see the unity of the plant. The remarkable thing is that when you build exact pictures over and over, moving from one characteristic to the next, patterns emerge. You begin to recognize how the characteristics express a whole—the unity begins to reveal itself. When you go back to characteristics you have studied before, they may suddenly express the unity you have discovered through another part. You have an “aha” experience in which you recognize connections between what previously appeared to be separate facts. You see a common watery, bud-like quality in the form and consistency of spathe, flower head and leaves. Skunk cabbage reveals the fluid quality of water in the way it unfolds and decays, as well as in its undulating, flowing forms. And in all of these characteristics you can see a vivid picture of early spring—a plant that is bud-like in so many ways and yet unfolds to bring the first life and movement to a still slumbering habitat.

While you have to work hard to get to such insights, you cannot force them. If you try to, you can be pretty sure they won’t come. This is a stage of knowing where you have to learn patience. You prepare the ground, but the moment of seeing always involves an act of grace. Or maybe we could just say: we have to wait till the world speaks. As Goethe describes:

I persist until I have discovered a pregnant point from which much may be derived, or rather—since I am careful in my work and observations—one which yields several things, offering them up of its own accord. If some phenomenon appears in my research, and I can find no source for it, I let it stand as a problem. This approach has proven quite advantageous over the years. When I found I could not solve the riddle of the origin and context of some problem, I had to let it lie for a long time; but at some moment, years later, enlightenment came in the most wonderful way. (Goethe, 1823, in Miller, 1995, p. 41; translation modified by CH)
Once you’ve come to understand a plant in this way, you never encounter it with the remark, “oh, that’s just a skunk cabbage.” Rather, you meet it with expectation and interest, wondering what else it has to show you. And this attitude begins to inform your overall orientation toward nature. Any other plant, beetle, or bird you see appears immediately as a riddle and not a thing. You know that each carries within itself—as you’ve experienced in skunk cabbage—a whole, unique world that’s just waiting to be disclosed.

Doing Goethean Science
One of the problems with talking about doing Goethean science is that the essence is in the doing itself. That’s why I have given what might seem to be undue attention to a concrete example. I’ll conclude this essay by pulling back from the example and presenting some of the key features of a Goethean approach as I’ve described it.

Preparing the ground—A new attitude of mind
All science has its roots in human questioning and the search for understanding. As far as I can see, most people who are drawn to Goethe’s approach to science recognize in it a way of understanding nature that can take them beyond the boundaries of what has developed as mainstream science. At the heart of the Goethean approach is the realization that as a scientist I must develop new capacities in order to do nature justice in my work. It’s not just a matter of developing new instruments or refining the intellect, but developing new ways of knowing that can illuminate to the phenomena in ways that science has largely neglected (or even deemed unscientific).

Out of this awareness arises the striving to develop a gentle sensibility that does not violate the phenomena in the process of getting to know them. It’s an active conversation, but one in which I hope the other—as something in its own right—can reveal itself. As Goethe writes, the scientist strives to “find the measure for what he learns, the data for judgment, not in himself but in the sphere of what he observes” (Goethe, 1792, in Miller, 1995, p. 11). This is the attitude that Goethe suggests with his expression “delicate empiricism” and that I’ve described above through the metaphor of conversation. As a kind of underlying intentionality it permeates all the work one does and grows as a capacity the more one works.

For this attitude of mind to actually inform every fiber of one’s work means removing many obstacles—habits of the mind that have us search for single causes, for general theories, for reductive explanations. In the end it means, in the words of Owen Barfield, ridding oneself of all “residues of unresolved positivism” (in Sugarman, 1976, pp. 13-15). This is not an easy task, and one that never ends. Yet the striving (and some success!) is absolutely necessary if the world is to show herself from new sides. In his description of phenomenology as a new way of viewing, Edmund Husserl is speaking out of the soul of the Goethean scientist:

That we should set aside all previous habits of thought, see through and break down the mental barriers which these habits have set along the horizons of our thinking, and in full intellectual freedom proceed to lay hold on those genuine philosophical problems still awaiting completely fresh formulation which the liberated horizons on all sides disclose to us—these are hard demands. Yet nothing less is required. What makes the appropriation of the essential nature of phenomenology, the understanding of the peculiar meaning of its form of inquiry, and its relation to all other sciences so extraordinarily difficult, is that in addition to all other adjustments a new way of looking at things is necessary, one that contrasts at every point with the natural attitude of experience and thought. To move freely along this new way without ever reverting to the old viewpoints, to learn to see what stands before our eyes, to distinguish, to describe, calls, moreover, for exacting and laborious studies. (Husserl, 1913/1962, p. 39)

Practicing Goethean science
The Riddle. This is the beginning of any investigation. I am drawn to a particular phenomenon and want to get to know it better. I’ve met something in the world that is a riddle I want to attend to. And because each person has a different biography—carries a unique world within herself—and is drawn to different features of the world, there is an endless and beautiful array of possible questions and areas of focus. I have colleagues who are physicists, chemists, ecologists, botanists, and zoologists. They are not only investigating different realms of phenomena, but take somewhat different approaches based on who each of them is. This does not make the work “subjective,” but merely points to the fact that in any scientific endeavor the subject as a particular being is actively at work. And the riddle that draws a particular person is the beginning of a pathway into the world that is specific, but can be shared with others. (We live, after all, in one world.)
Into the Phenomena. This is exploration, getting to know the phenomena. As Goethe wrote in connection with his work in optics:

The greatest accomplishments come from those who never tire in exploring and working out every possible aspect and modification of every bit of empirical evidence, every experience. (Goethe, 1792; in Miller, 1995, p. 15)

You really have to get to know the phenomena you’re dealing with from as many sides as possible. If you’re doing experiments, then it’s a matter of varying them in a methodical way to build up a rich picture. It’s not about proving (or falsifying) a particular hypothesis (Ribe and Steinle, 2002). In studying a living organism, you want to gain a many-sided picture of the life of the organism and its relation to its environment. In this work you make your own observations, but you also interact with and utilize the work of others (which may entail doing a good amount of separating out of theory and interpretation). Here is where a research community evolves. As Goethe writes,

What applies in so many other human enterprises is also true here [in science]: the interest of many focused on a simple point can produce excellent results…. I have always found working together with others so advantageous that I have every reason to continue doing so.
(Goethe, 1792, in Miller pp. 12-13; transl. modified by CH)
Since the phenomena are endless, this work is also without end. I can never get “all the facts,” but my goal is also not an encyclopedic totality of information. It’s more that I never cease to be interested in what the phenomena—perhaps some unassuming, seemingly esoteric detail—may reveal to me about the world. In my own work I often find that we don’t know nearly enough about the animal or plant I’m studying. I do extensive literature searches and speak with experts, am enriched by all I find, but am usually left feeling I’d love to know much more. I also discover how theory-burdened so much of science is, with a small number of facts being marshaled to apparently support grand ideas.

Exact picture building. While getting to know the phenomena, I intensify my experience through exact picture building—Goethe’s exact sensorial imagination. At first this may be a completely separate activity from being out and observing. I retreat from observation and quietly build up a precise inner picture of what I’ve experienced. The more I’ve done this, the more I find that my observing and perceiving becomes dynamic and full of life. I become active while perceiving, following inwardly the shapes, colors, smells, or tones as I observe. I sculpt the shapes while looking. This is where you notice how the picture-building as an exercise becomes integrated into your concrete interaction with the phenomena. You begin to see more intensely.
This work helps me to enter more deeply into the phenomenal world. It also gives my experience of the organism more continuity. The connectedness of all the details within the organism itself also becomes a connectedness within me.

I have come to see this activity of exact sensorial imagination to be the counter pole to theory building in traditional science. In both cases one is inwardly very active. But in exact sensorial imagination, the work of picturing—building images and letting the one transform into the other—keeps us close to the phenomena. We close the gaps that are given through our discrete observations and in this sense go beyond what perception gives us, but our whole intention is to take in the world. In theory building, I construct a picture or concept out of myself which fits the phenomena to a greater or lesser degree. Often, because we can know our own thoughts in such a transparent way, we become more interested in the theory than in the things the theory is supposed to explain. The tendency to reify concepts—which Whitehead called the fallacy of misplaced concreteness—is widespread in contemporary science (Whitehead, 1925/1967; especially chapters three and four). Theories tend to take on a life of their own and we may begin to see only the theory in the things. In this way a theory can become, in Goethe’s words, “lethal generality.” Concrete picture building has the cathartic effect of re-orienting our attention to the phenomena, while dissolving hard-and-fast ideas through mental molding and remolding.

Seeing the Whole. This is the “step” that we’ve been preparing for in all the other work. Or, stated more accurately, this is what can reveal itself in the course of one’s striving to get to know the phenomena. As I said above, it is an experience of seeing unifying relations, which may or may not happen during any investigation. When it occurs, it fills you with the greatest joy and you realize: “now I am knowing.” We can use the word intuition here as long as we don’t think of something vague, but rather a nondiscursive form of seeing connections that is comparable to the experience one can have most purely in mathematical insight.
In the example of skunk cabbage I showed how you can see a bud-like, watery quality in various characteristics of the plant. Its wholeness speaks through its parts and its relation to the environment. If you imagine this mode of cognition applied on a larger scale, you come to what Goethe writes about as the “archetypal phenomena” in his color work, or the “type” (Typus) and the archetypal animal or plant (Urpflanze; Urtier) in his biological studies. (He also speaks of “entelechy,” or “idea.”) What term one uses is much less important than the quality of knowing itself. Here’s how he describes the whole process, brilliantly condensed into a few sentences, that leads to a seeing that goes beyond, but is fully rooted in, empirical observation:

If I look at the created object, inquire into its creation, and follow this process back as far as I can, I will find a series of steps. Since these are not actually seen together before me, I must visualize them in my memory so that they form a certain ideal whole.
At first I will tend to think in terms of steps, but nature leaves no gaps, and thus, in the end, I will have to see this progression of uninterrupted activity as a whole. I can do so by dissolving the particular without destroying the impression….
If we imagine the outcome of these attempts, we will see that empirical observation finally ceases, inner beholding of what develops begins, and, at last, the idea can be brought to expression. (Goethe, 1795; in Miller, 1995, p. 75; translation modified by CH)
If you don’t pay attention to the process and context out of which Goethe speaks about bringing an idea to expression, you could imagine “idea” to be something abstract or bloodless (“just another theory”). But it’s not. It has much more the nature of seeing a being. That’s why Goethe was so distraught when Schiller reacted to his description of the archetypal plant by stating, “that is not an observation from experience. It is an idea. ” Goethe responded: “Then I may rejoice that I have ideas without knowing it, and can even see them with my own eyes”(Goethe, 1817; in Miller, 1995, pp. 18-21).

So when Goethe says there is “delicate empiricism which makes itself utterly identical with the object, thereby becoming true theory” (Goethe, 1829, in Miller, 1995, p. 307), then “theory” is to be understood in the sense of the ancient Greeks as a “seeing of the mind” or “beholding” and not as the abstract “theory” as we know it from modern science. If we use the term “idea” then we must think of an idea that Goethe could, in the end, see sensibly/supersensibly in every plant. Reflecting on his botanical studies, Goethe writes near the end of his life, A challenge…hovered in my mind at that time [1787] in the sensuous form of a supersensuous plant archetype [Urpflanze]. I traced the variations of all the forms as I came upon them. In Sicily, the final goal of my [Italian] journey, the conception of the original identity of all plant parts had become completely clear to me; and everywhere I attempted to pursue this identity and to catch sight of it again.…. Only a person who has himself experienced the impact of a fertile idea…will understand what passionate activity is stirred in our minds, what enthusiasm we feel, when we glimpse in advance and in its totality something which is later to emerge in greater and greater detail in the manner suggested by its early development. Thus the reader must surely agree that, having been captured and driven by such an idea, I was bound to be occupied with it, if not exclusively, nevertheless during the rest of my life. (Goethe, 1831; in Mueller, 1989, p. 162).

So finding the fertile idea is at once a completion of a process and the beginning of a new one. As an end, it brings us full circle to a more conscious glimpse of the being—the riddle—that formed the starting point of the investigation. As a beginning, it is the soil for further work and vital new insights. Goethe’s approach to science is itself a fertile idea that still has ample life to unfold.

______________
References
Bortoft, H. (1996). The wholeness of nature. Great Barrington, MA: Lindisfarne Press.
Cahill, J., et al. (2001). The herbivory uncertainty principle: Visiting plants can alter herbivory. Ecology, 82, 307-312.
Gelbart, W. (1998). Databases in genomic research. Science, 282, 659-661.
Goldstein, K. (1939/1995). The organism. New York: Zone Books.
Holdrege, C. (1996). Genetics and the manipulation of life: The forgotten factor of context. Great Barrington, MA: Lindisfarne Press.
Holdrege, C. (2000). Skunk cabbage (Symplocarpus foetidus). In Context, 4, 12-18. Available online: http://www.natureinstitute.org/pub/ic/ic4/skunkcabbage.htm
Holdrege, C. (2003). The giraffe’s short neck. In Context, 10, 14-19. Available: http://www.natureinstitute.org/pub/ic/ic10/giraffe.htm
Holdrege, C. (2004). Science evolving: The case of the peppered moth. In D. Rothenburg & W. Pryor (Eds.), Writing the future: Progress and evolution. Cambridge: MIT Press. (An earlier version of this paper is available online: http://www.natureinstitute.org/txt/ch/moth.htm)
Husserl, E. (1913/1962). Ideas: General introduction to pure phenomenology. London: Collier Books.
Miller, D. (Ed). (1995). Goethe: Scientific studies (Collected Works Vol. 12). Princeton: Princeton U. Press.
Mueller, Bertha (Trans.). (1989). Goethe’s botanical writings. Woodbridge, CT: Ox Bow Press.
Ribe, N., & Steinle, F. (2002). Exploratory experimentation: Goethe, land, and color theory. Physics Today, July 2002. Available: http://www.physicstoday.com/pt/vol-55/iss-7/p43.htm
Steiner, R. (1894/1999). The philosophy of freedom. London: Rudolf Steiner Press.
Sugarman, S. (Ed). (1976). The evolution of consciousness. Middletown, CT: Wesleyan U. Press.
Talbott, S. (2004). Toward an ecological conversation. In S. Talbott, In the belly of the beast: Technology nature and the human prospect. Ghent, NY: The Nature Institute. Available online: http://www.netfuture.org/2002/Jan1002_127.html
Dassow Walls, L. (Ed). (1999). Material faith: Henry David Thoreau on science. New York: Houghton Mifflin Co.
Whitehead, A.N. (1925/1967). Science and the modern world. New York: The Free Press.
Author’s note: Correspondence concerning this article should be addressed to Craig Holdrege, The Nature Institute, 20 May Hill Road, Ghent, New York 12075. E-mail: craig@natureinstitute.org.
Craig Holdrege 53
“Scotland” Photograph by Syrie Kovitz

Man or Matter, Chapter II: Where Do We Stand To-day?

In the year 1932, when the world celebrated the hundredth anniversary of Goethe's death, Professor W. Heisenberg, one of the foremost thinkers in the field of modern physics, delivered a speech before the Saxon Academy of Science which may be regarded as symptomatic of the need in recent science to investigate critically the foundations of its own efforts to know nature.1 In this speech Heisenberg draws a picture of the progress of science which differs significantly from the one generally known. Instead of giving the usual description of this progress as 'a chain of brilliant and surprising discoveries', he shows it as resting on the fact that, with the aim of continually simplifying and unifying the scientific conception of the world, human thinking, in course of time, has narrowed more and more the scope of its inquiries into outer nature.

'Almost every scientific advance is bought at the cost of renunciation, almost every gain in knowledge sacrifices important standpoints and established modes of thought. As facts and knowledge accumulate, the claim of the scientist to an understanding of the world in a certain sense diminishes.' Our justifiable admiration for the success with which the unending multiplicity of natural occurrences on earth and in the stars has been reduced to so simple a scheme of laws - Heisenberg implies - must therefore not make us forget that these attainments are bought at the price 'of renouncing the aim of bringing the phenomena of nature to our thinking in an immediate and living way'.

In the course of his exposition, Heisenberg also speaks of Goethe, in whose scientific endeavours he perceives a noteworthy attempt to set scientific understanding upon a path other than that of progressive self-restriction.

'The renouncing of life and immediacy, which was the premise for the progress of natural science since Newton, formed the real basis for the bitter struggle which Goethe waged against the physical optics of Newton. It would be superficial to dismiss this struggle as unimportant: there is much significance in one of the most outstanding men directing all his efforts to fighting against the development of Newtonian optics.'

There is only one thing for which Heisenberg criticizes Goethe: 'If one should wish to reproach Goethe, it could only be for not going far enough - that is, for having attacked the views of Newton instead of declaring that the whole of Newtonian Physics-Optics, Mechanics and the Law of Gravitation - were from the devil.'

Although the full significance of Heisenberg's remarks on Goethe will become apparent only at a later stage of our discussion, they have been quoted here because they form part of the symptom we wish to characterize. Only this much may be pointed out immediately, that Goethe - if not in the scientific then indeed in the poetical part of his writings - did fulfill what Heisenberg rightly feels to have been his true task.2

We mentioned Heisenberg's speech as a symptom of a certain tendency, characteristic of the latest phase in science, to survey critically its own epistemological foundations. A few years previous to Heisenberg's speech, the need of such a survey found an eloquent advocate in the late Professor A. N. Whitehead, in his book Science and the Modern World, where, in view of the contradictory nature of modern physical theories, he insists that 'if science is not to degenerate into a medley of ad hoc hypotheses, it must become philosophical and enter upon a thorough criticism of its own foundations'.

Among the scientists who have felt this need, and who have taken pains to fulfil it, the late Professor A. Eddington obtains an eminent position. Among his relevant utterances we will quote here the following, because it contains a concrete statement concerning the field of external observation which forms the basis for the modern scientific world-picture. In his Philosophy of Physical Science we find him stating that 'ideally, all our knowledge of the universe could have been reached by visual sensation alone - in fact by the simplest form of visual sensation, colourless and non-stereoscopic'.3 In other words, in order to obtain scientific cognition of the physical world, man has felt constrained to surrender the use of all his senses except the sense of sight, and to limit even the act of seeing to the use of a single, colour-blind eye.

Let us listen to yet another voice from the ranks of present-day science, expressing a criticism which is symptomatic of our time. It comes from the late physiologist, Professor A, Carrel, who, concerning the effect which scientific research has had on man's life in general, says in his book, Man the Unknown: 'The sciences of inert matter have led us into a country that is not ours. … Man is a stranger in the world he has created.'

Of these utterances, Eddington's is at the present point of our discussion of special interest for us; for he outlines in it the precise field of sense-perception into which science has withdrawn in the course of that general retreat towards an ever more restricted questioning of nature which was noted by Heisenberg.

The pertinence of Eddington's statement is shown immediately one considers what a person would know of the world if his only source of experience were the sense of sight, still further limited in the way Eddington describes. Out of everything that the world brings to the totality of our senses, there remains nothing more than mere movements, with certain changes of rate, direction, and so on. The picture of the world received by such an observer is a purely kinematic one. And this is, indeed, the character of the world-picture of modern physical science. For in the scientific treatment of natural phenomena all the qualities brought to us by our other senses, such as colour, tone, warmth, density and even electricity and magnetism, are reduced to mere movement-changes.

As a result, modern science is prevented from conceiving any valid idea of 'force'. In so far as the concept 'force' appears in scientific considerations, it plays the part of an 'auxiliary concept', and what man naively conceives as force has come to be defined as merely a 'descriptive law of behaviour'. We must leave it for later considerations to show how the scientific mind of man has created for itself the conviction that the part of science occupied with the actions of force in nature can properly be treated with purely kinematic concepts. It is the fact itself which concerns us here. In respect of it, note as a characteristic of modern text-books that they often simply use the term 'kinetics' (a shortening of kinematics) to designate the science of 'dynamics'.4

In the course of our investigations we shall discover the peculiarity in human nature which - during the first phase, now ended, of man's struggle towards scientific awareness - has caused this renunciation of all sense-experiences except those which come to man through the sight of a single colour-blind eye. It will then also become clear out of what historic necessity this self-restriction of scientific inquiry arose. The acknowledgment of this necessity, however, must not prevent us from recognizing the fact that, as a result of this restriction, modern scientific research, which has penetrated far into the dynamic substrata of nature, finds itself in the peculiar situation that it is not at all guided by its own concepts, but by the very forces it tries to detect. And in this fact lies the root of the danger which besets the present age.5

He who recognizes this, therefore, feels impelled to look for a way which leads beyond a one-eyed, colour-blind conception of the world. It is the aim of this book to show that such a way exists and how it can be followed. Proof will thereby be given that along this way not only is a true understanding achieved of the forces already known to science (though not really understood by it), but also that other forces, just as active in nature as for example electricity and magnetism, come within reach of scientific observation and understanding. And it will be shown that these other forces are of a kind that requires to be known to-day if we are to restore the lost balance to human civilization.

*

There is a rule known to physicians that 'a true diagnosis of a case contains in itself the therapy'. No true diagnosis is possible, however, without investigation of the 'history' of the case. Applied to our task, this means that we must try to find an aspect of human development, both individual and historical, which will enable us to recognize in man's own being the cause responsible for the peculiar narrowing of the scope of scientific inquiry, as described by the scientists cited above.

A characteristic of scientific inquiry, distinguishing it from man's earlier ways of solving the riddles of the world, is that it admits as instruments of knowledge exclusively those activities of the human soul over which we have full control because they take place in the full light of consciousness. This also explains why there has been no science, in the true sense of the word, prior to the beginning of the era commonly called 'modern' - that is, before the fifteenth century. For the consciousness on which man's scientific striving is based is itself an outcome of human evolution.

This evolution, therefore, needs to be considered in such a way that we understand the origin of modern man's state of mind, and in particular why this state of mind cannot of itself have any other relationship to the world than that of a spectator. For let us be clear that this peculiar relationship by no means belongs only to the scientifically engaged mind. Every adult in our age is, by virtue of his psycho-physical structure, more or less a world-spectator. What distinguishes the state of man's mind when engaged in scientific observation is that it is restricted to a one-eyed colour-blind approach.

*

'Death is the price man has to pay for his brain and his personality' - this is how a modern physiologist (A. Carrel in his aforementioned book, Man the Unknown) describes the connexion between man's bodily functions and his waking consciousness. It is characteristic of the outlook prevailing in the nineteenth century that thinking was regarded as the result of the life of the body; that is, of the body's matter-building processes. Hence no attention was paid at that time to the lonely voice of the German philosopher, C. Fortlage (1806-81), who in his System of Psychology as Empirical Science suggested that consciousness is really based on death processes in the body. From this fact he boldly drew the conclusion (known to us today to be true) that if 'partial death' gave rise to ordinary consciousness, then 'total death' must result in an extraordinary enhancement of consciousness. Again, when in our century Rudolf Steiner drew attention to the same fact, which he had found along his own lines of investigation, showing thereby the true role of the nervous system in regard to the various activities of the soul, official science turned a deaf ear to his pronouncement.6 To-day the scientist regards it as forming part of 'unknown man' that life must recede - in other words, that the organ-building processes of the body must come to a standstill - if consciousness is to come into its own.

With the recognition of a death process in the nervous system as the bodily foundation of consciousness, and particularly of man's conceptual activities, the question arises as to the nature of those activities which have their foundation in other systems, such as that of the muscles, where life, not death, prevails. Here an answer must be given which will surprise the reader acquainted with modern theories of psycho-physical interaction; but if he meets it with an open mind he will not find it difficult to test.

Just as the conceptual activity has as its bodily foundation the brain, with the nervous appendages, so it is volitional activity which is based on processes taking place in the muscular region of the body and in those organs which provide the body's metabolism.

A statement which says that man's will is as directly based on the metabolic processes of the body, both inside and outside the muscles, as is his perceiving and thought-forming mind on a process in the nerves, is bound to cause surprise. Firstly, it seems to leave out the role commonly ascribed to the so-called motoric part of the nervous system in bringing about bodily action; and secondly, the acknowledgment of the dependence of consciousness on corporeal 'dying' implies that willing is an unconscious activity because of its being based on life processes of the body.

The first of these two problems will find its answer at a later stage of our discussion when we shall see what entitles us to draw a direct connexion between volition and muscular action. To answer the second problem, simple self-observation is required. This tells us that, when we move a limb, all that we know of is the intention (in its conceptual form) which rouses the will and gives it its direction, and the fact of the completed deed. In between, we accompany the movement with a dim awareness of the momentary positions of the parts of the body involved, so that we know whether or not they are moving in the intended manner. This awareness is due to a particular sense, the 'sense of movement' or 'muscular sense' - one of those senses whose existence physiology has lately come to acknowledge. Nothing, however, is known to us of all the complex changes which are set into play within the muscles themselves in order to carry out some intended movement. And it is these that are the direct outcome of the activity of our will.

Regarding man's psycho-physical organization thus, we come to see in it a kind of polarity - a death-pole, as it were, represented by the nerves including their extension into the senses, and a life-pole, represented by the metabolic and muscular systems; and connected with them a pole of consciousness and one of unconsciousness - or as we can also say, of waking and sleeping consciousness. For the degree of consciousness on the side of the life-pole is not different from the state in which the entire human being dwells during sleep.

It is by thus recognizing the dependence of consciousness on processes of bodily disintegration that we first come to understand why consciousness, once it has reached a certain degree of brightness, is bound to suffer repeated interruptions. Every night, when we sleep, our nervous system becomes alive (though with gradually decreasing intensity) in order that what has been destroyed during the day may be restored. While the system is kept in this condition, no consciousness can obtain in it.

In between the two polarically opposite systems there is a third, again of clearly distinct character, which functions as a mediator between the two. Here all processes are of a strictly rhythmic nature, as is shown by the process of breathing and the pulsation of the blood. This system, too, provides the foundation for a certain type of psychological process, namely feeling. That feeling is an activity of the soul distinct from both thinking and willing, and that it has its direct counterpart in the rhythmic processes of the body, can be most easily tested through observing oneself when listening to music.

As one might expect from its median position, the feeling sphere of the soul is characterized by a degree of consciousness half-way between waking and sleeping. Of our feelings we are not more conscious than of our dreams; we are as little detached from them as from our dream experiences while these last; what remains in our memory of past feelings is usually not more than what we remember of past dreams.

This picture of the threefold psycho-physical structure of man will now enable us to understand the evolution of consciousness both in individual life and in the life of mankind. To furnish the foundation of waking consciousness, parts of the body must become divorced from life. This process, however, is one which, if we take the word in its widest sense, we may call, ageing. All organic bodies, and equally that of man, are originally traversed throughout by life. Only gradually certain parts of such an organism become precipitated, as it were, from the general organic structure, and they do so increasingly towards the end of that organism's life-span.

In the human body this separation sets in gently during the later stages of embryonic development and brings about the first degree of independence of bones and nerves from the rest of the organism. The retreat of life continues after birth, reaching a certain climax in the nervous system at about the twenty-first year. In the body of a small child there is still comparatively little contrast between living and non-living organs. There is equally little contrast between sleeping and waking condition in its soul. And the nature of the soul at this stage is volition throughout. Never, in fact, does man's soul so intensively will as in the time when it is occupied in bringing the body into an upright position, and never again does it exert its strength with the same unconsciousness of the goal to which it strives.

What, then, is the soul's characteristic relationship to the world around at this stage? The following observations will enable us to answer this question.

It is well known that small children often angrily strike an object against which they have stumbled. This has been interpreted as 'animism', by which it is meant that the child, by analogy with his experience of himself as a soul-filled body, imagines the things in his surroundings to be similarly ensouled. Anyone who really observes the child's mode of experience (of which we as adults, indeed, keep something in our will-life) is led to a quite different interpretation of such a phenomenon. For he realizes that the child neither experiences himself as soul-entity distinct from his body, nor faces the content of the world in so detached a manner as to be in need of using his imagination to read into it any soul-entities distinct from his own.

In this early period of his life the human being still feels the world as part of himself, and himself as part of the world. Consequently, his relation to the objects around him and to his own body is one and the same. To the example of the child beating the external object he has stumbled against, there belongs the complementary picture of the child who beats himself because he has done something which makes him angry with himself.

In sharp contrast to this state of oneness of the child's soul, in regard both to its own body and to the surrounding world, there stands the separatedness of the adult's intellectual consciousness, severed from both body and world. What happens to this part of the soul during its transition from one condition to the other may be aptly described by using a comparison from another sphere of natural phenomena. (Later descriptions in this book will show that a comparison such as the one used here is more than a mere external analogy.)

Let us think of water in which salt has been dissolved. In this state the salt is one with its solvent; there is no visible distinction between them. The situation changes when part of the salt crystallizes. By this process the part of the salt substance concerned loses its connexion with the liquid and contracts into individually outlined and spatially defined pieces of solid matter. It thereby becomes optically distinguishable from its environment.

Something similar happens to the soul within the region of the nervous system. What keeps the soul in a state of unconsciousness as long as the body, in childhood, is traversed by life throughout, and what continues to keep it in this condition in the parts which remain alive after the separation of the nerves, is the fact that in these parts - to maintain the analogy - the soul is dissolved in the body. With the growing independence of the nerves, the soul itself gains independence from the body. At the same time it undergoes a process similar to contraction whereby it becomes discernible to itself as an entity distinguished from the surrounding world. In this way the soul is enabled, eventually, to meet the world from outside as a self-conscious onlooker.

*

What we have here described as the emergence of an individual's intellectual consciousness from the original, purely volitional condition of the soul is nothing but a replica of a greater process through which mankind as a whole, or more exactly Western mankind, has gone in the course of its historical development. Man was not always the 'brain-thinker' he is to-day.7 Directly the separation of the nerve system was completed, and thereby the full clarity of the brain-bound consciousness achieved, man began to concern himself with science in the modern sense.

To understand why this science became restricted to one-eyed, colour-blind observation we need only apply to the human sense system, in particular, what we have learnt concerning man's threefold being.

Sharply distinguished by their respective modes of functioning though they are, the three bodily systems are each spread out through the whole body and are thus to be found everywhere adjacent to each other. Hence, the corresponding three states of consciousness, the sleeping, dreaming and waking, are also everywhere adjacent and woven into one another. It is the predominance of one or other which imparts a particular quality of soul to one or other region of the body. This is clearly shown within the realm of sense activity, itself the most conscious part of the human being. It is sufficient to compare, say, the senses of sight and smell, and to notice in what different degree we are conscious of the impressions they convey, and how differently the corresponding elements of conception, feeling and willing are blended in each. We never turn away as instinctively from objectionable colour arrangement as from an unpleasant smell. How small a part, on the other hand, do the representations of odours play in our recollection of past experiences, compared with those of sight.8 The same is valid in descending measure for all other senses.

Of all senses, the sense of sight has in greatest measure the qualities of a 'conceptual sense'. The experiences which it brings, and these alone, were suitable as a basis for the new science, and even so a further limitation was necessary. For in spite of the special quality of the sense of sight, it is still not free from certain elements of feeling and will - that is, from elements with the character of dream or sleep. The first plays a part in our perception of colour; the second, in observing the forms and perspective ordering of objects we look at.

Here is repeated in a special way the threefold organization of man, for the seeing of colour depends on an organic process apart from the nerve processes and similar to that which takes place between heart and lungs, whilst the seeing of forms and spatial vision depend upon certain movements of the eyeball (quick traversing of the outline of the viewed object with the line of sight, alteration of the angle between the two axes of sight according to distance), in which the eye is active as a sort of outer limb of the body, an activity which enters our consciousness as little as does that of our limbs. It now becomes clear that no world-content obtained in such more or less unconscious ways could be made available for the building of a new scientific world-conception. Only as much as man experiences through the sight of a single, colour-blind eye, could be used.9

*

If we would understand the role of science in the present phase of human development, we must be ready to apply two entirely different and seemingly contradictory judgments to one and the same historical phenomenon. The fact that something has occurred out of historical necessity - that is, a necessity springing from the very laws of cosmic evolution - does not save it from having a character which, in view of its consequences, must needs be called tragic.

In this era of advanced intellectualism, little understanding of the existence of true tragedy in human existence has survived. As a result, the word 'tragedy' itself has deteriorated in its meaning and is nowadays used mostly as a synonym for 'sad event', 'calamity' 'serious event', even 'crime' (Oxford Diet.). In its original meaning, however, springing from the dramatic poetry of ancient Greece, the word combines the concept of calamity with that of inevitability; the author of the destructive action was not held to be personally responsible for it, since he was caught up in a nexus of circumstances which he could not change.

This is not the place to discuss why tragedy in this sense forms part of man's existence. It suffices to acknowledge that it does and, where it occurs, to observe it with scientific objectivity.

Our considerations, starting from certain statements made by some leading scientific thinkers of our time, have helped us not only to confirm the truth inherent in these statements, but to recognize the facts stated by them as being the outcome of certain laws of evolution and thereby having an historic necessity. This, however, does not mean that man's scientific labours, carried out under the historically given restrictions, great and successful as these labours were and are, have not led to calamitous effects such as we found indicated by Professor Carrel. The sciences of matter have led man into a country that is not his, and the world which he has created by means of scientific research is not only one in which he is a stranger but one which threatens to-day to deprive him of his own existence. The reason is that this world is essentially a world of active forces, and the true nature of these is something which modern man, restricted to his onlooker-consciousness, is positively unable to conceive.

We have taken a first step in diagnosing man's present spiritual condition. A few more steps are required to lead us to the point where we can conceive the therapy he needs.
________________
1 This address and another by the same author are published together under the common title, Wandlungen in den Grundlagen der Naturwissenschaft ('Changes in the foundations of Natural Science'). Heisenberg's name has become known above all by his formulation of the so-called Principle of Indeterminacy.
2 See, in this respect, Faust's dispute with Mephistopheles on the causes responsible for the geological changes of the earth. (Faust II, Act 4)
3 See also Eddington's more elaborate description of this fact in his New Pathways in Science. The above statement, like others of Eddington's, has been Contested from the side of professional philosophy as logically untenable. Our own further discussion will show that it accords with the facts.
4 Both words, kinematics and kinetics, are derivatives of the Greek word kinein, to move. The term 'kinematic' is used when motion is considered abstractly without reference to force or mass. Kinetics is applied kinematics, or, as pointed out above, dynamics treated with kinematic concepts.
5 These last statements will find further illustration in the next two chapters.
6 First published in 1917 in his book Von Seelenrätseln.
7 Homer's men still think with the diaphragm (phrenes). Similarly, the ancient practice of Yoga, as a means of acquiring knowledge, shows that at the time When it flourished man's conceptual activity was felt to be seated elsewhere than in the head.
8 This must not be confused with the fact that a smell may evoke other memories by way of association.
9 For one who endeavours to observe historical facts in the manner here described, it is no mere play of chance that the father of scientific atomism, John Dalton, was by nature colour blind. In fact, colour blindness was known, for a considerable time during the last century, as 'Daltonism', since it was through the publication of Dalton's self-observations that for the first time general attention was drawn to this phenomenon.



--Ernst Lerhs

http://www.gutenberg.org/cache/epub/5641/pg5641.html