Rescher y Gadamer: dos visiones complementarias de los límites de la ciencia

Rescher and Gadamer: two complementary views on the limits of science

Author: Alfredo Marcos. University of Valladolid (Spain)
Text exhibited at: seminar of group of research Science, Reason and Faith
Date of publication: April 16, 2013

Recording of the second part and the final colloquium (WATCH VIDEO)

"It was Kant who distinguished between frontier (Grenz) and boundary (Schrank), thus highlighting an important nuance in identifying the scope of a theory. Boundaries are, by their very definition, movable; in the course of time, a boundary can be shifted and the confines drawn by an empirical research exceeded. In contrast, the boundary has a structural epistemological connotation. In other words: the limit sample the point of arrival fixed by the transcendental framework of our knowledge; hence it cannot be displaced by the improvement of the tools of research or by the accumulation of empirical data ".

summaryThe question of the limits of science leads us to think about the metaphor of the limit itself. We discover that there are different types of limits, as well as different possible actions in relation to them. In short, there are limits that shape science and others that constrain it. The former must be respected, the latter overcome. We explore these limits following the suggestions of Rescher and Gadamer. The views of both are complementary. The former establishes the limits of science by looking from within science; the latter from the outside, looking at the cultural position of science.

Palabras core topic: Limits of science; Nicholas Rescher; Hans-Georg Gadamer; Hans-Georg Gadamer

Abstract: The question concerning the limits of science leads us to think about the very metaphor of the limit. We found that there exist different kind of limits and different possible actions in relation to them. In sum, there are limits that configure science and others that constrain it. The first ones must be respected, the latter overcome. We explore these limits following the suggestions of Rescher and Gadamer. These authors present complementary views. The first one approaches the question from inside the science, while the second one sees the limits of science from an exterior cultural point of view.

Key words: Limits of science; Nicholas Rescher; Hans-Georg Gadamer

Author's biographical note: Alfredo Marcos is Full Professor of Philosophy of science at the University of Valladolid (Spain). He teaches courses and lectures at other universities in Spain, Argentina, Italy, France, Mexico, Colombia and Poland. He has recently published the books: Ciencia y acción (F.C.E., Mexico, 2012; translated into Italian and Polish) and Postmodern Aristotle (Cambridge Scholars Publishing, UK, 2012); as well as the chapter: 'Bioinformation as a triadic relation', in G. Terzis & R. Arp (eds.), Information and Living Systems (M.I.T. Press, 2011).

E-mail address and website:
amarcos@fyl.uva.es
http://www.fyl.uva.es/~wfilosof/webMarcos

1. Introduction

The question of the limits of science asks us to think about the metaphor of the limit itself (section 2) and suggests that we then try to think science from that metaphor (section 3). To think science from the metaphor of the limit, I propose that we dialogue with Nicholas Rescher (1928-) and Hans-Georg Gadamer (1900-2002). Both authors have written lucidly about the limits of science. Both were born in Germany, but Rescher has developed his degree program in Pittsburgh and can be considered an "Anglo-Saxon" philosopher of science, in the pragmatist tradition; whereas Gadamer is a "continental" one, influenced by Heidegger and focused on hermeneutics. Their two perspectives are complementary. The former thinks about the limits of science from the inside, from the point of view of one who is primarily interested in science. The second, on the other hand, looks at science from the outside, from a more general interest in civilisation as a whole. Finally, I will include a concluding summary to recapitulate the main ideas acquired during the research (section 4).

2. Thinking the limit

The word "boundary" comes from the Latin limes-limitis. In this language it refers to the path that separates one farm from another. This origin provides semantic features that are worth considering. The boundary shapes and constitutes the estate. Without it, there is no proper entity. The boundary of an estate distinguishes it from others, separates it, but at the same time communicates it. As a road, it has a certain physical, geographical thickness, it is not a mere geometric line. Its thickness allows us to think of it with blurred zones, ambiguous territories, suitable for partnership or for conflict. The boundary is not simply an entity "in sight", ready for our contemplation. It is also an entity "at hand", inviting action to walk, to explore, to go beyond.... What is more, it is an entity that arises from our action. As Antonio Machado's well-known verse says: "the path is made by walking" *(1).

A boundary, in principle, is not an abstract line, but a concrete entity immersed in a context of action, relative to an agent. The agent provides a space of possibilities, of possible actions, of attitudes, of goals and duties, of feelings and values that depend on a certain ontology. One can feel at ease within certain limits that do justice to the nature of things. Or, on the contrary, we may feel the limits as wrong constrictions, perhaps unfairly imposed. In the latter case, our attitude impels us to push the boundaries. The boundary is seen either as something positive, valuable, which contributes to the constitution of an entity, or as something negative, which unjustly constrains it.

As can be seen, along with the concept of limit comes a universe of attitudes, feelings and values connected with ontological presuppositions. That is why some limits are experienced by the agent as self-realisation or perfection - "become what you are", wrote Pindar- and others as constriction or frustration. We will see later how important these considerations are when we discuss the limits of science.

The boundary is originally a spatial entity. But the word undergoes various metaphorical displacements, firstly towards the sphere of time. Thus, the Dictionary of the Royal Spanish Academy (DRAE) includes this meaning: "The extreme at which a certain time arrives". Like "end" or "term", the word "limit" soon begins to play in the fourth dimension. It also undergoes a similar shift into the world of abstractions. There, paths lose their geographical thickness, their "actuation", to remain either mere impassable geometrical lines, or mathematical limits by definition unreachable. Finally, let us note the shift of the concept of limit towards the sphere of capacities. Again, the DRAE: "Extreme that the physical and mental can reach". Now the capacities of any entity can also have limits.

If the journey we have made along the Latin limes has been enriching, we can expect the same from a journey along the Greek horion, which dictionaries translate as "limit" or "frontier". Now, the temporal metaphor is obvious. Our counted hours are limits of time. The Hours (Horai), in Greek mythology, were the goddesses who brought order and regularity to nature, who managed the beneficial succession of the seasons. From this perspective, it is precisely the limits that separate us from chaos and confusion.

Not far from horion are the Greek words horama and horasis (sight, vision), as well as horizo(limit), in which we already glimpse our "horizon". The horizon is an elusive, unreachable visual limit, but it affects our action as goal. Our sight marks the direction in which we walk, that is, towards the front. This is why the notion of horizon is not only visual, but also "agential", in the same way as the notion of frontier, which is related to that of front. Both, horizon and frontier, have been applied profusely to science.

To give just a couple of significant and mutually contradictory examples: Vannevar Bush, then director of the Office of Scientific Research and Development, sent a report to the President of the United States in July 1945 with the significant degree scroll deScience, The Endless Frontier. association For his part, Bentley Glass, who was president of the American Association for the Advancement of Science (AAAS), sent a speech graduate "Science: Endless Horizons or Golden Age?" to the AAAS in 1970, which concluded with the assertion that there are no longer infinite horizons for science.

Once passed through Latin, Greek words such as horion have served to name other forms of limitation, such as shores. When we look back to the limits of science, we might do well to remember Newton's famous quotation : "I have only been like a child playing on the sea-shore [...] while the great ocean of truth lay undiscovered before me" *(2). Or the well-known account of St. Augustine from which this image could well come.

But before we focus on science, let us explore yet another source of our notion of limit. It is the Greek word peras. This term, probably related to "period" and "perimeter", brings to mind the concept of apeiron (the unlimited), itself surrounded by connotations of all kinds subject. Negative, because of its indeterminacy and difficult understanding. Positive, because of its fertility and potency.

We now see that thinking about the limits of science is not simply a question of whether or not science has limits. Perhaps, in one sense of the metaphor, it does, and in another, it does not. Rather, the question of the value of limits will have to be asked. And the answer to the latter question will inexorably open up new questions about our actions. We will try to approach all this work in dialogue with Rescher and Gadamer.

3. Thinking science from the metaphor of the limit

3.1. The limits of science: an inside view

Nicholas Rescher devoted a book to the question of the limits of science * (3). In it he argues, first of all, that science is not everything, that outside science there are perfectly valid and rational forms of knowledge and praxis. There are areas in which we have cognitive and practical interests that fall completely outside the province of science. The author speaks here of "territorial restrictions" *(4). In his own words: "There is no doubt that natural science is subject to incapacities outside its domain. We must recognise that a number of important evaluative and cognitive problems lie completely outside the realm of science as we know it" *(5). To think otherwise would be to subscribe to scientistic ideology.

As we see, Rescher thinks in terms of territories, with their limits or boundaries. Science occupies one of these domains, but beyond it there is also life. This subject of boundaries, which form the profile of science on the background of the world of life (lebenswelt), we could call constitutive boundaries *(6). In my opinion, these are blurred boundaries, more "geographical" than "geometrical", because there will always be cognitive contents and actions of dubious assignment. The constitutive boundary is traversable, for there must be passage from science to life and vice versa, and positive, for rather than constraining, it shapes science.

Regarding the speech between science and the rest of the world of life, Rescher's position is only partially satisfactory. He admits that there must be outlets from science to life. In fact, he states on several occasions that the criterion for assessment of science can only be its usefulness internship. But he does not accept that there is traffic in the opposite direction, from the world of life to science. Science, he states in a curious in crescendo, is autonomous, self-sufficient, sovereign *(7). But if the boundary delimits as much as it communicates, if it is to be permeable and passable, it will have to be so in both directions. Science will have to accept requests, support and also, why not, restrictions from other spheres.

Finally, let us note that science has to be seen within the world of life, as an integral part of it, and not in juxtaposition to it. Science is a part of human action. This integration and connection with the rest of the lifeworld will be discussed below in dialogue with Gadamer. But it is already an important step that Rescher allows us to take, by recognising the existence of the constitutive limits of science, and the legitimacy of some knowledge and practices situated outside them.

Let us now turn to the second boundary subject . To do so, we will only consider problems specific to science, leaving aside those that fall outside its territory. Well, one could think that there are theoretical reasons to affirm that science will never give a complete solution to all the problems in its domain. These theoretical reasons would mark a second limit subject , the theoretical limits of science. Outside these would be the scientific problems that science, for theoretical reasons, will never be able to address, let alone solve. On the inside we would have the problems that science, at least in theory, could successfully address. This would be the territory of theoretically possible science. Well, Rescher argues at length that such theoretical limits do not exist *(8). For Rescher, theoretically possible science would simply be identified with science.

However, this thesis has been widely debated. Let us take an example that illustrates the subject of objections that can be raised against it. Chaitin has shown, inspired by the work of Gödel and Turing, that the randomness of a mathematical sequence cannot be proved, that it is undecidable. This has consequences for the natural sciences, as the physicist Fernando Sols*(9) of the Complutense University of Madrid recently pointed out. From Chaitin's demonstration, it could be inferred that the question of the presence or absence of purpose in nature is also undecidable, since we will never know whether a sequence of natural phenomena occurs randomly or is directed towards an end. What would we say then, that the problem of chance and purpose is alien to the natural sciences, or that it belongs to the domain of the natural sciences but is unapproachable for purely theoretical reasons? If we opt for the former, Rescher's position holds, but not if we affirm the latter.

In any case, what interests us here is not so much whether or not theoretical limits exist, but the very concept of theoretical limits. These limits, if they exist, would be more restrictive than constitutive, and probably more "geometrical" than "geographical". Rescher helps to clarify this by distinguishing theoretical limits from practical limits. These constitute the third subject of the limits of science. Science falls short of many of the problems within its domain for practical reasons. For example, the capabilities of CERN's Large Hadron Collider mark a practical limit. It is the largest and most powerful particle accelerator in the world. If an experiment exceeds the capabilities of this facility, it simply cannot be done for the time being.

Often such limits, Rescher suggests, can be reduced to economic terms. But practical limits are not always translatable into money. For example, there are historical moments when the mathematics that would require a part of natural science is not available at available. This implies a practical limit for natural science that does not depend only on economic investment. Practical limits may also be linguistic, moral, social, political, ecological.... These limits must in some cases be overcome and in others respected.

In short, theoretically possible science is limited by practical factors, which determine what science is practically possible. We know that some of the theoretically possible science will never be possible on internship. But we cannot know in advance what that part will be. According to Rescher, there is no way of specifying which concrete problems will be left out of the scientific development . Problems that are beyond the practical limits today may not be so tomorrow. We are facing a limit of subject horizon. It is always there, but it shifts as we move forward. This subject limit responds in its displacement to what Rescher calls the Kantian principle of the propagation of questions: "Science is born as a project of self-transcendence. It embodies an inner impulse that always pushes beyond the limits of the capacity of the present time" *(10). We are faced with blurred and shifting limits. They function as challenge and frontier. They invite us to transgression, but this transgression is never fulfilled. Practical limits occupy an intermediate place in terms of their positive (constitutive) or negative (constrictive) condition. They are between constitutive and theoretical limits, on the one hand, and fallibility limits, on the other.

Limits due to fallibility include our inoperability staff, organisational and institutional shortcomings, our lack of attention or work or honesty, the mistakes we inevitably make given our human nature - all too human! These are what separate practically possible science from effective science. The difference between the one and the other is constituted by shortcomings or defects, while effective science is constituted by achievements. These limits of fallibility also have the aspect of a horizon. They are not surmountable in their totality, even if each of them individually is surmountable. In other words, science will always be fallible and unfinished, but none of its concrete errors is dictated by fatality. The attempt to overcome this class of limits is a requirement, since they are limits in a purely negative sense: they are constrictions on the development of science that generate deficiencies.

We have already passed through constitutive limits, theoretical limits, practical limits and limits due to fallibility. The time has come to remember Pindar's maxim: "become what you are". If science is what its constitutive limits mark, but has only become what its fallibility limits mark, the difference between the one and the other could be called Pindar's difference. Bridging this gap is the ultimate, unrenounceable and unattainable task of the scientific business .

All the limits we have encountered so far have a common origin. They derive from science itself, i.e. from its constitution, and from the subject that produces it, the human being and his concrete circumstances, his life-world, his environment (Unwelt). But science has an intentional character, it refers to something outside itself, to something outside even the world of life, and produces knowledge about that something. That something to which science aspires to refer is the world itself (Welt). And this pole goal of science also imposes limitations on science. Let us call them objective limits. Objective limits are insurmountable and have a positive character. They cannot count as shortcomings of science. That is, our science can neither be more precise nor more complex than nature itself, its depth and extent cannot exceed the dimensions of nature itself. If in nature there is indeterminacy, in our science there will be uncertainty. We will never be able to predict what nature itself has not determined.

Objective limits cannot be exceeded. Moreover, they cannot be reached either. This is due to the rest of the limitations we have identified. For example, science is constitutively conceptual, and the real - as Rescher states - can never be conceptually exhausted *(11). Or, to put it in Shakespeare's classic terms: "There are more things in heaven and earth, Horatio, than your Philosophy dreams of" *(12).

It is true that a creative and metaphorical use of language can reduce the distance between physis and logos, but without ever reaching the point of identity. We can only aspire to increase the similarity between being and thinking *(13). The identity between our conceptual system and reality is one of those unreachable and elusive limits.

Let us add a final remark about objective limits that has implications for the relationship between the natural and human sciences. The attempt to subsume the entire knowledge about human beings into the natural sciences may lead to an undesirable overreach. According to Rescher, "to exaggerate the aspirations of science to the point of claiming to have 'all the answers' about the human condition, the meaning of life, or matters of social policy, is to take a dangerous step [...] This inflated view of capacities invites scepticism and hostility as a consequence of the frustration of expectations that is its inevitable consequence" *(14). This frustration is caused by the attempt to overcome limits which may be of a character goal, for "man", Rescher continues, "is a member not only of the natural order of things, but also of the specifically human order" *(15).

3.2. The limits of science: an outsider's perspective

"My whole philosophy is nothing but phronesis" *(16)

Technoscience has constitutive limits, as we have seen * (17). And beyond them there is intelligent life. Beyond technoscience we encounter the rest of the spheres of the sphere of knowledge, such as art and morality. And the sphere of knowledge is, in turn, outlined against the background of the world of life, to which it undoubtedly belongs, and in which it has to coexist with other respectable human realities, as committed as technoscience can be to the true knowledge and to rational action. Cultural traditions, emotions, Philosophy, religion, politics, Education, speech and many other areas of human life, ranging from everyday experience and common sense to, for example, sport, are, like technoscience, part of the world of life.

According to the scientistic conception, the limits of science coincide with the limits of rationality: "Rationality," Gadamer summarises, "is seen in the context of science and confined within its limits" * (18). This is what the neo-positivists of the Vienna Circle called, in their manifesto, "the scientific worldview" ("Wissenschaftliche Weltauffassung") *( 19).

But if what we are looking for is more truthful and nuanced thinking about the relations between technoscience and the rest of human life, we have to look for it far from scientism, in some philosophical tradition that shows more respect for other territories of the human other than technoscience.

We have to overcome the modern idea of an absolutely autonomous science. We must reintegrate the sphere of knowledge, and technoscience in particular, into the world of life. Technoscience must interact with its environment. It then needs a healthy environment, made up of entities worthy of respect. It is only one facet of our life, bordering on many others. In other words, one of the functional limits of technoscience is that it is not enough on its own to provide the basis for an entire civilisation, a complete way of life.

To a large extent, this is the message of Gadamer's Philosophy . This limit does not imply a deficiency of technoscience, it is nothing negative, except for those who, with a scientistic mentality, pretend to base everything on technoscience. Gadamer's Philosophy is not anti-scientific. It is, on the other hand, anti-scientific. This is the main reason for choosing Gadamer as an interlocutor in the present text. But there are more reasons. Gadamer's arguments are very close to those of many other contemporary philosophers, whose echoes we will perceive alongside the Gadamerian voice. I am referring to other contemporary thinkers, both in the Anglo-Saxon and continental tradition, and mainly to Heidegger, Arendt, Husserl, Dewey, Wittgenstein, Popper, Kuhn, Polanyi, Toulmin, MacIntyre, Putnam, Habermas and Ricoeur. To dialogue with Gadamer will therefore be, in a way, to dialogue with many of these authors' ideas as well. Like most of them, Gadamer identifies the limits of technoscience, points out its insufficiency as the sole basis for a civilisation, denounces the excesses of scientism, and does so without falling into the ranks of the anti-scientific mentality, without indulging in relativism or irrationalism, without entering into what he himself calls "the shadow of nihilism" *(20).

Gadamer's hermeneutics can be read as a theory about the limits of science, according to Stefano Marino*(21). Science does not exhaust the territory of truth, of knowledge or of experience, not everything can be achieved by its means*(22). His thought is a critique of the scientistic hybris that seeks to push science beyond its constitutive limits. As a complement to this pars destruens, a pars construens appears in his work, which seeks the revaluation of other areas of "human experience of the world in general", which "go - according to Gadamer - beyond the limits of the concept of method established by modern science"*(23)."One cannot ignore - he states - such a 'knowledge' in whatever form it is expressed: in religious or proverbial wisdom, in works of art or in philosophical thought"*(24). It is a matter of "understanding the variety of experiences, whether of aesthetic, historical, religious or political consciousness"*(25). These experiences are beyond the limits of science and science must not attempt to colonise them. They must be respected and considered in themselves, since they are by nature irreducible to the methods of science.

We may now ask on what concept of science Gadamer's proposed delimitation of boundaries depends. In order to characterise modern science, Gadamer refers to a few concepts with clearly Cartesian and Baconian roots. The first of these is the concept of method. Modern science is primarily a method. It is a method with a vocation for universality, automatism and certainty. This notion is so important to Gadamer that it becomes part of the degree scroll of his landmark work, Truth and Method. Around the 17th century, a new form of civilisation was born, a new form of life (Lebensdform), defined practically univocally by the emergence of a new notion of science*(26). The essence of this notion is contained in a single word: method. The effect of method is objectification, i.e. the shaping or delimitation of the object, the transformation of (part of) reality into an object.

It is therefore an objectifying method. It objectifies by delimitation. Science, then, not only has limits, but, at a deeper level, it is a limit, it is born of a process of delimitation. In the first place, we intend to draw a boundary between subject and object, in the style of the Cartesian separation between res cogitans and res extensa. We thus divide reality into two parts, and place one in front of the other. One of them is an object for the other, and inevitably also an obstacle. It is that which is not subject and which resists the subject. From this disposition of things, the attitude of control, domination and planning immediately emerges as procedures to reintroduce the subject into objective reality. This is the new way of life of the modern subject. The subject that has been separated from the object returns to it as dominator. "The 'Objekt' or 'Gegenstand'," says Gadamer, "is defined by a 'method' that prescribes how reality is to be turned into an object. Thus, the aim of the methodological research of the object consists essentially in the breaking of the resistance of 'objects' and the mastery of their processes"*(27). Thus science, which supposedly knows with Cartesian objectivity, and technology, which will bring Baconian control over the object, are intertwined. This opens the way to today's technoscience.

The possible overreach or hybris of technoscience can already be clearly sensed. This will occur when we try to impose the objectifying method and the attitude of domination on the whole of reality. We overreach ourselves, we push technoscience beyond its constitutive limits, when we accept - in Gadamer's words - that "nothing can be scientifically investigated or truly understood unless it conforms to the procedures of the method. Objectivity thus specifies in this sense the very limits of our knowledge: what we cannot objectify we cannot know either"*(28). This movement can be seen, indistinctly, as an unjustified extension of technoscience, or as an unjustified reduction of reality. Scientism and reductionism go hand in hand. Whether we look at it one way or the other, the result is the same, the identification of the limits of the human knowledge with those of the scientific method, and the consequent attempt to base all our action, all our relationship with reality, on the application of the scientific knowledge .

We know the consequences. In epistemic terms, there is a replacement of objective truth by subjective certainty*(29). In the practical, an attempt to artificialise all that is natural. What begins as an objectifying movement becomes an immense subjectivisation of reality. We have already set up instructions for the malaise of our culture. But before going in depth into the chapter on the malaise of modernity, allow me to make the idea of objectification by limitation a little more precise. The methodical splitting of the real into subject and object is concretised in other splits. We separate the primary from the secondary qualities, we carefully delimit the quantitative from the qualitative, we leave aside, of course, any emotional evocation, any aesthetic quality, we focus on the "facts" apart from the values. The scientific method seems to require it. We leave in brackets, from the real, all that is aesthetic, emotional, qualitative, axiological... in order never to recover it, to deny it or simply to exclude it. Or to try to forcibly reduce it to the parameters of the objectifying method.

Let us turn now to the question of the malaise of our Western civilisation in its modern version, which manifests itself through multiple symptoms, which became visible especially during the last century. I will cite some of them, identified explicitly by Gadamer himself, although each reader could certainly add a few more. We could identify as the main symptom of modern pathologies what Gadamer calls the shadow of nihilism*(30). Under this poetic formula we can include the atmosphere of anxiety that dominates life today, as well as the lack of hope and meaning in life, which techno-science is unable to alleviate. We must also include the void that remains after the dissolution of religion, brought about by the scientistic mentality; a void that technoscience is incapable of filling *(31). According to Gadamer, "the contribution of the scientific Enlightenment reaches an insurmountable limit in the mystery of life and death" * (32). Similarly, Gadamer identifies as pathological symptoms modern voluntarism and relativism *(33), which lead to moral subjectivism *(34) and aesthetic irrationalism *( 35). Alongside these, we have fragmentarism and specialism * (36), individualism, lack of solidarity, the breakdown of the sense of community * (37), and others such as consumerism * (38) or historicism *(39).

We have tried to base our way of life on technoscience, but that means clearly going beyond its constitutive limits. Technoscience is not capable of sustaining a way of life. That is why the modern West has been afflicted by various ailments. If we want to heal our civilisation, if we want our way of life - including techno-science - to survive in its postmodern version, we have to give it other forms instructions. Gadamer proposes the rehabilitation, alongside technoscience, of other areas of knowledge, of human experience and action, as well as a dialogue between all of them.

But especially the German philosopher focuses on the rehabilitation and autonomy of wisdom internship, irreducible to episteme or science. The role of the expert is always important, but the final decision in all our actions, even in those that make up research tenoscientific, belongs rather to wisdom internship *(40). This is the source of our undelegable responsibility. Wisdom internship is formed through one's own practices. For example, it is essential for its training to participate in a certain tradition: "We produce ourselves", Gadamer argues, "to the extent that we understand and participate in the evolution of the tradition" *(41).

In a way that may be provocative for the modern mindset, Gadamer argues for the rehabilitation of the authority of tradition. To avoid any misunderstanding, let us recall that tradition, for Gadamer, is an event at development, not a static entity. Therefore, there is not by any stretch of the imagination a defence of the status quo. What is certain is that the continuity of a given background tradition clearly benefits the rationality of technoscience. Scientific paradigms, as Thomas Kuhn has shown, follow one another in an apparently discontinuous, revolutionary and to some extent disjointed fashion. The commensurability and comparability between them, and consequently the possibility of justifying the rationality of scientific decisions and the progress of science itself, can be questioned. However, the different scientific paradigms are, in fact, comparable in a rational way, as Kuhn himself argues *(42), and they are so thanks to the permanence, underneath the changes, of a communal current of values, practices and wisdom that we may well call tradition and that goes far beyond the limits of science.

All this suggests an inversion: it is not only that technoscience alone is incapable of rationally founding our way of life, but that, on the contrary, technoscience itself maintains its aspiration to rationality thanks to its reliance on certain practices, values and traditional knowledge, proper to a certain way of life. There is a certain way of life that makes the appeal to reason possible. Gadamer proposes, at this point, a reversal of the modern approach, especially Kantian, which proposed founding ethos on reason. In exchange, he recovers the Aristotelian perspective: it is a certain human ethos that allows the development of rationality, including scientific rationality *(43)."The rationality of reason internship -Gadamer argues- receives its normative power not so much from arguments as from what Aristotle called ethos"*(44).

The wisdom internship to which Gadamer refers is thus situated in the Aristotelian tradition, and it would not be at all unfair to identify it with the intellectual virtue of phronesis. In fact, Gadamer states, in clear harmony with the Nicomachean Ethics (1106b 36 ff.; 1144a 35-6), that "there is no phronesis without ethos and no ethos without phronesis" *(45).

4. summary conclusive

We have tried to think science from the metaphor of the limit. To do so, first of all, we have had to delve into the metaphor itself. We have seen the different levels of metaphoricity of metaphor, from the most conventional, which appeal to spatial and temporal limits, to the most metaphorical, which refer to functional limits. The connotations that the different versions of metaphor carry with them are also very diverse. The limit is in a certain sense a positive concept, since it constitutes entities, but it is also negative, insofar as it constrains them, it can be sharp or blurred, fixed or dynamic, permeable or not.

The metaphor of the limit is very useful for thinking about science. But we have seen that it is not really sufficient for this task on its own. It is a metaphor that brings lucidity especially when it is inscribed in a network of metaphors. Ideas such as border, exploration, road, shore, hour, horizon, link, nexus or pore are in the immediate vicinity of the idea of limit, they belong to the same network of metaphors. But we are particularly interested in metaphors of an agential nature. The idea of limit brings us immediately to that of a subject who does things with that limit: respects it, crosses it, reaches it, goes beyond it, pursues it, builds it and explores beyond it.

Once we have explored the metaphor of the limit and its connections with other nearby metaphors, we are now in a position to apply it to the task of thinking about science. To this end, we have made use of the dialogue with Rescher and Gadamer, as complementary authors.

With Rescher, we have identified the limits of science from within science. Thus, we have distinguished, in this order, constitutive, theoretical, practical and fallibility limits. Within the constitutive limits is science as reality and possibility, i.e. all of science. Within the theoretical limits we would find theoretically possible science. Within the practical limits we have practically possible science, and within the fallibility limits we have effective science. The scientific business as a whole is guided by the aspiration to reduce the difference between effective science and science without further ado. It is a matter of bridging what we have called Pindar's gap, so that science becomes what it is. This is an unrenounceable and impossible task, as tragic as it may sound. There are, moreover, objective limits to science, which are marked by the very nature of things and, in particular, by human nature.

The constitutive limits of science, as outlined by Rescher, leave room for other equally respectable human realities. In other words, science is embedded in the world of life, which turns out to be much broader than science itself. We are therefore interested in thinking about the relations of science with its environment, with other aspects of human life. We are interested in looking from the outside to the limits of technoscience. This part of the journey we have made in dialogue with Gadamer.

According to him, technoscience is not enough to found a civilisation, to provide the basis for a way of life. This is one of its limitations. However, Western civilisation, in its modern version, tried to seek its foundation mainly in technoscience, both epistemically and practically. It is from the failure of this attempt that the malaise of the modern West derives. This malaise is expressed in a multitude of symptoms with which we are all familiar, and which Gadamer poetically summarises in the expression "shadow of nihilism". We have broken down some of these symptoms, and we have asked ourselves how to heal them, or at least how to alleviate them.

The most promising of the strategies consists in the rehabilitation of other areas of knowledge, of human action and experience, and especially in the revaluation of wisdom internship or phronesis. Wisdom that supports an ethos, a way of life, in which the appeal to reason and experience has value, and in which techno-science has a place. But it must be remembered that this wisdom internship is, in turn, based on the ethos that it itself contributes to founding. No one should be shocked by such a (hermeneutic?) circle, a reciprocal support between the wisdom internship and the sensible internship . Technoscience, for its part, far from founding a way of life, which is beyond its limits, receives its foundation from wisdom internship.

5. Bibliographical references

Brewster, D. (1855). Memoirs of the Life, Writings, and Discoveries of Sir Isaac Newton. Edinburgh: Thomas Constable and Co.

Carnap, R., Hahn, H. and Neurath, O. (1929). Wissenschaftliche Weltauffassung: der Wiener Kreis. Vienna: Artur Wolf Verlag.

Gadamer, H. G. (1979). Historical Transformation of Reason, in T. F. Geraets (ed.), Rationality Today. La rationalité aujourd'hui. Proceedings of the International Symposium. Ottawa: University of Ottawa Press, pp. 3-14.

Gadamer, H. G. (1983). Reason in the Age of Science. Cambridge, MA and London: MIT Press.

Gadamer, H. G. (1985-1999). Gesammelte Werke. Tübingen: Mohr Siebeck.

Gadamer, H. G. (1986). The Relevance of the Beautiful and Other Essays. Cambridge: Cambridge University Press.

Gadamer, H. G. (1987). The Relevance of Greek Philosophy for Modern Thought. South African Journal of Philosoohy, 6 (2), 39-42.

Gadamer, H. G. (1989). Das Erbe Europas. Beiträge. Frankfurt: Suhrkamp.

Gadamer, H. G. (1993). Über die Verborgenheit der Gesundheit. Aufsätze und Vorträge. Frankfurt a. M.: Suhrkamp. M.: Suhrkamp.

Gadamer, H. G. (1996). The Enigma of Healing: The Art of Healing in a Scientific Age. Stanford: Stanford University Press.

Gadamer, H. G. (1998). Praise of Theory: Speeches and Essays. New Haven-London: Yale University Press.

Gadamer, H. G. (1999). Hermeneutics, Religion and Ethics. New Haven-London: Yale University Press.

Gadamer, H. G. (2000). Towards a Phenomenology of Ritual Language. In L. K. Schmidt (Ed.), Language and Linguisticality in Gadamer's Hermeneutics. Lanham-Oxford: Lexington Books, pp. 19-50.

Gadamer, H. G. (2003). A Century of Philosophy. Hans-Georg Gadamer in Conversation with Riccardo Dottori. London-New York: Continuum.

Gadamer, H. G. (2004). Truth and Method. London-New York: Continuum.

Kuhn, Th. (1977). The Essential Tension. Selected Studies in Scientific Tradition and Change. Chicago: The University of Chicago Press.

Machado, A. (2001). Poesías completas. Barcelona: RBA.

Marcos, A. (2010). Philosophy of human nature, Eikasia. Revista de Philosophy, VI (35), 181-208. [available at: www.revistadefilosofia.com]

Marcos, A. (2010a). Science and action. Una Philosophy internship de la ciencia. Mexico D. F.: Fondo de Cultura Económica.

Marcos, A. (2012). Postmodern Aristotle. Newcastle: Cambridge Scholars Publishing.

Marino, S. (2011). Gadamer and the Limits of the Modern Techno-Scientific Civilization. Bern: Peter Lang.

Misgeld, D. and Nicholson, G. (eds.) (1992). Hans-Georg Gadamer on Education, Poetry and History: Applied Hermeneutics. Albany: SUNY Press.

Rescher, N. (1994). The limits of science. Madrid: Tecnos. [Original edition: The Limits of Science. Berkeley: University of California Press, 1984. Revised edition: The Limits of Science. Pittsburgh: University of Pittsburgh Press, 1999].

Sloterdijk, P. (1998-2004). Sphären: Eine Trilogie. Frankfurt: Suhrkamp.

Sols, F. (2011). Heisenberg, Gödel y la cuestión de la finalidad en la ciencia. discussion paper presented at the Simposio Internacional Ciencia y Religión en el siglo XXI: ¿diálogo o confrontación? Madrid, 10-11 November 2011, Fundación Ramón Areces.

6. Notes

  1. Machado (2001, p. 186).

  2. Brewster (1855, vol. 2, p. 407). The etymological works I have been able to consult do not connect the English term "shore" with the Greek "horion". Their obvious similarity, then, seems to be due to mere coincidence, poetic justice perhaps.

  3. Rescher (1994). The original English edition is from 1984. There is a revised English edition of 1999. I will quote from the 1994 English translation.

  4. Rescher (1994, ch. 12).

  5. Rescher (1994, p. 247).

  6. I will now allow myself to modify the terminology used by Rescher. He speaks of disabilities, limits, incapacities and deficiencies (1999, p. 3). In reality all these terms refer to different kinds of limits. It would be interesting, therefore, if the term itself indicated which class limit it refers to. The terminological change I am adopting is intended to fulfil this function.

  7. Rescher (1994, pp. 246-247).

  8. Rescher (1994, chaps. 6 and 7).

  9. Sols (2011).

  10. Rescher (1994, p. 53).

  11. Rescher (1994, p. 92).

  12. Hamlet, Act I, scene V.

  13. This topic is discussed further in Mark (2012, ch. 6).

  14. Rescher (1994, p. 243).

  15. Rescher (1994, p. 245). For a more extensive treatment of this point, see Marcos (2010).

  16. Gadamer (2003, p. 54).

  17. Science and technology are historically and conceptually distinguishable realities. Degree Nowadays, however, they have reached such a degree of symbiosis that we can properly speak of science and technology as detectivescience. For a philosopher of science such as Rescher, the conceptual distinction between science and technology is very important, which is why we have stuck to it up to this point. For Gadamer, on the other hand, whose interest lies more in hermeneutics than in the Philosophy of science, the relevant entity is rather the conglomerate of science and technology, which justifies that from here on we speak of technoscience (Cf. Gadamer, 1985-1999, vol. 4, p. 247 and Gadamer, 1996, p. 6).

  18. Gadamer (1985-1999, vol. 4, pp 29-30). Unless otherwise specified, Gadamer's literal quotations are his own translations from the English or German version referred to.

  19. Carnap, Hahn and Neurath (1929).

  20. Gadamer (1985-1999, vol. 9, p. 367).

  21. Marino (2011, p. 33, n. 37).

  22. Gadamer (1993, pp. 127-8).

  23. Gadamer (2004, p. xx).

  24. Gadamer (2004, pp. 565-6).

  25. Gadamer (2004, pp. 84-5).

  26. Gadamer (1983, p. 6).

  27. Gadamer (1998, p. 127).

  28. Gadamer (1987, p. 41).

  29. Gadamer (1996, p. 148).

  30. Gadamer (1985-1999, vol. 9, pp. 367-382; vol. 3, p. 407; Misgeld and Nicholson, 1992, p. 114).

  31. Gadamer (1993, p. 197); (1996, p. 159).

  32. Gadamer (1985-1999, vol. 4, p. 293); (1996, p. 67).

  33. Gadamer (1985-1999, vol. 10, pp. 236, 263).

  34. Gadamer (1985-1999, vol. 7, pp. 398-399).

  35. Gadamer (1986).

  36. Gadamer (1985-1999, vol. 10, p. 263).

  37. Gadamer (1985-1999, vol. 10, pp. 235-6); (1989, p. 157).

  38. Gadamer (1985-1999, vol. 4, p. 256); (1996, p. 18).

  39. Gadamer (1985-1999, vol. 10, p. 263).

  40. This idea is discussed in depth in Marcos (2010a).

  41. Gadamer (1985-1999, vol. 1, p. 298); (2004, p. 293).

  42. Kuhn (1977, chap. XIII).

  43. Gadamer (1985-1999, vol. 4, pp. 187-188); (1999, pp. 29, 34-35).

  44. Gadamer (1985-1999, vol. 8, p. 437); (2000, pp. 48-49).

  45. Gadamer (1985-1999, vol. 7, p. 390); (1999, p. 155); cf. Mark (2012, ch. 2). As Gadamer suggests, if we were to look for a Kantian correlate of this subject of wisdom internship, we would have to look first to the Critique of Judgment rather than to either of the other two great critical works (Gadamer, 1985-1999, vol. 10, p. 278).