Вы находитесь на странице: 1из 211

The Material Realization of Science

A Philosophical View on the Experimental Natural Sciences, Developed in Discussion with Habermas

Hans Radder

The Material Realization of Science

ii

The Material Realization of Science


A Philosophical View on the Experimental Natural Sciences, Developed in Discussion with Habermas by Hans Radder

2008 Hans Radder, Amsterdam Originally published in 1988 by Van Gorcum, Assen (ISBN 90 232 2399 3) No parts of this electronic edition may be reproduced in any form, by print, photoprint, microfilm or any other means without written permission from the author.

Translation: Tineke van Putten, Dawn Wolfswinkel

iv

Contents

Preface Introduction

vii 1

Part I

1 1.1 1.2 1.3 1.3.1 1.3.2 1.4 1.4.1 1.4.2 1.5 1.5.1 1.5.2 1.5.3 1.5.4 1.6

Philosophical views of Habermas with respect to the natural sciences Introductory remarks The intention of Habermas epistemology Two fundamental distinctions Purposive-rational and communicative action Communicative action and discourse: two forms of communication The constitution theory and the role of the experiment Objectivity of experience and the categorial structure of the objectdomain The experiment in the natural sciences The consensus theory of truth What is truth? Grounded consensus as the criterion of truth The formal characteristics of discourse The ideal speech situation Objectivity and truth

7 7 8 11 12 14 16 17 21 22 23 26 30 32 34

2 2.1 2.2 2.2.1 2.2.2 2.3 2.3.1 2.3.2

Analysis and criticism Introductory remarks The transcendental method and the role of the experiment On the constitution of objectivity The role of the experiment in Habermas A criticism of Habermas theory of truth On the meaning of truth The inadequacy of the criterion of truth

37 37 37 38 43 45 45 47

Part II

59

3 3.1 3.2 3.3 3.4

Experimentation in the natural sciences Introductory remarks The theoretical description of experiments Experimentation as material realization Experimental production and the possibility of realism

59 59 59 69 77

4 4.1 4.2 4.3 4.4 4.5

Verifiability and reference, relativism and realism Introductory remarks Verifiability Conceptual discontinuity and scientific realism A criterion of (co-)reference A realistic realism

80 80 81 89 102 115

5 5.1 5.2 5.2.1 5.2.2 5.2.3 5.2.4 5.2.5 5.3 5.3.1 5.3.2 5.3.3

Specification and application: two case studies from the history and philosophy of quantum mechanics Introductory remarks The correspondence principle and the historical development of quantum mechanics Bohr 1913: correspondence as numerical agreement Correspondence and conceptual continuity: 19161922 Numerical and formal correspondence: 19231925 Correspondence and material realization Philosophical conclusions The measurement problem in quantum mechanics Measurement problem and realism The measurement problem as a problem of correspondence Quantum mechanical measuring process and communication

124 124 124 126 130 136 144 148 153 157 161 163

Conclusion Notes Bibliography Index of names

170 176 191 201

vi

Preface

Now that this book is finished I would like to thank everybody who has contributed to its realization. While studying in the section Philosophy of the exact natural sciences I was able to profit greatly from the broad and profound knowledge and understanding available in the field of philosophy of science. Working as an editor of Krisis I grew to understand the importance of the ever present social and political presuppositions and implications of philosophical views. Accordingly I have also tried in this work to explicitly take into account the fact that philosophizing will always imply an interested position. Philosophy is, however, also a profession which has its own professional rules and criteria. In the Krisis work group Alternatives in science we tried to combine both aspects, the interested and the professional, and to work them out in various directions. The cooperation within this interdisciplinary group has had a most stimulating effect on me. Philosophy of science has in addition its own, more specialist requirements. From debates in the Dutch Study group on the foundations of physics I have learned a good deal, particularly about the measurement problem in quantum mechanics. The material realization of this book has been made possible thanks to the facilities of the Department of Philosophy and the financial support of the BRO-fund of the Free University of Amsterdam. Peter Kirschenmann, my supervisor, has played an important role in the realization of its contents. His tolerance towards other than his own opinions and his always relevant and profound comments and criticism have been of invaluable influence on the final result. Less direct, but not less important, has been the support I received from Francis, Niels and Wouter, especially concerning the question of the conditions and limits of working as a theoretical scientist.

vii

Finally I would like to thank in particular: Ben Bakker, Hans van den Berg, Peter van den Besselaar, Ren Boomkens, Roger Cooke, Jon Dorling, Maarten Fransen, Peter Groenewegen, Jan Hilgevoord, Dick Hoekzema, Jean Jaminon, Wim de Jong, Bas Jongeling, Jozef Keulartz, Harry Kunneman, Chunglin Kwa, Peter van Lieshout, Annemarie Mol, Willem de Muynck, Pieter Pekelharing, Arie Rip, Willem Roos, Bert van der Schaaf, Hanneke Stasse, Karen Vintges. Amsterdam, 12 March 1984

Preface to the English edition This English edition is an almost literal translation of the Dutch text. The revisions are largely of a stylistic character. Occasionally I have added an explanatory remark, stimulated by the comments of, among others, Roger Cooke, Wim de Jong and Bas Jongeling. It is a satisfying development in science studies that the interest in experiment and experimenting has been definitely on the increase in the last few years. Nevertheless most studies published in this field did not prove directly relevant to the subject of this work. An exception is Ian Hackings Representing and Intervening, which I discuss briefly in this edition. I would like to suggest the following to the reader. The two parts of the book constitute obviously a coherent unity. Still, I can imagine that readers more interested in natural science could start with Part II and then continue with Part I, while those who take a more philosophical interest can follow the natural order. Finally I would like to thank the Netherlands Organization for the Advancement of Pure Research (Z.W.O.) for making available a translation grant, the translators Tineke van Putten and Dawn Wolfswinkel for their contribution and the pleasant cooperation, and Susan van Putten and Tini Tuijp for typing out the final text. Amsterdam, 21 January 1988

viii

Introduction

The question whether objective truth can be attributed to human thinking is not a question of theory, but is a practical question. In practice man must prove the truth, that is, the reality and power, the this-sidedness of his thinking. The dispute over the reality or non-reality of thinking which is isolated from practice is a purely scholastic question.1

In this work I would like to discuss a number of philosophical problems related to the experimental natural sciences. By these I mean physics, chemistry and parts of biology, geology and medical science.2 These sciences will be considered as non-monolithical, complex practices in which one can at least distinguish theoretical-conceptual work, formalmathematical activities and experimental action and materialization. Contrary to what is usual in the mainstream of philosophy of science I will not restrict myself to analyses of (formal or conceptual) theoretical aspects, but will also and foremostly pay attention to the experimental production process, to the actions of scientists and to the material and social realization of experimental effects and substances. From a historical point of view this focussing on the experimental production process needs no explaining. Since the Renaissance the characteristic feature of the modern natural sciences is after all experiment or experimental method and not in the first place theoretical interpretation and argumentation.3 We also find theoretical explanations in the Greek philosophers of nature; and theoretical argumentation forms an important part of the Roman jurisprudence. I do not wish to deny at all that theories likewise play a most important role in the modern sciences also when it comes to carrying out and interpreting experiments. Therefore I will discuss the theoretical aspect, too, in great detail in this study. Nonetheless, characteristic of the modern natural sciences remains the fact that this theoretical aspect is combined with the experimental. Considering this state of affairs it is really no less than shameful that philosophy of science has paid so little explicit attention to the

process of experimental production up to now. Experimentation is either not analyzed at all or only to the extent in which it plays a part as a supplier of data for use in argumentations about theoretical statements. This neglect of experimentation by the philosophy of science is not only incomprehensible in the light of the historical development of the natural sciences themselves, but also because of its social importance. In society the productive character of the experimental natural sciences has certainly not gone unnoticed: through the development of a highly scientified technology especially in this century human beings, society and nature have been radically influenced and changed in many respects. From the historical and social importance of experimentation I cannot but conclude that it is high time philosophy of science becomes aware of the existence of the experiment as a productive instance and allows it a more prominent place and function in the discussion of philosophical problems than up to now has been the case. The most important philosophical problem I will be concentrating on in my discussions will be that of scientific realism. In particular I will try to answer the question whether, and if so how, concrete propositions from the experimental natural sciences can relate to a human-independent reality. The questioning of realism with regard to the experimental natural sciences has a concrete basis in factual historical developments. This is in contrast with the much older philosophical problem of scepticism. Many sceptical arguments are based on a general, abstract doubting of truth, reality, perception etc. without wondering whether there exist concrete reasons for such a systematic doubt. Thus, the ensuing debates (e.g. on the question: this glass is on the table) often resemble the sophistic, while they moreover usually excel in dulness. With regard to the possibility of a realist interpretation of natural scientific propositions the situation is quite different. In the past 25 years the philosophy of science has become more empirical, in the sense that it tries to justify its conclusions also in the light of historical and, more recently, sociological studies of science. One of the main results of these studies in relation to the problem of realism is that natural science is always historically and socially situated, and that the theories, paradigms and the like can differ markedly in philosophically relevant aspects from one situation to another. For example: when there exist several good theories about electrons (e.g. such as Lorentz and Diracs) then the question arises which theory is true or what electrons really are.

In other words contrary to the case of the above-mentioned problem of scepticism, there are at any rate at first sight in the experimental natural sciences concrete reasons for doubting the truth and the relation to reality of theoretical propositions. How the experimental natural sciences are exactly situated in their historical and social context is still being debated in the philosophy of science. In this work I will concentrate on the conceptual discontinuities which frequently occur in the theoretical development of the experimental natural sciences. A major conclusion from my observations will be that, because of these conceptual discontinuities, the most common form of realism namely convergent realism, which holds that theories in the course of the development of science will approach truth more and more is no longer tenable. This raises the question whether there are other ways in which natural scientific propositions are linked to reality. The most obvious way in which to answer this question is to look for possible continuities in the development of science as the vehicles of the mentioned conceptual discontinuities. At this point, as we shall see later, the philosophical importance of the experimental materialization process, which is such a continuity, becomes clear. Besides this, formal or mathematical intertheoretical relations will often play a part in the experimental natural sciences. They also represent an important type of continuity. On the basis of these continuities it proves possible to defend a restricted form of realism, in which the theoretical terms in natural scientific propositions in general do refer to elements of a humanindependent reality, but in which we have to abandon the idea that these propositions in some sense picture, represent or reflect reality. In this realism there is reference and possibly co-reference to reality, but no correspondence. Thus the pretensions of this realism are moderate. More ambitious realist interpretations (such as convergent realism) however cannot, according to me, be tenable if we take philosophically serious the general conclusions of the historical and social studies of science of the past 25 and 15 years respectively. And this seems to me a necessary condition for a serious philosophical study of science. In spite of all this the most important thing at stake in this study is not an internal philosophical defense of a realist conception of science. Of much greater relevance is the implied philosophical stance. In this the view of science as a concrete production-practice takes a central position. By turning the perspective around, from the realist conclusion to the philosophical premisses, we can avoid lapsing into a scholastic dispute about reality or non-reality which is isolated from practice (see the

motto of this work). In the present study I have chiefly worked out these philosophical views in the direction of the problem of realism. Other directions are also conceivable: in my conclusion I will briefly touch on these. As has been stated, experimentation plays a major role in the natural sciences, and I would like to begin by taking this role into account in philosophical studies. But what is a beginning? Of course there have been other authors who have tried their hand at a philosophical analysis and account of experimenting. A good example is Jrgen Habermas.4 It is certainly true that Habermas has never developed a detailed philosophy of the experimental natural sciences. But he did develop a number of important insights which are relevant for such a philosophy. First of all there is his intention of social criticism and the ensuing refusal to analyze the sciences as an isolated and autonomous practice. Secondly, he pays explicit attention, particularly in his early philosophy, to the action and production aspect of the natural sciences. Finally, he also includes the role of theories in his observations and sees the experiment as the point where theoretical argumentation and instrumental action merge. In this work I will use Habermas views on the experimental natural sciences both as a starting point and a spring-board. The first chapter offers a survey of these views. The general intention of Habermas epistemology, his philosophical premisses and conceptions are discussed. Subsequently I scrutinize his account of the objectivity of experience (including the place taken by instrumental and experimental action) and that of the truth of propositions (including the role of discursive argumentation). There are several kinds of possible criticism of Habermas views. Firstly, a direct criticism of the actual text of his discussion is given in chapter 2. I analyze and criticize his quasi-transcendental treatment of the objectivity of experience. The emphasis of the discussion, however, is on the criterion of truth proposed, i.e. grounded consensus. It is concluded that this criterion is inadequate. It does not apply to paradigmatic cases of true scientific propositions. Furthermore it is not sufficiently worked out in the sense that it cannot support the notion of convergence towards the truth in its role as a regulative idea. Secondly, Habermas views generally lack the detail and precision required for tackling specific problems in the philosophy of science. As he puts it himself:

During the past decade I have been only cursorily occupied with questions in the theory of science and have lost the feeling for the relevant discussions.5

Thus, if one wants to make use of some of his ideas for purposes of philosophy of science, one must make a fresh start. This is in fact what I will do in part II of this study. In chapter 3 I carry out an analysis of natural scientific experimentation which, to some extent, arises from Habermas views. The main objective is to introduce and make explicit a distinction between the theoretical description of experiments and their material realization which can be described in everyday language. I also make use of some of Von Wrights ideas in analyzing the notion of reproducibility of experiments, a central aspect of which concerns the production of closed experimental systems. Finally, the arguments which Bhaskar derives from experimentation with respect to the possibility of realism are briefly discussed. In the fourth chapter I use this analysis of scientific experimentation for my own philosophical purposes. A detailed exposition is given of the meaning and implications of conceptual discontinuities which frequently occur between successive theories in the historical development of the experimental natural sciences. It turns out that these discontinuities are carried by two types of continuity. The first concerns stable aspects of the process of material realization, while the second has to do with formal or mathematical intertheory relations. With the help of these insights I develop alternatives to Habermas views on truth and objectivity. I formulate and make explicit criteria for the verifiable truth of propositions and for the reference of theoretical terms. Here the focus is on reference; the problems surrounding verifiable truth are treated more generally. The analysis is epistemological, in the sense that it claims that the criteria mentioned are applicable to paradigmatic cases of verifiably true propositions and referring theoretical terms. The differences between Habermas views and mine may have become clear from the terminology I have been using. Instead of speaking of founded truth, I speak more in a pragmatic sense of verification and verifiability; among other things this shift facilitates a better coordination between epistemological considerations and the results of the modern sociology of scientific knowledge. Instead of speaking of objective (that is, for Habermas, intersubjective) experience,

I speak more in a realist sense of reference and, especially, coreference of theoretical terms. The resulting realist position, however, is rather moderate, because I argue that any plausible criterion of reference can only tell us that a term refers, and not what the referent really is. On the other hand, in this view we do have at our disposal a genuine intertheoretically applicable notion of co-reference. This is, in my opinion, an important result. In this context in particular the views of Putnam (and his followers) are examined. On the one hand, his earlier convergent realist interpretation of science is contested. On the other hand, in one essential respect my realist position appears to be stronger than Putnams recent internal realism. A final point of criticism regarding Habermas is that his work completely lacks relevant examples, let alone case studies of scientific practice. This is why chapter 5 contains two extensive case studies with the aim of further clarifying and evaluating the preceding philosophical considerations. First I analyze the role of the correspondence principle during the genesis of modern quantum mechanics between 1913 and 1925. This yields a detailed illustration of the above-mentioned interplay between conceptual discontinuities and experimental and formalmathematical continuities in the development of the experimental natural sciences. Secondly, I deal with the so-called measurement problem in quantum mechanics. In this context, I again make explicit use of the notions of conceptual discontinuity and co-reference. From the philosophical viewpoint of this study we see that the problem, as far as its philosophical aspects are concerned, can be relatively easily analyzed and subsequently solved. In my conclusion I look back at the road I have traveled. I discuss the connection between part I and part II by specifying agreements and disagreements between Habermas ideas and the philosophical conception as developed in part II of the book. I also suggest possibilities for further elaboration of this conception, especially in the direction of the problem of science, technology and society. This makes a certain evaluation possible of the critical potential of the philosophical standpoint which is advocated in this study.

PART I

Philosophical views of Habermas with respect to the natural sciences

1.1

Introductory remarks

I would like to start off this study in the philosophy of the experimental natural sciences by a confrontation with Habermas ideas on the subject. In this chapter I will therefore first expound his views on this matter.1 Inevitably, in such an account, some more general aspects of his philosophy which form the framework in which fit his views on the natural sciences, will have to be touched on briefly. In section 1.2 I give a broad outline of the intention of Habermas epistemology, such as it is advanced in Knowledge and Human Interests. The central proposition in this book, written in 1968, is that a radical critique of knowledge is only possible in the form of a social theory. In 1.3 I then introduce two major distinctions used by Habermas: that between purposive-rational and communicative action and that between communicative action and discourse. The latter differentiation represents a new phase in Habermas thought (round about 1970). The relation between a critique of knowledge and a critical social theory proved to be more problematical than had been expected or hoped for. In this new phase the notion of communication becomes explicitly prominent. In relation to epistemology this change brings Habermas to make a clear distinction between the objectivity of experience and the truth of propositions. Therefore I base the systematic discussion of his views on epistemology and natural science on this distinction. In 1.4 I will discuss the constitution theory of experience and the role that experiment plays in the natural sciences according to Habermas. Next, the consensus theory of truth will be treated in section 1.5. Finally, I will comment upon Habermas reasons for introducing the distinction objectivity truth.

It is with a somewhat ambivalent attitude towards it that I make use of Habermas philosophy: on the one hand it is so abstract and treats of so much (the social evolution of humanity, communication theory, ethics, the theory of science, sociology, politics), that one feels inclined to believe it to be all about nothing; on the other hand such an impressive system-building has its own charms which might tempt one into docility. The best way of coping with such an ambivalence is, I think, by testing a number of quite specific ideas of Habermas in a detailed analysis, in this case of the experimental natural sciences. Although we will see that the resulting views (see part II) differ in some essential respects from Habermas, it also becomes clear that a number of his general ideas have a high heuristic value.

1.2

The intention of Habermas epistemology

In this section I would like to describe briefly a number of Habermas ideas in which the general intention of his epistemology becomes clear. In this I use in particular his earlier work. Habermas holds that epistemology can never find or create a final absolute foundation for knowledge, even though many great thinkers have tried this again and again in their philosophical systems.2 Habermas postulates first and foremost the contingent nature of our knowledge. According to him epistemology is only possible as reflection on the conditions of possibility for this, ever contingent, knowledge. Following Kant he then looks for the transcendental conditions the knowing subject must fulfill in order for (scientific) knowledge to be at all possible. He phrases the overall idea governing such an approach to epistemology as follows:
The idea underlying transcendental philosophy is to oversimplify that we constitute experiences in objectivating reality from invariant points of view; this objectivation shows itself in the objects in general that are necessarily presupposed in every coherent experience; these objects in turn can be analyzed as a system of basic concepts.3

This constitution, however, does not take place from a transcendental consciousness, as in Kant, but from human action possibilities, i.e. purposive-rational and communicative action. On the one hand, this refers to empirical characteristics of the action-oriented physical and social organization of human beings, which have contingently come into

existence in the biological evolution of humanity. On the other hand, these action frames are so static, that they can be regarded as constant in relation to the social-cultural evolution of humanity in which the growth of knowledge should be situated. They are invariant anthropological structures. Because of these two aspects, the contingent-empirical and the invariant, Habermas calls the conditions of possibility of knowledge, purposive-rational and communicative action, quasitranscendental. He can combine the transcendental and naturalistic approaches to quasi-transcendental, because he extends Kants constituting subject to the human kind as a whole.4 Even though Habermas rejects the absolutist pretensions of a philosophy of origins (as if this could establish knowledge once and for all) he does not want to restrict himself either to a naturalistic analysis of scientific statements and their relations, as the positivists did; i.e. an analysis in which the factual development of science itself is proclaimed the only norm by which to evaluate and criticize scientific knowledge.5 In contrast with positivism and absolutism Habermas advocates reflective philosophy: knowledge is always knowledge produced by subjects and thus its status depends on subjective conditions of possibility; reflection then means that the subject again bends about his or her knowledge and becomes aware of its constitution-context. Such a reflective attitude clears the way for a critique of knowledge, in particular of positivism. Habermas typifies this philosophical movement aptly as follows:
That we disavow reflection is positivism.6

According to him the basis of this positivism is scientism, the philosophical doctrine that identifies knowledge with science, much impressed as it is by the overwhelming progress of modern science. Scientific knowledge is seen as an autonomous body of systems of statements, and the pretension of a reflective, philosophical critique of knowledge in the original Kantian sense is rejected. Philosophical criticism can never put science in its (right) place, because the right place of science is the one it factually occupies. Epistemology is reduced in positivism to a study of the whole complex of rules according to which theories are built and tested. The knowing subject, which has formed from Kant to Marx the frame of reference for a possible criticism of unjustifiable pretensions to knowledge, has disappeared in positivism. Instead the fact of scientific progress itself has become the normative foundation of a judgement of scientific knowledge and its development.

Habermas rejects this scientistic ideology on the basis of his reflective epistemology. A particularly important result of his critique of knowledge is the distinction between empirical-analytic and historicalhermeneutic sciences. The quasi-transcendental condition for the possibility of the first (this includes the natural resp. the empirical social sciences as far as they produce nomological knowledge) is instrumental (resp. strategic) action. Communicative action forms the condition of possibility of historical-hermeneutic knowledge (history, literature, jurisprudence etc.).7 At the very least this insight in the differential constitutioncontexts of the various sciences changes the self-consciousness or the self-understanding of the reflecting subject with respect to the knowledge he or she acquires. In the second place it may be so that certain (mental or social) regularities have been reified into natural laws because of their specific context of origin, while in fact they only rest on fossilized habits or contingently limited patterns of action. A reflective process of knowledge can lead to emancipation in such cases, i.e. to a cancellation, or rather a suspension of these natural laws. As examples of such critical or emancipatory sciences Habermas mentions psychoanalysis and the (marxist) critique of ideology. The critique of knowledge in these examples does not only reflect on the context of constitution, but also on the context of application of the acquired knowledge. In other words, psycho-analysis and critique of ideology do not only perceive how and why certain, apparently lawful, mental and social mechanisms have come into existence, they moreover play an essential part in eventually suspending these mechanisms. When they are successful, they will through their own application make themselves redundant. According to Habermas the frames of action mentioned above establish general rules both for the building and testing of theories as well as for the way in which these can be used in practice. Consequently, he proposes a systematic relation between the pre-scientific frames of action, the scientific methodologies and the possible applications of the acquired knowledge. This cohesion, he argues, presupposes a number of interests that guide knowledge: the technical interest for the empiricalanalytic sciences, the practical interest for the historical-hermeneutic sciences and the emancipatory interest for the critical sciences.8 These interests take shape as universal cognitive strategies of humanity, which safeguard the survival of the species. Thus instrumental action, or work, is of direct vital importance for natural reproduction. The cognitive interests are also quasi-transcendental and invariant. They are not external influences on science, which then could be criticized, for

10

instance in the name of the objectivity of science. Therefore the analysis of the cognitive interests is not psychological, sociological or historical, but philosophical. It concerns the categorial meaning statements acquire taken the fact that they belong to a specific kind of science. In his article Technology and Science as Ideology Habermas links this reflection on the constitution-contexts and cognitive interests of the various kinds of science to a social criticism of the predominance of the instrumental reason, of thinking one-dimensionally:9 in the modern Western societies domains which were traditionally subjects of historical-hermeneutic understanding and study become increasingly subjected to the instrumental empirical-analytic sciences and related technology. Because of this, practical characteristics related to communication will disappear automatically to leave room for technical characteristics linked to instrumental action. In this trend Habermas does not criticize the existence or the effects of instrumental reason as such, but its exclusive application, at the expense of practical, communicative reason. Only the latter makes science and technology degenerate into ideology. On the other side Habermas does not reject outright alternative (e.g. esthetic or ethical) approaches to nature either, but only in as far as they claim to represent theoretical knowledge.10 I would now like to conclude this outline of the general drift and intention of Habermas philosophical views on the various sciences. Characteristically his approach is epistemological and his aim is social critical. For the sake of brevity I have restricted myself to Habermas early work. But an epistemological approach and a social critical intention, although interpreted in a partly different sense, are also typical of his more recent philosophy. In so far as this later philosophy is relevant to the problems in the experimental natural sciences I will hereafter deal with it in detail. Those aspects concerning Habermas more recent theories of history and society I will however ignore because they go much beyond the scope of this study.11

1.3

Two fundamental distinctions

In this section I will introduce a number of concepts which will play a large role in the later discussion. I will do this by means of a discussion of two fundamental distinctions made by Habermas. Firstly, he distinguishes two domains of communication, namely communicative action

11

and discourse. The second important differentiation is that between communicative action and purposive-rational action. The point is to distinguish between two different aspects: between forms of communication and types of action. Both will be introduced in this section in a rather sketchy way. I intend to make them be more substantial later on when I concentrate on the problem of objectivity and truth.

1.3.1

Purposive-rational and communicative action

As we have seen above Habermas distinguishes between various kinds of science, each one supported by different knowledge-guiding interests. Moreover, he links his analyses to a social criticism of the present predominance of one of these interests, namely the technical one. To develop these insights in a more systematical philosophical manner he uses the distinction between purposive-rational action (or work) and communicative action (or interaction). The first he defines as follows:
By work or purposive-rational action I understand either instrumental action or rational choice or their conjunction. Instrumental action is governed by technical rules based on empirical knowledge. In every case they imply conditional predictions about observable events, physical or social. These predictions can prove correct or incorrect.12

In this context Habermas calls the action based on rational choices or decisions strategic. The point is to find out which way of action is the best, i.e. in this case the most useful or advantageous, to reach certain goals, given a number of preference-rules or value-systems. However in my discussion it is especially the role of instrumental action and its contrast to communicative action which is important and so I leave strategic action aside.13 Next Habermas introduces the concept of communicative action:
By interaction, on the other hand, I understand communicative action, symbolic interaction. It is governed by binding consensual norms, which define reciprocal expectations about behavior and which must be understood and recognized by at least two acting subjects.14

Essential in both Habermas earlier and later work are his discussions of distorted and unconstrained communication. The main objective of communicative action is the maintenance or the (re-)establishment of

12

mutual understanding between acting subjects. This implies that communicative action in Habermas should certainly not be taken as individualistic, as the action of isolated subjects. The most important element of distinction between interaction and work is the difference between dialogical action based on mutually binding norms on the one hand and on the other the monological following of anonymous technical or strategic systems of rules. In Habermas view both aspects should be done justice to in the social development. Thus he argues that in order to reach a rational compromise between technological progress and practical-normative action, we should replace politics based on a technocratic ideology by one founded on a general and public, unconstrained debate. I must add two clarifications to this distinction between purposiverational action and communicative action. The first one concerns the epistemological status Habermas ascribes to these action-frames. On the one hand he is particularly opposed to Marcuses view of the historical possibility of a completely alternative natural science and technology (instrumental action, which is the condition of possibility of natural science and technology, is for Habermas an invariant anthropological structure); on the other hand he opposes what he sees as the instrumentalistic reduction of production-relations to production-forces in Marx (the dimension of interaction has its own evolution and cannot be reduced to the dimension of work). To make these criticisms effective Habermas claims a quasi-transcendental status for instrumental and communicative action, which is neither completely empirical nor completely transcendental.15 Especially in his more sociological work, however, Habermas uses the distinction in a purely empirical manner. Apart from the problem of whether and if so how this empirical-sociological approach can be reconciled with a quasi-transcendental approach, the question arises what do the notions of purposive-rational and communicative action exactly refer to. Particularly in his early work Habermas often seems to suggest that two different, empirically identifiable, types of action are meant. Such an interpretation calls for the just criticism that this kind of pure action types do not actually occur: concrete instrumental actions always form part of a network of interactions, and concrete communicative actions are always accompanied by and make use of instrumental actions. McCarthy has therefore proposed a re-phrasing in order to avoid these problems16: on the one hand there is an analytic difference which in itself concerns aspects of concrete empirical actions; on the other hand empirical actions can also be characterized by the dominating aspect.

13

Thus, from an empirical point of view, the difference between instrumental and communicative action is one of degree.17

1.3.2

Communicative action and discourse: two forms of communication

As said above Habermas has clearly taken a new road in his later philosophy.18 Of the problems urging him to do so I will mention only two. First of all Habermas has come to understand that making the of the validity of scientific statements explicit only in terms of pre-scientific frames of action, such as instrumental and communicative action, is in itself insufficient. Besides action, theoretical or discursive argumentation plays a central role in science. To do this point justice Habermas, in his later work, distinguishes the objectivity of experience from the truth of propositions. While objectivity is still constituted from the frames of action, Habermas analyzes the problem of truth in terms of the new concept discourse. Secondly the idea of founding his critical social theory by putting the emancipatory interest as an anthropological invariant on a par with the practical and technical interest appeared no longer tenable. In Knowledge and Human Interests Habermas tried in vain to squeeze two meanings of self-reflection into a straitjacket: reflection on the general constitution context, on the conditions of possibility of empirical-analytic and historical-hermeneutic sciences on one side, and ideology-critical reflection on specific contexts of origination and application of socially and historically determined laws on the other side. But the gap between a general critique of knowledge and a specific social criticism is larger and harder to bridge than Habermas initially thought. That is why the emancipatory interest acquires a different status than the technical and the practical in his later work. It is incorporated in Habermas communication theory, in which it is especially linked with the notion of unconstrained discourse. Both the differentiation between truth and objectivity and Habermas later ideas about the basis of critical social theories make use of a second fundamental distinction, namely that between communicative action and discourse. Habermas describes the difference between these two kinds of communication thus:
With the keyword action I introduce that field of communication in which we tacitly presuppose and acknowledge the validity claims implied in utterances (also in assertions) in order to exchange information (i.e. action-oriented experiences). With the keyword discourse I introduce that kind of communication characterized by

14

argumentation, in which questionable validity claims are subjected to discussion and are examined as to their justifiability. In order to have a discourse we must in a certain sense leave the context of action and experience; here we do not exchange information, but arguments which serve to justify (or reject) the problematic validity claims.19

By the utterances and assertions mentioned in this quotation Habermas does not mean well- or not well-formed sentences but, following Austin and Searle, he means correctly or incorrectly performed speech acts which refer to the use of sentences in utterances. In his universal pragmatics these speech acts form the basic entities of language and thus of verbal communication. To perform succesfully a concrete communicative action grammatical validity (well-formedness) of the utterance is not sufficient. As we will see other, pragmatic, requirements are also important. These pragmatic requirements concern the question whether a sentence, when uttered, is embedded in a practical context in such a way that the speaker obtains the communicative meaning he or she intends. In both kinds of communication non-verbal actions and expressions do play an important part besides verbal utterances. Habermas believes, however, that at least with respect to the main goal of communication, the maintaining or (re-)establishing of mutual understanding, the verbal aspect is of much greater importance than the non-verbal. He even holds that in order to create a situation capable of unconstrained communication expressions, actions and experiences (the day-to-day worries) should be set aside for a while, thus removing all sorts of possible, systematic interferences in communication. In his words, we have to leave the level of communicative action in order to enter that of discourse. In this discourse only verbal utterances really count, although, of course, in actual fact these will always be accompanied by actions and expressions. The basic point now of Habermas universal pragmatics is that all (normal) communication is achieved because speakers make, usually implicitly, four mutually acknowledged claims to validity in their speech acts.20 The claim on comprehensibility linked to grammatical wellformedness as the basic condition for the establishment of communication; the claim on the truth of the propositional content of the utterance; the claim on the rightness of the illocutionary aspect of the utterance (Austin) measured by contemporary norms; and the claim on

15

the genuineness of the speaking subject, which emerges from his or her actions. A successful speech act has to be embedded in the pragmatic context in such a way that not only the comprehensibility in a grammatical sense is guaranteed, but also that (true or false) propositions can be thus exchanged, that the relation speaker-hearer expressed in it agrees with (is right on the basis of) contemporary norms and conventions, and that the speaker phrases his/her intentions in such a genuine way that she/he says what is meant. Within the frame of communicative action these validity claims are usually tacitly presupposed and accepted. But failures or obstructions will of course occur in communication. When this happens the existence of validity claims becomes explicitly manifest. Problems in or with communication can be rectified in two ways according to Habermas. Small failures, which one can characterize probably best as misunderstandings, can usually be rectified within the context of communicative action itself. For example, depending on the area in which the communication has been disturbed, by making meanings explicit, by filling in gaps of information, by making up for errors (which have damaged mutual confidence), or by appealing to accepted norms or conventions. More serious obstructions, those which make the reaching of a mutual understanding in a systematic way impossible, cannot be solved adequately within the frame of communicative action, but only within that of discourse. Here Habermas distinguishes theoretical discourse, in which the testing of the claim on truth of propositions is possible, from practical discourse, in which the claim on the rightness of norms can be debated.21 For Habermas it is a practical hypothesis (i.e. one which may possibly be realized in practice) that forms of life can be developed in which a rationally grounded consensus can be achieved on the validity claims with respect to the truth of propositions or the rightness of norms, once they have become debatable in communicative action. What this view exactly entails with regard to the definition and function of the concept discourse I will go into in section 1.5 in the discussion of Habermas consensus theory of truth. For the time being I will be content with this short introduction of the couple discoursecommunicative action. 1.4 The constitution theory and the role of the experiment

As said above Habermas, in his later philosophy, makes a clear distinction between the objectivity of experience and the truth of the

16

propositions in which the said experience is expressed. In the next section I will discuss the consensus theory of truth. Here I focus on Habermas constitution theory of the objectivity of experience and on his views on the role of instrumental and experimental action.

1.4.1

Objectivity of experience and the categorial structure of the object-domain

Habermas distinguishes two kinds of experience both of which are also means to acquire knowledge: sensory experience (or perception) of things and events, and communicative experience of persons and their utterances. It is sensory experience or perception which is especially important for the philosophy of the natural sciences with which I am now primarily concerned. With respect to the question of objectivity Habermas shares the general standpoint of transcendental philosophy, which is that this problem of the objectivity of (knowable) reality can be solved by seeing it as directly connected with the question of the (inter)subjective conditions for the objectivity of our experience of this reality:
Every form of transcendental philosophy claims to identify the conditions of the objectivity of experience by analyzing the categorial structure of objects of possible experience.22

This possibility of a simultaneous explication of the categorial structure of objects of possible experience and of the conditions of possibility for the objectivity of sensorial experience is directly connected with the typically human unity of and the systematic interplay between sense reception, action and linguistic, inter-subjective representation on which cognition rests. Habermas transcendental philosophy is, as we will see, pragmatic, that is, it is action which fulfills the role of condition of possibility (in the form of instrumental action). It follows that this instrumental action is both constitutive of the categorial structure of the objects of possible sensorial experience, and of the objectivity of this experience. Hence, two different, though coherent, lines of reasoning are recognizable in Habermas argumentation. The first holds that all experience which gives us knowledge of the categorial structure of objects is always structured in advance by

17

language. Habermas calls the language through which we a priori structure sensorial experiences or perceptions the thing-event language. Its fundamental concepts lead to a specific structure of the object-domain of things and events, adjusted to the specificity of the sensorial as compared with the communicative experience:
When we analyze the syntax of this language we will come across the categories which structure the object-domain of possible experience in advance. In order to form certain hypotheses about objects of experience, which may be refuted by experiences we must presuppose that which constitutes objects of possible experience in the first place: the general structure of an object-domain. We presume an object-domain of bodies-inmotion preceding our sensorial experiences.23

In the constitution of the categorial structure of the object-domain of things and events a number of specific semantic categories play a role. These determine the structure of the objects of possible experience a priori; they form the condition of possibility of a successful identification of and reference to these objects:
When we identify objects about which we state something (on the basis of the experience we have had), we do so either ostensively or by means of names or characterizations. These must contain or imply predicative determinations. It is true, predicative determinations are not used predicatively in the context of denotative expressions. But a properly functioning system of reference has to have a certain propositional content. This minimum content of properties which objects as such have is the categorial framework for objectivating experienceable happenings as happenings. In this respect, Piagets cognitivistic developmental psychology has confirmed Kants analyses: the basic notions of substance, space, time and causality are the minimum conditions for determining a system of reference for objects of possible experience.24

Thus in Habermas view, a correct usage of the categories substance, space, time and causality is the condition for successful reference to and identification of things and events.25 He claims that the objects of possible sensorial experience have to be identified, for instance, within a Euclidean space and within an abstract continuum of points in time. Habermas calls the categories mentioned semantic because they establish the general meaning of objects of possible sensorial experience and because they make successful reference to these objects possible. They are not concepts from semantics, such as meaning, truth, reference, but concepts with a semantic function. The link between the general structure of the thing-event language on the one hand and the categorial structure of things and events on the other hand is effected according to Habermas by action, by

18

instrumental or (in the case of the natural sciences) by experimental action, which is analogous in structure. Parallel to Piagets views on the cognitive development of individuals Habermas regards the semantic categories as the result of instrumental action schemes which have proved successful in the course of evolution and which have settled into language.26 The universality of categories such as substance, space, time and causality is founded on the anthropologically invariant character of the action frame of instrumental action. Thus it is this universal instrumental action frame that makes objective reference through the thing-event language possible. Habermas second line of reasoning deals with the link between objective experience and its condition of possibility, instrumental action. According to Habermas perceptions or sensorial experiences cannot but be objective. Their objectivity consists precisely in the fact that they can be shared and communicated inter-subjectively. If this proves impossible, then it follows that the claimed experience was not an experience, but a fantasy, a hallucination or something like that. The fundamentally subjective sensation can never be the basis of an epistemology; sensation has only a cognitive meaning when it has been transformed into an objective experience: when it can be shared and communicated intersubjectively.27 Habermas develops his discussion by advancing the view that the intersubjective character of sensorial experiences consists in the fact that they can be verified, namely, once they are sufficiently problemized and specified, in the form of an experiment: perceptions are quasi-experiments.28 This experimental verifiability Habermas interprets as the essential repeatability of experiments on the basis of an instrumental manipulatability of all relevant experimental conditions. The fact that experiments are essentially repeatable is not just a fact from experience, but an a priori necessity. This a priori repeatability implies that every single experiment has at the same time a general meaning:
The singular event is at the same time a general phenomenon, the reason being that it guarantees that all operations carried out in the future and repeating the initial experiment under exactly the same conditions must lead to the same effect.29

Singular empirical propositions in which we express the experimentally obtained experience imply always a general categorial proposition expressed in the form whenever x (under normal conditions n) then y. The reason for this is that the instrumental or experimental action, which is constitutive of the objectivity of this experience, a priori

19

necessarily entails that whenever we do x (under normal circumstances n) then y will be the result. We conclude from this: the objectivity of sensorial experience and the categorial meaning of the empirical propositions expressing this experience are constituted in one movement from the quasitranscendental action frame of instrumental action. The objectivity reveals itself in the verifiable success of the actions concerned, when we base our acts on this experience. And the categorial meaning of empirical propositions shows their origin, structure and applicability as intrinsically bound up with the action frame of instrumental action. This common structure of empirical propositions and instrumental actions reflects on one side the origin of statements about things and events: empirical propositions can only be discovered in an instrumental or experimental context. On the other side, this common structure restricts the possibility of applying empirical-analytic knowledge to objectdomains sharing the same general structure: empirical things and events will always be at once instrumentally manipulable things and events.30 Finally I would like to consider a misunderstanding arising probably from Habermas early phrasing. In his book Knowledge and Human Interests Habermas does not speak about the categorial meaning, but about the meaning of the validity of judgments or propositions.31 Such a formulation easily leads to a confusion with the problem of truth or validity. But the question which really mattered to the early Habermas, was: what do empirical propositions mean with respect to reality, if they are valid? His answer is that the form of these propositions structures reality a priori in such a way that only that emerges as objective experience, which can be manipulated instrumentally. Especially the later Habermas would never say that the truth of empirical propositions and the existence of things and events should depend solely on instrumental action. Firstly, the truth of a proposition such as this particle has a speed of 100 m/sec is certainly not determined completely by the categorial meaning of the terms used for things and events. A discursive argumentation based on the theoretical meaning of these terms is also needed. Secondly, Habermas also agrees with the anti-idealist assumption that the existence of things and events is an existence-in-itself, independent of human thought and action:
We do reckon with the existence of a reality that is independent of men who can act instrumentally and arrive at a consensus about statements. But what the predication

20

of properties catches of this reality is a matter of fact that is constituted only in the perspective of possible technical control.32

Thus Habermas on the constitution of sensorial experience. The entire theory is rather sketchy and abstract and certainly needs to be worked out in more detail and subjected to close examination. On the one hand Habermas ideas on the categories space, time etc. do no more than touch a problem and indicate a general direction. He admits this himself and for a further elaboration refers to the protophysics by Dingler, Lorenzen and their followers, who give a more detailed theory of physical measurement.33 In this study, however, I will not go into this more specific elaboration and its problems any further. On the other hand the transcentental-pragmatist point of departure of Habermas needs to be clarified and criticized. In the following chapters I will return regularly to the objectivity question broached by Habermas and its relation to the experimental practice of the natural sciences.

1.4.2

The experiment in the natural sciences

In research in the natural sciences instrumental action takes the shape of experimental action or of perception as quasi-experimental. This experimental action is according to Habermas only a precise form of instrumental action in general that has been made possible by operations of measurement.34 In this sense scientific research (via the constitution of its object-domain) remains bound to an action frame originating in daily life. Habermas does observe, however, that this experimental action, as a learning process, is only in an indirect way under pressure of the urge to survive. This makes it possible to carry out in concrete, individual experiments only freely chosen actions. Habermas describes the role of the experiment in the research process of the natural sciences thus:
In an experiment we bring about, by means of a controlled succession of events, a relation between at least two variables. This relation satisfies two conditions. It can be expressed grammatically in the form of a conditional prediction that can be deduced from a general lawlike hypothesis with the aid of initial conditions; at the same time it can be exhibited factually in the form of an instrumental action that manipulates the initial conditions such that the success of the operation can be controlled by means of the occurrence of the effect.35

Apart from description and performance perception of course also plays a role in an experiment. As we have noticed Habermas sees experimental 21

controllability as the guarantee for the objectivity of the perceptions or experiences concerned. In his later philosophy Habermas allows an experiment a second function on the basis of his new distinction between action and discourse: besides the acquisition of objective experience, experimental action also serves to transform these experiences into theory-laden data, that can be used in discursive argumentations. This leads to a certain relativization of the similarity between the technical instrumental action and the discourse-dependent experimental action. On the one hand both forms of action have analogous structures (so the transcendental characterization described in 1.4.1 remains unscathed). On the other hand they have different functions and objectives. In this context Habermas borrows a distinction made by H. Schndelbach:36 the function of technical instrumental action is in the first place the application of undebated technical means to reach a given goal. The difference then with experimental instrumental action lies in the fact that here in discursive argumentation about the truth of propositions these means can be problemized. When an experiment yields discursively useful data, one can speak of epistemic success. But this does not alter the fact that scientific arguments finally should be testable in an experiment. This test remains based on instrumental action, on operational success. This is why Habermas maintains his instrumentalist theory of the constitution of daily and scientific sensorial experience, despite the relativization induced by the further distinction between technical and experimental instrumental action.

1.5

The consensus theory of truth

So far the possibilities for scientific activity have been linked fairly directly in Habermas view to the action-context of instrumental action, even though the activity proper to natural science, experimentation, is temporarily detached from work necessary for direct daily needs.37 Within the context of Knowledge and Human Interests this last addition comes quite out of the blue. Apparently Habermas realized this himself, for in his later work the introduction of the concept theoretical discourse is an attempt to take into account the distance between daily and scientific activities. Apart from action in the form of experimentation and communication, theoretical argumentation occupies a major and

22

relatively independent place in science. This argumentation takes place within the communication-area of theoretical discourse, which has the same structure and function for all sciences. The discourse is experience-free because, viewed thematically, all urge for action (and action in general) is suspended. What counts in discourse is the force of the better argument, this being the manner whereby we try to reach a founded consensus on the truth or falsity of propositions. Habermas thus introduces a marked distinction between two aspects of knowledge, viz. the objectivity of sensorial experience, which is constituted from the action-frame of instrumental action; and the truth of propositions for which a founded consensus is sought within the discourse as soon as their claim to validity is or has been doubted. In this section I will try to describe as accurately as possible what Habermas means by these related concepts truth and discourse.

1.5.1

What is truth?

A first important question for every theory of truth is: what do we assign truth to? In this Habermas follows Strawson, who proposes that truth should be assigned to propositions or statements (Aussagen).38 On the one hand it cannot be denied that propositions can only be uttered in datable (or indexical) assertions or as the propositional content of datable speech acts in general (e.g. the command leave the light on states implicitly that the light is on and therefore has the propositional content the light is on). On the other hand what matters in truth will always be the representation of a state of affairs, independent of the kind of speech act in which the propositional content is embedded and, moreover, independent of the fact that the performance of speech acts is always situated in space and time. If a proposition is true, then it will always be true. In Habermas view truth always claims invariance.39 The predicate true refers to what we assert, and not to that (the speech act) through which we perform the assertion. A second preliminary remark concerns the difference between what we assert and about which we assert this. Following Ramsey and Strawson Habermas calls what we assert in a proposition a state of affairs (Sachverhalt) or, if the proposition concerned is true, a fact (Tatsache), i.e. an existing state of affairs. That about which we assert something are the things and events (or persons and their utterances) which we can experience.40 Habermas always reproaches correspondence theories of truth for trying again and again but constantly in vain to

23

find a link (correspondence, representation, isomorfism etc.) between linguistic propositions which assert facts and non- or pre-linguistic things and events which exist in reality:
When we cannot ascribe any other meaning to the term reality than that which we associate with propositions about facts and conceive of the world as the sum total of all facts, then again can the correspondence relation between propositions and reality only be determined by propositions. The correspondence theory of truth tries in vain to break away out of the sphere of language in which only the validity claim of speech acts can be explained.41

After these introductory remarks I would now like to first of all examine what Habermas holds to be the meaning of truth in the light of some quotations of his:
Truth is a claim to validity which we attach to propositions by asserting them.42 A claim to validity is equivalent to the assertion that the conditions for the justifiability of an utterance have been fulfilled.43 In action, the factually raised claims to validity, which form the underlying consensus, are assumed naively. Discourse, on the other hand, serves the justification of problematic claims to validity of opinions and norms.44

From these quotations we can extract the following definition: Truth is a claim to validity which people attach to propositions when they assert them and which, although tacitly accepted in communicative action, can be founded in a theoretical discourse always and everywhere by means of a valid argumentation. The following points help to clarify this definition. 1. Typical of Habermas theory is that also the meaning of truth is made explicit with the help of epistemological terms (e.g. founded claim to validity). This does not mean that only those propositions, whose truth we have in fact determined through a discursive argumentation, are true. After all, a basic assumption in Habermas is that the truth of propositions, which is explicitly expressed only in discourse, is tacitly presumed or naively accepted in the context of communicative action. As we have seen truth always claims invariance. If a proposition is true then it will also be true before and after its truth has been determined in discourse. From this it follows that also those propositions which we assert in the context of communicative action are either true or false. 2. The linking of truth to the validity of an argumentation which could be realized in discourse is essential to Habermas theory.45 Taken his rejection of any notion of correspondence between individual propositions and reality he sees the meaning of the truth of a proposition in terms of a

24

cohesion with other propositions and the context in which this cohesion is realized. His position is specific in these aspects: (1) the cohesion is that of argumentative validity; (2) within these argumentations data supplied by experiment are used; (3) a consensus on validity should be realized within the social context of a discourse. Thus in this conception formal, empirical and social factors play a part. The exact relations between them will be discussed in detail later on. 3. Finally I would like to focus on another important aspect of the theory, the pragmatic side. Although truth is invariant and assigned to propositions, even so the latter can only be distinguished analytically from the contextual, datable speech acts in which they are embedded. Thus, according to Habermas, the meaning of truth can never be clarified by discussing only propositions, as e.g. the formal-semantic approach proposes. And so the contextual, the pragmatic aspect of speech acts plays an important part in his definition of truth. The meaning of truth shows itself in the performance of (in particular constative) speech acts in which we assert facts. We should therefore distinguish clearly, where implied claims to validity are concerned, between assertions made in the context of communicative action and those in discourse. In communicative action assertions serve to supply action-oriented information, the truth of the propositional content of the assertions being tacitly presupposed and accepted. In discourse, on the other hand, what matters is the finding of argumentations for the truth or falsity of propositions lifted from the action-context for discussion. There is therefore no syntactic or semantic but a pragmatic difference between propositions about objects of experience on the one side, and information about experience with these objects on the other side. One and the same sentence can have the pragmatic meaning of an expression of experience or of a proposition correlated to a debatable fact yet to be founded in discourse. To explain this pragmatic difference a bit more I use as an example the sentence this water is boiling. Lets take it first as a communication about the cooking of vegetables in a household situation. In that case this sentence would often mean: Can you turn down the gas, or it will boil over. Such communications or information have, under normal circumstances, a high degree of reliability, which can be deduced from the success of the action-instructions to be followed:
The reliability of a communication can be measured by the probability with which the behavioural expectations derived from this communication (in action contexts) are fulfilled.46

25

In this situation, the question: have you checked with a thermometer whether the temperature really is 100C? would in general be absurd. People systematically showing such a doubt within the action context usually end up quite quickly on the psychiatrists couch: systematic scepticism reaches here its practical limits.47 But in another context doubting the truth of the sentence the water is boiling could indeed be justified, e.g. in a scientific debate on the boiling points of liquids. One could imagine a discussion beginning after the following explanation: the water is not yet boiling because the temperature is only 80C; the air bubbles are produced by the vegetables for they contain hollow spaces filled with air which break open exactly at 80C so that the air is set free.

1.5.2

Grounded consensus as the criterion of truth

Now I would like to examine what Habermas proposes as the criterion of truth. The discussion will be based on this quotation:
Only the agreement, which we can reach in discourses, is a grounded consensus. This counts as a criterion of truth; the meaning of truth, however, lies not in the fact that in general a consensus is reached, but: that at any time and in any place, whenever we enter a discourse, agreement can be reached under conditions which show it to be a grounded consensus.48

Thus for Habermas the criterion of truth is grounded consensus. In more precise terms: The claim to truth of a proposition is justified, if a grounded consensus has been reached within a theoretical discourse. As I explained in the previous section and as the quotation above shows, a grounded consensus actually achieved is sufficient to establish the truth of a proposition but is not necessary. Also propositions asserted within the context of communicative action can be true, namely in those cases in which, if their claim to truth is doubted, a grounded consensus within a theoretical discourse at any time and in any place could be reached. A criterion of truth phrased in such a way of course raises the question at once: when is a consensus grounded? Habermas does not accept an accidental or forced consensus between scientists as an adequate criterion of truth. He condemns those forms of pragmatism in which truth refers to nothing more than factual consensus, reached no matter how. In this respect he agrees with the advocates of correspondence theories, who also hold with the grounding of truth.

26

According to Habermas consensus should be reached rationally, by means of valid argumentations. To understand what these words rational and valid can mean here, he examines the structure of the arguments which we adduce to make true the claim to validity for a certain proposition:
The consensus theory of truth claims to explain the typically unforced force of the better argument by the formal characteristics of discourse and not by something that either, like the logical consistency of sentences, underlies the structure of the argumentation, or like the evidence of experiences, enters the argumentation as it were from the outside. The outcome of a discourse can neither be determined by logical nor by empirical force alone, but by the power of the better argument. This power we call rational motivation. It should be explained within the framework of a logic of discourse49

What this force of the better argument signifies Habermas explains in a number of steps. His criterion of truth proves to have both communicative and material aspects. The former he analyzes in terms of the formal characteristics of discourse and of the ideal speech situation (see 1.5.3 and 1.5.4); the latter in terms of the adequacy of theoretical language systems with respect to their object-domains.50 I would like to start with the material aspect. Let us suppose a controversy about the truth of the following proposition C (=Conclusion): C: this water is boiling In order to reach a consensus on the truth of this we need the following argumentation. We supply data D relevant to the occurrence of C, e.g.: D: this water has a temperature of 100C Then we make use of an empirical regularity or lawlike hypothesis W (=Warrant) by which we can deduce C from D, e.g.: W: all water will boil at 100C Finally, an acceptable regularity W should be supported by empirical evidence B (=Backing). In this example the support will be given by a number of situations in all of which water at 100C was boiling. In his commentary on this argumentation scheme, taken over from Toulmin, Habermas stresses that the most relevant case for the empirical sciences is the one in which the truth of W does not follow deductively from B. In such an informative argument an inductive step arises from B to W, a step not logically compelling but having a certain amount of plausibility. It is here, he argues, that the pragmatic aspect of the logic of discourse is to be found: an argumentation does not simply consist of a chain of sentences governed by deductive relations of inference, but of a chain of speech acts, governed by among other things an inductive inference relation of plausibility.

27

Giving an argumentation for the truth of proposition C consists then in giving a deductive-nomological explanation for C in the sense of Hempel-Oppenheim, which makes use of a plausible, empirical regularity or lawlike hypothesis W.51 Concerning the interpretation of these laws Habermas agrees with the empiricist or positivist view, both in his early and in his later work. Laws can be interpreted as statements about the covariance of observable events52; they are empirical uniformities supported by conclusions about the repeatedly observed covariance of quantities.53 This is the classical Humean view of laws as constant conjunctions of in principle observable events. At this point Habermas certainly does not provide an alternative for (epistemological) positivist interpretations. The only thing he does is restrict their (metaphysical) pretensions, their self-knowledge, to the empirical-analytic sciences in order to clear the way for the historicalhermeneutic and critical sciences.54 As said, a good argumentation on the truth of propositions will contain plausible inductive steps. According to Habermas an inductive line of argument will ultimately derive this plausibility from the adjustment of the whole underlying theoretical language system to the domain of things and events concerned. On the one hand this theoretical system precedes experience, which it a priori structures; on the other hand it once originated in a confrontation with things and events and was found adequate. Eventually the pragmatic character of the logic of discourse lies in this that the used inductive argumentations are only plausible in so far as they are exemplary (and successful) repetitions of exactly those kinds of experience which have once led to the shaping of the fundamental predicates contained in the theoretical language system. These fundamental predicates of the language concerned result from the application of cognitive schemata (see 1.4.1). Following Piaget Habermas interprets these thus:
Cognitive schemata are the results from an active interaction of the personality and social system with nature: they are developed in assimilation- and simultaneous accomodation-processes. The bedrock of these schemata enters into the personality structure and forms the cognitive apparatus; but also the less fundamental and changeable schemata, which appear as fundamental concepts in theories and other interpreting systems, play a constitutive part in the construction both of objectdomains and interaction structures.55

A language system is adequate if reality has (ever) let itself been caught in a network of concepts. Within this network Habermas distinguishes

28

fundamental and theory-specific cognitive schemata: the former, which do not change in the course of the cultural evolution, are the categories of substance, space, time and causality discussed above. Habermas holds that the adequacy of that part of the language system corresponding to these categories has been proved by Piagets theory of cognitive development.56 The adequacy of that part of the language system corresponding to the theory-specific categories depends in the example used above on an affirmative answer to questions such as: can the concept of a boiling point be applied, i.e. do we always find the same fixed point on the scale of a thermometer for one liquid; do the methods of temperaturemeasurement give generally reproducible and consistent results; can we experimentally produce pure water, etc.?57 An affirmative answer to these questions can in general only be given with hindsight, looking back on certain theoretical developments. And even then such an answer remains tentative as a consequence of the ever present possibility of new information and insights in the future. This state of affairs regarding the material aspect lends a fallibilistic character to Habermas criterion of truth. Although each answer to the above questions depends on experience, an application of Habermas criterion of truth to the proposition this water is boiling requires an argumentation in which an answer to these questions is already presupposed. We can never reach consensus on the truth of a proposition when we start doubting all relevant presuppositions at once.58 Concrete discursive argumentations which take place within a language system for the time being accepted as adequate only indirectly make use of objective experience, Habermas argues. In testing validity claims of propositions what matters is not comparing individual sentences with reality but their justifiability by means of an adequate theoretical language system. We conclude from this that Habermas holds the being-grounded of a consensus on (the truth of) a proposition in the first instance to be based on the proven adequacy of the theoretical language system which is used to formulate the proposition, so that the truth of the proposition concerned depends also only indirectly on experience. A necessary condition for the being-grounded of a consensus on the truth of a pro-

29

position C is the existence of a deductive-nomological explanation of C, with the help of data D and an empirical regularity W which is thought plausible on the basis of the empirical evidence B, in which C, D, W and B are formulated in terms of an adequate theoretical language system. I call this condition material, because a relation to the material reality, although indirect, plays a part in it.59

1.5.3

The formal characteristics of discourse

Up to this point Habermas theory of truth could at best be adequate for a static or normal science, in which the adequacy of a once accepted theoretical language system is no longer debated and in which all progress takes place within this language system (within a fixed paradigm).60 But for Habermas, too, only part of the problem has been treated. Scientific progress is possible because there are degrees of adequacy of theoretical language systems. Progress should be seen as a critical development of theory languages which interpret the prescientific object domain more and more adequately.61 In the course of the history of the sciences we regularly see that theoretical language systems come to be considered less or even no longer at all suitable, and are replaced by new ones. The theory-specific cognitive schemata are then changed (whereas the fundamental, unchangeable schemata stay of course the same). Habermas idea now is to formulate a number of structural, defining characteristics for theoretical discourse, which make it possible to also introduce alternative theoretical language systems and discuss them in an argumentation on the truth of a proposition. If in a certain case, due to contingent reasons (e.g. the power relations in an industrialized science), an alternative theoretical interpretation of a proposition and the corresponding theoretical language system is systematically excluded from consideration in a scientific discussion, any reached consensus is in Habermas eyes not rationally motivated and simply sham; truth has not come into it at all. To enable him to take into account this distinction between genuine and false consensus in his theory of truth Habermas insists that within theoretical discourse what he calls a radical critique of knowledge should be possible; i.e. he suggests that within discourse it should be possible at any time and in any place to argue at four different levels62:

30

1. It should always be possible to question the acceptability of a proposition and introduce it into discourse. 2. Within discourse it should always be possible to supply, with the help of a chosen theoretical language system, an argumentation for or against the proposition concerned. 3. It should always be possible to subject this particular language system to discussion, and to argue for or against other theoretical interpretations of the proposition concerned. 4. It should always be possible to debate the (necessarily normative) concept of knowledge underlying the science concerned by opposing to it another conception, an alternative science or an alternative scientific practice and plead for this. At this last level one can, in an unconstrained discourse aiming at the attainment of consensus, also debate the use of the fundamental cognitive schemata (e.g. of causality in sociology) and one can advocate other approaches (e.g. a hermeneutic sociology). These defining characteristics of discourse which make it possible to revise a once chosen theoretical language system and replace it by a better one have been phrased by Habermas on purpose rather ideally. They do not simply refer to the practice of the sciences. Only under a number of particular social conditions, i.e. those of the ideal speech situation, can we be certain of the existence of a (real) discourse with the above-mentioned possibilities of a radical critique of knowledge. Because the criterion of truth is a grounded consensus within the discourse the truth of a proposition is likewise dependent on the realization of this ideal speech situation:
A consensus reached through argumentation is an adequate criterion for the redemption of discursive claims on validity if and only if on the basis of the formal characteristics of the discourse the freedom to move between the levels of discourse is guaranteed. But which are the formal characteristics which fulfill this condition? I propose: the characteristics of an ideal speech situation.63

Thus not only material aspects but also communicative ones, play a part in Habermas criterion of truth. The criterion of truth, in the sense of a sufficient condition for a claim to truth to be justified, is the grounding of a consensus reached through argumentation. A grounded consensus requires (1) the deductive and inductive steps in the given argumentation to be correct or plausible and the theoretical language system used to be adequate to the object-domain concerned (the material condition); and (2) the argumentative consensus to have resulted from a real discourse, i.e. that it was possible to argue freely at and between the four levels of the discourse (the communicative condition). I therefore

31

interpret the last quotation from Habermas as follows: the material and the communicative condition are each separately necessary and in combination necessary and sufficient for the being grounded of an (actually reached) consensus on the truth of a proposition, while grounded consensus is in its turn sufficient to justify the claim to truth of that proposition.64 We should not forget that the distinction between the material and communicative condition is of an analytic nature. According to Habermas theory these conditions can only function in combination and never separately as a criterion of truth. Sometimes people have overlooked the fact that both elements, the material and the communicative, play an essential role in Habermas theory. Thus Keuth for instance comes up with a couple of problems for the consensus-theory precisely because he regards the material condition in itself as sufficient for the being-justified of a truth-claim.65 While Puntel on the other hand suggests an interpretation of Habermas concept of truth in which the material aspect is not explicitly taken into account.66

1.5.4

The ideal speech situation

A next and final step in Habermas explication of his truth theory consists of an explanation of the meaning and function of the ideal speech situation. The communication structure of the ideal speech situation should be thus that systematic impediments and failures in communication do not occur. This situation should make it possible that at any moment a discourse can be embarked upon and that within such a discourse at any time one can move between the four levels of discussion, so that no presumption can remain exempt from thematising and criticism.67 Habermas translates this requirement in his theory of speech acts as follows: all potential participants in the discourse (and they are in principle all people in the past, present and future68) should have an equal share of the chances to perform speech acts. This general requirement is then further specified in these four conditions for the ideal speech situation69: 1) All potential participants should have within the discourse the same communicative possibilities, to speak, oppose, question, answer, etc. 2) All potential participants should have within the discourse equal chances to formulate statements, predictions, explanations etc. and to problematize, substantiate or repudiate their claim to validity.

32

3) All potential participants should have the same material and social possibilities within the domain of communicative action, so that they are able to suspend the forces of reality and to enter into the experience-free and action-free domain of the discourse.70 4) All potential participants should have equal chances to express honestly their attitude, sentiments, intentions etc. within the domain of communicative action in order to guarantee that those acting are true to themselves also as participants in the discourse and make their inner nature transparant.71 The difference between the first two and the last two conditions is of vital importance. Whereas the first two concern requirements relating directly to the discourse, the last two concern the domain of communicative action, the entire social context in which science functions. If these four conditions have been fulfilled, Habermas argues, we can be sure that we are dealing with an ideal speech situation and that in such a discourse a grounded consensus is possible. With regard to the status of this ideal speech situation Habermas himself notes an obvious problem: that is, whether such a situation has ever been or can be realized. To further determine the status of the ideal speech situation Habermas uses a number of terms taken from the philosophical tradition.72 The ideal speech situation is neither an empirical phenomenon nor purely a construction; it is not merely a regulative but also a constitutive condition for every verbal communicative action here and now. In pursuing or maintaining mutual understanding Habermas holds that we do in fact presume the generally counterfactual presence and operation of the ideal speech situation. Whether we can do something with this terminology we will have to look into later on. Now I would like to conclude by pointing out two aspects of Habermas theory of truth. Firstly there exists no discourseindependent external criterion to determine whether at a given moment we are holding a (real) discourse, or a sham discourse, so that we can never decide with certainty whether a certain proposition is true or false. Therefore Habermas criterion of truth is fallibilistic not only with respect to the material, but also to the communicative aspect: every consensus on the truth of a proposition accepted for the time being as grounded can later prove to be accidental or coerced. Secondly, by bringing material and communicative aspects together, Habermas links truth and rationality in his criterion of truth.

33

All those propositions are true on which a rational consensus can be reached between all potential participants in the discourse.73 Habermas theory of truth, or rationality, is an attempt on the one hand to take seriously the fact that all acquisition of knowledge takes place with the help of historically and socially situated theoretical language systems without on the other hand slipping into the relativist conclusion, which often follows this, that truth is identical to a factual consensus, however reached. Habermas claims to have found a universal trait in human communication in the notion of the ideal speech situation, with which it is in principle possible to distinguish between an accidental and a grounded, a forced and a rational consensus on the truth of propositions.

1.6

Objectivity and truth

I would like to start by putting right a misunderstanding with regard to the distinction made between objectivity and truth, to which early formulations by Habermas may also have contributed. In the discussions of the constitution theory of experience and its difference from the truthproblem I have mainly used the articles Theorie der Gesellschaft oder Sozialtechnologie and Wahrheitstheorien and the postscript to the book Knowledge and Human Interests and deliberately not Vorbereitende Bemerkungen zu einer Theorie der Kommunikativen Kompetenz. The latter contains namely a number of opaque and incorrect remarks in the section on the consensus theory of truth, which later have been corrected by Habermas. It concerns a still insufficiently clear distinction between the objectivity of experience and the truth of propositions or, more specifically, between the categorial and the discursive meaning of propositions. In this article there is the suggestion that perception, too, can be a criterion of truth, while moreover the exact meaning of criterion remains unclear. Habermas regards perception as a nonconventionally chosen method for testing empirical observations and for forcing a consensus on the validity of the corresponding statements.74 Such non-conventional testing is possible, Habermas proposes, because observation predicates can in principle be operationalized and the objects to which they apply be identified by these operations. In their turn these operations can then be reduced to basic, constitutive action schemes on the basis of which we assign substance, space, time and causality to things and events.

34

The premiss of this argumentation which interests me here is the reductionist view that the (complete) meaning of observation predicates is given by and merges into the way in which they are operationalized. This position, which has already been superseded in the philosophy of science, has been criticized justly in the case of Habermas by Beckermann.75 I hope, however, it will have become clear that this position is also incompatible with Habermas own sharp distinction between the categorial and the discursive meaning of propositions. The categorial meaning concerns, just as in every transcendental approach and as the term itself denotes, only a limited number of categorial characteristics of possible objects of experience in general (spatial, temporal, substantial and causal for instance) and certainly not all its material characteristics (e.g. resistance, temperature, colour). The a priori of experience never determines the contents, only the general form of experience. A description of experience as regards contents takes place with the help of discursively justifiable propositions, on the basis of theoretical language systems which are in a way conventional and changeable and which supply the discursive meaning of the propositions. It is exactly this insight which brought Habermas to distinguish objectivity from truth. Therefore, in concluding this chapter, I would like to make some remarks on his reasons for introducing this distinction. Historically, it dates back to Peirce, who also differentiated his pragmatic theory of meaning from his consensus theory of truth.76 Habermas main epistemological argument for the introduction of this distinction follows on his recognition of the importance and the relative autonomy of theoretical argumentation in the development of science. He sees scientific progress as development of discontinuously changing theoretical language systems which interpret the prescientifically constituted objectdomain of things and events in an ever more adequate way.77 Theoretical progress does not only take place on the basis of the production of new experiences, but also and primarily on the basis of reinterpreting the same experience. This view of progress as re-interpreting identical experiences implies that it is impossible to have the conditions for the possibility of experience (the a priori of experience) coincide with the conditions for the possibility of discourses (the a priori of argumentative reasoning), something Kant could still do by regarding Newtonian physics as definitive. Habermas view is that both a prioris fulfill an essential but different task in science:

35

However, both combine to define the limits of empirical theories . Theories can only be constructed, and progressively reconstructed, in the context of conditions pertaining to the nature of argumentation and within the limits of prior objectivation of experienceable occurrences. Conditions pertaining to the nature of argumentation refers to the discursive verification of systems of propositions. Within the limits of prior objectivation of experienceable occurrences means: in a theoretical language whose fundamental predicates are always related to the independently constituted objects of possible experience.78

Habermas considers it the task of, what he calls, a non-objectivist theory of science to show in detail that the logic of science exactly is the logic of the relation between the a prioris of experience and of argumentative reasoning. I think that distinguishing objectivity from truth follows from a correct insight, which makes it worthwile to examine the idea in more detail. But after the expositions in this chapter, which summarize the main points Habermas makes on this subject in sufficient detail, it will be clear that greater accuracy and further specification is needed. In the following chapter I will therefore subject Habermas ideas to precise analysis and criticism, the aim being to find out which of these ideas are useful for the second part of this study in which I explain my own philosophical view on the experimental natural sciences.

36

Analysis and criticism

2.1

Introductory remarks

In this second chapter I intend analyzing critically Habermas views described above with respect to the experimental natural sciences. I do not pretend to offer a criticism as exhaustive as possible. My primary concern is to analyze and evaluate some aspects of Habermas theory which appear at first sight useful in relation to the philosophical problems in part II. In structuring this chapter I have followed the same order as in the previous one in describing Habermas view. In section 2.2 a number of critical comments will be made on his (quasi-)transcendental approach and on his ideas on the role of the experiment in the natural sciences. In 2.3 I first attempt to make a closer analysis of Habermas point of view by means of a sharper distinction and a more accurate elaboration of the questions of the meaning and of the criterion of truth. This then enables me to show in a detailed argumentation the inadequacy of the criterion of truth proposed by Habermas. For all my criticism, however, I will not lose sight of those elements from Habermas theory which seem to me fruitful for the second part of this work.

2.2

The transcendental experiment

method

and

the

role

of

the

I start this section with a critical discussion of Habermas constitution theory. I will focus on two points. First of all on the problem of the universality of the relation between instrumental action and the objectivity of sensorial experience; and secondly on the structure and the possibilities of Habermas transcendental arguments. In section 2.2.2 I will make two brief remarks on Habermas ideas on experimenting in the

37

natural sciences. These concern the idea of perception as quasiexperimentation and the diffference between technical and experimental instrumental action.

2.2.1

On the constitution of objectivity

For the sake of clarity I will again summarize point by point the structure of the argumentation as proposed by Habermas and as I described it in 1.4.1. First of all he assumes that sensorial experience or perception will always be inextricably linked with language, in this case with the so-called thing-event language. Vice versa we can read from this language the categorial structure of the objects of possible experience. The point of the transcendental-philosophical approach is that the categorial structure of the objects of possible experience is made explicit in a direct relation to the conditions of possibility of the objectivity of experience. Therefore instrumental action is both constitutive of the categorial structure of objects of possible sensorial experience and of the objectivity of the experience. To render this true two distinct, but closely connected, argumentations are required. The first goes like this: 1a: the language of things and events establishes the categorial structure of objects of possible sensorial experience. 1b: this language appears to contain semantic categories (causality etc.), which make the reference to things and events possible. 1c: these semantic categories have (ever) been formed as the sediment from the application of successful instrumental action schemes. The second argument, that instrumental action is at the same time constitutive of the objectivity of experience, should be seen according to Habermas like this: 2a: the objectivity of experience consists in its capacity to be intersubjectively shared. 2b: the intersubjectivity of sensorial experience ultimately rests on its instrumental controllability. 2c: instrumental action guarantees this controllability in two ways, viz. empirically (instrumental action is anthropologically invariant) and transcendentally (the instrumental controllability is a priori necessary). One of the problems of this argumentation is that it is rather sketchy. As Habermas admits himself, his statements so far on the tran-

38

scendental relation between (objects of) sensorial experience and instrumental action do not pretend to be a systematic, conceptual analysis. He regards his philosophical interpretation of Peirce and Piagets cognitive psychology as mere indications that he is with his transcendental-pragmatic approach on the right track.1 As we will see below even this can be doubted. For a start there is the problem of the relation between instrumental action and the objectivity of sensorial experience. In Habermas eyes instrumental action is anthropologically invariant because there is an intrinsic link with the universal interest of humanity to safeguard survival by means of labour. This idea that all people (must) act instrumentally seems plausible enough. But Habermas clearly goes further. His earlier work implies that the relation between instrumental action and objective experience as it is expressed in the language of things and events is also of a universal character. But this is surely incorrect. The universality of instrumental action does not after all exclude that in other ages and cultures the daily contact and also the instrumental intervention with nature was interpreted in, for example, magical or expressive terms.2 In his later work Habermas recognizes this problem. His solving it boils down to on the one hand dropping his claim of universality with respect to the relation between instrumental action and sensorial experience of things and events as a historical claim. In other words: he admits the existence of societies in which there is instrumental action, but in which this action is not felt in terms of things and events. On the other hand he does stick to his view of a systematic link between instrumental action and the objectivity of experience. In other words: objectivity can be assigned only to the instrumentally controllable experience of things and events.3 Habermas now describes the genesis of the objectivizing attitude as an evolutionary process of rationalization, in which the objective world of things and events is distinguished from the social and from the subjective world.4 This solution of the problem is, in any case for me, far from satisfactory. To begin with, I do not see how Habermas can so easily deny objectivity to magical or expressive experience, precisely because he himself identifies objectivity with a certain capacity of being intersubjectively shared. Furthermore, Habermas analysis has strong normative overtones. On the one hand we have the mythical world view which he characterizes in terms of illusion, confusion, conceptual blending, defectiveness etc.; opposed to this stands the modern world view, in which the right distinctions are made and for which

39

objectivity and rationality are characteristic.5 At best this seems to me to be naive, at worst imperialistic and ethnocentric.6 More in general I find these views problematic because they offer no advantages. I can find no good reasons (not in Habermas either) for believing them. I will now let this matter rest7 and pass on to an evaluation of Habermas horizontal reconstruction of the systematic relation between instrumental action and the objectivity of sensorial experience.8 Habermas does not pretend to give an absolute, metaphysical foundation of (scientific) knowledge. He bases his position on the actual existence of a language of things and events and of the objectivity of experience, and then looks for the conditions of its possibility. He attempts to find these conditions on the ground of a conceptual analysis of this language and experience, with the aim to make explicit the implicitly applied, anonymous (i.e. not individually bound) systems of rules, such as for instance in instrumental action. All this suggests that Habermas regards transcendental argumentation as an analytic activity, i.e. a purely conceptual explication of fundamental categories as conditions for the objectivity of our experiences and their verbal expression. There is thus no question of a transcendental deduction, of a metaphysical founding of these categories in some sort of necessary structure of a transcendental consciousness. The only justification is the reference to the functioning of these categories in contingently existing verbal experience.9 Nevertheless Habermas claims that instrumental action is constitutive of the objectivity of experience and of successful reference to objects of possible experience. Because he definitely does not wish to return to Kants synthetic activities of a transcendental consciousness, one wonders what he does mean with this concept constitution. This question is not sufficiently answered in the transition from transcendental to quasi-transcendental. Lacking a clear answer from Habermas it seems sensible to have another look at this problem from the analytic tradition, a connection which Habermas has tried in other fields. In the modern, analytic reception of Kant transcendental arguments are usually analyzed in terms of necessary or sufficient conditions.10 A first interpretation could then be that the possibility of the instrumental controllability of an experience is a sufficient condition for the objective character (in terms of things and events) of this experience.

40

I consider this interpretation untenable. A major objection is that de facto instrumental controllability is not even sufficient to ensure the occurrence of experience in terms of things and events. Because: even if people act instrumentally, they do not necessarily have to interpret or describe the accompanying experience in terms of things and events. Therefore instrumental action can certainly not be sufficient for the objectivity of such an experience. Besides, other conditions may also suffice for the objectivity of sensorial experience. Moreover, opponents of Habermas could argue, they can be more relevant or fundamental with respect to the objectivity of experience than instrumental controllability. I will deal with this problem of the uniqueness of the condition of possibility in greater detail below. The second possibility is an interpretation in terms of necessary conditions. In the light of the intuitive meaning of the term condition of the possibility of such an interpretation seems at once more satisfactory. In the case of Habermas, moreover, the idea of instrumental controllability as a necessary condition for the objective experience of things and events does not seem wrong straightaway. I would grant Habermas the benefit of the doubt if he were to work out at this point his germinal argumentation and make it plausible. But even if he were to succeed the problem remains that also in the case of necessary conditions uniqueness cannot be guaranteed. There is in principle always, and in practice often, more than one necessary condition conceivable or present as a condition of possibility of the objectivity of experience. To make acceptable that his explanation of the objectivity of experience in terms of instrumental action is the only right one, Habermas would have to analyze all existing and imaginable alternatives and prove their inadequacy. An impossible task. This is why philosophers such as Rorty and Bhaskar stress the ad hominem and critical character of transcendental philosophical arguments.11 Philosophy does not take place in a vacuum or an ivory tower. It cannot simply justify its conclusions by means of logic or facts. Philosophy is also a confrontation with and a criticism of alternative views. An important function of philosophical argumentations is to criticize with a transcendental refutation at any rate the extant alternative views as wrong, implausible, insufficient, inadequate, inconsistent, superfluous or undesirable. The question then is how does Habermas transcendental refutation stand up to alternative, and especially realist views on the objectivity of experience.

41

First of all there is the correspondence theory of truth, which tries to make the objectivity of experience explicit in terms of a correspondence with reality. Habermas argues that this theory is incoherent.
When we cannot ascribe any other meaning to the term reality than that which we associate with propositions about facts and when we conceive of the world as the sum total of all facts, then again can the correspondence relation between propositions and reality only be determined by propositions. The correspondence theory tries in vain to break away out of the sphere of language, in which only the validity claim of speech acts can be explained.12

But this is certainly not a transcendental refutation of realist correspondence theories: the argument contains a presupposition which is expressly rejected by the realists themselves as an epistemic fallacy.13 According to realists, the meaning of the term reality is not restricted to the sum total of all facts which we (as human beings) can possibly assert in propositions.14 The reality of nature can very well be conceived of as existing independently of human existence and knowledge, and thus also independently of the existence of propositions and facts. Such a reality without people is not only imaginable, it can also in a time of nuclear overkill be realized quite simply. Thus Habermas argument certainly does not show the incoherence of correspondence theories. I do, however, agree with the intention of Habermas criticism. For there are other and more poignant arguments against a correspondence interpretation of the truth of propositions in the natural sciences. These, however, require a discussion much more detailed than Habermas gives here. Later in this work I offer an extended criticism of correspondence theories of truth.15 In spite of this sympathy with Habermas intention I cannot concur with his transcendental approach. In part II of this book I will defend a realist view of scientific propositions, which on the one hand does not suffer from the drawbacks of the correspondence theory of truth and on the other hand has some obvious merits compared with Habermas view. I will already mention two of these advantages. First of all, this view will not be based on the problematic claim of a universal relation between instrumental action and the objective experience of things and events. Secondly, the epistemic fallacy will be avoided by differentiating between what I call the material realization of experiments and the description of this material realization via the language of things and events. In my eyes this description plays a role only in relation to the acquisition of knowledge about the object-domains in the natural sciences and not in

42

relation to their constitution. The crucial point being that this constitution takes place exclusively on the basis of the specific theoretical language systems. As we will see, in this way the relation between objective experience and true propositions, or more specific, between fundamental and theoretical cognitive schemata, which is problematic in Habermas, can be clarified satisfactorily.

2.2.2

The role of the experiment in Habermas

In the present section I will briefly touch on Habermas views on experimental action, which I summarized in 1.4.2. I will limit myself to two questions, viz.: is a characterization of perception as quasi-experimentation adequate; and: what are the consequences of Habermas further distinction between technical and experimental instrumental action? To answer the first question we need of course to know what exactly is meant by the prefix quasi. Although Habermas does not give an explicit description anywhere, this quotation will show what he more or less intends:
The nature of perception guarantees intersubjectivity in that it can be controlled, when it is problemized to a sufficient extent, in the form of an experiment. This interesting aspect is clearly related to the fact that observational predicates can in principle be operationalized and the objects to which observational predicates refer can in principle be identified with the help of operations. Perceptions are methodically secured in so far as they are founded on experiences of the success of repeatable operations. These again can be led back to operations, which are effected according to rules of physical measurement.16

Habermas seems to mean by his viewing perceptions as quasiexperiments that perceptions obtain their methodically secured character guaranteeing intersubjectivity from the same state of affairs as experiments do, i.e. from experiences of the success of repeatable operations. Thus instrumental, repeatable action is both a condition of the possibility of objective (i.e. intersubjective) perception and of the possibility of scientific experimentation. I will not again go into this argumentation in great detail but will be content to state that Habermas (at best) shows that perception can be called quasi-experimental with regard to one aspect only that of its objectivity. This does not imply, however, that there cannot be important differences between perceiving and experimenting. Thus Janich for instance a defendent of protophysics and therefore a man after 43

Habermas own heart distinguishes explicitly between observation instruments, measuring instruments and other experimental devices. He assumes that measuring instruments and other experimental devices always produce artificial phenomena, whereas observations concern phenomena which are not the result of active human intervention.17 Because I also consider this or a similar distinction necessary for an adequate philosophical discussion of the natural sciences, I cannot agree with Habermas in the identification of empirical with experimental sciences, as he suggests it. Hence, I have restricted the claims of my discussion explicitly and on purpose to the experimental natural sciences, to chemistry, physics and parts of biology, geology and medical science. The second question I would like to discuss briefly here concerns the distinction between technical and experimental instrumental action. What matters in technical action is that a given goal is reached with the help of certain, also given, technical means. The experiment, on the other hand, serves in the later Habermas for the acquisition not only of objective experience, but also of discursively usable data. This entails that in the experiment the means used and the goals set can themselves be problemized. After all: both the description of the experimental means and the hypothesis (goal) to be tested will have to be interpreted with the help of a certain theoretical language system, and such an interpretation can subsequently always be disputed and replaced by another one. This implies that the experiment has also an epistemic meaning apart from an instrumental or pragmatic meaning, that it also aims at the acquisition of true knowledge. In other words: the success of an experiment lies not only in the actual feasibility of the instrumental actions needed, but just as much in the usefulness of the obtained data for discursive argumentations concerning the truth of propositions:
Experimental action does not lie on the same level as instrumental action of naive or scientised practice. In its function of producing data, which is always gathered with a view to testing hypothetical validity-claims, experimental action is related to discourse from the start. This is a vague expression: I readily admit that the evidential dimension of the concept of truth is badly in need of further clarification.18

A major objective of the second part of this study is to offer such a clarification of the role of experimentation with regard to the problems of objectivity and truth. But before we can start with this we need a detailed evaluation of Habermas theory of truth.

44

2.3

A criticism of Habermas theory of truth

The main aim of this section is to subject the consensus theory of truth to a careful examination and a detailed criticism. In order to achieve this it is necessary to find out what kind of questions Habermas theory tries to answer. In philosophical discussions of truth it is often and justly so pointed out that we can ask various questions with respect to this problem, which all need a different answer and which moreover have different philosophical relevance.19 As a start I pose the following three separate questions: (1) What is the meaning of truth? Assuming that truth is assigned to propositions, this boils down to the question of the meaning of proposition C is true, in which C stands for an arbitrarily chosen proposition from the experimental natural sciences. (2) Under what kind of conditions can we correctly assert C is true? What is at stake here is not the conditions for the being-true of propositions (often taken as an explanation of the first question), but the kind of conditions for a correct use of the concept of truth. To answer this question we must offer one or more sufficient conditions for correctly assigning truth to arbitrarily chosen propositions C from the natural sciences. I will call the sum total of these sufficient conditions the criterion of truth. (3) How do we find out for concrete propositions C, that C is true? Here we are concerned with the question of how to track down the actual conditions for the truth of specific propositions (such as this liquid boils at 120C); in other words: how can we verify or falsify these propositions in practice. Taken the fact that the subject of this study is of a philosophical and not of a scientific nature this third question does not play a systematic role in the further discussion.20 What I will attempt is a clarification and, when necessary a criticism of Habermas theory with the help of the first two questions.

2.3.1

On the meaning of truth

The question of the meaning of truth in Habermas theory can be answered quite unambiguously on the basis of what has already been said in the previous chapter (see 1.5.1). Truth is, Habermas argues, the claim to validity which people attach to propositions when they assert

45

them. This claim which is accepted tacitly in communicative action, can be grounded in a discourse with the help of valid argumentations, always and everywhere. Whether such a definition is satisfactory depends in the first place on whether the defining concepts are sufficiently clear and clarifying (the most important are: discourse and claim to validity). I will not again go into this, but refer back to the previous chapter. In the second place, one can judge a definition on what it does not denote and on its consequences. On this I would like to make some remarks in this section. The most striking aspect of the definition is that the question of the meaning of truth is answered by Habermas in (pragmatically coloured) epistemological terms right from the beginning. He does draw a formal distinction between the truth and the truth-claim of a proposition, but simultaneously he proposes that an explanation of this distinction can only take place in epistemological terms:
The truth-claim advanced for p is certainly not identical with the truth or validity of p. P is true if the truth conditions of p are satisfied; but to raise the corresponding validity-claim is to offer to defend in argument against opponents the assertion that the truth conditions for p are satisfied. The one cannot be equated with the other. The point of the discourse theory of truth is that it attempts to show why the question of what it means for the truth conditions of p to be satisfied can only be answered by explaining what it means to redeem or to ground with arguments the claim that the truth conditions for p are satisfied.21

According to Habermas we cannot explain the fact that the truth conditions for p are satisfied in the sense that there is something in reality which makes p true. Semantic notions such as correspondence, reference etc. do not enter into his theory of truth. In Habermas true propositions do correspond with facts, but these facts are not something in the world; they are neither things nor events. The major problem with which correspondence theories wrestle the relation between reality and verbal representation is not so much solved by Habermas as pushed aside as a fake problem, on the basis of the argument that we can only have access to reality through language. As we saw above this argument in itself is not sufficient to refute correspondence theories. What has not been shown is the impossibility of making the relation between language and reality coherently and unambiguously explicit in language. But I think that there are better arguments, at any rate in the case of scientific propositions, against the correspondence theory. These are based on the fact that (even though there be only one reality) there are many scientific languages which in

46

general cannot be reduced to one basic language for instance because they contain mutually incompatible conceptual characteristics , but which nevertheless can be accepted as applying to one and the same domain of reality. When such a situation is rather rule than exception a notion of correspondence with reality becomes very problematic. In part II of this study (in particular in section 4.3) I will work out such a criticism of correspondence views in more detail. In his universal pragmatics Habermas defines truth in the first place as an implicit presupposition in ordinary language use, and his point of view is characteristic in that it does not only have factual, but also universal, counterfactual moments. Generally speaking I do not disagree with Habermas idea that the meaning of concepts such as truth should be made explicit in close connection with its ordinary practical usage. The specific elaboration of this idea by Habermas, however, does not seem tenable to me. In the following section I will treat the criterion of truth as proposed by Habermas in extenso and show its inadequacy. On the other hand, such a search for a connection with the practice of, in my case, the experimental natural sciences does not imply that merely because of this attempts at a realist interpretation of true propositions have become impossible. My criticism of Habermas is exactly this that he does not reach beyond epistemology to realist questions on the relation between language and reality. Why can we not enquire into the realist aspects of a concept of truth clarified in the first place epistemologically (and also methodologically)? In part II I will try to deal with these questions, focussing primarily on the problem of reference of theoretical terms.

2.3.2

The inadequacy of the criterion of truth

In this final section I analyze and criticize Habermas criterion of truth in six points. 1. The second question I asked concerned the criterion of truth, in the specific sense of the sufficient condition(s) under which we can justly assert that C is true. In the sections 1.5.2 and 1.5.3 I gave in detail Habermas answer to this question.22 In his view truth is justly attributed to a proposition if a grounded consensus on it has been reached within a theoretical discourse, a consensus being grounded if and only if both the material and the communicative conditions have been fulfilled.

47

The material condition ensures that the deductive and inductive steps in the argumentation needed are respectively correct or plausible and that the theoretical language system used is adequate to the object-domain concerned. The communicative condition requires that the consensus has been reached under the conditions of the ideal speech situation, so that the possibility of freely exchanging arguments at or between the four levels of discourse is guaranteed. On the one hand, Habermas criterion of truth differs from his definition of truth primarily in that in the criterion a grounded consensus should factually be reached whereas in the definition of truth what matters is the possibility of a grounded consensus.23 On the other hand the formulation shows that this criterion does not indicate any specific methods or procedures by which we could verify the truth of concrete propositions from the natural sciences.24 Below I will criticize this criterion of truth proposed by Habermas in detail. In my criticism I take the line that an adequate criterion of truth should be required to be either applicable to paradigmatic cases of true scientific propositions or as a regulative notion to give meaning to the idea of convergence towards truth. This does not entail that a criterion of truth should or could give us certainty on the (approximate) truth of propositions. It should, however, give us good reasons for believing in their (approximate) truth. If the criterion is so fallibilistic that it is not relevant to concrete scientific propositions in this sense, then I believe it to be inadequate in an explication of the problem of the truth of these propositions. 2. Is the condition which Habermas proposes in his criterion of truth indeed sufficient for justly assigning truth to a proposition C? As we have seen both a material and a communicative condition for the being justified of claims to validity play a part in this criterion of truth. In the first place I will evaluate the material condition here. This imports that there should be a consensus on a deductive-nomological explanation of C, with the help of data D and an empirical regularity or law W, which is considered plausible on the basis of empirical evidence B, and this should all be phrased in terms of an adequate theoretical language system T. It is, however, clear that this condition, combined with the communicative condition of the ideal speech situation is certainly not sufficient for justly assigning truth to C. In fact Habermas makes at this stage a simple, formal mistake, but one with far-reaching con-

48

sequences.25 In the material condition he only requires the plausibility of W:


The consensus oriented power of an argument rests on the transition from B to W justified through induction, on a cognitive development which guarantees the adequacy of the description system and which precedes every single argumentation.26

The fundamental problem is that in the material condition no demands are made upon the singular propositions, i.e. D and B. Habermas forgets that before we can speak of a criterion of truth as a sufficient condition there should be a consensus on the being-justified of the claim of truth for data D and backing B. After all, the deduction of C with the help of law W and data D is in general only valid if data D are justly called true; and for the plausibility of law W in general the truth of backing propositions B is necessary. This fundamental problem cannot be solved within Habermas theory, neither when we stay within a certain theoretical language system nor when we enter another, more adequate language system. I will show this below with regard to the problem of the truth of data D. First of all: a grounded consensus on data D can never be reached only on the basis of an adequate theoretical language system plus the accompanying plausible laws. In the example in which D is the singular proposition this water has a temperature of 100C of course the adequacy of the theoretical language system concerned and the plausibility of the laws used are also required for a grounded consensus. But this in itself is not sufficient to make the truth of D acceptable. However long we analyze our concepts and theories on the phenomenon of boiling, we will never only on the basis of this reach a consensus on the fact whether this water has a temperature of 100C. In some way or other an explicit relation to the objective experience has to play a role in reaching a grounded consensus on this proposition.27 We cannot solve this problem either by starting a completely new argumentation within the same theoretical language system with D for a conclusion. For in that case the same difficulty will arise with respect to data D, which are needed for this new argumentation. Ad infinitum. This means that we will either get an infinite regression, so that Habermas criterion is not practically applicable; or that if we stop somewhere the assigning of truth to D and so to C is not justified.28 Thus the point is that Habermas does not come up with epistemological arguments on the basis of which we could for good reasons put a stop to the infinite regression at a certain point.

49

In the second place I would like to consider whether Habermas could solve this problem by radicalizing the critique of knowledge, or in other words, by re-interpreting data D with the help of a more adequate theoretical language system and more plausible laws. This aspect of Habermas criterion of truth implies a clear distinction between the role of data D and of laws W. This follows from the fact of the material condition being sufficient for justly assigning truth only when combined with the communicative condition. If the conditions of the ideal speech situation are also fulfilled, the most adequate theoretical language system and the most plausible law will prevail in the end this is at least Habermas hypothesis. One can of course doubt this (and I will do so later on), but formally there is no fault.29 The difference between the role of W and that of D can be given in exact terms. If we were to systematically doubt the plausibility of W in a real discourse, we would obtain, according to Habermas, more and more adequate theoretical language systems and more and more plausible laws. This reinterpretation with the help of more adequate language systems of course implies also a reinterpretation of D, e.g. to D . But the problem, as I also pointed out above, is that general characteristics such as (greater) adequacy and plausibility can never be sufficient for the justly assigning of truth to singular propositions. In other words: D is not more adequate than D. We could carry on the radicalization of the knowledge of data indefinitely but it would get us nowhere! From this criticism we can conclude that Habermas criterion of truth has a major flaw. This flaw is not the fact that in an argumentation about the assigning of truth to C the truth of D can be doubted, so that a new argumentation is required with D as a new C. After all the same goes for the plausibility of W. But whereas Habermas obtains from the ideal speech situation the conditions which will enable us finally to distinguish a factual from a grounded consensus on the adequacy of the chosen theoretical language system and the plausibility of the general laws linked to it, such a criterion is completely lacking when one has to take a decision on the assigning of truth to the singular propositions D and B. From the above follows that in that case a factual consensus suffices as a criterion of truth, contrary to an important assumption of Habermas theory.30 A direction in which the solution of this contradiction could be found is the following. Habermas assigns a certain materialist or realist meaning to a plausible law, which is formulated with the help of a

50

theoretical language system which is accommodated to the domain of objects it describes: reality has been caught in this language system (although we should continue our search for more adequate theoretical language systems). It seems to me that in order to come out of the deadlock such a material condition should be formulated also with respect to the truth of singular propositions such as D and B. In part II and in particular in section 4.4 I will try to solve the problems indicated here, problems which are in my view inherent to and unsolvable within Habermas theory of truth. The central element in the criticism of the criterion of truth as given up till now concerns the fact that for Habermas the problem of truth lies (primarily) in the universal propositions:
Paradigms of knowledge with which the meaning of truth can be made explicit are not the singular propositions, with which observations can be communicated, but universal, negative and modal propositions; they express the specificity of knowledge: that is the conceptual organization of the material of experience.31

This does not mean that Habermas only treats truth-questions in relation to universal propositions. On the contrary, in the examples he gives C is always a singular proposition. However: he holds that the core of the problem of truth of singular propositions such as C lies in the fact that an argumentation is needed which uses a universal law W. I criticize this because in such an argumentation true or acceptable data D (and empirical evidences B) should also be made use of and Habermas does not discuss this problem at all. In recent philosophy of science it has been shown convincingly that also observations and observation statements are conceptually organized in an irreducible way. Even singular observation statements have inherently a general aspect, which does not imply, however, that because of this all differences with universal propositions fade away. Present philosophy of science should therefore account for these nuances accurately and not lump everything together. An important consequence of the fact that Habermas only proposes requirements for (the revision of) general theoretical language systems and not for data D is that his theory of truth closely approaches a pure coherence theory.32 3. A second criticism of Habermas criterion of truth and again of its material aspect concerns the required adequacy of the theoretical language system T in which both propositions C and D and the plausible law W and the background support B are formulated.33 As I explained in section 1.5.2 Habermas analyzes the adequacy of theoretical language

51

systems in terms of cognitive schemata, distinguishing the fundamental from the less fundamental. At first sight this difference might appear to be one of degree. But after closer examination this does not prove to be the case.34 The difference between fundamental and less fundamental is after all linked to that between objectivity and truth. The fundamental schemata determine the objectivity of experience and stand only in an indirect relation to the truth of propositions. This truth depends on among other things the adequacy of less fundamental, theory-specific cognitive schemata. In Habermas view for example the truth of propositions from the general theory of relativity (implying the non-Euclidean nature of space) can be compatible with the fact that the objectivity of experience is dependent on a cognitive schematizing of space in a Euclidean sense. And with respect to the interpretation of quantum mechanics he writes:
The identity of experiences in the manifold of the interpretations we produce of them is assured because of the conditions of possible objectivation. The particular view of quantum theory developed at Copenhagen provides considerable support for this position: It is argued there that the classical conceptions needed to describe a measuring apparatus point up the limits of the autonomous object domain of bodies in motion. The nonclassical theories of modern physics can interpret this domain differently but they cannot put a new object domain in its place.35

Whatever we might think of this, at this point I only want to observe that in Habermas theory the question of the fundamental cognitive schemata is only indirectly relevant to the problem of truth. I will therefore restrict myself to the problem of the adequacy of theoretical language systems. Habermas argues that these on the one hand fix the description of the phenomena to be explained, with the help of a coherent network of concepts, on the other hand they determine which kinds of empirical evidence are permitted in an argumentation. The observation data are theoretically (re-)interpreted experiences and as such dependent on the categorial framework of the chosen theoretical language system. Habermas uses the term adequacy in two ways: a: the plausibility of a law rests on the adequacy of the theoretical language system T in which the law is formulated. b: theoretical progress can be defined as re-interpretation of experience with the help of increasingly more adequate theoretical language systems. I am troubled by both of these points. The first one rests on a misconception analogous to the one criticized in the previous point. In the

52

same way as the plausibility (or the truth, too) of W is insufficient for justly assigning truth to C, so the adequacy of T is insufficient for the plausibility of W. Boyles law for instance could have been formulated pV = RT2 instead of pV = RT in terms of exactly the same adequate language system. The problem is that the first formula is not and the second is plausible. Habermas has forgotten the specificity which laws still have given a fixed theoretical language system.36 This specificity depends on the nature of the background support B, that is on the fact that certain singular propositions (supporting the law pV = RT) are true or acceptable, and others (for instance leading to pV = RT2) are not.37 Again the problem of the truth of singular propositions raises its head. The plausibility of laws also rests on the truth or acceptability of the singular propositions which together constitute the background support B and cannot be reduced to the adequacy of theoretical language systems. The trouble with the second point is the question of how such a relative evaluation of rival theoretical language systems is possible within Habermas theory given his view that these determine both the description of the phenomena as well as the selection of permissible data. In other words, how can Habermas reconcile theoretical progress with theses tending towards incommensurability?38 An appeal to the communicative condition is not sufficient to master this problem. The social conditions of the ideal speech situation are at best necessary (for the possibility of proposing alternative theoretical interpretations), but certainly not sufficient for a concrete relative evaluation of the degree of adequacy of these alternatives. In the latter cognitive factors will after all also play a part. The introduction of a distinction, as proposed by Kunneman39, between methodology and theory of truth (as part of a theory of rationality) does not do away with the fact that Habermas himself includes the adequacy of T in his criterion of truth. If this criterion is to be applicable at all (e.g. to propositions like this is gold), then it must at least in some way indicate what it might mean for one theoretical language system to be more adequate than another. Until such is the case it is impossible to obtain clarity about the worth of Habermas theory in this question. I will also come back to this problem of theory change in part II in more detail. The main issue in this is the question, and the possible specification, of continuities and discontinuities which might play a role in theoretical progress in the experimental natural sciences.

53

4. Next I would like to discuss in more detail the role of the communicative condition in Habermas criterion of truth. To put it quite bluntly, this condition seems to me to be irrelevant and, moreover, misleading (at least with respect to the problem of truth; I will not consider here the role of the ideal speech situation in a practical discourse on the rightness of norms). The terminology offered by Habermas for purposes of clarification, such as constitutive illusion, counterfactual effectiveness, operational fiction etc. clarifies precious little when we are concerned with answering concrete questions which should be put to a theory of truth. For example the question whether the criterion of truth is adequate with respect to simple propositions such as this liquid is boiling. It seems to me that this question should be answered for Habermas theory in the negative. My main objection is the following. Take any proposition C from the natural sciences about which it is asserted on the ground of a valid discursive argumentation that it is just to regard C as true. Given this C, this argumentation, under these social circumstances, it is not difficult to show (or make it highly probable) that C is unjustly called true. I would do this with the aid of a small-scale sociological or psychological examination of the question whether we are really dealing with an ideal speech situation. In fact, up to the present the conditions for a real discourse almost certainly have never been fulfilled and not even been approached. For these conditions not only demand severe requirements of the community of scientists but also of the organization of society as a whole in which this scientific community is embedded. The idea that by means of suitable institutional measures the ideal speech situation could be approached within the scientific realm is already problematic, considering the existence of power relations, dogmatism and prejudices. But the requirements Habermas asks of the ideal speech situation are in fact even more severe. They also concern the communicative action of all potential discourse-participants and thus the entire social context in which these potential scientists live. Not only within the scientific institutions, but also within the entire social frame within which these institutions function, power relations should have ceased and the inequality of chances should have disappeared. (see 1.5.4, point 3). Moreover, scientists should not only within the theoretical discourse, but also in many other social situations, be true to themselves and make their inner life transparent for others and for themselves (1.5.4, point 4). We should apparently free all of society from power-pressures before we can even begin to approach a scientific discourse.40

54

In the light of these impossibly stringent demands, which are necessary conditions for the being grounded of a consensus on simple propositions such as this liquid is boiling, the supposition that we are not in a position to mention one proposition to which we can justly attribute truth seems to me most plausible. Habermas criterion of truth is much too fallibilistic. Not only can we never be sure of correctly calling a specific proposition C true, but on the basis of his criterion, it certainly seems to be possible to reject any claim to truth (unless people suddenly changed into angels, to put it in his own words41). Moreover, it is important to note that Habermas criterion of truth is not concerned anymore with Cs being true in itself but with justly assigning truth to C, thus with a weaker, a rather epistemological question. One should at least require of a criterion of truth that it be really applicable in such an epistemological context. In any case I find it too much to accept that it is almost certainly never applicable in paradigmatic practical situations. Thus far the criticism of the idea of the ideal speech situation rests on the actual impossibility of even approaching it at present. Habermas, however, sometimes seems to mean by this ideal situation a form of life to be realized in the future.42 However this may be he is not always very clear about this in both cases his position amounts to a form of rationalism: if humanity and society were free of constraints and rational, then our knowledge would be the true knowledge. I share neither Habermas (possible) belief in progress nor his rationalism. In the light of the course of history up till now, particularly when not seen from a Western point of view, the former does not seem a plausible supposition to me. And about the latter: when we are considering knowledge of the complex reality of nature, people and societies, I do not think that a rational society in which unconstrained discourses have been institutionalized will suffice to guarantee the discovery and acceptance of theories better and better adjusted to their object-domain. Of course we can continue to propose again and again alternative theoretical language systems, but the problem is that these must prove to be adequate, and that in our argumentations data D and background support B must prove to be true. Even in an ideal speech situation we cannot make bricks without straw. 5. A last point of criticism concerns the relation between true propositions and objective experience. Habermas admits that up to the present

55

the problem of how empirical limitations operate in the process of justifying truth-claims connected with descriptive statements is still unsolved in his theory of truth.43 I think that what he has said about this up to now can explain why this is so: in the direction in which he is looking for an answer no solution to this problem can be expected. As we have seen above the fundamental cognitive schemata, determining the objectivity of experience, stand only in an indirect relation to the truth of propositions, for the formulation of which theoretical, non-fundamental, cognitive schemata are used. This raises the question in which indirect way do the theoretical schemata remain related to the independently constituted object-domain of things and events.44 With regard to this question Habermas argues:
also the less fundamental and changing schemata which appear as basic concepts in theories play a constitutive part in the structuring of object domains .45

This quotation seems to imply a certain conceptual continuity, or at least a compatibility, between the theoretical and the fundamental cognitive schemata: for if not, how could they both, simultaneously, play a constructive part in the constitution of the object-domain of things and events? Hesse interprets this in a similar way.
We understand the natural-language reference of ball under most circumstances without having to ask whether a speaker is a Newtonian or an Einsteinian or indeed a flat-earther. It is because scientific theory is continuous with these natural-language meanings, that theoretical explanation can be translated into technical knowledge.46

But I find such a continuity or compatibility between fundamental and theoretical cognitive schemata highly doubtful. How can for instance a non-Euclidean and a Euclidean concept of space simultaneously play a constitutive part with respect to the categorial structure of the objectdomain of things and events? Or, how can the daily constitution of objects as moving particles be compatible with the basic quantum mechanical concepts which explicitly exclude such a notion of a moving particle (a path)? The requirement of conceptual continuity or compatibility between the fundamental and theoretical cognitive schemata seems to introduce again some sort of transcendentalism in Habermas theory of truth. After all, this requirement entails that the way in which the constitution of sensorial experience takes place imposes specific limitations on the pos-

56

sibility of variation in the basic concepts of our theoretical (re-) interpretations of this experience.47 From a realist point of view Habermas falls prey to the epistemic fallacy: aspects of the contents of the fundamental cognitive schemata which play a part in acquiring empirical knowledge impose because of the requirement of compatibility concrete restrictions on the basic predicates being used in the description of the internal-theoretical ontologies. In part II I will come back to this important and complicated problem of the relation between theory and experience and will then try to solve the problems I have noted with respect to Habermas views here. I will then distinguish clearly though without disregarding what has been said at the end of section 2.3.1 the reality-relatedness of theoretical terms (or basic concepts) from characteristics of the contents of our daily experience, which plays a part in the process of acquiring scientfic knowledge. In Habermas words: the fundamental and the theoretical cognitive schemata will not be linked anymore through their contents. 6. Thus far the main objections against the criterion of truth as proposed by Habermas.48 For him the criterion for the justification of a claim to truth is grounded consensus. The consensus is grounded if the argumentation supporting the claim to truth is valid and based on a theoretical language system which fits the object-domain concerned and if also the consensus has been reached under the conditions of the ideal speech situation. Habermas assumes that such a consensus on the truth of a proposition can always and everywhere be reached again and again, whenever the conditions for its being grounded are fulfilled. In this sense he believes truth to be invariant. I will summarize two points of my critique which are most important in that they touch the core of Habermas theory and cannot be solved by small modifications. First the assigning of truth to the singular propositions D and B remains unfounded, so that a grounded consensus in the sense of Habermas criterion still has a factual aspect generated by the data and background support actually used. In the second place: the requirement of the ideal speech situation implies that almost certainly all truth claims are invalid and will remain so, unless people suddenly change into angels. The class of known true propositions is and stays empty. If, moreover, this ideal speech situation is insufficient for obtaining more and more adequate theoretical language systems, in other words for an approximation of truth, then truth cannot play a regulative role in Habermas theory either. I cannot

57

but conclude that a criterion such as offered by Habermas, one which cannot be applied to paradigmatic cases of true propositions and which is neither worked out in enough detail to substantialize the notion of convergence in the direction of truth as a regulative idea, is inadequate. Nevertheless, Habermas theory has some interesting and stimulating points which I will gladly use in the next part of this book.

58

PART II

Experimentation in the natural sciences

3.1

Introductory remarks

Commendable in Habermas is his explicit acknowledgment of the philosophical relevance of the process of experimenting in natural science. This experimental process, if it is discussed at all, is generally treated in a stepmotherly fashion in the philosophy of science. Following Habermas, experimentation will also play a major role in the philosophical discussion of the next chapter. But before I can start with this we need a more precise analysis (than Habermas provides us with) of experiment and experimentation in the natural sciences. I will give this analysis in the present chapter. In doing so I distinguish three aspects: the theoretical-scientific description of experiments (3.2), experimentation as action and materialization (3.3); and the epistemic aspect of experimentation (3.4).

3.2

The theoretical description of experiments

In an experiment we deal with an object to be studied (in a general sense, e.g. a process, a thing, a state of affairs) and with experimental equipment, likewise divisible into various categories. Thus Janich for example distinguishes between (1) observation instruments such as eyes and ears plus corresponding amplifiers such as glasses, electron microscopes, sound amplifiers (the senses of touch, smell and taste are unusual but not on principle excluded as observation instruments); (2) measuring equipment, i.e. all equipment directly aimed at producing a quantitative result on a metric scale, such as voltmeters and thermometers; and (3) other experimental devices, i.e. all necessary

59

items not included in (1) or (2), such as electromagnets, ion sources, basic solutions and photographic plates.1 Now, experimentation involves the realization and the theoretical description of a number of manipulations of the object and the equipment, brought into mutual interaction. I understand an experimental system to be this complex of object and equipment within a specified spatial area and during a fixed interval of time. It goes without saying that this system does not float in a vacuum. It is always embedded in the rest of reality. In successful experiments in the natural sciences we aim at reproducibility. I concentrate here on successful experiments because of the philosophical use I wish to make of the discussions at issue in the next chapter. A central aim of this section is to make a further analysis of reproducible experimental systems. In this I am not so much concerned with an abstract definition as with the question of how we are able to produce and maintain these systems in the experimental natural sciences. As we will see reproducibility appears to be on the one hand dependent on the chosen theoretical description, and on the other hand on the actual possibilities present in the material realization of experiments. In the next section experimentation as action and materialization will be discussed. Here I give a closer analysis of some aspects of the theoretical description of successful experiments. The theoretical description of an actually performed experiment contains first of all a specification of the experimental episodes or situations that occurred. In general these theoretical descriptions will refer to very diverse kinds of episodes: to the existence of objects or of their properties, to the (non-)occurrence of processes and events, etc. Moreover, such descriptions need not exclusively refer to episodes occurring within the spatio-temporal location of the experimental system S. Knowledge of the external setting and of its possible interaction with S is also of importance for a correct theoretical description of experiments. And finally an experimental episode p is in general built up out of various subsituations: p = p1 p2 pn.2 Secondly certain relations always exist between the experimental episodes theoretically described. I shall analyze these relations here in terms of sufficient respectively necessary conditions, noted as follows: p q (whenever description p is true, then q is also true), respectively p q (whenever description q is true, then p is also true). These relations apply in the first place to descriptions or propositions, they do not always refer to causal relations between the experimental

60

episodes themselves. As will be shown the validity of these relations follows from the (claimed) truth of the theories with which we describe or interpret the experiments concerned. In order to make all this more concrete I give the following example. Suppose we wish to determine by experiment the boiling point of a particular liquid x and in order to do so we observe an experimental system S during an interval of time [t, t]. S contains a heat source, a vessel with the liquid x, a thermometer and various kinds of auxiliary equipment needed to carry out the experiment accurately and correctly. This experiment to determine the boiling point of x is theoretically described as follows: p q. Here p = p1 p2 p3 a, and: p1 is the proposition: the temperature of this liquid x at time t is 123C. p2 is the proposition: the liquid x is supplied with a net quantity of heat during every second of the interval [t,t], which is at about the same magnitude as the heat capacity of x. p3 is the proposition: the temperature of this liquid x is during the interval (t,t] 123C. a is the description of all relevant experimental circumstances, comprising at least these statements: a1: the thermometer used has a well-defined scale, based on a fixed melting and boiling point of a standard liquid; the effect of the thermometer on the temperature of the liquid is negligible. a2: the liquid is pure. a3: the experimental system S is during the measurement thermically isolated from the external setting. a4: the air pressure around S is constant during the measuring process. a5: all other known factors which could possibly interfere with the temperature of the liquid or that of the thermometer during measurement have been eliminated; and by this we presume that all disturbing factors are ruled out. q is the experimental result: this liquid boils at 123C in the interval [t,t]. This example shows that the experimental result is in fact a conclusion of an argumentation whereby, from description p, with the help of the necessary physical theories, we infer result q. But why does a theoretical description of a successful experiment in the natural sciences in fact require such an inference or argumentation in terms of necessary

61

and/or sufficient conditions? Why do we not simply perceive the situations p and q? To find the answer to this question we should first of all realize that a successful experiment in the natural sciences is always a reproducible experiment, that reproducibility is required for the success of experiments.3 Non-reproducible effects are called artefacts or noise, instead of facts or true peaks. (That reproducibility is required seems hard to refute. In concrete cases, however, to decide whether an experiment has in fact been reproduced or is reproducible may pose a delicate practical problem). If we require every single experiment to be reproducible, we will have to define firstly what must be reproducible. For this we will always need an interpretation or description of the experiment, which is supplied by the scientific theories concerned. In my eyes the requirement of reproducibility is accounted for in the theoretical description of (successful) experiments by establishing conditional relations between experimental episodes such as p and q with the help of available theoretical concepts and laws. These relations will always allow q to be inferred from p (resp. p from q), whenever the conditions specified in p (resp. q) have been fulfilled. This proves that every individual experiment always has a universal aspect, too, through its claimed reproducibility.4 This does not, however, mean universal without qualification. Even if the experiment described above is reproducible, then this does not imply as yet that liquid x will boil always and everywhere at 123C. This will after all only be the case when the experimental episodes described theoretically by p obtain. Thus we are dealing with universality within a theoretically defined domain of phenomena. In the example above the theory determines what we understand by boiling phenomena and under which circumstances we are dealing with these phenomena. This all will become still clearer in the elaboration of this example in the context of a discussion further on of the closedness of experimental systems. The description of the experimental circumstances summarized in p is mainly taken from theories and auxiliary theories on boiling phenomena. It testifies to a number of stages in the theoretical development of this field.5 It goes without saying that these theories and auxiliary theories do not need to be in apple pie order before an experiment can be sensibly started. The example analyzed here has made this quite explicit. Experiments on the basis of solely p1, p2, p3 and a1 can certainly be carried out and moreover can produce at a certain moment reproducible experimental results (by chance we might say afterwards), so that the experiment at issue can in this sense (and at that moment) be called successful. In other words: in scientific practice 62

the (auxiliary) theories needed for the inference may vary between the vague and the specific and in general they may also be very different in nature. The above experiment appears paradigmatic for the natural sciences. In terms of Janichs distinctions, mentioned at the beginning of this section, there are observation instruments, e.g. for temperaturereadings; there is measuring equipment, such as the thermometer; and there are other experimental necessaries such as the liquid itself and the heat source. Basing myself on this paradigmatic character I hold it plausible to generalize the above analysis thus: every successful assigning of an experimental result q no matter how direct this may seem to be will always require an inference or argumentation based on a number of (auxiliary) theories about the experimental phenomena concerned. In the remainder of this section I will focus on an important, necessary condition for experimental reproducibility, that is the closedness of the experimental system. In doing this I will in the first instance make use of Von Wrights ideas on this subject. Therefore I will discuss these first, using his own terminology.6 I use his work, which primarily aims to give an explanation of the notion of causality, only for an analysis of experimenting in the natural sciences.7 Von Wright does not analyze experiments in terms of descriptions of episodes of (closed) systems, but in terms of temporarily ordered states of (closed) systems, using for the sake of clarity a model with a discrete temporal ordering. In his analysis the necessary and sufficient relations occur between succeeding (initial, intermediate and final) states of experimental systems. On closed systems in general Von Wright says this:
There are several senses in which a system, when it is instantiated, can be said to be closed to causal influences from outside the system. One sense is that no state (or feature of a state) at any stage in the system has an antecedent sufficient condition occurring outside the system.8

He adds that this meaning of closedness is very common and that he will therefore restrict himself to it in his further analyses. The definition of closedness as proposed in the quotation above is worked out in more detail as follows. Suppose we have a system S localized in space and time with initial and final states a and b. We now examine the role of state a0,

63

which immediately precedes a and is therefore outside S. If system S is to be closed in the above sense, then firstly a0 must not be sufficient for a, and, secondly, not sufficient for all next states of S up to and including final state b. Thus for closedness a first condition is that the system will not by itself move from state a0 to state a. Furthermore a0 must not influence S through one of the intermediate states or the final state, i.e. a0 must not be sufficient for one or more of these states. How do we get closed systems in scientific practice? According to Von Wright these have to be produced nearly always artificially through experimental interference with nature. If only passive observation of the course of events were possible, then we would never know which factors to incorporate in the description, because we could never know whether the experimental system concerned was truly closed:
It is only the characteristic operation of active interference, of changing a state, which would otherwise not thus change, into the initial state of a system, that can give us this assurance.9

Von Wright answers the question of the possibility of the actual producing of closed systems by drawing into the argument the necessity of active human interference through experimental action. As we saw above a closed system should not by itself move from state a0 to state a. How do we then produce such a system with initial state a? Von Wrights answer is: only by an active interference that changes a0 into a. Secondly, for a closed system a0 should never suffice to reach one of the intermediate states and final state b. Again the question arises: how do we arrive at such a system? Von Wright: we should (actively) renounce changing a0 into a and leave the system be. If the states occurring then (i.e. a series of states certainly not having a as an initial state) differ from the intermediate states and the final state at those moments that we should have expected these states, then a0 is certainly not sufficient for the occurrence of final state b, nor for any other of the preceding states of the system. Thus: with the help of these two procedures we can also produce in practice experimental systems, which reasonably fulfill the above definition of closedness.10 In this way Von Wright stresses the essential importance of human action for the experimental sciences. In the case of a relation between states in the form of a sufficient condition he describes the action aspect as follows: by doing a we can bring about b. In the case of a necessary

64

condition: by suppressing a we can remove or prevent b. Thus successful experimenting by means of closed systems rests on the capacity of people to do or suppress certain things by a direct intervention in nature. Therefore Von Wrights discussion could be used to make Habermas concept of instrumental or experimental action more precise. On the other hand Von Wright does stress explicitly that it is not the experimenting scientist but the state a which is the cause of the experimental effect b. By doing a we achieve the occurrence of b; in order to bring the system into state a we will in general have to interfere artificially. But this does not mean that we (the actors) are thus the cause of b. If the system is in state a, a itself causes the occurrence of effect b.11 The image of experimenting Von Wright sketches is thus as follows: first we produce an initial state a by artificially interfering with the course of events and subsequently we wait (passively, as he says12) to see whether certain intermediate states and a certain final state b do or do not occur. It remains to be seen, however, whether this view of experimenting is tenable. I do not think it is and I will argue against it with the help of an argument taken from Bhaskar. Von Wright fails to see (as I will illustrate below with an example) that human interference is necessary not only for the production of initial state a, but also for that of final state b. It is certainly not so that once we have produced a we can sit back and relax until b does or does not occur. It is of vital importance to keep the experimental system closed during the entire interval of time the measurement occupies. In general active interference is necessary to achieve this. For b it holds therefore equally that without our causal activity, given a, b may not, and in general, will not occur.13 This entails that Von Wrights view on the closedness of experimental systems, which is that state a even if this is artificially produced will subsequently always and automatically lead to state b, is not adequate. The question now arising is what is an adequate analysis of closed experimental systems. Bhaskars definition of a closed system as one in which a constant conjunction of events occurs is itself riddled with a number of unclarities and disadvantages.14 This is why I prefer to include Bhaskars criticism in a revised analysis of closedness la Von Wright. For this purpose I return to the terminology of descriptions of experimental episodes or situations introduced earlier on. In general the theoretical description of experiments does refer both to situations within and outside the spatio-temporal location of the experimental system S. With regard to the closedness of S itself, however, we should

65

draw a clear distinction between situations within and outside S. The aim is after all to prevent situations outside S from disturbing situations within S. Our main concern now is to determine more accurately when such disturbances on the experimental system do or do not occur. But before I discuss this I would like to observe the following. The concept closedness is often used in the sense of a complete isolation of the system concerned with regard to all conceivable (physical, but also other) influences. Since this definition is only met in truth by the entire universe, it is usually weakened to approximate isolation. Whatever the merits of such a weakened concept of closedness may be, it is still not useful for the analysis of successful experimenting in the natural sciences which I intend to give here. We have to relativize this concept even more: a) As explained above a reproducible experiment in the natural sciences is always interpreted in a certain theoretical mode: p and q are theoretical descriptions of experimental situations. These descriptions are always theory-specific, and likewise when we are dealing with the closedness of experiments in the natural sciences system S not being influenced is specific, too. The theories used determine in some general way which factors outside S could possibly influence the experimental situations p and q described like this. The closedness of experimental systems will therefore always be relative with respect to the chosen theoretical description. b) We should also take into account the fact that every experiment tries to answer a specific question. To obtain a useful answer the system does not need to be closed (in the relativized sense as defined in a) with regard to all experimental situations occurring in that particular interval of time. Only those external factors have to be excluded or controlled which make the obtaining of a useful answer to the specific question phrased for the experiment involved impossible. In other words: we can limit ourselves to relevant experimental situations. From now on I will take an experiment to be a series of situations actually occurring in a particular spatio-temporal location. I will argue that experimental closedness is a necessary condition for the success of experiments; the question is what this experimental closedness exactly consists of. In order to arrive at an explanation of the concept closed experimental system I will first take another look at the example of the experiment to determine the boiling point of the liquid x: p q, in which p and q are the theoretical descriptions, as given above, of experimental situations which have actually occurred (for the sake of brevity I will often call p, etc. a situation of S).

66

In the first place I will again take Von Wrights definition of closedness as a starting point. In my terminology this boils down to none of the experimental episodes of S having sufficient conditions outside S. Suppose in the example z1 is this situation: there is a net heat current from liquid and thermometer to the external surroundings of S, a current which is of exactly the same size as the heat supplied to the subsystem of liquid and thermometer within S. Now the following holds good: taken that the descriptions p1, p2, a1, a2, a4 and a5 are right, then z1 p3 (the temperature remains constant). Therefore the situation z1 outside S is sufficient for an experimental episode of S, and so the system is not closed. If we would in this case not include closedness as a condition for successful experimenting, then we would conclude from the experiment that the liquid x boils at 123C. The experiment seems successful, but in fact it is not because in general it will not be reproducible: its success depends after all on the fortuitous and delicate equilibrium between supply and release of heat and on the equally accidental circumstance that the initial temperature is already at boiling point at time t. Generally it will obtain that p3 is not sufficient for q. It is true that p q, but because condition a3 has not been fulfilled, description p is not justified in this case. In the second place I will allow for Bhaskars criticism of Von Wright. This means in my terms that in general for any experimental system S there will be situations outside this system, which are necessary conditions for the occurring of experimental episodes within S; that these conditions are usually produced artificially and have to be maintained; and that this causal activity influences the experimental result q.15 Let us take as an example condition a4, which requires the air pressure around S to remain stable during the measurement. Now it so obtains that: taken the descriptions p1, p2, a2, a3 and a5 to be correct, then a4 p3. If a4 is not fulfilled, then (if all other descriptions are correct) the temperature will not remain constant. The experimental result can then not be described as q = this liquid x boils at 123C. We will not succeed in determining a fixed boiling point for x, which we will blame in this case on the fact that the system is not closed. It is important to note that these necessary conditions are indeed required in an analysis of closedness. Sufficient conditions are not enough. It is of course true that nota4 notp3 (if the other

67

descriptions are correct). But the concept of closedness refers to the experimental episodes of a system which factually occurred, also according to Von Wright: It is important to note that closedness, as here defined, is a property of a system on a given instantiation of it.16 This means: given a system S with a series of factually occurred situations, Von Wright will then call S closed if these situations do not have sufficient conditions outside S. That this is insufficient for the closedness of experimental systems is proved straightaway by the fact that in order to experiment successfully it should also in general be required that there are no sufficient conditions outside S for (some) situations which in fact do not occur (as for instance notp3). In other words: if we restrict ourselves to conditional relations with respect to the situations factually occurred, we will have to include also necessary conditions in the definition of closedness.17 This example shows plainly that, if not closed, an experiment can fail for widely varying reasons. In the one case we conclude a fixed boiling point does exist, but when repeating it the result proves to be irreproducible. In the other case we do not even succeed in finding a fixed boiling point, in just this one experiment. This entails that the experiment is not suited to answer the question of the boiling temperature of liquid x. Again it becomes clear how important theory is when establishing the closedness of experimental systems. The usual theory states that every liquid has a determined boiling point. But if we had a different theory, e.g. one in which the boiling point is a probabilistic quantity, then we could in the first case after repetition of the experiment still regard the experimental system involved as closed (certainly if we know nothing about situation z1). And if the theory did not give a boiling point but a boiling interval then we could also regard the second case as a closed system certainly as long as we do not know anything about the non-occurrence of a4. Moreover, it will be obvious that in this experiment not all experimental episodes are relevant. Thus for instance are the details of the way in which heat is supplied to the liquid within S irrelevant for the success of the experiment. In concluding I would like to generalize these observations to arrive at a general criterion for the closedness of experimental systems.18 As explained above, we are concerned with a relativized concept of closedness, i.e. given a theory to describe the experimental episodes and given a specific experimental or theoretical problem.

68

The experimental system S is closed during the interval [t,t] if and only if these two conditions have been fulfilled: (1) the relevant experimental situations within S which have factually occurred do not have sufficient conditions outside S. (2) the conditions outside S have been fulfilled which are necessary for the occurrence of those experimental situations within S which make possible in a reproducible way an answer to the theoretical questions posed. If one of both, or both, of these conditions have not been fulfilled, then the experiment will break down. Either for the reason that the result is not reproducible, because it has been effected by a fortuitous concurrence of circumstances, by a whimsical and uncontrollable interplay between situations inside and outside S. Or because the resulting experimental episodes are such that the theoretical questions asked are not answerable anyway. In that case we conclude that given this theoretical phrasing of the question the experimental system is not closed. Finally, it will be obvious that the assigning of closedness to experimental systems is always of a provisional nature: we can never be completely sure whether we have eliminated all sufficient conditions and produced all necessary conditions.

3.3

Experimentation as material realization

In the last section we were primarily concerned with giving a philosophical reconstruction of a number of important aspects of theoretical descriptions of experiments in the natural sciences. It goes without saying that this is only one aspect of experimenting, there also has to be something done. Duhem describes both aspects, the action and the theoretical description in the following way:
When a physicist does an experiment, two very distinct representations of the instrument on which he is working fill his mind: one is the image of the concrete instrument that he manipulates in reality; the other is a schematic model of the same instrument, constructed with the aid of symbols supplied by theories.19

In this section I will focus on the actual performing of experiments, as distinct from their theoretical description.20 The main problem can be illustrated with this example. Suppose we want to determine experimentally the mass of an object a which is at rest in relation to the measurement equipment. Two scientists each

69

carry out such an experiment in the same way, i.e. with the help of the same actions. Nevertheless one gives as the experimental result mN(a) = 10g and the other mE(a) = 10g. In other words: one interprets the actions performed as a measurement of the Newtonian mass, the other as a determining of the Einsteinian mass of a.21 Still both did the same actions and thus in a certain sense the same experiment. Therefore if we want to describe experimental action unambiguously, we have to find some sort of abstraction of these various specific theoretical interpretations. On the other hand it is an indisputable fact that concrete experimental action is always action on the basis of certain theoretical ideas: without theoretical ideas no experiments. In order to give a more detailed analysis of the difference between the theoretical description and the actual performing of experiments, we cannot simply describe what happens in the concrete experimental practice. In this section I will develop a terminology with which to tackle later on, in the next chapter, this problem of the identity of experiments under various theoretical descriptions. I propose the following manoeuvre to make the distinction between the theoretical description and the actual performance of experiments more explicit and manageable. Let us suppose that in the example of the experiment to determine the boiling point of a liquid the experimenter A orders B, a complete layperson in the field of heat theory, to actually carry out all necessary actions. A tells B for example: take this thing (here) and put it in the holder (over there), in that way that its lower part hangs in the liquid; then watch the rising silver column; when it does not rise anymore for a bit, then write down (here) which point (which number) is reached by the upper end of the column; and B does all this correctly according to A. The general supposition is that experimenter A can transfer the actual carrying out of experiments to one or more laypersons B. We are referring to actions like: the installation of various experimental apparatus, the preparation of the necessary substances and material, the reading of pointer positions on graduated scales, of colour changes, of traces on photographic plates etc. This all happens under As supervisions and with perhaps As correcting instructions to guarantee that the experimental result is obtained in a right way according to the experimenter. With the help of this manoeuvre I can now define the concept of the material realization of an experiment like this: the whole of the experimental actions carried out by B in a correct way according to A and which are described in As instructions to B in the language through which A and B communicate with each other.

70

Two central ideas behind this definition are: (1) not only experimenters but all human beings are themselves part of nature (are themselves also nature); therefore they are able to produce certain situations in experimental systems because of the material interaction with nature surrounding them (thus the name material realization). (2) there exists some sort of language, everyday language, in which laypersons and experimenters can communicate with each other about certain fundamental aspects of this production process in a usually reliable and successful way. In the remainder of this section I shall try to explain the meaning of this concept material realization and what I intend to do with it. 1. First of all we must ask ourselves whether it is actually possible to carry out experiments in the way described in the definition of material realization. In the first instance this definition refers to a thought experiment (with experimenters and laypersons) with the intention to clarify certain ideal-typical aspects of experimental action in a philosophical way. On the other hand we will certainly have to be able to formulate, at least more or less definite, descriptions of the material realization of concrete experiments. If not, then we are not dealing with an ideal-type but with a philosophical fiction. The best way to make this thought experiment plausible is in fact to carry it out. This has been done already in quite a different context, independent of the discussions held here in an ethnomethodological study by Schrecker.22 The experiment referred to is this:
In Schreckers study, a methodological set-up was used for the purpose of perspicuously identifying the mutual dependence of chemical reasoning and embodied action. Schrecker volunteered his services to aid a handicapped student in his laboratory work for an undergraduate chemistry course. He then received the students permission to use his experimental work as a research topic. There resulted a division of labour and responsibility between Schrecker, who was largely ignorant of the field of chemistry and ignorant as well of the specific lab assignments he bodily assisted and Gordon, the chemistry student, who because of a spinal injury was paralyzed from the neck down with very limited use of his hands. Gordon depended upon Schrecker to bodily perform the work at the bench for the weekly lab experiments assigned to students in the course on Quantitive Analysis. The isolation of Schreckers handiwork from its theoretical basis in chemistry necessitated that Gordon and he make explicit for one another how the experiment was progressing as a witnessable production of chemistry. Gordon provided instructions for Schrecker on what to do next, while at the same time he relied upon Schreckers developing work to show him what next meant in terms of where they stood in the course of the experiments events.23

71

This study is most suitable for further clarification of a number of important aspects of the notion of material realization introduced above. a) In the study there is also clearly question of a manoeuvre: this is not the usual and neither the most efficient way of carrying out chemical experiments. In fact how big would the chance be for a partially paralyzed chemistry student and a sociologist interested in ethnomethodology to meet in this way? b) In such a situation the instructions and corrective indications given by the (theoretically informed) student Gordon are indispensable. A simple action such as for instance the cleaning of the glass turns out to be quite problematic: sufficiently clean in one phase in a chemical experiment is not necessarily so in another phase. Only on the basis of theoretical knowledge about the reaction sensitivity of the chemicals to be used in various phases can we know when the glass is sufficiently clean.24 Thus even if the description of the material realization of a successful experiment itself is put in ordinary language terms, theoretical scientific insight will certainly be necessary for arriving at such a description. c) Schreckers research deals with so-called standard experiments, which every student must learn to carry out. In this, too, it ties in with my discussions which after all aim at giving a philosophical reconstruction and evaluation of the manifold standard experiments in the natural sciences which are not considered controversial and which are unproblematically performed, such as e.g. the determining of the boiling point of a liquid. And in one way the thought experiment in the definition of material realization is even somewhat easier to carry out than Schreckers. In the former case we are after all dealing with an experienced experimenter A, so that possible complications arising from Gordons learning process will not occur. d) As the quotation above shows Schrecker was indeed a layman, both in the field of chemical theory and with respect to the laboratory equipment. Notwithstanding the fact that our daily language and experience is permeated increasingly with scientific knowledge and concepts, there still exist, as I have called them complete laypersons. This is because usually this knowledge is too general and these concepts too ill understood to serve as a basis for a successful communication with experimenter A in a concrete experiment. Even though one may have heard of an electron microscope, it may still be quite impossible even to point it

72

out, surrounded as it would be by the baffling abundance of apparatus especially in modern laboratories. Then the instructions of the experimenter cannot be: switch on the electron microscope (and certainly not: determine the spatial structure of DNA with the electron microscope). They will have to be phrased in clearly non-theoretical language as: press that black button on the left upper side of that apparatus over there in the corner. e) A possible problem might be lurking in the fact that the carrying out of the necessary actions in the material realization of experiments requires certain practical expertise including certainly a number of tacit skills mastered by the experimenter but not yet by the layperson.25 This can be solved in two ways. On the one hand we can select from the complete laypersons in the field concerned a sub-group who are by nature more handy than the rest. On the other hand there can be nothing against laypersons B acquiring certain necessary practical skills in the laboratory before starting out: time and money do not play a role in this philosophical thought experiment.26 Schreckers case shows that this is possible without the layperson learning the theory involved: he was recognizably doing chemistry, whether he knew it or not.27 f) For philosophical reasons I analyze the carrying out of experiments here as a material realization, which can be described in the daily terminology indicated above. This implies a dual abstraction, namely not only of the theoretical interpretation of the experimenters, but also of the social aspects of the practice of the experimental natural sciences. In fact the material realization will always be embedded in a social context and is thus at the same time a social realization. This social realization of experiments may be analyzed further from various perspectives. In my conclusion I will briefly touch on this. Now I would like to restrict myself to one aspect of this social realization which seems most relevant in the present context, that is the division of labour when carrying out experiments in the natural sciences. This division of labour is an increasingly frequent phenomenon certainly in the modern scientific laboratories.28 In order to produce for instance a certain experimental chemical effect intensive cooperation is in general needed between, among others, salesmen of industries (the chemicals and instruments needed must be bought), lab technicians (the chemicals must be prepared), general technicians (the equipment must be placed, maintained and operated), graphic designers (the results must be transformed into diagrams, slides or photos) and scientists (the final

73

result must be interpreted). These professions correspond with certain phases in the production process of experimental effects. In these phases which rest on (an enormous amount of) unproblematic background knowledge, it is really only the final result of the phase concerned which counts. For the experimenters the exact procedure by which the result has been obtained is not of direct importance; for them it is just a black box (but of course not for the technician involved who is responsible for this step). Conversely, it is the technicians job to operate the equipment in such a way that two peaks on a display screen for instance will be separated as much as possible. The theoretical meaning of these peaks is to them irrelevant and in general does not have to be understood at all. The communication between experimenter and technician can in such a case be established with the help of notions on the height, the intensity, the separation etc. of the peaks. Within this communication an experimental effect is simply a signal distinct from the background of the field and the noise of the instruments.29 In the case of the cooperation between Gordon and Schrecker we are dealing with a division of labour as required by the definition of material realization. As said above, however, this is not the most economical way of experimenting. That is why in practical situations of division of labour we are not exclusively dealing with complete laypersons. The communication in the various stages of an experiment will partially be in the language of the material realization but will also make use of scientific background knowledge considered more or less unproblematic (e.g. the lab technician may talk about chemically pure water). 2. Next I would like to say something more about my reasons for the introduction of the notion of material realization. With this notion I want to draw explicit attention to experimental natural science as a material production process: experiments need not only be interpreted theoretically, but are also realized materially. Because of the work of people such as Popper, Hanson, Kuhn, Feyerabend and Lakatos especially the recent philosophy of science has focussed predominantly on the prime importance of the linguistic or theoretical side. Here science appears as a theoretical, interpretative activity. The productive aspect is either totally absent or at best a derivation, an application of theory.30 At the same time the fact that science as a materialization process is always embedded in a social context has disappeared from the range of vision of philosophy of

74

science. Precisely to counteract this trend I introduced laypersons in the definition of material realization. The experimenter gives his instructions and the laypersons carry these out, exactly analogous to the situation in the modern natural sciences based on division of labour and hierarchical power relations where the scientists (usually men) tell the lab technicians (often women) which routine jobs they should carry out next.31 By referring not only to the theoretical, but also to the material aspect in our discussions we will get a more adequate view of the experimental natural sciences. It would not be surprising if this more adequate view, in which both the theoretical and the material are taken into account, could be profitably used for philosophical goals. In the next chapter I will try to prove this. But for the sake of clarity it might be better to already touch briefly on the philosophical relevance of the notion of material realization. As said above, without theory there can be no experiment. It is therefore certainly right that the theory-ladenness of observations and experiments is pointed out by modern philosophy of science. The description of the material realization is not a description of the observations of scientists, it is not what is traditionally understood by observation sentences and the like. But neither do I intend to cut every link between the theoretical interpretation and the description of the material realization. When I say that the description of the material realization is theory-independent, I mean specifically this: in order to draw up such a description theory is certainly needed, that is in the form of the (theoretically informed) instructions and corrections of experimenter A. But the resulting description of the material realization is phrased only in daily terms. The latter makes it possible that another scientist A carrying out the experiment again by following this description may, for instance, arrive at the experimental result mN(a) = 10g, while A on the other hand regarded the experiment as a measurement of mE(a) = 10g. The instructions in the description of the material realization are of course never so complete and exact that they can determine the actions of A. To transform the written instructions correctly into situated [experimental] procedures32 theoretical insight is again needed. The crucial point is, however, that taken the fundamental underdetermination of theories by data see section 4.3 , this theoretical insight of A is not necessarily the same as that of A. We have now found an empirical mechanism which makes it possible to step from one theoretical description of an experiment to

75

another. I will come back to this application of the notion of material realization (and to possible objections to it) in the philosophical discussion in the next chapter. 3. Finally I would like to draw some comparisons with Habermas views. In this context it is important to observe that the material realization by layperson B (also for B her- or himself) does not simply consist of random acts, but that these actions too are interpreted with the help of (parts of the) daily language which is the shared language enabling A and B to communicate. Thus the material realization is one of the ways in which the experimental natural sciences are tacitly embedded in our daily experience and thus in an encompassing social setting. The cooperation between A and B is after all based upon a shared action-frame and communication structure. Moreover and this is quite important even if A carries out the experiment him- or herself (s)he must perform similar actions in which (s)he will implicitly assume the reliability of a number of fundamental semantic categories or cognitive schemata. I agree with Habermas that fundamental categories are implicitly made use of (see 1.4.1). The exact structure and contents of the language of material realization is not however laid down by this, though it seems plausible that notions such as permanent object, time, space, number, place and motion play a part in such a language.33 As can be expected after what I said in section 2.2.1, I do not claim a (quasi-)transcendental status for the material realization. Nor do I assume an exclusive relation between objectivity and a description of the material realization in terms of things and events. What does seem plausible is that the above account of the concept of material realization is adequate and useful at least for the period following the rise of the modern experimental tradition in the Renaissance. Such a view does not exclude, however, that in other days and cultures the daily interaction with nature was primarily interpreted in magical or expressive terms. In the next chapter I will go into the consequences of all this for the philosophical use of the notion of material realization. Then I will, as already indicated in section 2.3.2, point 5, draw a relation between the description of the material realization and the theoretical-scientific propositions which is markedly different from Habermas.

76

3.4

Experimental production and the possibility of realism

In the next chapter I will try to give a certain realist interpretation of propositions from the natural sciences in which extensive use is made of the analysis of experimenting given up to now. In section 3.2 it was explained how in successful experiments not only the initial situation, but also the final situation is produced artificially by human beings. This fact seems to raise an important problem for realist interpretations. If indeed not only the knowledge of experimental situations, but if also these situations themselves are in general dependent on the existence and causal activity of people, how can one then ever substantiate the claim that reality as studied in the natural sciences exists on its own, free from human interference? To overcome this difficulty I will follow in the first instance Bhaskars account. He discusses this question in the context of a critique of the empiricist view of laws as constant conjunctions of observed events:
Now as constant conjunctions must in general be artificially produced, if we identify causal laws with them, we are logically committed to the absurdities that scientists, in their experimental activity, cause and even change the laws of nature!34

The epistemic meaning of experimentation lies according to Bhaskar far more in the fact that with its help we can produce knowledge of real mechanisms which are operative independent of the existence, the actions and the knowledge of human beings. He shows this with a transcendental argument, the main premiss (the major) of which goes like this: if experimental action is to be intelligible, then there has to exist in reality an ontological difference between the things and events in an experimental system which the experimenter by his intervening in nature does produce himself (in the sense that they would not have happened without his intervention) and the causal links or mechanisms between these things and events which he does not produce himself. With the help of this and the second premiss (the minor) that experimental action is in fact intelligible one can conclude that there exists an ontological difference between things and events on the one hand and active mechanisms existing independently of human beings on the other hand. Thus according to Bhaskar an important goal of knowledge in experimenting is the identification of these real mechanisms and the making explicit of their operation.

77

One may, however, wonder whether Bhaskars conclusion is generally right. It seems to me that his argument only holds good with regard to empirical realists who, as for example Von Wright does, argue that causal laws, in the sense of constant conjunctions between events, also exist independent of human intervention.35 If it is so that constant conjunctions, apart from some exceptions, only occur in closed experimental systems and that both the initial and the final situations of an experimentally closed system in general have to be produced artificially, then one cannot both cling to an empiricist notion of causal laws and interpret these same laws realistically. From this phrasing it is at once apparent that Bhaskars argument is not applicable to transcendental pragmatists such as Habermas, who conceive of the universal causality of laws as constituted from the quasitranscendental instrumental action. An experiment seems after all also intelligible (in any case as long as this notion of intelligibility is not made more explicit) if one on the one hand agrees with Bhaskars position that the situations or events involved (have to be or) are produced artificially by the experimenter, but on the other hand argues that the description of a causal link must be conceived of as (e.g.) instrumental. In other words: if one does not wish to assign independent reality to causal relations. Generally speaking Bhaskar shows a difference between events and mechanisms, but not that this difference is of a necessarily ontological character. The latter only obtains within realist interpretations. But Bhaskar has given a second transcendental argument to show the real existence of mechanisms and this seems to me to be more convincing, also with regard to Habermas. This second argument uses this main premiss: if the practical use of science in open systems (which are according to Bhaskars definition systems in which no constant conjunctions of events occur) is to be intelligible, then there has to be an ontological difference between events on the one hand and the causal mechanisms which are active in open systems without generating a regular series of events on the other hand. Taking into account the fact that a regular series of events nearly always has to be produced artificially in the laboratory, then the argument goes the assumption of the real existence of mechanisms is required in order to understand or explain the meaning of the use of science, also outside a laboratory context:

78

For if causal laws are, or depend upon, constant conjunctions, then we must ask: what governs phenomena in open systems, that is in the vast majority of cases? The empiricist is now impaled on an acute dilemma for he must either aver that nothing does, so that nature becomes radically indeterministic; or suppose that, as yet, science has discovered no laws!36

It will be obvious that these two arguments of Bhaskar concern a complicated problem, so that what has been said above is certainly not exhaustive.37 I would like to conclude with some limited and preliminary observations, which are important for my own aims. To start with I will leave aside all claims specifically referring to the status of laws. In other words: Bhaskars arguments are used only in defence of the possibility of realism, as a refutation of idealist conclusions which could follow from an analysis of the process of experimental production. Besides this Bhaskars argumentation advocates (the importance of) a philosophical ontology, in the sense of a categorial classification of reality, in e.g. things, events, mechanisms, tendencies, structures etc.38 This philosophical ontology must be distinguished from, and completed by, scientific ontologies linked to specific theories. For the question whether concrete scientific propositions can be interpreted realistically cannot be answered on the basis of a philosophical ontology. The situation is rather this that if we can interpret concrete propositions realistically, then we can categorize the elements of reality which these propositions refer to with the help of terms from philosophical ontology. For instance: if electrons exist, they exist as things. The epistemological question of whether, and if so how, we can obtain this kind of knowledge (e.g. that electrons exist) has, however, not been answered satisfactorily by Bhaskar. The next chapter will be primarily about this epistemological problem. This will result in a realist interpretation of scientific ontologies which is much less pretentious than Bhaskars.39 But Bhaskars philosophical ontology seems to me to be basically useful as an important complement to the realist interpretation proposed below of propositions from specific scientific theories.40

79

Verifiability and reference, relativism and realism

4.1

Introductory remarks

In part I I criticized Habermas views on truth and especially his criterion of truth, and announced that I would develop an alternative view. I will try to do this in the present chapter, by making use of the concepts of theoretical description and material realization of experiments that have been introduced in the previous chapter and which were inspired by Habermas distinction between instrumental action and discursive argumentation. I assume that a concept such as scientific truth should be made explicit in relation to the practice of the experimental natural sciences. Since, according to me, this practice does not contain any references to an ideal speech situation, the result of this explanation will in the first place turn out to be more pragmatic than Habermas. For the sake of clarity I will therefore in section 4.2 speak of a criterion of verifiable truth for propositions instead of a criterion of truth. These pragmatic aspects of verifiable truth will be discussed rather broadly. The main goal of this chapter is to find out whether such a criterion of verifiable truth has not in the second place also some semantic aspects which make a certain realist interpretation of scientific propositions possible. For that purpose I will first analyze in section 4.3 a number of the most important problems realists have to face. The focus is mostly on those problems arising from the occurrence of conceptual discontinuities in the development of the natural sciences. This analysis will show that in fact only a realism where reference, not representation, correspondence or the like, occurs, seems defendable. In 4.4 I will work this out in the form of a criterion of reference and co-reference for theoretical terms. The final section 4.5 consists of a general discussion of the resulting realist interpretation of the experimental natural sciences. Finally: the examples given in this chapter are only meant as illustrations. For a more detailed case study, which elucidates and supports the philosophical discussions given below, I refer to the next chapter.

80

4.2

Verifiability

In this section I would like to propose a criterion of verifiable truth for propositions from the experimental natural sciences. At the beginning of 2.3 criterion has been defined in terms of sufficient conditions for justly attributing truth to scientific propositions C so as to fit the analysis of Habermas theory of truth. Now I will use criterion in a slightly modified sense, i.e. in the sense of sufficient and/or necessary condition(s) for propositions being verifiably true. When I say C is verifiably true I mean in principle the same as when I say truth is justly attributed to C. Nevertheless, I prefer the first statement to the second because it expresses more clearly that the act of attributing truth is based on the concrete possibility of (experimental) verification.1 In the previous chapter this experimenting was analyzed in more detail. On the one hand the importance of materially realizing experiments was stressed. On the other hand, it so turned out that the theoretical description of experiments will always, implicitly or explicitly, contain an argumentation which is necessitated by the requirement of reproducibility and which makes use of one or more general laws accepted for the time being. As a criterion of verifiable truth for propositions from the experimental natural sciences I would like to propose the following: A singular proposition C is verifiably true, if and only if either C is the theoretical description of a situation of an experiment, of which (1) the material realization is possible in a reproducible manner and (2) the general laws needed for this description are plausible; or if C can be deduced from other verifiably true propositions with the help of logical rules and/or plausible general laws. Trying to make the meaning of this criterion explicit is the aim of the remainder of this section. Firstly, I would like to stress the limitations of the range of this discourse. As has been said above, the discussion of the pragmatic aspects of verifiably true propositions is of a general and provisional nature. The next two points will confirm this. To begin with there is the problem of the plausibility of general laws. Concerning the problem of truth I will, for the sake of simplicity, distinguish only between plausible

81

and non-plausible laws. Full justice to scientific practice would, however, require the introduction of degrees of plausibility. And then there is the question of the nature of the relations of logical deducibility mentioned in the criterion. A philosophical analysis of this is a complicated matter which I will leave for the present. I do want to state, however, that it does not necessarily have to concern formal logic, and certainly not first order predicate logic. Secondly the criterion has been so formulated that it is of no importance whether the experiment concerned has already been carried out at a given moment or not. What does matter is that this experiment can be carried out in the required manner in concrete laboratory practice. (When the experiment has been carried out like this, then we can call C verified true). So the criterion makes explicit in general terms when C is verifiably true and at the same time how we can find this out. Thirdly: to avoid possible confusion I will point out some essential differences between the above criterion of verifiable truth and the verification principle of the early logical positivists. This principle states that it holds true for all propositions that they only have cognitive meaning if they are in principle verifiable. In comparison: the criterion formulated above can only be applied to singular scientific propositions2, so that well-known problems such as that of the verifiability of universal propositions do not occur. Moreover, the criterion requires a concrete answer to the question of the verifiable truth of propositions. We are not concerned with verifiability in principle (with all ensuing problems), but with the actual (possibility of) describing and carrying out of the experimental verification. The fact that the latter will not always succeed just like that is with regard to the verifiable truth of propositions much less of a problem than in the case of their meaningfulness. Finally, the criterion does certainly not imply that verifiably true propositions are restricted to the (direct) results of concretely possible experiments. The formulation allows also singular propositions which are derivable through logic and/or plausible theory to be verifiably true. A fourth point is the practical applicability of the criterion. Suppose we want to verify proposition C which is the example given in section 3.2 (C = this liquid x boils at 123C in the interval [t,t]). As we have seen C is itself the conclusion of an argumentation. What we really want to verify when we give a theoretical description of a reproducible experiment

82

are data D, with the help of which we deduce C (the plausibility of the necessary laws we will discuss later on). It will be clear that verifying the truth of these data presupposes again the being verifiably true of other propositions D. If one of the data D is: the temperature of this liquid x is 123C within (t,t], then its verifiable truth certainly does not follow from direct observations of the thermometer. D again is a conclusion of an implicit or explicit argumentation in which among others proposition D = the temperature in this liquid x is homogeneous plays a part. In other words: to establish the temperature of the entire liquid x it will suffice to measure the temperature at one random point of that liquid. If we now require that in order to verify D, D also needs to be verified then we will lapse into the well-known infinite regression. From a logical point of view such a regression is inevitable, because of the fact that the verification of propositions always demands an inference from other (verifiably true) propositions. Given the logical inevitability of such an infinite regression a theory of verification is set the task to supply plausible reasons for the fact that in practice we can (and have to) put a stop to this regression. It seems quite natural to look for these reasons in the conditions of the verification process. An important requirement which I incorporated in the criterion of verifiable truth was that the material realization of the experiment should be reproducible. As explained in the previous chapter, a theoretical interpretation is indispensable for arriving at a reproducible material realization. Suppose C is verified to be true on the basis of the criterion. Then we can apparently realize the experiment materially in a reproducible way with the help of among other things this theoretical description D (this appears from the invariance of the description of the material realization). This fact forms, I think, a good reason for calling a halt to the regression. Under these circumstances Ds use as a premiss in the derivation of C is justified. And this is so even though Ds being verifiably true has not been checked on the basis of the criterion. In section 2.3.2 (point 2) Habermas criterion of truth has been criticized precisely on this point of infinite regression and it has been suggested that this problem could be solved by formulating a material condition also with respect to the truth of singular propositions. With the help of the concepts introduced in the previous chapter we can now express this more precisely. In practice this regression is stopped with an appeal to the reproducibility of the material realization; in other

83

words: with an appeal to experimentally produced experience. In section 4.4 I shall develop this semantic aspect of verifiably true, singular propositions into a criterion of the referring and co-referring of theoretical terms. A fifth point: the notion of reproducibility is because of all this a crucial element in the proposed criterion of the verifiable truth of propositions. Experiments should be repeatable, if intersubjective verification is to be possible. Two famous illustrations of the fact that reproducibility in the experimental natural sciences is a virtue in itself, are the experiments of Holt in 1973 and of Miller in 19251926. The first one supplied experimental evidence against the quantum theory and pro the so-called local hidden variable theories;3 the second against the special theory of relativity and for the existence of the ether as an absolute reference system.4 This means that these experiments led to propositions which were contradictory to what was predicted as a result of those experiments on the basis of the quantum theory resp. the special theory of relativity. But since both experiments turned out to be irreproducible, they were not regarded as falsifications or anomalies of the theories in question. This reproducibility gives (just as plausibility does by the way) a fallibilistic character to the proposed criterion. Experiments which were reproducible up to now, are for the time being accepted as a means of verification. But since they may prove to be irreproducible tomorrow, we can never be absolutely certain whether propositions are verified true.5 Finally I will discuss with respect to this reproducibility of experiments a possible objection which is based on the occurrence of chance processes. Characteristic of these processes is that in general they will not give the same result in a series of identical experiments, neither in the theoretical description, nor in the description of the material realization. In these cases we are not dealing with a straightforward reproducibility, so that the criterion formulated above has to be slightly adapted. Fortunately this adaptation is quite natural. When we deal with probabilistic processes we certainly want the properties of the statistical distribution of the results of a great number of identical experiments to be reproducible. In other words: we require the material realization of an ensemble of experiments determining these properties to be possible in a reproducible way. In fact the problem is purely verbal. When we redefine the ensemble of experiments as one large experiment, the criterion in its

84

original form is also applicable to the experimental verification of probabilistic propositions. In the sixth place the application of the criterion of verifiable truth of propositions presupposes that we dispose of plausible general laws or theories. By plausibility I understand the degree of acceptance of the laws or theories concerned at a given moment, within the scientific community of a given discipline or specialism. It goes without saying that this description is general and preliminary. I do not, however, intend to give a precise definition and an extensive explanation of the notion plausibility here. This is because the main subject of this chapter is the question of scientific realism. A detailed discussion of the problems relating to plausibility and a comparison with all sorts of other points of view relevant to this subject would lead me too far astray from the main thread of the argument. Therefore I will just mention a series of factors which are (have proved to be) important for the plausibility of laws, without ordering them in any way or giving priority to some of them. The concepts necessary for this (e.g. evidence, explanation, prediction, testing etc.) will be used in an informal and pragmatic sense. In scientific practice both cognitive and social aspects play a part when assigning plausibility to laws and theories.6 As far as the cognitive aspects are concerned I will limit myself to three factors which most certainly contribute to laws and theories being plausible. (a) First of all, a law should be directly supported by the facts about the object-domain for the explanation or prediction of which the law was proposed in the first place. If Boyles gas law is to be plausible in some degree, we need experiments in which the values of pressure, volume and temperature will prove to cohere in the way postulated in the law. (b) This direct evidence is in general necessary but certainly not sufficient to give a hypothesis a reasonable degree of plausibility. It is also important to test a postulated law in completely different (independent) object-domains, e.g. the testing of Newtons law of gravitation with the help of the phenomena of falling, the motion of celestial bodies and the tides.7 The larger and more varied the domain of application of a law becomes, the greater the increase in plausibility. To illustrate this point I shall give two further examples. In 1900 Planck published his hypothesis on the discrete character of the energy of heat radiation emitted by black bodies. This hypothesis did not become plausible among physicists until 1907, after Einsteins application of the quantum concept in a new field, i.e. that of the temperature-

85

dependence of the specific heat in the molecular, kinetic theory.8 Another example is the light quantum hypothesis, which Einstein proposed in 1905 to explain the photo-electric effect. This radically unclassical hypothesis claiming that free light and X-rays consisted of particles (quanta) and which was based in the first years after 1905 on only one domain of phenomena i.e. the photo-electric, failed at first to gain any plausibility: Physicists in 1908 had no reason whatever to accept the light-quantum hypothesis on the grounds of X-ray and photo-electric phenomena.9 Only between 1921 and 1924 did the hypothesis become more plausible by successful application in other fields, such as the Compton effect, the Doppler effect etc.10 (c) A third important cognitive factor related to the plausibility of laws is the proven intrinsic fertility of the model which is associated with or forms the basis of the theoretical laws.11 A well-known example of this is Bohrs atomic model which, as for instance Lakatos has pointed out, showed a number of spectacular successes on the basis of its internal dynamics, notably in the first years after 1913, especially in the field of atomic spectra (so, always in the same domain of phenomena; this in contrast to the factor of independent testing just mentioned).12 Also Derksen mentions the same example, following McMullin, but for another reason, namely that of an argument for a possible realist interpretation of the model. To highlight the difference between claims of plausibility and of realism I will briefly discuss his arguments. Derksen proposes: An argument based on proven fertility is in certain spectacular cases intuitively acceptable for supporting a claim to a realist interpretation of the model in question.13 But even this careful description is too loose for Bohrs atomic model (which Derksen considers to be a spectacular case14). This model is after all itself without a doubt inconsistent15, and that while Derksen argues thus explicitly a model can never be a candidate for reality when it has contradictory characteristics.16 All this does not, however, form an argument against the much weaker claim to plausibility: inconsistent or not, Bohrs model was certainly in its early days most successful and this was enough of a reason to hold it plausible, even with reserve, as was the case. By this I do not deny that consistent models (and theories based on them) will in general be more plausible than inconsistent ones. The point is that inconsistent models and theories are not necessarily totally implausible.17 As will become clear from the remainder of this chapter I think that even inconsistent theories can be interpreted realistically in a certain way.

86

So much for cognitive factors which may play an important part in the assigning of plausibility to laws and theories. There are also social factors. Especially during the past years there has been an increasing flow of literature containing explicit and accurate analyses of the way in which social factors exert influence at a cognitive level, or better still: of the way in which social and cognitive factors are intrinsically interwoven in the development of the natural sciences. As one of the spokesmen of the so-called strong programme in the history and sociology of knowledge, Shapin puts it:
The number of concrete studies which analyse the production, evaluation and acceptance of scientific knowledge as social processes is large and growing larger. The continued assertion that scientific knowledge is autonomous and transcendent is now more often founded on ignorance of this new literature than on considered criticism thereof.18

In this brief discussion of plausibility of laws and theories I will confine myself to one illustrative example of a case study by Harvey, entitled: Plausibility and the Evaluation of Knowledge: A Case-Study of Experimental Quantum Mechanics.19 This study concerns a series of experiments carried out in the 1970s to solve a theoretical dispute between the generally accepted and very successful quantum theory and an alternative, a class of so-called local hidden variable theories. We should note that this alternative did not attack, from an empirical point of view, the entire quantum theory, but only its predictions concerning a specific domain, i.e. that of long range correlations. Harvey describes a number of stages in the controversy. First of all (up to 1976) two experiments had been performed, one with a positive and the other (Holts) with a negative outcome for the quantum theory. Still, this does not mean that the situation at the time was symmetrical in other words it is certainly not just the direct evidence which determines the plausibility of hypotheses. In fact a number of unscientific or at least non-empirical arguments, as Harvey calls them, proved to work against the hidden variable theories. He mentions, amongst others, inductive arguments (the quantum theory is for a large domain very successful and will thus be right in this dispute, too), ad hominem arguments (Holts supervisor, Pipkin, had a tainted past, in the sense that in earlier days he had laid an unjustified claim to finding an experimental counter-example against quantum electrodynamics), expertise arguments (if many competent experts agree that the quantum theory is the most rational one, then this is probably the case).20 The plausibility which the sum total of these

87

arguments lends to the theories concerned depends, according to Harvey, most of all on the micro-social context in which the scientific dispute is set. In the second stage of the dispute a number of results of other experiments all confirming the quantum theory are introduced. Although this certainly contributes to the plausibility of the quantum theoretical explanation of the long range correlations, it does not refute, Harvey argues, the hidden variable alternative in a logically compelling manner. In every experiment some tacitly accepted, but unverified auxiliary hypotheses or observation theories play a role. In the discussions about the experiments at issue two assumptions in relation to the functioning of the experimental equipment were acknowledged, which were of great importance in drawing the conclusions (pro the quantum theory). But in this second stage the possibility of rejecting one of these two (or both) was not taken seriously. Logically this is possible, but factually it would require the postulating of such pathological mechanisms that serious experimenters would not waste their time on it, it was said. In the third stage the plausibility of one of these auxiliary hypotheses was decreased after all simply because a competent and highly esteemed experimenter (the Frenchman Aspect) took the trouble and managed to get the money to test this pathological hypothesis. Partly owing to his plans the interest in the hidden variable theories (which certainly within the larger community of physicists were regarded as refuted by the earlier experiments) was revived and its plausibility increased a bit again.21 Harvey concludes that in this (and other controversies) the winning side does not possess truth, but rather that it has monopolized plausibility.22 In general this modern sociology of science has given us most interesting and valuable insights in many social aspects of the production, evaluation and acceptance of scientific knowledge. The results of these sociological studies are of direct importance to both philosophical and social and political discussions on the status, method, role etc. of the various sciences. With the above example I have tried to illustrate the proposition that this kind of micro-social factors can influence the increasing or decreasing of the plausibility of general laws and theories. At the same time it will be obvious that the relevance of these micro-sociological studies for philosophical, (macro-)sociological, social and political debates and controversies on science is also subject to limitations.23 Firstly, the influence of local micro-social factors (such as Pipkins tainted past and Aspects competence) on the plausibility of laws and theories will often be short-lived; and secondly, plausible laws

88

(and the verifiably true singular propositions based on them) play a part in many more social contexts than just those, in which they are produced. From this follows that the plausibility of laws and theories is certainly not completely determined by local social factors, which operate only for a short time and within a limited context. Arguments which deny the possibility of philosophical realist interpretations of science on the basis of the results of this kind of research in the sociology of science and this happens quite often do therefore not hold true. They do not exclude after all that there exist more general aspects of the long term development of the natural sciences which we could interpret realistically.24 Nevertheless there are other, weighty arguments putting drastic limits to the possibilities of such realist interpretations. In the next section I shall expound these arguments which rest on the occurrence of conceptual discontinuities in the historical development of the natural sciences in great detail. In the context of these expositions plausibility will again crop up. In section 4.4 and 4.5 in particular I will therefore come back to this concept.

4.3

Conceptual discontinuity and scientific realism

The aim of this section is to show that two kinds of realism, i.e. naive and convergent realism, are irreconcilable with the fact that the theoretical development of the experimental natural sciences is regularly interrupted by conceptual breaks. This means that if we take such historical data seriously these kinds of realism are not tenable. In the next two sections of this chapter I shall therefore search for a form of realism which is compatible with the occurrence of conceptual discontinuities between theories. Necessarily the claims of this kind of realism will have to be more limited than those of naive or convergent realism. The problems to be dealt with in this section are complex and so is the accompanying argumentation which consists of a number of separate steps. First I make some general remarks on the approach to the problem of realism (1). Then I explain extensively the concept conceptual discontinuity (2). Next I treat the consequences following from these discontinuities with respect to naive and convergent realism

89

(3). Finally, I conclude with a brief discussion of the various kinds of relativism (4). (1) If scientific realism is to hold a philosophically interesting and socially or ideologically relevant position, then it should at least acknowledge these two statements: a. The existence of the world (of nature) and its general structure is, in a way which can be specified in more detail, independent of the existence of human beings and the process by which they acquire knowledge. b. Generally speaking, concrete scientific propositions provide us with knowledge about this human-independent reality, in a way which can be specified in more detail. The first claim belongs to the field of philosophical ontology, a field notably neglected or despised by the philosophy of science.25 I have touched on this in section 3.4 and I will come back to it briefly in 4.5. In this chapter I shall be mainly concerned with the second claim, with the problem of a realist interpretation of concrete scientific propositions. This problem of a realist interpretation of scientific propositions can be viewed both semantically and epistemologically. The first approach analyzes the relation to reality of propositions, or of entire theories, chiefly with the help of concepts such as meaning, reference and truth, the theories and propositions mostly being taken as abstract and isolated products of knowledge. The epistemological approach on the other hand aims to include an account of science as a human activity and as a historical and social process and tries to find good arguments which would make a realist interpretation of knowledge acceptable, given certain properties of these activities and of this process. Referring to what I proposed in my critique of Habermas theory of truth in 2.3.1 I would like to stress once more that the distinction semantical-epistemological is in itself quite useful, but that it is wrong to isolate the two approaches from each other (as happens far too often). An example of this is Cookes rhetoric as he uses it to support his purely semantic approach. Cooke wants to fix the relation to reality of (theoretical) terms in a formalized theory by interpreting them with the help of stable semantical rules. On the one hand he writes this:
The stability of these rules has nothing whatever to do with observation, or with any other epistemological considerations. The degree of subjective conviction with which we attach an interpretation to a term is absolutely irrelevant.

And on the other hand:

90

The existence of stable semantical rules depends upon nothing else than our ability to say what things physical theory is about, and the stability of these rules with respect to the theory consists only in the fact that our commitment to talk about those things is stronger than our commitment to the theory in question.26

The stability of semantical rules and with it the relation to reality of theories is in this view completely independent of the fact whether we have good reasons for believing the truth of the theories in question. The ability and commitment of the philosopher have apparently nothing to do with the scientists conviction. The question arises how such semantic discussions can ever be relevant for obtaining an adequate picture of what science is and what it does. My criticism is therefore not that the distinction between semantical and epistemological aspects is nonsensical. The point is that they are inextricably bound together. For the claimed stability of semantical rules, too, good epistemological arguments have to be supplied. And good epistemological arguments are based on, or make at least extensive use of, analyses of the factual historical development of the natural sciences, within which the (subjective or not) convictions of scientists have their place. All this holds good most of all for the correspondence theory of truth. If this theory cannot be proved also epistemologically (e.g. through convergent realism) it seems to me uninteresting and irrelevant with regard to the philosophical problems of the experimental natural sciences. (2) An important result of the analyses of the factual historical development of the natural sciences is the insight that there have been big and radical conceptual discontinuities. It is these discontinuities in particular which have produced crucial problems for realist interpretations of the natural sciences. Let us start by examining an example. In Lorentz electron theory electrons have at least these central properties:
Electrons exist within the interiors of molecules and atoms. They have a definite internal distribution of charge density, falling off continuously from their centers. They are clearly discrete particles, the continuity of charge density at their boundaries functioning as a mathematical simplification. They move as discrete particles, oscillating about equilibrium positions in the atoms of dielectrics in response to electrical forces exerted on them by the ether, and moving freely in conductors. And their identification, on the basis of Zeemans work, with cathode ray particles indicates that they are negatively charged and permits the calculation of their charge to mass ratio. Their mass and radius can then be calculated on the basis of the theory by treating the mass as resulting entirely from self-induction.27

91

It is quite instructive to comment on these central properties in Lorentz theory, one by one, from different theories developed after his: 1. If we want to speak about electrons as discrete particles, the waveparticle duality in modern quantum theory will always cause a radical break in such an interpretation. Electrons are both particles and waves in this theory. 2. If we cannot speak simply about localized particles it will also be unclear what the notion of an internal continuous distribution of charge within the electron actually means. 3. In connection with this the motion of electrons as localized particles also becomes problematic. According to the quantum theory we cannot speak consistently of the path which individual electrons follow. 4. The classical atom model of oscillating electrons could not explain the spectra and was therefore replaced by Bohrs (inconsistent) model. Then again this model was superseded by the abstract quantum theoretical description of atoms in terms of state functions in Hilbert space. 5. According to the special theory of relativity electrons do not have a fixed (invariant) mass, the latter is dependent on the relative velocity between observer and object (c.q. the electron).28 6. The concept of an electron as a rigid particle with a finite extension (so: no point mass) was already problematic in the Lorentz theory. The hypothetical Poincar pressure had to be introduced ad hoc to explain why the negative charge in the electron stays together. Within the relativistic classical theory of electrons such a rigid electron of finite extension (with a certain radius) is even inconsistent: in a rigid particle of a finite size forces will propagate at an infinite speed, this in contrast to a basic postulate of the special theory of relativity. On the other hand, the model of a point electron has, in both theories, the unpleasant property of an infinite self-energy. 7. The value, too, of the electric charge of the electron in Lorentz theory simply cannot be compared with that in modern physics: the bare charge of the electron is much greater than the measured charge. Indeed, in quantum electrodynamics the bare charge is assumed to be infinite.29 Mutatis mutandis the same holds true for the value of the electrons mass. Thus after Lorentz theory a radical conceptual change took place with respect to the concept of electron.30 This implies that, if we want to

92

maintain an unambiguous terminology, we should speak about Lorentz electron, Dirac electron etc. instead of electron in itself. Three points, all philosophically most relevant, become apparent from this example: I. To learn and master various conceptually discontinuous theories apparently does not involve any fundamental problems. II. The conceptual differences between these theories can apparently be made explicit in language. III. The example concerns conceptually different theories on the same electrons. After all, if it is to be relevant in the context of a discussion about realism, then one must suppose that the various theories in one way or another refer to one kind of entity. The fact that Freudian psycho-analysis and the general theory of relativity are conceptually discontinuous does not pose of course any problems to realist interpretations. From an intuitive point of view these observations seem right. Still needed is a precise philosophical explication of the notion of conceptual discontinuity in support of this intuition. First of all a remark about terminology. Thus far I have used the term theoretical in a general sense. But there are in fact two aspects which play a part here, the conceptual and the formal, or more in particular, the mathematical. From now on I shall, where necessary, distinguish between theoretical-conceptual and theoretical-mathematical. This distinction is important because theoretical discontinuities are usually of a conceptual nature. Mathematically there often exist continuities between successive theories in the form of formal, intertheoretical relations. Notably in the sections 5.2.3 and 5.2.5 I will dilate upon this. To return to the explication of conceptual discontinuity. This, too, should be done in several steps. Only after introducing the notions of underdetermination and formal correspondence can we say what conceptual discontinuity really means. First of all I will make use of the notion of underdetermination of theories by the experimental data.31 I shall define underdetermination here, in my own words, like this. Suppose we are given the descriptions of the material realizations of a class K of experiments for a domain D of phenomena.32 It is now logically possible and (what is much more important) in practice this occurs regularly, that two or more different theories exist which are all equally adequate in predicting and explaining the material realizations from this class K. I would like to make two comments on this definition. First of all, it does not use the questionable distinction, let alone the contrast, between

93

observational and theoretical scientific propositions. As stressed before, the description of the material realization is not an account of scientific observations during the experiment, it does not refer to what is traditionally meant by observation sentences.33 It is, however, true that every definition of underdetermination requires the use of two different kinds of language: the theoretical languages and the language in which the experimental data have been described. If one handles, as I do, a notion of underdetermination and simultaneously accepts the existence of a radical theory-ladenness of observation sentences, then there has to be another language, apart from the theoretical languages, by which the experimental data can be described. In our case this is the language of the material realization of experiments mentioned in 3.3.34 Secondly: this definition concerns the underdetermination of two (or more) theories with respect to all experiments carried out up to now in a specific domain of phenomena. For other domains these theories may give different results. This is why underdetermination is usually defined more inclusively by requiring the theories involved not merely to be equally adequate with respect to all experiments carried out within a certain domain up to the present, but in fact with respect to all possible experiments.35 Although arguments in favour of this more inclusive version are welcome, they are not really necessary in the context of the discussions to be held here as we will see below. The kind of underdetermination which is most relevant for my argument concerns those instances in which one theory in a certain domain is a limiting case of another theory. This means that there exists a certain domain for which they are both adequate, but of which they give simultaneously a different theoretical description. Some wellknown examples from the development of physics are: geometrical and wave optics (in the limit of short wavelengths), Newtonian and Einsteinian mechanics (in the limit of low velocities), classical and quantum mechanics (in the limit of large quantum numbers). In the next chapter I will discuss this last example in great detail. Thus the concept of underdetermination allows us to specify the idea that we can interpret and describe theoretically one and the same class of (experimentally produced) phenomena in different ways. Please observe that at this point I talk about different ways and not yet about conceptually discontinuous ways. What the latter means I was after all just about to explain with the help of the notion of underdetermination.

94

We are now therefore dealing with the differences between theories in a preliminary and informal fashion. Two theories, for instance, differ if the number of quantities in both theories are not equal, or if the types of mathematical symbols and equations are clearly different, etc. If on further consideration theories, although differing in these properties, prove to be continuous from a conceptual point of view, then these differences were apparently only formal. In other words: the differing of theories in the sense explained above is in itself not sufficient for conceptual discontinuity. Quite important in all this is the actual fact that we can learn (know and apply, that is: live) various theories, whether discontinuous or not. Precisely because we can learn new things, changes can occur. And since we can learn to know other theories, we can be aware of similarities and differences between them. In this there is an obvious difference between the culture of science and the more embracing, total culture. Generally speaking it is certainly not completely impossible to learn to understand other cultures. I do feel, however, that this learning process will only be partly successful when the cultures involved are far apart: learning to know and live a completely alien way of life seems not feasible in general. The point is that such a radical strangeness between different scientific cultures is not at all the case. It goes without saying that Einstein still understood classical mechanics after 1905, after the introduction of the special theory of relativity (probably even better than before). For an adequate explanation of the notion of conceptual discontinuity we will, apart from the idea of underdetermination, also need the concept of a formal correspondence between terms from different theories. Because this concept will only be defined in detail in section 5.2 I must here suffice with a general indication. The point about conceptual discontinuity is that we are of course not dealing with arbitrary concepts from different theories. It is not very interesting to know that Newtonian mass is a quite different concept from Einsteinian momentum. What is significant is that the terms mN and mE are conceptually different. In other words: there exist conceptual differences between terms corresponding with each other. In the next section, but even more in the next chapter, I try to make it plausible that this correspondence is of a formal, or more specifically, of a mathematical nature. Moreover, I try to show with the help of examples how we can achieve knowledge of such correspondences.

95

By means of the concepts formal correspondence and underdetermination we can now define conceptual discontinuity as follows. Suppose two theories contain terms which correspond formally with each other. Suppose then that we are dealing with two different theories, which are both just as adequate for one and the same domain of experimental phenomena. Then these theories are conceptually discontinuous in a philosophically interesting way if the terms concerned are conceptually incompatible, i.e.: if a simultaneous application of these terms to the phenomena in the domain in question results in a conceptual incoherence. For instance: if we assume for the moment that the terms eL and eD from Lorentz and Diracs theories of the electron correspond formally with each other, and that there exist domains for which both theories are equally adequate, then the terms eL and eD are clearly conceptually incompatible. After all, they assign incompatible properties to the entities they refer to: eL is for example a localized particle, but eD is not, etc.36 Thus Lorentz theory of the electron and that of Dirac are conceptually discontinuous. Concluding, we can say that we have thus far reached a precise explanation of the notion of conceptual discontinuity between adequate theories with a shared domain. This discontinuity is carried by (or even better: made possible on the basis of the existence of) two continuities, namely of the description of the material realization (see the definition of underdetermination) and of the continuity implied by the occurrence of formal correspondences. I will come back repeatedly to these aspects of the development of science later on. Precisely on the basis of these kinds of continuities conceptually discontinuous theories are not incommensurable, in the sense of completely incomparable. A more fitting term is irreconcilable. Conceptually discontinuous terms cannot be unified into one coherent body of knowledge. The fact that the discontinuous change of theories also presupposes continuities has of course been pointed out before. If we leave it like this as in fact happens quite often we content ourselves with merely stating the problem. A second step is to solve it, i.e.: to point out these continuities precisely and to clarify them philosophically. (3) Thus far the notion of conceptual discontinuity. The next step in the argumentation of this section is to expound the consequences of the occurrence of conceptual discontinuities for naive and convergent realism.

96

Naive realism assumes that concrete scientific propositions or theories generally correspond in a direct or literal way with the states of affairs in a human-independent reality. There is no need to dwell extensively on this viewpoint, because it is hardly advocated any more (and moreover because the arguments against convergent realism hold good a fortiori for naive realism). For years and years philosophers of science have realized, also on the basis of the insight that scientific knowledge is changing knowledge, that such a naive-realist interpretation of concrete scientific propositions or theories is not tenable. If theories from the natural sciences are always subject to changes and remain so, then any selection of specific theories as being correspondent with reality is haphazard. In the light of these evident problems faced by naive realism, attempts have been made to show that science can at least approach truth closer and closer; that theories approximate truth in some important aspects, even if they are improved upon by their successors. In this view the evolution of science produces increasingly more points of contact with reality and thus converges in the direction of truth.37 Elements of such a convergent realism we find, though sometimes developed very differently, in philosophers such as Peirce, Popper, Sellars, Boyd, Putnam and Newton-Smith. The evaluation of naive and convergent realism will be based on two considerations. Firstly: above the possibility of formulating formal correspondences between terms from conceptually discontinuous theories was pointed out. I explicitly avoided the term translation in doing this. In my opinion this term is used in far too many senses in philosophical discussions, with the results of opaqueness and confusion. Apple we can, in a lot of cases, translate into the Dutch word appel. This is, because these two terms (a) have about the same extension and (b) have in both languages more or less the same meaning. Of course there is much more to be said about the notion of translation. I am, however, not so much concerned with criteria of translatability, as with nontranslatability. From the apple-example I only want to conclude that when terms differ radically in meaning a translation is in any case impossible. For example, consider the terms Lorentz electron and Dirac electron occurring in two conceptually discontinuous theories both of which we could master properly: we can formally relate each to the other, but we cannot translate the one into the other.38 A second important point when evaluating naive and convergent realism concerns the actual occurrence or non-occurrence of conceptual

97

breaks. For claritys sake I would like to stress that my view is certainly not that every change in the experimental natural sciences implies a conceptual discontinuity. Only historical and philosophical research can prove when this is the case. In the next chapter I discuss in great detail one such example, namely that of the discontinuity between the classical physical theories (mechanics and electrodynamics) and the nonrelativistic quantum theory, to illustrate the philosophical argumentation. Of course there are many more examples.39 In classical mechanics and electrodynamics we are dealing, however, with a paradigmatic example of mature scientific theories, which play a very important part in much research in the experimental natural sciences. Consequently even this one example poses enormous problems for naive and convergent realism. But there is another conclusion that can be drawn from the occurrence of conceptual breaks in the history of the experimental natural sciences. Like Hesse, I also start from the principle of no privilege, according to which our own scientific theories are held to be as much subject to radical conceptual change as past theories are seen to be.40 This means that one should not assign a privileged place to present scientific theories in philosophical argumentations, and in particular not in those about realism. We should not suppose that just at this very moment the development of science has been completed or that exactly from now on this will be of a different character. This no-privilege principle seems to be most plausible in the light of what we know about human history in general and history of science in particular. I cannot see any good arguments for denying its validity. Thus I assume that it is not at all unlikely that even our best present theories will be replaced in the future by theories which are radically different from a conceptual point of view. After all this preliminary defining of concepts and viewpoints we finally reach the central problem of this section, i.e. the evaluation of specifically convergent realism.41 To evaluate the convergence-view, we first must know in which sense one can speak of convergence; in other words: in which sense does a theory T2 describe reality (or approach truth) better than its predecessor T1? We can explain this better thanrelation in generally speaking two different ways. (a) The most obvious way would be that in convergent realism we are concerned with a conceptual convergence. One could explain this for example thus: the corresponding concepts of T1 and T2 refer to the same entities, but the list of correctly ascribed properties to the shared entities is longer for T2 and incorporates that of T1. Putnam formulates this as:

98

There is nothing in the world which exactly fits the Bohr-Rutherford description of an electron. But there are particles which approximately fit Bohrs description: they have the right charge, the right mass, and they are responsible for key effects which BohrRutherford explained in terms of electrons; for example, electric current in a wire is a flow of these particles.42

Note that Putnam speaks about the right mass and the right charge, which apparently means: right on the basis of our present theory. He really should have added explicitly that in order to have a truly better than-relation (and he doubtless has this in mind), this present theory is an improvement on the Bohr-Rutherford theory: our theory on electrons says the same and more. It is obvious that this view of convergence denies exactly those two positions which I defended above. On the one hand conceptual change is often much more radical than Putnam wants us to believe. As shown in the first part of this section indeed also referring to the electron, the mass and charge of electrons are not exempt from conceptual change. This means that in this case we cannot speak of a convergence whereby, at any rate if we take the term literally, the differences between successive theories would decrease in the course of time. On the other hand, Putnams argument is explicitly inconsistent with the noprivilege principle: what in the Bohr-Rutherford theory is right is judged from our present theory. From a methodological point of view there is nothing against this, but as an argument for realism (in which right is interpreted as corresponding with an independent reality) this is clearly not plausible.43 (b) The second way of explaining in which sense a theory T2 is better than its predecessor T1 is pragmatic: with T2 one can make more and/or more precise predictions and explanations of experimental data than with T1.44 There are two difficulties with this interpretation. First it should be observed that this kind of progress in which the new theory can explain or predict all phenomena which the old theory could just as well or even better is often only possible in principle. Thus we ought to be able in principle to describe all macroscopic phenomena adequately with quantum mechanics. In practice this is almost always impossible. This is why precisely in a pragmatic sense the older theories remain just as indispensable as before. Of greater importance when evaluating convergent realism is, however, the second problem: that the increase of pragmatic possibilities

99

does not imply that the theoretical ontologies correspond better and better with reality. Still, many realists claim such a correspondence. In Laudans words:
any genuine realist must insist that T1s underlying ontology is preserved in T2s, for it is that ontology above all which he alleges to be approximately true.45

The conclusion of this criticism of convergent realism is that in any case those forms of realism claiming that science can provide us with an at least approximately true, conceptual image of reality or that the internal ontologies of successive mature scientific theories correspond with reality turn out to be not plausible.46 In the next sections I examine the possibilities for other kinds of realism (which apparently differ from what Laudan calls genuine realism), given the failure of convergent realism. (4) In concluding these discussions on conceptual discontinuity and scientific realism I would like to add some remarks on the related problems of relativism. My views as explained up to now imply a conceptual relativism with respect to theoretical propositions from the experimental natural sciences.47 Concrete scientific propositions are always phrased in specific theoretical language systems, even though this may be obscured by the language we use. In order to be able to judge whether something is or is not verifiably true, we have to speak about mN and mE instead of m, and about eL or eD instead of e.48 The concepts used in formulating these propositions are unable to tell us, however, what reality in itself looks like. On the one hand this is because of the occurrence of conceptual discontinuities between theories we can learn to master and live, but which we cannot translate into each other; while, on the other hand, we cannot assign a privileged position to any theory, as far as claims to representation of reality are concerned, on the basis of the plausibility of the no-privilege principle. This conceptual relativism should be clearly distinguished from other types of relativism. In the first place from the relativism of truth, which argues that a scientific proposition which may be true in one context (or for one group) can be false in (or for) another. This relativism of truth assumes quite the opposite of what I defended here. For it presupposes that we first have propositions, which apparently have the same meaning in all contexts (or for all groups), but which nevertheless can subsequently be deemed true or false, depending on the context or group. In other words: this kind of relativism presupposes the existence of general conceptual continuities between the various theoretical

100

language systems in which the scientific propositions are formulated.49 The existence of such general conceptual continuities is not very plausible. Moreover, its assumption is very often inconsistent with other statements by the advocates of this relativism of truth, so that understandably enough Newton-Smith has little difficulty in showing the incoherence of this viewpoint.50 Nor does my point of view imply a perceptual relativism, in the sense that scientists, whose observations are laden by conceptually discontinuous theories, live one way or another in different worlds. On the contrary: as will be shown in the next section the conceptual incompatibility of theoretical terms does not exclude that these terms may co-refer to the same elements of a common domain, and so to the same world. Since the main subject of this chapter is realism I restricted the systematic discussions here primarily to conceptual relativism. One could, however, go further in the direction of a context-dependence or relativism of the reasons which scientists give for the accepting of various theories.51 In the previous section we briefly touched on this when discussing the role of plausibility in the criterion of verifiability. In my eyes this problem, which does deserve the attention of philosophers, is, however, mainly important with respect to the larger question of scientific rationality. For an adequate treatment of the problem of realism the discussions about conceptual discontinuity and conceptual relativism given above seem to suffice. Finally, in the previous section a criterion of the verifiable truth of propositions was introduced. I talked about verifiably true and not about true in itself, because I did not want to run ahead of the results of the present section. In other words, I did not want to preclude the possibility that a proposition C can be verifiably true in one theoretical description T1, but not in another description, which we obtain by a translation of T1 into a better theory T2. In the meantime we have seen that such a translation is generally impossible and that the notion of a theory which is better than its predecessors can at best be used in a pragmatic sense. This implies that in fact we cannot expect anything more from concrete scientific propositions than that they are verifiably true. But if this is the case, we can then just as well leave out the prefix and speak simply of true. This is what I shall do from now on. It should, however, be noted that of course the difference between verifiably true and verified true (in the old terminology) has not disappeared. That is to say, it will always remain possible for a proposition to be true, while we do not know its truth.

101

4.4

A criterion of (co-)reference

In the previous section we saw that naive and convergent realism are incompatible with the occurrence of conceptual breaks and the noprivilege principle. Because I take the arguments derived from the historical-social nature of the development of science seriously I concluded that such forms of realism are untenable. This means that if realism is at all possible its pretensions should in any case be quite restricted. As will soon become apparent, my resulting point of view will be that we can say of terms in scientific propositions that they refer to elements in a human-independent reality, but not that these propositions describe reality as it is in itself (approximately or not). Thus this kind of realism implies reference, but not correspondence, picturing, reflection, representation or something like that. In this section I will be concerned with the question of the reference of terms occurring in the theories of the experimental natural sciences. Its aim is formulating and defending a concretely applicable criterion of reference. We shall again make extensive use of the analysis of scientific experimentation given in chapter 3. Suppose a theoretical description of an experiment, e.g. of the form p => q. Generally the descriptions of the situations p and q and of the laws which guarantee for closed systems that p is sufficient for q will contain all kinds of descriptive terms, e.g. for properties, relations, things, events, processes, mechanisms etc. The intention is to formulate a criterion of reference which will apply to all these kinds of terms. In other words we are not only concerned with referring to entities, but also and this might be less usual to properties, mechanisms and the like. In general I will therefore speak of reference to elements of reality. As a criterion of reference I would like to propose this: the descriptive terms occurring in the theoretical description of an experimental situation refer to elements of a domain in a humanindependent reality, if the material realization of this experimental situation is possible in a reproducible manner. This criterion is phrased purposively in terms of a sufficient condition. The point is to specify epistemically relevant conditions under which we can claim justly that the theoretical terms at issue refer; i.e.: under which we can obtain knowledge of the reference of these terms. In

102

this case these circumstances refer to the material realization of closed experimental systems. This does not exclude, however, that terms in other circumstances, for example in what Bhaskar calls open systems, may also be referring.52 The criterion of reference thus defined presents us with a number of questions: (1) what does the requirement of reproducibility mean? (2) which role does theory play in this criterion? (3) how is co-reference possible? (4) what do referring terms refer to? (5) how is reference fixed for the first time? (6) how is change of reference of a term possible? The aim in answering these questions will be to show the adequacy of the criterion and to arrive at such a degree of plausibility for it that it can be used successfully in the proposal for a realistic realism in the next section. (1) First of all the aspect of reproducibility: if an experiment in the natural sciences is to continue to be successful in the long run, it has to be at least reproducible. This requirement of reproducibility causes claims of reference to be always fallible. We will never achieve absolute certainty whether certain terms do or do not refer: the fact that the material realization is reproducible up to the present day, does not guarantee that this will also be the case tomorrow. But an up to now proven reproducibility does give us good reasons to assume that the term in question refers. In the criterion of reference we are not concerned with the reproducibility of experiments as such, but, more specifically, with reproducibility of the material realization of experiments. This does of course raise the question of how we can find out whether a certain material realization is or is not reproducible. As the definition given in 3.3 shows, the description of a material realization is an account worded in ordinary language of the tasks set by experimenter A for layperson B, which B carried out correctly according to A. The material realization is now reproducible if also other laymen B, B, etc. succeed in carrying out the experiment again on the basis of the description of the material realization supplied by A and B; and this in the way A holds to be correct and again according to A, with the same theoretical result. Suppose, for example, that A concludes from the material realization carried out by B that liquid x boils at (123 0,1)C, then this material realization is reproducible if A arrives at the same

103

conclusion from the material realizations carried out by B, B etc.53 It seems feasible to verify the reproducibility of material realizations in this (I admit somewhat laborious) way; certainly if we make the description of these material realizations minute enough and moreover select our layperson a bit, e.g. on the basis of possessing a reasonable degree of skilfulness. (2) As has been stressed in section 3.3, theory is an essential requirement if we want to carry out an experiment successfully.54 The theoretically informed assignments and corrective indications of the experimenter are indispensable, also for arriving at the description of the material realization, even though this description itself is put in everyday language. Consequently, specific theory is needed for the application of the criterion of reference to concrete cases. The criterion states explicitly that it is the terms of that theory which refer. This theory is especially needed for making the reproducibility of the material realization possible. In order to know what we must reproduce and how we can do this, we have to fix the experimental situations by a theoretical description. Of course it may happen that it is not (yet) clear whether a certain experiment is reproducible. In the light of the theory-dependency of this reproducibility, controversies on this will mostly be caused by the fact that an acceptable theory on the phenomena of the domain involved has not yet emerged, so that a fortiori no consensus can be reached about the right materialization of that theory, that is: about which experimental devices should be used and how we should use them.55 In view of the fact that the consensus on the reproducibility of by far the most experiments in the natural sciences is widely spread, the occurrence of controversies does not supply a general argument against the possible applicability of the criterion of reference. As we saw in 3.2 reproducibility will never obtain absolutely, in the sense of: under all circumstances and with all kinds of different objects. It is the experimenter who decides whether a material realization has been carried out and described correctly, as the definition clearly shows. (Division of labour has nothing to do with democracy). (S)he selects the objects to be examined and the measuring equipment on the basis of his/her theoretical views and with the very aim of possible reproducibility of the material realization. By this choice a theoretically defined domain of phenomena is determined, to which this and identical experiments refer.56 Thus, reproducibility is always reproducibility within a theoretically defined domain.

104

Because of this, theoretical terms do not simply refer to elements of reality itself, but to elements in a domain of reality and this seems to me to be characteristic for the criterion of reference proposed here.57 It is most important to observe that the criterion of reference does not concern a translation (or a reduction) of the theoretical description into (or to) the description of the material realization, whereby theoretical terms are first linked to certain terms from the description of the material realization and then owe their reference to the reference of those latter terms. Such an empiricist approach faces (at least) these two enormous problems: (a) The failure of the logical empiricist reduction of theories to either a phenomenalist or a physicalist description is now generally regarded as quite definite. (b) But also: even if such a reduction were successful, the problem of reference would only be partially solved. We must still show that the terms from the description of the material realization refer to a humanindependent reality and how they do this. Apart from the specific difficulties arising because of this, we would inevitably wind up in an infinite regression. Therefore the criterion of reference proposed here is not based on such a founding empiricist strategy. I do not presuppose a possible translation or reduction, but regard the theoretical description as a direct interpretation of the material realization itself (not of its description!). This description has only been introduced as an artificial manoeuvre, in order to get an empirical mechanism with which to establish the reproducibility of a material realization, and also (see point (3) and (4) below) its identity under various possibly conceptually discontinuous theoretical descriptions. This objection to the idea of translating or reducing the theoretical description into a description of the material realization can also be phrased as follows.58 The relation between the theoretical description and the material realization is in my conception of a causal nature, in the sense that there is a material reality which is (in part) the cause of our theoretical terms referring. Those philosophers who try to translate the theoretical description into a description of the material realization attempt the impossible transformation of a causal relation into an epistemological one. In this way the theoretical description would be reduced to, and thus based on and justified by, the description of the material realization. However, as explained under (b), this will not lead to a solution but rather to a shift of the problem.

105

In contrast, the problem of the reference of terms from the language of the material realization is independent of that of the reference of theoretical terms in the view defended here. Required is merely that the description of the material realization remains (approximately) constant. Whether the terms occurring in this description actually refer or not does not matter in this case. Of course there will be a transformation from the experimenters theoretical notions to the instructions formulated in the language of the material realization. The crucial point of the criterion of reference is, however, that this transformation is only required to lead to some description of the material realization which, firstly, makes it intersubjectively reproducible (see (1)) and, secondly, makes it possible to determine co-reference (see (3)). Whatever other characteristic contents the language of material realization may have is not important for the applicability of the criterion of reference. It goes without saying that in our society this language will have much in common with the language of things and events. But in principle one can also imagine the communication about the material realization between the experimenter and the laypersons to take place in another daily language, e.g. a magical or an expressive language. That is why at this point an essential difference from Habermas views emerges. In my view there is a systematic relation between (our knowledge of) the reference of theoretical terms and the material realization of experiments (in Habermas words: the instrumental action), but not between these theoretical terms and the description of the material realization (for Habermas: the description in terms of things and events; see 2.2.1). The latter means that aspects of the contents of this description of the material realization do not impose any limits on the possible theoretical interpretations and ontologies (cf. 2.3.2, point 5). (3) An important, if not the most important, advantage of the criterion of reference proposed here is that it implies a full, in the sense of a truly intertheoretical, notion of co-reference. Philosophers and historians of science usually assume, implicitly or not, an intertheoretical coherence in the form of co-reference (see the examples in section 5.2.5). However, quite often they neglect to explain what this coherence exactly consists of and how it can be integrated into their other views on science. As explained in the previous section, we have to prove that corresponding terms, even if they are conceptually incompatible, in

106

certain cases do refer to the same elements in reality. Thus there are two questions to be answered: (a) what are corresponding terms? (b) when do these terms co-refer? The first question I can only briefly answer here with the help of an example.59 We expect the Newtonian mass mN and the Einsteinian mass mE to be corresponding terms. That this is indeed the case we can understand from the following. We start from the philosophically unproblematic assumption that we understand the two theories in question. We then know that the terms mN and mE occur in all sorts of mathematical equations. For purely mathematical reasons we can understand that, in some cases60, E-equations unambiguously transform into N-equations when we take the limit of v/c 0 and at the same time identify mN and mE (as formal symbols) by naming them both m. We are here dealing with a mathematical continuity.61 From a physical point of view we can now say on the basis of this that in some cases the theoretical terms mN and mE formally correspond. In these cases there exists a formal as opposed to a conceptual or qua content correspondence between the terms Newtonian mass and Einsteinian mass. From now on we will note this as mN mE. In the next chapter I shall analyze this subject in greater detail with the help of the example of non-relativistic quantum mechanics and classical mechanics and electrodynamics. I hold it plausible that such formal correspondences also exist in a lot of other cases. (Although it is true that only detailed examination can decide what these correspondences exactly look like and in which domains they are valid). The above would seem to have answered the first question for the time being. I have indicated a way of finding out which terms from (possibly conceptually discontinuous) theories correspond with each other, and thus also which terms are possible candidates for co-reference. The second question remains to be answered, which is: when do terms such as mN and mE co-refer? The answer is simple: when they both fulfill the criterion of reference concerning the same reproducible material realization for a certain domain of phenomena. Whether the latter is the case we can check for instance like this. We include an experimenter A in the group of laypersons B. A will then also carry out the instructions given by A in daily terms. But in contrast with the laypersons he interprets these instructions in a scientific-theoretical manner: he will finally conclude that he has measured mN. A, however, will say when the experiment is finished that she measured mE (for v/c << 1). In the light of the fact that A and A are modern physicists who

107

know both theories well, this procedure will not raise any communicative problems whatsoever. We now have co-reference of corresponding terms from theories, which may be conceptually discontinuous, on the basis of a common material realization in a domain of experimental phenomena. Since the description of the material realization remains the same, notwithstanding the shifting theoretical interpretations, we can conclude that we are dealing with one and the same domain of phenomena. And when, moreover, certain terms from these theoretical interpretations formally correspond with each other, then we know that these terms refer to the same elements of that domain. Characteristic of this view of co-reference is that no use is made of the idea of a translation of, for instance, the concepts mN and mE through the neutral language of the material realization. Here, too, applies, just as with ordinary reference, that the theoretical descriptions of the experimenters A and A are direct interpretations of the material realization itself and not a translation of the description of this material realization.62 Thus far the notion of co-reference as implied in the criterion of reference. I would like to conclude this subject with some remarks on the concept of correspondence used above. We have shown that we can relate the concepts Newtonian mass and Einsteinian mass to each other as a case of co-reference in a certain domain (mN mE). Moreover, we can identify the formal symbols mN and mE in this case and rename them as m, i.e. as mass. It is clear that the formal correspondence mN mE and the term mass introduced here do not belong from a conceptual point of view to one of the two theoretical languages in question. They belong to, as I shall call it, the correspondence language. We can now say that the terms Newtonian mass and Einsteinian mass in some concrete cases both refer to one and the same element within a domain of reality, and that we can describe this element with the term mass. Under the next point I go into the question of what it is, to which terms such as mN and mE refer. With this correspondence language I have now introduced three kinds of language into the description of the experimental natural sciences: the daily language with which we describe the material realization; the various theoretical languages with which we scientifically interpret the material realization; and the correspondence language which contains terms and phrases expressing the possibility of co-reference. With the help of these three kinds of language I think it is possible to give an adequate description of the relations, the continuities

108

and discontinuities in the historical development of the experimental natural sciences. The correspondence language introduced here aims at taking into account and at the same time explaining a facet in the language use of natural scientists as well as that of historians and philosophers of science. After all, scientists and historians or philosophers often talk or write freely and theoretically aspecifically about mass, electrons, frequency of spectral lines, etc. The idea of a correspondence language forms, thus, an important complement to the contents of the discussions held so far. However, at this stage I have to restrict myself to briefly mentioning the contents and function of this language. In section 5.2.5 it will be analyzed it in more detail. (4) The next question I asked myself was: what do referring terms refer to? From the above follows that for example a theoretical term such as eL in concrete cases refers to an element in a domain of reality, an element which we can designate with the term electron. The term eL is taken from a specific theory of the natural sciences, which has many specific conceptual characteristics. The most important characteristics of the term Lorentz electron I have listed in section 4.3, under point (2). But the term electron is a term from the correspondence language and has thus no specific theoretical meaning. This entails that the element in reality which eL in concrete cases refers to, does not have the properties (not even roughly) which are associated with the term eL in the Lorentz theory. Something analogous to this holds true for the term eD. This is why electrons cannot be seen either as elements in reality with those properties common to the terms eL and eD. The conclusion is: the criterion of reference does tell us that a term refers, but not what the referent in itself is. On the other hand, in this way and I think this is the only way we have at our disposal a truly intertheoretical notion of co-reference, a notion which is undeniably a fundamental element of every kind of scientific realism. Precisely because of this possibility of co-reference we can philosophically say something more about the elements of reality to which referring terms refer. In this view an electron, for example, is not simply a Ding-an-sich. After all, reality is apparently thus that electrons in certain domains possess such aspects that we can describe them adequately with various terms. However, we cannot assemble all these aspects together and form a coherent image, an unambiguous representation of reality.

109

The major argument in support of this has been given in the previous section: the occurrence of conceptual breaks in the history of the natural sciences together with the no-privilege principle. Please note: as explained earlier, I do not hold that after every theory change all terms from the new theory are conceptually incompatible with those from the older theory.63 But, if we consider the history of a term over longer stretches of time, I do think it realistic to assume that generally none of the conceptual features associated with such a term are safeguarded against radical change. The example of the electron forms a good illustration of this position. The usual views on reference clearly claim more than this. Thus Putnam for instance holds that we can get to know better what the referent really is through a process of convergence towards truth. He thinks that the question of what referring terms refer to can be answered in due time, approximately with a specific theoretical description of the referent. Mature theories of the natural sciences give us an approximately correct definite description of referents.64 This is convergent realism again, although now worded with respect to the reference of theoretical terms. One often comes across this point of view, differently formulated, in recent philosophy of science. Hardin and Rosenberg phrase it like this:
reference is severed from the detailed beliefs of the theorist and its success is accorded retrospectively in the light of subsequent further approximation to the truth.65

Another example is Derksen. He speaks of a cumulative growth in terms of knowing more about the same thing, while it is necessary for a referents remaining identical that certain central properties stay intact during a change of theory.66 Specific theories present an approximately correct characterization of referents, because as follows from this view generally speaking only the non-central, the peripheral properties of referents are described differently; or because merely the details of the characterization of referents change, whereas the essence remains the same. This position seems to me untenable in general. The view for instance that since the Lorentz electron only the peripheral properties of electrons have changed, or that only the details of the description of the Bohr atom have been rejected, is an absolute caricature of the history of the natural sciences. Convergent realists will admit this more or less from time to time. However, instead of drawing the right conclusion, they often lapse

110

into another strategy, that of an implicit or explicit rejection of the noprivilege principle. Co-reference is then judged according to our present theories.67 But such a notion of co-reference is not what I called truly intertheoretical, so that it cannot supply arguments in favour of a realist interpretation of propositions of the natural sciences. (5) The criterion of reference requires that for reference the material realization must be possible in a reproducible way. We can take cognizance of the referring of a term by factually finding out whether we are dealing with reproducibility. We call this fixing the reference. To clarify my ideas about fixing the reference a bit more, I would like to oppose them to Putnams.68 He roughly defines the referent of a term as that which is in some, possibly unknown, way causally responsible for certain observable or (literally) demonstrable effects, under certain specifiable circumstances; while he thinks the reference is fixed during one well-defined, paradigmatic event, the baptizing ceremony, and subsequently passed on to others via causal, historical chains of transmission. In Putnams view it is not necessary to have at ones disposal an explicit and completely right theory about how referents produce the observable effects in question for fixing the reference for the first time. The fact that the theoretical interpretations of the term are only partly, and often only roughly right, does not hamper a successful fixing of its reference, according to Putnam. Nor does this fact hinder the reference in remaining constant during shifting theoretical descriptions of the referent (neither, in other words, the co-reference of the terms from these various descriptions). This stability of reference (or this co-reference) is, according to Putnam, due to the process of the causal, historical transmission of terms from one term-user to the next; a transmission which finally goes back to the above-mentioned baptizing ceremony of the term. What is said under point (2) about the role of theory in the criterion of reference entails that in my view the fixing of the reference of a term is also theory-dependent. If we want to find out (for the first time) whether a term refers, then we must succeed in materially realizing the experimental situations concerned in a reproducible way,

111

and, in order to achieve this, theory is, as we saw in the previous chapter, indispensable. The same holds good for co-reference, where after all the specific theoretical terms are those co-referring.69 It has escaped Putnams notice that in the experimental natural sciences we do not postulate a theoretical entity for every random effect or phenomenon, but only for phenomena or effects that are experimentally testable in a reproducible way. In order to fix the reference we need theory, but the mechanism of co-reference described above does not presuppose conceptual continuity between successive theories. This also implies that the theories used for fixing the reference may very well be scarcely plausible from our present point of view, without endangering the reference of the terms in question. The case of Thales can serve as an illustration to this point. Suppose, and this seems quite plausible that Thales used at least some vague theoretical terms when interpreting his experiments with amber, and that these experiments, interpreted in these terms, were reproducible. From the criterion of reference it then follows that Thales referred to electricity with his theoretical terms (possibly as the first human being ever), electricity being a term from the correspondence language.70 It is a different question that from a modern point of view his theory is only slightly plausible and the domain of experimental phenomena to which it refers very restricted. That this is in fact another question, is exactly an essential property of the view of reference defended here. (6) This example naturally takes me to the next and final question, which is: how is change of reference possible? In Putnams view the causal historical chains of transmission guarantee the stability of reference. His view, however, has been criticized for the reason that within such a theory it seems impossible that terms for which the reference has once been fixed later turn out not to refer at all.71 The standard example is the case of Priestleys phlogiston, a term referring to oxygen in Putnams theory, but which according to his critics should not refer at all, because they think something like phlogiston does not exist and has never existed. What about this problem in the theory of reference proposed here? Can a shift occur with respect to the reference of terms to elements in a domain of reality? The answer to this is: yes and no. Yes, for in the future the reproducibility can for whatever reason turn out to fail in the domain concerned, so that we cannot maintain the claim to reference of the term in question, in contrast to what we thought at first. In a less

112

trivial sense the answer must, however, be no. Precisely because, in my view, the various theoretical interpretations which fulfill the criterion of reference in their common domain are of equal value for this domain, it cannot be true that a newer theory simply nullifies the reference of the term in the older theory, even although we consider the newer theory more plausible. I can briefly illustrate this with the case of the term phlogiston in Priestleys theory. In this case there certainly are experiments e.g. those relating to certain combustion phenomena , which we can materially realize on the basis of this phlogiston theory in a reproducible way. Thus the term phlogiston refers to an element in the domain of the phenomena concerned. We can also interpret these same experiments from the oxygen theory, e.g. with the help of the assumption that the weight of the (vanishing) gaseous combustion products exceeds that of the oxygen combined in the combustion process. In this domain the terms phlogiston and oxygen co-refer.72 These combustion experiments are, however, not the only reproducible experiments to which the phlogiston theory can successfully be applied. Another example concerns the pouring of vitriol on zinc, when according to Priestleys theory the highly inflammable substance phlogiston is produced.73 Here, too, the term phlogiston refers to an element in the domain of phenomena in question. And again we are dealing with co-reference: we can after all describe these same experiments with our present theory in a reproducible way as the pouring of sulphuric acid on zinc, which releases the highly inflammable gas hydrogen. Consequently, in this domain phlogiston and hydrogen co-refer.74 From this example I conclude that the term phlogiston does indeed refer to an element of certain domains in an independent reality. Please note that I do not assert: phlogiston exists, in the sense that in reality there is a substance with roughly such and such properties as described in the phlogiston theory. I do assert, however, that the term phlogiston refers in a certain domain to the same element to which in that domain the term oxygen (or hydrogen) also refers. Another important conclusion we can draw from this example is that, because we relativize reference and call it reference in a domain, co-reference is not unambiguous; in one domain phlogiston and oxygen co-refer and in another phlogiston and hydrogen do. In this case we can therefore speak of partial co-reference, analogous to a term of Fields.75

113

However, this possibility of partial co-reference may still lead to the occurrence of changes in the reference of terms in a somewhat different way. In the experimental natural sciences the same theoretical term is often used in different kinds of experiment (see the phlogiston case). Since reference always concerns reference of theoretical-conceptual terms, it seems reasonable to claim that in these cases the same theoretical-conceptual term generally refers to the same class of elements (in different domains) of reality. Yet, within the views on reference presented here, this claim has to remain fallible. This becomes apparent precisely when we are confronted with partial co-reference. In that case, on the basis of the later theory, a generally referring term from the earlier theory (e.g. phlogiston) after all turns out to refer to different kinds of elements of reality in different domains. Thus, in the light of these processes of partial co-reference we may speak, in a specific sense, of progress in the development of science: later theories refer to a more differentiated and richer structured world. In view of this there certainly are good scientific grounds for preferring for instance modern chemical theory to Priestleys phlogiston theory. We can after all get to know both theories and on the basis of this knowledge observe that, firstly, modern chemical theory can be applied just as well to the experiments which the phlogiston theory can interpret theoretically in a reproducible way. And, secondly, that there are a number of reproducible experiments which modern theory can, but the phlogiston theory cannot, describe theoretically in an adequate manner. One of the consequences of this applicability to a larger domain is that modern chemical theory has more explanatory power and (for closed experimental systems) a larger predictive capacity. As I argued above (section 4.2), it is exactly this extension of a theory to a larger and more pluriform object-domain, that can increase its plausibility to an important extent.76 All this, however, does not imply that the term phlogiston now no longer refers. With regard to the domain of phenomena in which both phlogiston and for instance oxygen fulfill the criterion of reference, phlogiston still refers, not to oxygen but to the same entity to which oxygen also refers in that domain. Thus far the discussion of the questions posed at the beginning of this section. I hope that I have attained my goal: a further explanation and vindication of the criterion proposed. As we saw, this criterion makes

114

intertheoretical co-reference possible, so that the following metainduction formulated by Putnam does not apply:
just as no term used in the science of more than fifty (or whatever) years ago referred, so it will turn out that no term used now refers.77

Putnams former78 theory of reference was an attempt at blocking this meta-induction. Above it was argued that this theory of Putnam is unsound. An answer to the meta-induction formulated above is provided in this section. The problems which seem to make this meta-induction plausible have been solved by distinguishing clearly between reference and (the approximation of) truth analogous to Habermas distinction between objectivity and truth, and by moreover restricting reference to reference to elements of a domain. It is important to note that my criticism of a theory of reference such as Putnams, which is often taken as the basis of convergent realism, is of a different nature than that of, for instance, Laudans. He bases his criticism primarily on the premiss that (we now know that) terms such as phlogiston do not refer at all.79 But this premiss implies a rejection of the no-privilege principle, so that also Laudan still seems to be implicitly stuck in the realist position which he explicitly states he is opposing. In this section I proposed a theory for the reference of theoretical terms. In doing this I mentioned several times the plausibility of specific theories of the natural sciences. As we saw in 4.2 this plausibility plays a role in the criterion of the verifiable truth of propositions. In the next section this link between reference and truth will be discussed more generally, in the context of the realistic realism point of view to be formulated there.

4.5

A realistic realism

In this chapter I have aimed at discussing the question of whether, and if so to what extent, concrete scientific propositions can supply us with knowledge about an independent reality. On the basis of the criteria made explicit above, this question can now be answered as follows: A singular proposition C is about elements in a domain of the independent reality if the descriptive terms of C refer. Moreover, it holds good a fortiori that the truth of C is also sufficient for its being about reality, because the referring of the descriptive terms of C

115

follows from Cs being true.80 These two conditions of reality form the core of my realistic realism.81 The realist aspect rests mainly on the distinction between reference and (verifiable) truth. While the notion of truth is linked to the pragmatic context of specific theoretical language systems in which the propositions concerned are formulated, (co-)reference is a concept which can be intertheoretically applied. The major difference between the criterion of reference and that of verifiability is that the latter requires the laws and theories needed to be plausible while with the former this is not the case. This implies that propositions not satisfying the criterion of verifiability may still contain terms referring to elements from a domain in the independent reality. In previous sections I mentioned as possible examples of this kind of theoretical terms Thales-electricity and Priestley-phlogiston. The resulting realism has modest pretensions and thus the scientific knowledge of reality is limited. That is why I used the neutral expression is about in the reality condition formulated above: it is out of the question that true propositions describe, represent, picture or reflect reality or anything like that. In this condition every notion of correspondence is absent. The realism advocated here holds that the terms of a true proposition C do refer to an independent reality, but also that this reality cannot be simply accepted in the manner asserted by C. What really exists is e.g. an electron, which we denote with a term from the language of correspondence, namely electron. But the way in which it exists is not as a Lorentz electron or a Dirac electron. Co-reference, however, is possible. In this case we can state that the elements in a domain of reality apparently possess such aspects that we can describe or interpret them adequately, in a reproducible manner, by means of different terms. Thus reality imposes important restrictions on those propositions eligible for truth. Experimental natural science is in this sense not only a matter of obtaining your right but also of maintaining it. It goes without saying that this view is also fundamentally different from instrumentalism: after all, the theoretical terms are those referring and co-referring. If it is tenable, this realism implies in any case a refutation of idealism. I think it offers a number of possible elaborations. First of all, we can try to combine it with a philosophical ontology, as suggested in section 3.4. We can for instance interpret an electron as a thing and its mass as a property. If then the theoretical terms eL and mN, on the basis of the condition of reality, refer to the same element in a domain of reality, then apparently there is in this element a coherence between

116

thing and property. This means that with the help of terms from the language of correspondence we can formulate the proposition this electron has mass. We can then interpret this proposition as a statement about (a domain in) the independent reality. Analogous to this we could also examine other propositions, such as this watermolecule consists of hydrogen and oxygen atoms or the ordinary carbon atom has six electrons. We must, however, keep in mind that terms such as electron, watermolecule and carbon atom belong to the correspondence language and have therefore no theory-specific meaning. This elaboration of the realism proposed here is still largely programmatic. In the next chapter I carry out one of the parts of this programme, namely a detailed examination of a concrete example from the experimental natural sciences. I now have to deal with the realistic aspect of the realism proposed here. This means that it is based upon the premiss that the historical-social context of the experimental natural sciences cannot be eliminated. And, more specifically, that the occurrence of conceptual discontinuities should be taken seriously and accounted for, and that the no-privilege principle should be accepted, in philosophical interpretations of these sciences. An important advantage of such a realism is that it makes philosophical reflection relevant to the problems of the relation between science, technology and society. This relevance will be made explicit in the conclusion. The realistic point of view has led me to a criterion, based on which true propositions in the experimental natural sciences cannot be regarded in terms of picture or correspondence, but in which truth depends on plausibility, a notion with both cognitive and social aspects. Suppose for instance that (1) mN(a) = 10g and (2) mE(a) = 10g are the conclusions of correct theoretical descriptions of an experiment to determine the Newtonian and the Einsteinian mass respectively of an object a, while the relative velocity between the object and measurement apparatus is zero. Up to the present, Newtonian mechanics is still regarded as a highly plausible theory. One of the reasons we can give for this is the fact that most, if not all, factors mentioned in 4.2 as contributive to a high plausibility are present in the case of this theory. In other words: the Newtonian theory is still generally accepted as a mature scientific theory. Because of this plausibility we can say that in the example under discussion (1) is Ntrue and (2) Etrue with respect to the same domain of phenomena. The latter is the case because of the fact that the experiments to determine the truth of (1) and (2) have an identical

117

material realization. Therefore (1) is exactly as adequate as (2) with respect to this domain. Thus I regard it as not reasonable to state that (2) is true and (1) untrue, or that (2) corresponds better to reality and because of this approaches truth more than (1). At this point I agree with Heisenberg when he argues:
Today we still assign truth-value to Newtonian mechanics, yes even a rigorous and general validity, but by adding where its concepts can be applied we point out that we consider the domain of applicability of the Newtonian theory bounded.82

As shown in section 3.2 this restricting of truth to truth in a domain is not an ad hoc manoeuvre to get rid of philosophical problems, but an essential requirement for understanding the genesis and development of scientific theories. We have seen that reference, the realist aspect, is necessary but not sufficient for truth and thus neither for plausibility. Even although Thales and Priestleys theoretical terms refer, we still do not consider their theories plausible. In this respect these theories seem to differ markedly from, for instance, Newtons and Einsteins. Given the impossibility of a convergent realism this difference can, however, only be of a pragmatic nature: the domain of application of Thales and Priestleys theories is very small and limited, in comparison with that of the Newtonian or the Einsteinian theory. How can we explain this pragmatic difference, which is apparently decisive for plausibility and thus for truth? One of the possible explanations is, I think, the hypothesis that we can ascribe this difference to a technical interest of the experimental natural sciences. After all, the larger and more varied a theorys domain is, the larger and more varied its technical potential will be. Conversely, Thales and Priestleys theories, for example, are not plausible, because they can hardly be implemented from a technical point of view. It would carry us too far to work this hypothesis out in more detail here. I only want to draw a brief comparison with Habermas notion of technical interest. As has been pointed out in section 1.2 Habermas regards the cognitive interests as quasi-transcendental and universal. They play a role in all natural science, which is after all guided by instrumental or experimental action. In contrast with this I regard the technical interest as a historically changing category to be specified sociologically. In my view the technical interest is not quasi-transcendental, not universal and not relevant to all natural science. This is at once evident from the fact that this interest did not play a part in Thales or Priestleys theories. (In

118

any case Priestley should certainly be included in the modern scientific tradition). A second difference, connected with this, is that for Habermas the technical interest is only of indirect importance for the truth of propositions just as are instrumental action and objective experience. In my eyes, however, the technical interest is of direct influence on the plausibility of laws and theories and thus also on the truth of propositions. A brief comparison with recent views of Hacking and Putnam may further clarify the realism proposed here. To start with Hackings view. He first distinguishes between realism about theories and realism about entities. The first form we have already come across. Its essence is that truth or falsity of scientific theories is decided by reality, and that our best theories contain a representation of reality which is in any case approximately true. As we saw above, such a form of realism is hard to defend. Realism with regard to entities holds that, in general, the entities postulated by the theories of the natural sciences really exist. Hacking specifies this viewpoint in his criterion of reality, applied to electrons:
We are completely convinced of the reality of electrons when we regularly set out to build and often enough succeed in building new kinds of device that use various well-understood causal properties of electrons to interfere in other more hypothetical parts of nature.83

This criterion of reality proposes to draw realist consequences from standardized experimental success: the theoretical entities, which we routinely use through standardized techniques are (therefore) real. These entities are thus instruments, not for thinking as the instrumentalist assumes but for (experimental) action. Hackings criterion is, however, subject to a number of problems concerning its exact meaning and implications. First of all his attitude is not consistent towards entities of which we, now, on the basis of our present theories, hold that they do not exist. He argues that caloric as postulated in the theories of heat of e.g. Berthollet, Lavoisier, Biot and Laplace does not exist.84 But it is in fact quite simple to regard standardized heat sources as producers of caloric and subsequently to use them in experiments. We can, for example, maintain that when boiling liquids with a Bunsen-burner the liquid becomes increasingly hotter because the flame supplies it with more and more caloric. Such an

119

experiment satisfies Hackings criterion and caloric must therefore indeed exist. A second question concerns the contents of Hackings realism. Does he only claim that entities exist or does he also claim to know what their properties are? If the latter is the case he will again end up with the problematic realism in terms of representation. But if the only claim is that entities exist, then, thirdly, the problem of co-reference arises: how will we then know, for instance, that Lorentz electrons and Dirac electrons refer to the same class of entities? In the criterion of reference proposed in the previous section these problems have either been solved or avoided. If an experiment which we can theoretically describe in terms of caloric can be materially realized in a reproducible way, then the theoretical term caloric refers to an element of reality. The fact that the caloric theory is now considered untrue does not alter this at all: untrue theories or statements may also contain referring terms. Moreover, the referential realism proposed is explicitly non-representational and contains a clear criterion of coreference, which we can add is not merely applicable to theoretical terms referring to entities, but to all kinds of (descriptive) terms. Next I would like to briefly discuss Putnams recent views.85 Putnam now defends an internal realism and this is a turning point in his thinking. An important aspect of this is that reference is theoryspecific:
What I am saying is that, in a certain contextual sense, it is an a priori truth that cow refers to a determinate class of things Adopting cow talk is adopting a version, in Nelson Goodmans phrase, from within which it is a priori that the word cow refers (and, indeed, that it refers to cows).86

Putnam arrives at this conclusion on the grounds of a criticism of metaphysical realism, as he calls it, which assumes truth to be a unique relation of correspondence between theoretical propositions and reality, truth being a radically non-epistemic notion.87 On the one hand, I agree (with Putnam and against metaphysical realism) that a realism in which the truth of concrete propositions of the natural sciences is made explicit in terms of a unique relation between these propositions and reality is neither tenable nor fruitful. This has been argued extensively, primarily in section 4.3. On the other hand, Putnams self-criticism goes too far when also reference becomes solely an internal-theoretical affair. That this is in fact a consequence of internal realism can be shown as follows. On the

120

basis of internal realism we can only speak of intertheoretical coreference of two terms from the internal point of view of either theory.88 In other words we are dealing with a Tintertheoretical co-reference. But this surrogate for truly intertheoretical co-reference, as I called it, is obviously insufficient for a realist interpretation of propositions of the natural sciences. To return to the example of the electron, as considered by three theories. Suppose that the theoretical term eL refers to an element of reality, namely xL, and similarly eD to xD and the quantum electrodynamical term eQED to the element xQED. The most we can now say on the basis of an internal realist interpretation is that, from the point of view of the Dirac theory, eL actually refers to xD, or that from the quantum electrodynamical theory eD refers to xQED. We can, however, not make it plausible that (for certain domains) holds that xL = xD, or that xD = xQED, or possibly that xL = xD = xQED, although this is required for truly intertheoretical co-reference. With the criterion of reference proposed here I hope to have shown how truly intertheoretical co-reference is possible in the experimental natural sciences. On the one hand metaphysical realism is backed up by the plausible intuition (which I also assume as a starting point) of the independency of reality and thus of the idea that human beings are not simply the measure of all things. On the other hand one can also, to a certain extent, consider the discussion given in this chapter as an argumentation for an internal realism which is still realist. Putnam may claim that every theory is a symbolic representation of the world, but this the is in fact inappropriate to his view. After all, without a really intertheoretical criterion of co-reference there is no reason whatsoever to believe that the symbolic representations of various theories may refer to the same world. Thus Doormans conclusion that Putnams position in its present shape can hardly be described as realism is still too weak.89 Strictly speaking, internal realism la Putnam boils down to non-realism. It might be more sensible, however, to present this question of realism somewhat differently. I suspect that for hard-core realists a realistic realism will be too minimal. In that case it is always convenient to be able to pass this objection on to Putnam, who cherishes a viewpoint which is in any case less realist than mine but which is nevertheless still being recommended as realist by himself and others! The more so because I attach less importance to realism itself than to the philosophical position which forms its basis.

121

The referential realism defended here is grounded on the idea that the experiment, in which theory, action and materialization come together, is constitutive of the rise of the modern natural sciences (physics, chemistry, and parts of biology, geology and medical science), roughly from the 17th century up to the present. As the historian Koyr puts it:
observation and experience in the meaning of brute, common-sense observation and experience had a very small part in the edification of modern science; one could even say that they constituted the chief obstacles that it encountered on its way. It was not experience, but experiment that had nourished its growth and decided the struggle: the empiricism of modern science is not experiental; it is experimental Far from being opposed to each other, experiment and theory are bound together and mutually interdetermined, and it is with the growth of precision and refinement of theory that grow the precision and refinement of the experiments.90

The analysis given of experimentation in the natural sciences distinguishes the theoretical description from the actual carrying out, the material realization. The main goal of this chapter was to find out whether, and if so how, knowledge produced by the experimental natural sciences is about a human-independent reality. In order to be able to produce scientific knowledge there has to be a tradition of experimental natural science. I am concerned with the status of this knowledge, where such a tradition does exist. That is why the required invariance of the description of reproducible material realizations only refers to societies with such an experimental tradition. The extent to which elements or basic features of this tradition occur outside these societies, is not important in this context. It does, however, seem fairly obvious that a group of laypersons partly originating from a magical and partly from a scientific society will have great difficulty in reaching a consensus about a description of a certain material realization. An important characteristic of the philosophical interpretation of the experimental natural sciences explained here is its non-monolithic nature. Scientific practice forms a complex body, in which theoreticalconceptual work, experimental action and materializing, and more formal-mathematical activities play a part. These three kinds of activities fulfill essentially different functions in the experimental natural sciences. This is shown most clearly, for instance, by the fact that theoretical propositions are not translatable into the language of the material realization, but that they are direct interpretations of this material realization itself. This implies that the non-linguistic level of action and materializing has its own defined place in the experimental natural sciences, and that it should philosophically be taken into account

122

as such. Science is not simply a verbal practice, as so many modern philosophers of science who have experienced the linguistic turn in philosophy have tried to make us believe.

123

Specification and application: two case studies from the history and philosophy of quantum mechanics

5.1

Introductory remarks

In this chapter I would like to work out two cases from modern physics, which may serve to throw more light on the philosophical discourse of the previous chapter. The cases do, however, vary in kind. First, I discuss a number of historical aspects of the genesis of quantum mechanics, stressing the part played by the correspondence principle. Thus I intend to clarify the philosophical argumentation given up to now by illustration with a concrete historical case. Then, in section 5.3, I go into the so-called measurement problem in quantum mechanics. This problem is viewed as an example of an application of the philosophical arguments: here I intend to demonstrate that the philosophical conclusions obtained allow another view to the measurement problem. The main result shows this problem to be much less problematic than has been suggested in a part of the literature dealing with it.

5.2

The correspondence principle and the historical development of quantum mechanics

The realistic realism, presented in the previous chapter, is based on two fundamental postulates. The first one holds that, because of the regular occurrence of conceptual discontinuities combined with the noprivilege principle, an intertheoretically applicable notion of truth as representation, reflection etc. is untenable. The second one is that even so a relation of co-reference between terms from conceptually discontinuous theories is possible. Here the distinction between the

124

theoretical description and the material realization of experiments plays a crucial part. In order to elucidate and test these two basic premisses I shall analyze the history of the correspondence principle in this section. Its relevance to the philosophical argumentation will become evident with the two major views on correspondence. On the one hand, application of the principle could induce a certain conceptual continuity between the classical physical theories, Bohrs atomic theory, and modern quantum mechanics. On the other hand, we could be dealing with a correspondence more in the sense of a numerical agreement between the values of some of the quantities in the theories concerned. This agreement is therefore restricted to a certain domain of experimental phenomena, and it is obvious that in this view correspondence can very well be linked to the concepts material realization and co-reference as I use them. It is certainly true that both aspects have played a part in the historical development of the correspondence principle. This section, however, aims at a precise analysis of this development in order to arrive at an evaluation of the two aspects. The conclusion will be that the development as analyzed supports the philosophical position of the previous chapter. The sections 5.2.1 up to and including 5.2.3 will show that that particular understanding of the correspondence principle which sought for conceptual continuity was after all untenable (even though it proved temporarily to be of great heuristic value). What was tenable, however, was the idea of correspondence as numerical agreement of the values of quantities in a domain of experimental phenomena. In section 5.2.4 I link up this kind of correspondence with my realistic realism, through the notion of material realization. Finally I draw some philosophical conclusions in the last section (5.2.5), primarily with respect to the meaning and function of the correspondence language, which was discussed only very briefly in chapter 4.1 It so turns out that in this language intertheoretical relations can be formed which in themselves are of great importance to the progress of science but that these relations are of a mathematical or formal nature, so that they cannot represent a conceptual intertheoretical continuity. The at first sight paradoxical conclusion of this section is that indeed the history of the correspondence principle demonstrates how wide the theoretical-conceptual gap between the theories at issue is. It is exactly this fact that forms an insurmountable problem when one would like to come to a more ambitious form of scientific realism than defended here.

125

In this section I certainly do not intend to write a complete history of the development of the correspondence principle. This would be impossible and unnecessary within the context of this philosophical study. I only want to draw a number of major lines in its development and to learn a philosophical lesson from this. Detailed historical research has been done on the history of the quantum theory between 1913 and 1925. I will therefore make extensive use of this in order to support the philosophical lesson to be learnt with as much and as sound historical evidence as is possible. Before starting I must mention that initially the terms correspondence principle and correspondence are used in a broad sense. While these terms were first introduced by Bohr in 1920, there are earlier instances of correspondence-like arguments (Bohr then often speaks of an analogy between the quantum theory and the classical theories). The intention is to analyze the development of all these kinds of argument. In the analysis I shall especially examine how such arguments were used in concrete physical research.2 Between about 1913 and 1923 this will mainly concern Bohrs work. He was, after all, not only the originator of the correspondence principle, but also the leading man in the field of atomic theory, or the old quantum theory, in those days. In the decisive period, between 1923 and 1925, Bohr, however, fell a bit into the background (relatively spoken). I will then focus more on the role played by correspondence in the work of mainly Kramers, Born and Heisenberg. Generally, the correspondence principle is studied not so much for its heuristic value within the (old) quantum theory, but rather as a principle which leads to intertheoretical relations between the various theories involved. For claritys sake one should notice that in these developments we are dealing with three kinds of theories: the classical (electrodynamics and mechanics), Bohrs atomic theory, and modern quantum theory in the form of Heisenbergs matrix mechanics.

5.2.1

Bohr 1913: correspondence as numerical agreement

In the development of the correspondence principle three stages are to be distinguished, which I discuss in this and the next two sections. The first takes place roughly between 1913 and 1915 and concerns correspondence as a numerical agreement of values of certain physical quantities in classical electrodynamics and mechanics and in Bohrs atomic theory, for a specific domain of phenomena. In this period we come across the first correspondence-like argument in Bohrs famous trilogy On the

126

Constitution of Atoms and Molecules of 1913.3 Here Bohr assembled a number of elements from physical research at that time and worked them into a provisional but extremely fruitful quantum theory of the atom. These elements were: the atomic model of Rutherford, notions from the quantum theory of Planck and Einstein, and the combination laws for atomic line spectra based on experimental data.4 In Rutherfords model an atom consists of a small, positively charged nucleus around which a number of electrons revolve, each with a negative electrical charge e. The number of electrons varies with the elements in the periodic table, in such a way that for each element the electrons exactly compensate the nucleus charge, so that the atom as a whole is always electrically neutral. The simplest example is hydrogen: this consists of a nucleus with charge +e and one electron with charge e, tracing elliptical (Kepler) orbits around the nucleus. In this model an atom resembles a miniature solar system. Bohr made this model more precise and extended it with two important postulates, which also introduce quantum notions into the atomic model5: (a) Not all electronic orbits are physically possible, but only a limited number, the so-called discrete stationary states each corresponding to a well-determined energy. The dynamic equilibrium of the electrons in these states can be described by classical (particle) mechanics. (b) Contrary to what one would expect on the basis of classical electrodynamics, the electrons do not continuously emit radiation during their motion in a stationary state. Emission of radiation only takes place when the electron jumps from one stationary state into another. The frequency of the radiation emitted during a transition from state 2 into state 1 is proportional to the difference in energy between the states and is given by Bohrs frequency rule: E2 E1 = hv1 2 (1)

(E2 > E1; h is Plancks constant). Classically we would, however, expect a radiation frequency in state 2, equal to the frequency of the motion of the electron in that state. Subsequently, Bohr linked up this atomic model with the spectra, in the first place with that of hydrogen. The lines of these spectra proved to allow a general ordering in series of term differences. Thus the Balmer series for hydrogen can be described as:

127

2 v n = Rc(1/22 1/n2)

(2)

2 ( v n is the frequency of the light; R is the so-called Rydberg constant; and

n is an integer larger than 2). When we multiply (2) by h and compare the result with (1), we see that we can interpret the Balmer formula in terms of the differences in energy between stationary states. To connect (1) and (2) also theoretically, Bohr goes to work as follows. Classically the electron (with mass m) was thought to move in an elliptical Kepler orbit with the nucleus in one of the focuses. The relation between the binding energy E and the frequency of the motion of that orbit is given by: = (2/m)(E3)/e2 (3)

To derive the Balmer formula theoretically Bohr needs a quantum condition, which tells which of the classically possible orbits are quantum-theoretically permitted. In part I of his trilogy he initially made use of an analogy with Plancks quantized oscillators. This analogy was, however, rather ad hoc and could not really serve as a fundamental assumption in his theory. That is why Bohr quickly changed his strategy. He no longer tried to derive the Balmer formula, but instead used it as a premiss for a theoretical calculation of the Rydberg constant.6 For this calculation he explicitly uses correspondence arguments and not the analogy with Plancks theory. The only theoretical quantum notion is now in his frequency rule, equation (1).7 The argumentation is quite simple: the Balmer formula, used as an empirical quantum condition, supplies En = Rhc/n2 for the energy of the n-th stationary state. With the help of (3) follows that the classical frequency of the motion in that state is: (n)2 = 2R3h3c3/2me4n6
n 1

(4)

It will be clear that n will generally be quite different from the quantum frequency v n . But, Bohr argues:
On one point, however, we may expect a connection with the ordinary conceptions; namely that it will be possible to calculate the emission of slow electromagnetic oscillations on the basis of classical electrodynamics.8

128

For small v n or, what in this case boils down to the same thing, for large n two successive stationary states differ only very little in energy. That is why Bohr expects that in this domain the value of the quantum frequency and that of the classical frequency will be approximately equal:
n vn 1 = Rc(1/(n1)2 1/n2) 2Rc/n3 = n

n 1

(5)

Filling out n from (4) will then supply the theoretical value of R: R = 22e4m/ch3. Rs value calculated thus turns out to correspond very well with the value we can derive directly from the spectral series. Moreover, Bohr at once extends this agreement between the frequencies n and v n
n 1

in an important way. Classically a system

moving with frequency n does not only emit radiation of that frequency, but also higher harmonics with frequencies n ( = 1,2,). Bohr, now, makes it plausible that also for n >> 1 and << n the correspondence
n v n = n is valid. (For later usage I note down the harmonic component

n as n,). The quantum frequency resulting from a transition between states n and n agrees in the domain concerned with the frequency of the -th harmonic component of the radiation emitted by a classical electron in the n-th stationary state.9 It is this more general form of correspondence which is to play an important role in the next stages of the development. Bohrs correspondence assumption in this first stage is that:
the frequency of the radiation emitted during the passing of the system between successive stationary states will coincide with the frequency of revolution of the electron in the region of slow vibrations.10

To justify this assumption he refers to Lorentz and Jeans calculations with respect to heat radiation. On the one hand their classical calculations tally quite well in the domain of slow vibrations. On the other hand Planck had shown as early as in 1906 that his radiation law for 0 passed into the classical law of Rayleigh-Jeans.11 Remarkably, Bohr always speaks of agreement of calculations. The correspondence only concerns the numerical results of calculations of the values of quantities (in a certain domain). For, despite the existence of this analogy between classical electrodynamics and quantum theory, Bohr was very well aware of the conflict between both theories.12 To mention a number of important differences: the mecha-

129

nical and electrodynamic stability of the atomic model used is completely incomprehensible from a classical point of view. Furthermore, the radiation mechanism in both theories is essentially different: according to the classical theory the basic frequency and the higher harmonics are emitted in one atomic process, while in Bohrs theory every quantum frequency originates from a separate atomic transition. And finally, the correspondence between n, and v n
n

holds only good in a limited

domain: for small quantum numbers the numerical values of both quantities differ radically. 5.2.2 Correspondence and conceptual continuity: 19161922

In this second period the theory about atoms and their line spectra becomes more profound and is extended in two important ways. This simultaneously entails a change in use and status of the correspondence principle. First of all: we have seen that Bohr had not as yet succeeded in completely deriving the Balmer formula in 1913. He lacked plausible theoretical quantum conditions with the help of which the permitted stationary quantum states could be selected from all electron orbits possible according to classical mechanics. Bohrs theory of 1913 did, however, contain a first step towards such quantum conditions.13 In 1915 it was, among others, Sommerfeld who formulated general quantum conditions which had to be met by the mechanical orbits calculated in the classical way.14 Moreover, the powerful Hamilton formalism from classical mechanics with which to compute the classically possible orbits became available after Schwarzschilds introduction of the so-called action-angle co-ordinates. The method is, in brief, this: in the general case of a physical system with s degrees of freedom we can describe the motion with the help of generalized coordinates and momenta, qk and pk (k=1,2,,s). As Schwarzschild showed, the generalized action-angle co-ordinates (Jk,wk) are better suited to the singly or multiply periodic systems, with which we often have to do in the case of atoms. The wks can be chosen as linear functions of the time (wk = kt + k), while the Jks are related to the qks via the transformation: + qk = 1s = Ck1s exp [2i(1w1 + ... + sws)] (6)

130

The Ccoefficients in (6) depend only on the action co-ordinates Jk. Sommerfelds quantum conditions, now, hold that for such a multiply periodic system the Jks which describe a permitted stationary quantum state can be selected from the classically possible states with the help of the relations: Jk = nkh (7)

(the nks are positive integers). With this in any case the theoretical tools had been created for a more general atomic theory which would not have to be restricted to elliptical orbits in hydrogen and hydrogen-like atoms, as Bohrs in 1913 had had to be. Two important results in which more complex atomic motions were related to detailed characteristics of the spectra, constituted Sommerfelds explanation of the fine structure of the frequencies of the hydrogen lines, as caused by the relativistic corrections of the classical Kepler motion; and, subsequently, the theoretical explanations of the Stark effect by Epstein and Schwarzschild.15 (This Stark effect is the splitting up of the hydrogen lines into various components under influence of a homogeneous electric field). The reason why I discuss this development at some length is that Bohr combined it with his ideas of 1913, and especially with those about correspondence. In his important article On the Quantum Theory of Line Spectra Bohr again treated the correspondence (for large quantum numbers) between the quantum frequency of the light emitted and the classical frequency which is associated with the motion of the electrons in their orbit, but now in general for multiply periodic systems.16 This correspondence between the frequencies and could now be proved also theoretically, contrary to the more empirical argumentation of 1913, thanks to the introduction of adequate quantum conditions. Bohr does, however, stress again that the radiation mechanisms in the two theories are very different, despite this correspondence for large quantum numbers. According to classical electrodynamics the basic frequency and all higher harmonics are emitted simultaneously in one process; while, according to Bohrs atomic theory, a separate transition is needed for each quantum frequency v n
n

(for each value of ). Since the

transitions in the case of one atom certainly do not take place simultaneously, the general correspondence between and is only statistically valid, in other words, only when we are dealing with a great number of atoms, so that all the kinds of transition contribute to the

131

radiation emitted as a whole. Thus the difference with the classical conception is clear. But what the radiation mechanism exactly looks like for individual atomic transitions, about that, Bohr stresses repeatedly, his quantum theory cannot as yet say anything. A second important theoretical step forward in this stage concerns the intensities and polarizations of the spectral lines. So far Bohrs atomic theory was clearly incomplete, and, from a conceptual angle, inferior to the classical one. After all, the light emitted by atoms has not merely a frequency, but also an intensity and in many cases it is polarized in certain directions. In classical electrodynamics these characteristics of the emitted light can generally be calculated on the basis of the exact state of motion of the radiating particle; the only problem was that the results of these calculations proved to be wrong in the domain of the atomic spectra. In the years 1917 and 1918 Bohr, however, succeeded in fitting in intensities and polarizations, too, into his theory, largely with the help of an extension of the correspondence principle. His idea was to also formulate a correspondence referring to the intensities and polarizations in both theories, analogous to the correspondence between the frequencies. If one, however, wants to do justice to the quantum character of the stationary states, it is not possible to relate the intensity of the light directly to the values of the coefficients Ck from (6), which classically determine the intensity. Analogous to the energy in (1) here, too, a quantity is needed for linking two different stationary states (the initial and the final state of the radiation process). For this quantity Bohr chose the transition probability introduced into the theory by Einstein in 1916. We define A2 as the a priori probability of an electron to change from state 2 into state 1. Bohr, now, suggested the existence of a correspondence for large quantum numbers between the coefficients Ck and these probabilities of spontaneous transitions. Moreover, he expected calculations based on such a correspondence to give reliable estimates of the intensities and polarizations also for smaller quantum numbers and this is an important point which I discuss fully later on.17 In Kramers dissertation these ideas of Bohrs were worked out in greater detail and applied in concrete cases to the problem of the intensities. For large quantum numbers the correspondence is, according to Kramers18:
n An = gCn, 2 (n,)3/h

(8)

132

(For the sake of perspicuity I use a somewhat different notation from now on. The index k of (6) is left out and the classical Ccoefficients of the same equation are noted in analogy to the classical frequencies n,). Following Bohr, Kramers expects to be able to give reasonable estimates of the transition probabilities for smaller n, too, with the help of this correspondence. On the basis of (8) he calculates the intensities of the spectral lines for a number of cases and compares these with Starks and Paschens experimental results. Generally speaking the correspondence is certainly not exact, but nevertheless he judges it to be suggestive and convincing.19 On the basis of these successes on the one hand, the introduction of quantum conditions and the methods for determining the stationary states, and on the other, the proven possibility for theoretically incorporating intensities and polarizations and obtaining a reasonable amount of agreement with the experimental data Bohr was fairly optimistic about the future possibilities of his theory in these years (1917 and 1918):
It seems now really possible to a certain degree to overlook the theory with all its different applications from a uniform point of view. At the present I am myself most optimistic as regards the future of the theory.20

This entire development went hand in hand with a quite different interpretation of the correspondence principle, an interpretation which can generally be characterized by the view that, disregarding large differences, important conceptual similarities exist between Bohrs quantum theory and the classical theories.21 To elucidate this I discuss in some more detail Bohrs argumentation and its elaboration by Kramers. Bohr argues that the correspondence between the transition probabilities and the coefficients of the Fourier development of the motion may clearly be expected to be of a general nature. Although, of course, we cannot without a detailed theory of the mechanism of transition obtain an exact calculation of the probabilities, unless n is large, we may expect that also for small values of n the amplitude of the harmonic vibrations will in some way give a measure for the probability of a transition 22 Kramers specifies this statement as follows. We can express the quantum frequency v2 corresponding to the transition of state 2 to state 1 exactly, also for small quantum numbers, as an average over all classically possible states with energies between E2 and E1.23
1

133

Consequently, in order to carry out an exact calculation with respect to the intensities, the quantity Cn, of (8) should really be replaced by a suitable average to be chosen, C n, , for the Cn,s of all mechanically possible states between E2 and E1. However: on the one hand there is no theory about the radiation mechanism which could tell us the correct way of averaging; and on the other, probing obvious methods of obtaining an average would lead to (too) laborious calculations. For these reasons Kramers renounces such calculations and continues to work with the original correspondence (8).24 So: on the one side, one assumes at this stage the hypothesis that correspondence in itself is generally valid for all quantum numbers; on the other side, due to the ignorance of the radiation mechanism it is still impossible to make exact calculations. The confidence in this extended notion of correspondence stretches so far that Bohr even speaks of a general law in 1923:
this circumstance leads us to consider the connection between the harmonic components of the motion and the various processes of transition traced in the region of large quantum numbers, as evidence for a general law holding for all quantum numbers. According to this law, the so-called correspondence principle25

Summarizing: in this second stage the correspondence principle holds that a far-reaching analogy exists between the quantum theory and the classical theories. This analogy means that in quantum theory, just as in classical theory, there is an intimate connection between the Fourier components of the motion of the electrons and the properties of the radiation emitted by these electrons and that this connection applies to all quantum numbers. The quantum theory is a rational generalization of the classical theories, Bohr argues. It is a generalization because the new concept transition probability has been introduced; and this generalization is rational because the relation between motion and radiation is analogous to that in classical electrodynamics. As yet, however, nothing is known about the quantum radiation mechanism (except that it is essentially different from the classical one). This is why we still do not know the exact form of the, in itself unambiguous, correspondence between An
n

and Cn,. In con-

sequence accurate numerical results corresponding to the experiments are still not obtainable. But: it is hoped and expected that a more exact calculation of the transition probabilities, also in the field of small quantum numbers, will be achieved following these theoretical lines, in the future.26

134

The shifting of the views about correspondence is clearly shown by the difference between these two formulas27:
n An = gCn,2 (n,)3 /h, for large n and << n n An = g Cn, 2 ( n, )3 /h, for all n and

(8a) (8b)

The first formula can be used for calculating the numerical values of the intensities in the domain of large quantum numbers. The classical and quantum theoretical radiation conceptions remain, however, essentially different. Formula (8b) is an attempt to bridge this conceptual gap as best as possible. The quantities Cn, and n, form the bridge between classical and quantum theoretical discussions. The only problem is yet to find the correct way of averaging the Fourier components and the frequencies of the classical orbits situated between n and n. Seen in this light these quantities Cn, and n, represent an element of conceptual continuity between the two theories. And indeed Bohrs entire atomic theory (the old quantum theory) rests on the assumption that the description of classical mechanics remains valid for the stationary states, as real orbits which electrons trace. Once we would know exactly in which (mechanical) way an electron changes from one orbit to the other, we would also know how to average for the orbits in between, so that exact calculations of intensity and polarization would be possible. (Of course there would remain a difference between classical electrodynamics and Bohrs atomic theory due to the discrete character of the stationary states). The conceptual continuity mentioned above would have been philosophically important if it could have proved tenable in the later developments. To maintain the link with the philosophical discussion I will briefly explain this in the light of the exposition about reference in section 4.4 (and focus on point 5). Within the discussions about correspondence in this second stage we can fix the reference of, for example, the term hydrogen atom not only by giving, la Putnam, a specification of certain observable effects (such as the properties of the hydrogen spectrum), but also by pointing out, la Nola, Leplin and Derksen, its theoretical central properties (such as the specific form and the frequency of the motion of the electron orbits) which determine this spectrum. Next one might observe that even though the classical theory is empirically less successful than its successor, there is conceptual continuity in the sense of the preservation of an important central property of hydrogen atoms: the electrons trace

135

orbits which can be described in both theories in the same way, that is with the help of the Hamilton formalism of classical mechanics. Moreover, these mechanical orbits form the basis on which the radiation processes can be explained, albeit that the exact formulas in the classical theories differ from those in the quantum theory (cf. (8a) versus (8b)). In this view these shared theoretical properties (the quantities Cn, and n,) justify the conclusion that in both of the theoretical interpretations of the intensities and polarizations of the spectral lines we are dealing with the same hydrogen atoms (co-reference). Unfortunately this argumentation is erroneous. The strategy of this second stage proved far less successful than expected, and the second was followed by a third (and last) stage.28 In this last stage the transition of Bohrs atomic theory to Heisenbergs matrix mechanics took place, from the old to the modern quantum theory. In that stage, too, the correspondence principle played an important, though changed, role. The reason for this change is, first of all, that the description of the stationary states with the help of the classical Hamilton formalism was increasingly regarded as generally impossible; while, secondly, even the idea of a kinematics of the electrons in terms of orbits was abandoned. In the next section I try to find out what correspondence in this third stage did signify.

5.2.3

Numerical and formal correspondence: 19231925

First of all I would like to discuss briefly several aspects in the development of the quantum theory during these years. I shall (again) restrict myself to those aspects which are important for studying the role of the correspondence principle, so that the larger context will only be treated in rough outline. As we have seen, the correspondence principle in its second stage was linked to Bohrs atomic model with its mechanical orbits which could be described as multiply periodic motions.29 In the course of 1922 and 1923, however, the foremost atomic physicists (Bohr, Kramers, Heisenberg, Born, Pauli, Van Vleck) had reached the conclusion that application of classical mechanics to the problem of stationary states was out of the question in the case of atoms with more than one electron (which display highly complicated spectra) and in the case of interaction between radiation and atoms (fluorescence, dispersion, and so on).30 Especially the failure of classical mechanics in the simplest case of a many electron system, the helium atom with two electrons (treated by

136

Kramers, Van Vleck, and Born and Heisenberg), contributed a lot to the plausibility of that conclusion. This state of affairs certainly increased the readiness to explicitly work with other atomic conceptions deviating from the Bohr model. In this respect the so-called virtual field model, which was introduced in the well-known Bohr-Kramers-Slater theory, was very important.31 This model was successfully applied by Kramers, and later on by Heisenberg, to the problem of the interaction between radiation and matter. In this virtual field model an atom, in a certain stationary state, is described by a number of virtual harmonic oscillators. These oscillators generate a virtual radiation field which (a) contains all frequencies meeting Bohrs frequency rule (eq. (1)) and of which (b) the amplitudes
n n C n determine the values of the transition probabilities An . This new

atomic model, which was introduced at first as a formal description for the interaction processes between radiation and matter, has a number of important properties I would like to draw attention to. First of all, the virtual field is present and effective during the entire life time of the stationary state in which the atom is and not only during the relatively brief period in which transitions take place. This fact makes it possible to conceive of the model as a full atomic model. Secondly, the field is virtual because it does not carry energy. The meaning of virtual remains somewhat opaque, but Jammer interprets this concept in my opinion quite adequately as effective but not factual.32 Thirdly, this virtual field is really a probability field.33 The field amplitudes do not determine the transitions in themselves but the probabilities of their occurring, and this in a fundamentally different way than in e.g. (8b). For fourthly, the model contains the essentially new quantities
n C n . These are the transition or probability amplitudes, which are

taken as necessarily dependent on two stationary states. This entails that the attempts of the second stage to link up more or less directly the Fourier components and frequencies of the mechanical motion ( C n , and n,) with the transition probabilities An quantities C n , and C n
n n

are now abandoned. There

does remain, as I show below, a kind of correspondence between the , but this is of another nature than that of the second stage. In fact, a similar radical step has been taken in this virtual field model with reference to the quantities Cn, as Bohr had taken for

137

the frequencies of the motion n, already in 1913. From the fact that these frequencies for small quantum numbers did not yield an adequate description, Bohr concluded that they should be replaced by the quantum frequencies v n amplitudes An
n

. In the virtual field model the Fourier components of . In Borns words:

the motion Cn, are now replaced analogously by the transition


n

We knew that frequencies of vibrations were proportioned to energy differences between two stationary states. It slowly became clear that this was the main feature of the new mechanics: each physical quantity depends on two stationary states, not on one orbit as in classical mechanics. To find the laws for these transition quantities was the problem.34

Born himself took the next important step towards the new mechanics. Now that direct, theoretical calculations of the stationary states with the help of classical mechanics had proved insufficient, attempts were made to find quantization rules directly in terms of quantum numbers guided by empirical and correspondence-like arguments. These numbers were now no longer thought of as characteristics of mechanical orbits but as formal properties of quantum states. In 1924 Born succeeded thus in discovering a generally applicable quantization rule, implying replacement of certain classical differentials by quantum theoretical differences, by differences of functions in two stationary states.35 This quantization rule constituted a generalization of Kramers dispersion theory and was also supported by Heisenbergs work on the anomalous Zeeman effect.36 Moreover, the rule satisfied the correspondence principle, in the sense that for large quantum numbers the classical and the quantum theoretical formulas become identical. Hendry adequately summarized this approach as follows:
given the failure of classical mechanics and the failure of the orbital model within this mechanics , the success of the [virtual] oscillator approach to dispersion suggested a search for new quantum conditions based upon this approach, interpreted in terms of difference equations.37

Finally these developments culminated in Heisenbergs quantum theoretical re-interpretation of classical mechanics and kinematics in the summer of 1925.38 In his article, which in fact contained the core of matrix mechanics, Heisenberg drew the consequences of the developments sketched up to now. He based his discussion on the virtual field model, with its probabilistic traits, and on Borns general quantization rule in which the atomic quantities are always conceived as differences

138

of functions of two stationary states. At the same time he added something new. For he radicalised these ideas by not only applying them to mechanical, but also to kinematic quantities, such as position and velocity. Concerning the latter he was primarily influenced by Pauli, who wrote for instance to Bohr in December 1924:
The relativistic doublet formula appears to me to show unquestionably that not only the dynamic concept of force but also the kinematic concept of motion of the classical theory shall have to undergo fundamental changes. (It is for that reason that I have avoided entirely in my work the designation orbit).39

It was the integration of the above-mentioned elements in Heisenbergs matrix mechanics which brought about the end of the old quantum theory, of Bohrs atomic theory. After this brief outline of the relevant developments in physics during the period in question, we should now treat the problem of meaning and role of the correspondence principle. As in the rest of this section I shall not try to offer a consistent reconstruction of all Bohrs statements about the principle, but rather examine how it worked in the developments sketched above.40 In the light of the fact that the correspondence principle, in the second stage, depended on the mechanical atomic model with its classical orbits, and also in the light of the developments in the third stage implying a replacement of classical mechanics and kinematics, it will be clear that there has to exist a difference in meaning and function of the correspondence principle between these two stages.41 Above I briefly touched on the relation between the virtual field model and the correspondence principle, which results in our having the following correspondence:
n C n, = C n , for n >> 1 and << n

(9)
n

While then for the purely quantum theoretical relation between C n and An
n

the following holds true: (10)

n n n An = g C n 2 ( v n )3/h, for all n and

The essential difference between the second and the third stage becomes obvious when comparing (8b) with (10). Generally speaking the quantum theoretical transition amplitudes C n
n

and frequencies v n

have no

longer anything to do with the form and frequency of motion of the

139

electron orbits. This means that these C n as the C n, and

and v n

cannot connect

classical conceptions from the Bohr model with the new quantum theory,

n , in (8b) did. A definitive conceptual discontinuity has n , and C n, are permanently replaced by
and C n
n

arisen. Bohrs initial optimism from the years 1917 to 1922 has been confuted. The situation with respect to the correspondence principle is this: the classical quantities the quantum quantities v n
n

, which are based on the radically

unclassical atomic model of the virtual probability field but whose numerical values for large quantum numbers agree with those of the corresponding classical quantities. So, from a methodic angle, we are back in the first stage. A most important consequence of all this is that in this third stage the values of the An
n

can no longer be calculated theoretically ab initio

with the help of the Hamilton formalism, while on the basis of equation (8b) this still seemed possible. The only other possibility is to take these quantities as experimentally definable parameters, as purely formal symbols. This is precisely what happened in, e.g., Kramers dispersion theory and in Heisenbergs matrix mechanics, both grounded on the virtual field model.42 I would like to briefly digress on the dispersion theory. The point is that, in fact, when Kramers articles were written, the experimental data to test his theory in its general form were not yet available. Thus neither a theoretical deduction nor an experimental proof of his dispersion formula were possible. When he finds himself in that predicament Kramers resorts to, apart from the correspondence principle, yet another legitimation, in the form of a positivist epistemology. He proposes that the fact that the dispersion theory only contains experimentally definable (observable) quantities in itself is reason enough to believe in its correctness.43 Elsewhere I have examined, following Formans analysis, the possible presence of social determination in the use of these positivist epistemological argumentations, which we come across not only in the case of Kramers but also in that of other leading physicists of the Weimar period (Pauli, Heisenberg, Bohr, Born).44 In section 4.2 it was argued that the plausibility of general laws or theories may depend on social factors. In this case the plausibility of (predecessors of) matrix mechanics seems to be partially dependent on general characteristics of the intellectual climate in the cultural sphere of the Weimar Republic.

140

In Borns work the above-mentioned correspondence-assumption of the virtual field model (eq. (9)) is further analyzed and generalized. In the classical formulas (which we want to quantize) we do not only come across quantities such as Cn, but also their differentials, and especially differentials to the Jks. Borns contribution refers to the quantization of this kind of differential. I will now go into some of its more important aspects.45 Suppose we have an atomic system with a number of classical electron orbits, of which the action co-ordinates are given by: J = h(n + ), 0 1 and > 0 (11)

The classical frequency of the -th orbit with Hamiltonian H is then: , = (H/J) = (dH/d)/h (12)

Quantum theoretically, however, only discrete stationary states (=0 and =1) are permitted. The frequency resulting from a transition between these two states is:
n vn + = (Hn+ Hn)/h

(13)

From (12) and (13) follows a relation between the classical and the quantum theoretical frequency:

n + n

=
0

(14)

The step which Born then takes is a generalization to arbitrary (classical and quantum theoretical) quantities 46:
+1

( / J ) d
,

n n = ( n+ n )/h

(15)

From (15) follows Borns quantization rule, which applies to all quantum numbers. This rule holds that we can obtain correct quantum mechanical descriptions by replacing classical formulas in the form of the integrand in the left side of (15) with the quantum theoretical differences of the right side. Born advances two main arguments to support his quantization rule.47 First of all he applies it to the problem of dispersion and shows that the result is exactly Kramers dispersion formula (which was also thought to be applicable to all quantum numbers). Secondly, he

141

examines the limit for large quantum numbers. In that case the states n and n+ are very close to each other. , is then virtually independent of and almost identical with n,. Therefore follows from (15): (n,/n) = ( n+ n ), for n >> 1 and << n
n n

(16)

When we now regard the left-hand side of (16) as a classical quantity, say Xcl(n,), and the right as a quantum theoretical one, say Xqu(n,), then we see that Xcl and Xqu, in the domain of large quantum numbers are the same mathematical functions of n and : Xcl(n,) = Xqu(n,), for n >> 1 and << n (17)

So: in (17) we encounter again the well-known case of correspondence as numerical agreement of the values of quantities from two theories in a certain domain of phenomena. The novelty of Borns approach should, however, not be sought for in this numerical correspondence, but in his equation (15), which applies to all n and . This rule implies another kind of correspondence, to which I would like to draw attention. The left side of (15) is a purely classical expression. , is some quantity (e.g. C, or ,) which refers to a classical electron orbit, with being a parameter allowing a continuous averaging or integration of all mechanical states between n and n+. In contrast with this the righthand side should be interpreted in terms of the virtual field model, in which generally all quantities depend on two different stationary states. I would like to point out two important aspects of (15). The first one is the conceptual discontinuity between the interpretation of the left side and the right side, originating in two divergent atomic conceptions. In this context it is instructive to look back once more at equation (8b) from the second stage. As we saw (8b) represented the hope to derive the right transition probabilities and thus the intensities and polarizations of the radiation through a suitable averaging process over the Fourier components and the frequencies of the motion, within the context of the classical orbits in Bohrs atomic model. In equation (15) a similar averaging is carried out via integrating over . Nevertheless, the interpretation of (15) is totally different from that of (8b). Since the n are based on the virtual field model a possible n reduction to classical quantities n, is excluded and the conceptual discontinuity continues to exist. Born puts it literally like this:

142

the present approach looks upon the virtual resonators as the real primary thing and upon the classical calculation methods only as a means to a rational tracking down of the true laws of quantum mechanics.48

Secondly: as said before, Borns formula (15) contains quantities from two different theories. It is an intertheoretical relation which, however, is of a formal or mathematical character in view of the conceptual discontinuity between these theories. The equality sign in (15) does not imply a conceptual relation between the quantities on the left and the right side. In other words: there is a correspondence between terms of type , and that of type n , but this correspondence is n formal. Besides the numerical correspondence, such as C n , = C nn for large quantum numbers, we also have a formal correspondence holding good for all quantum numbers in those cases in which the intertheoretical relation (15) is valid. Thus equation (15) makes it possible to decide which specific terms from the various theories (formally) correspond with each other. As we saw in section 4.4 we have to know what is exactly meant by corresponding terms in order to be actually able to ask whether a possible co-reference of terms from various theories exists. In section 5.2.5 I shall return to this philosophical problem, on the basis of the example discussed here. Summarizing we can say that this third stage proved the optimism of the second stage in the development of the correspondence principle to be unwarranted: a partial conceptual continuity between quantum theory and classical mechanics, grounded on the part played by the Fourier components and the frequencies of the motion in both theories, turned out to be impracticable. That is the reason why, on the one hand, a return in the third stage to the first stage was made: correspondence means again numerical agreement of the values of classical and quantum quantities in a certain domain (with these quantities themselves, however, being conceptually quite different). Nevertheless the correspondence principle proved to be a fruitful heuristic aid also in the second stage. This interpretation is supported by a quotation from Bohr, dating from 1926, in which he looks back on the entire development:
at the present stage of the development of the quantum theory we can hardly say whether it was good or bad luck that the properties of the Kepler motion could be brought into such simple connection with the hydrogen spectrum, as was believed possible at one time. If this connection had merely had that asymptotic character which one might expect from the correspondence principle, then we should not have been tempted to apply mechanics as crudely as we believed possible for some time. On the other hand it was just these mechanical considerations that were helpful in

143

building up the analysis of the optical phenomena which gradually led to quantum mechanics.49

On the other hand, a formal correspondence emerges from Borns work in this third stage which applies to all quantum numbers. Such a use of the correspondence principle still occurs very often in modern quantum theory.50 It represents an element of continuity in change. This continuity is, however, not of a conceptual but of a mathematical or more generally a formal nature. In section 5.2.5 I formulate the philosophical conclusions from the case discussed above more systematically. But first I will examine in more detail the idea of correspondence as numerical agreement in a certain domain, especially taking a closer look at the role of the experiment when establishing this agreement.

5.2.4

Correspondence and material realization

Previously we saw that in the historical period in question (apart from the formal) numerical correspondence between the values of quantities from various theories covering one and the same domain of phenomena proved of permanent importance. The problem of the identity of the elements in this domain under various theoretical descriptions has not as yet been touched on. In this section I would like to say something about it with the help of the notion of material realization introduced in the third chapter (to the role of formal correspondences in this problem I return in the next section). I propose that a (roughly) unambiguous description, which is necessary to establish whether experimentally produced phenomena are identical, cannot be obtained on the basis of theoretical languages, but only with the help of the language of material realization. I shall illustrate this postulate with two examples of intervening experiments, i.e. experiments which can be carried out successfully in a certain domain and which can be theoretically described in different and in this case conceptually incompatible ways. The first experiment can be interpreted in terms of classical electrodynamics and Bohrs atomic theory. It concerns the atomic line spectra and especially the series of spectral lines in the atomic spectrum of hydrogen. We can measure the frequencies of these series experimentally and represent them thus in a formula: the frequency of the light emitted = Rc(1/(n)2 1/n2) (18)

144

(this is a generalization of eq. (2), the Balmer formula). In the experimental domain of line spectra there is a sub-domain of those spectral lines given by conditions n >> 1 and << n. In this sub-domain there are then two interpretations of (18) possible, namely: n =

n , = Rc(1/(n)2 1/n2)

(19a) (19b)

n v n = Rc(1/(n )2 1/n2)

As we saw above these two theoretical interpretations originate in radically different conceptual contexts: the classical electrodynamical frequency against the quantum theoretical difference-frequency of the light emitted by atoms. At the level of the explanatory atomic models experiments in this domain cannot therefore be interpreted unequivocally (i.e. in a manner which would then supply the realist description in this domain): the theories are obviously underdetermined by the experimental data. At first sight we might think that on the level of equation (18) we are dealing with a more empirical description of the experimental phenomena in the domain, which would be unambiguous. The frequency of the light emitted would then be the constant explanandum for which two different theoretical explanantia are possible in the domain at issue. This continuity could then be used for a notion of (co-)reference via the lower, but still theoretical, levels. The problem is, however, that also via this lower level of the observation theories no unambiguousness and continuity is guaranteed. Historically the formulation frequency of the light emitted was interpreted theoretically in various ways, namely both in terms of classical wave optics and in terms of a particle conception of light. The latter is the light quantum hypothesis introduced by Einstein in 1905 which became more and more the issue of a heated controversy, in the same period of Bohrs successes in the field of the atomic spectra.51 Moreover, this controversy was not solved in favour of one of the fighting parties. Modern quantum theory itself supplies, with its wave-particle duality, a new interpretation of the lower theoretical levels which deviates from previous ones. In other words: to prove (19a) and (19b) to be interpretations of one and the same experiment we will apparently have to abandon the theoretical. And indeed as I explained in section 3.3, only the unambiguousness of the non-theoretical description of the material

145

realization can make it plausible that we are dealing with the same phenomena in the interpretations (19a) and (19b). A good example of a description of this kind of experiment measuring the frequencies of spectral lines can be found in the Undergraduate Physics Lab Manual (Free University, Amsterdam 1967, 31-32):
First one puts the collimator slit into the focal plane of the collimator lens ...; the position of the tube, for which this is the case is indicated by a line on it. The images of the slit will then be projected onto the focal plane of the objective lens of the telescope. This focal plane also contains the image of a graduated scale; it can be lit with the little lamp on the third tube. The ocular should be sharply focused on this graduation with the help of the screw mechanism. By another screw device the lens tube can be turned, so that one can observe the whole spectrum ... For emission spectra one measures the positions of the lines on the scale. The slit can optionally be taken wide or narrow. One should, however, always read the right side of the slit image. ... Read as accurately as possible the position of the lines and estimate, if possible, in tenths of the scale unit.

In itself this example does not exactly meet the ideal-type of the nontheoretical description of a material realization, because of the occurrence of theoretical terms such as focal plane and emission spectrum. Nevertheless it seems to be quite near it. I think it plausible that laypersons (if we adjust the description a bit more) could carry out experiments with these kinds of instructions and corrections in everyday language, and indeed without having the slightest idea of either their classical electrodynamical or Bohrs quantum theoretical interpretation. The second example concerns an experiment between Bohrs atomic theory and modern quantum theory, namely Heisenbergs matrix mechanics. This is Francks and Hertzs experiment of 1914, in which electrons were made to collide with hydrogen atoms.52 After these collisions the final energy of the electron current was measured, in relation to the initial energy. The voltage in the final stage of the experiment showed a number of sharp, discrete peaks. Again we can theoretically interpret these results in two different ways:
the values of the voltage of neighbouring peaks represent the values of the energies of neighbouring stationary states of the hydrogen atom in Bohrs atomic model. (20a) the measured peaks in the final voltage relate to the values of the discrete (20b) energies in the virtual field model of matrix mechanics.53

146

As we saw previously both theoretical interpretations are conceptually discontinuous. However: the numerical values of the quantities at issue agree in the common domain of, what one might call, the discrete energy states of the hydrogen atom. Here, too, we could try to look for an unequivocal description on the lower level of observation theories, just as in the previous example. We could, for example, describe Francks and Hertzs experiment purely in terms of electrical voltage and current: what is being measured is then a certain diagram of values of these quantities. But just as in the previous example that wont wash. After all: also at this lower theoretical level there is an electron current colliding with hydrogen atoms. And it is precisely the exact nature of these electrons and atoms which is at issue in the various theoretical explanations. Are we dealing with Bohr atoms and electrons or with the radically different entities from modern quantum theory? Since a choice between these two (based on the true nature of these particles) is impossible in the experimental domain concerned, we cannot obtain unambiguousness on the lower level of the observation theory either; also on this level there is no theoretical continuity which could form the basis for a more ambitious reference theory (than the one proposed here). A realistic reference theory and a realistic realism based on it should therefore return to the description of the material realization. For, despite the fact that very different theories are concerned in the examples of intervening experiments analyzed here, it proves possible to materially realize these experiments in the same, reproducible way; on the one hand the description of the material realization, formulated in daily terms, is the same; on the other, both experimenters judge the experiment as reproducible from their theory. Thus the realistic realism explained in the previous chapter can be applied to this kind of (frequently occurring) cases. In the next section I shall explicitly draw a number of philosophical conclusions from the considerations given up to now. To conclude I would like to briefly say something about the notion of domain in the two examples given. In section 3.2 it was argued that when experimenting, the domain is fixed on the basis of theoretical conceptions and in view of the possible reproducibility of the results. We must choose the experimental objects and apparatus in such a way that reproducibility is guaranteed. In the previous examples we were dealing with the line spectra and the stationary states of atoms, while theoretically presupposing that the atoms of each element from the periodical table can be characterized by a

147

specific spectrum and specific stationary states. This has its consequences for the performance of reproducible experiments with regard to these phenomena. In spectroscopy, for instance, the purity of the materials is an important factor. A striking historical example of such a theory-dependency of the domain is the so-called PickeringFowler series. Due to the impurity of the gas used at first this was mistakenly attributed to hydrogen, instead of helium.54 Apart from this we must screen the experiments from electric or magnetic fields, because they interfere with the electrons in the stationary states, and thus with the nature of the emitted light (Zeeman effect, Stark effect, etc.).

5.2.5

Philosophical conclusions

The example of the history of quantum mechanics analyzed above can be used more explicitly to illustrate and support the general philosophical argument in chapter 4. In the example we can distinguish: (a) the theoretical languages: of the classical theories (electrodynamics and mechanics), of Bohrs atomic theory and of modern quantum theory. An important conclusion of sections 5.2.1 to 5.2.3 was that the theories formulated in these languages display conceptual discontinuities in various respects.55 (b) the language of the description of the material realization: in 5.2.4 I gave two examples of the way in which this language (a sublanguage of ordinary language) can function to obtain an unambiguous description of experiments, despite the existence of different, and conceptually incompatible theoretical interpretations. (c) the correspondence language: this contains on the one side the formal correspondences between terms from the theories at issue, such as n , nn or, from an earlier example, mN mE. On the other this language includes terms such as frequency, mass, electron etc. These terms are the theory-independent names of elements in domains of an independent reality, to which theoretical terms, such as n,, mN, mE, eL, etc., can refer or co-refer.56 I have already discussed extensively a number of philosophical aspects of (the existence of) theoretical languages and of the language of the description of the material realization in chapter 4. But the correspondence language has only been treated briefly. Therefore I use the rest of this section to throw more light on the meaning and function of this language.

148

For this purpose the distinction between theoretical-conceptual and theoretical-mathematical made in 4.3 is of great importance. Generally, theoretical work in the experimental natural sciences shows conceptual and mathematical aspects. It can focus (more) on the development or clarification of concepts or (more) on explanation and extension of mathematical properties or structures of theories. Physics even has a separate specialty for such a mathematical approach, viz. mathematical physics. (1) The subject of this section, then, is the role of the language of correspondence in the philosophical argumentation for a realistic realism. The main issue here is the possibility of co-reference. In the first place let us look at Borns equation (15). Just as any other mathematical equation it lends itself to formal mathematical operations. However: the equation is not purely formal. We interpret , after all as a quantity from classical mechanics or electrodynamics and

n as one from modern quantum theory. In other words, Borns n


equation belongs to mathematical physics. It is a formal, intertheoretical relation.57 From (15) we conclude that a term or concept of type C, from the classical theory formally corresponds to a quantum theoretical term of the type C n
n

. From a philosophical viewpoint this means that we can

thus find out which terms from different theories are candidates for coreference. Secondly we can also via (16) and (17) relate, though indirectly, the theories concerned. Which is to say that we not only examine the formal correspondence, but also the numerical. This relation rests on a correspondence of the numerical values of certain quantities in a domain of experimental phenomena. For instance: the numerical values of the
n quantities C n , and Cn are identical for n >> 1 and << n.58 In this

domain Bohrs atomic theory and matrix mechanics are two equally correct interpretations of the same experiments. Thirdly: that we are dealing with the same experiments in this domain follows from the fact that the actual carrying out of both experiments takes place in the same (reproducible) manner, as is confirmed by the invariance of the description of the material realization of these experiments. Concluding we can say that the corresponding quantities

C n , and C nn co-refer in the domain of large quantum numbers.


Because of the conceptual discontinuity between the classical theories and modern quantum mechanics we cannot theoretically describe the referent of both these terms in an unambiguous and

149

coherent way. We can, however, give this referent a name, with the help of a term from the language of correspondence. Sometimes the names do not have to be coined: mass, frequency, electron etc. In the example above the finding of an adequate name is harder, and therefore I propose to use a symbol, namely C(n,). These and other terms from the language of correspondence have no specific theoretical meaning: they have no function in some theory. In other words: when compared with C n , from the classical theory and with C n
n

from the quantum theory the name

C(n,) is merely a label.59 But this label is not without meaning either. Its meaning is derived from the fact that it is the name of an element in a domain of reality, which we can describe theoretically in one or more clearly specifiable, reproducible ways. The terms from the correspondence language can be compared with numbers on the labels of bottles containing certain liquids. Via such a number we can refer to a certain bottle with a certain liquid. Moreover, we can identify the same liquid in various bottles when the labels are marked with the same number, however varied the form, size or colour of those bottles may be. The above is the essence of the realist position which I am defending. I have illustrated it with relations between classical electrodynamics and mechanics, Bohrs atomic theory and modern quantum theory. The general characteristics of this position are flexible and leave room for variation in other cases. What these other cases exactly look like can, of course, only be established by analyzing them in detail both historically and philosophically. (2) In the practice of scientific experimentation and theorizing these three languages are often mixed in an intricate way. All three are necessary for an adequate characterization of the activity of natural science. Theoretical languages supply the conceptual frameworks in which explanations and predictions are phrased; they establish internal ontologies and contain referring terms. The language of material realization describes the action and materialization aspect of the experimental natural sciences. Finally, the correspondence language contains formal, intertheoretical correspondences and names of referents to which we can refer by means of completely different theoretical terms. This language allows us to speak about different theories simultaneously. In the practice of science one daily encounters this way of speaking. Historians and philosophers of science also speak (partly) in this language, especially when they reconstruct the development of a science over longer periods. When Kuhn for example proposes that the term mass has different meanings in the Newtonian and the Einsteinian theory, this presupposes that there is

150

something to which both theories assign a different meaning.60 Only on the basis of this common something it makes sense to point out differences in meaning at all. For, the fact that terms referring to distinct somethings differ in meaning is not such a spectacular proposition. A major aim therefore of the complex of concepts introduced in this study, and primarily centred around the notion of co-reference, has been to provide a philosophical clarification of this something. For example: in my opinion this realistic realism makes it possible to give a satisfactory (re-)interpretation of MacKinnons view that science aims, among other things, at producing claims whose truth is independent of particular cultures and their limitations. Consider a theoretical claim involving inferred, rather than directly observed, entities: CAR: the normal carbon atom has six electrons.61 I would like to regard CAR as an expression formulated with the help of the language of correspondence. I can then agree with these statements of MacKinnons except on one point. In my view CAR is not a theoretical claim: it is for instance not clear whether it concerns Lorentz or Dirac electrons. (3) With the introduction of the language of correspondence an important problem has been solved which was mentioned but left undiscussed in the philosophical argumentation of chapter 4. There I argued that it is the theoretical terms which refer and (possibly) co-refer. The problem was that for co-reference via the mechanism of a common material realization we cannot deduce from its description which specific terms from both theoretical languages co-refer. The introduction of the language of correspondence solves this problem in the way specified under (1). This solution presupposes of course that scientists can learn both the classical theory and Bohrs theory and apply them to certain problems. But the correctness of this presupposition does not seem, philosophically, problematic and for the sake of clarity I repeat again that its denial is certainly not implied by the conceptual relativism advocated in chapter 4. (4) With respect to the notion of conceptual discontinuity the argumentation in this study rests primarily on the example of the history of quantum mechanics. Of course this is only one example. But philosophers such as Hanson, Kuhn and Feyerabend and their followers have pointed out several other cases of such conceptual breaks. Moreover, it is important to observe that my main thesis certainly is not that each scientific change involves a radical conceptual discontinuity. In fact this problem is treated only in so far it is relevant in the context of scientific realism. And the point is that this one example of the transition of classical theories to (old and new) quantum theory in the light of the enormous importance of classical and quantum theory in the practice of

151

science already raises huge and in my opinion unsolvable problems for more ambitious realist positions, such as naive or convergent realism. I think the only tenable realist positions are those which encompass reference and co-reference, but leave out notions such as representation or approximate truth. (5) In any case, one might expect it to be the task of philosophers of science who do propose conceptual continuity to prove this with detailed historical argumentations. In much literature, however, these arguments are lacking, so that continuity seems to function rather as a philosophical dogma. I give one example to illustrate this. MacKinnon postulates the existence of an intertheoretical continuity based on the fact that experimental data can continue to be used on a lower (a phenomenological) level, even after depth explanations of these data have been found in a theory on a higher level:
the phenomenological framework, and the problematic that emerges within this framework, sets constraints on the form of the depth solution. ... This requires a basic coherence between the phenomenological and depth accounts.62

Prima facie I could agree with this, but the crucial point is of course what this basic coherence exactly means. On the one hand MacKinnon speaks about a conceptual continuity between the phenome-nological and depth theory, and one of the examples he uses for this is the correspondence (when h is very small in relation to the macroscopic action) between the classical macroscopic theories and quantum mechanics.63 As we saw above there is, however, definitely no question of conceptual continuity in this case. On the other hand MacKinnon seems to move in that direction, too, when he writes: This coherence is often achieved by reinterpreting the phenomenological account.64 As a whole, however, the result is that it is still unclear what he means by the claimed continuity, by this basic coherence. The previous discussions are an attempt to make these and related concepts explicit. (6) Finally I would like to briefly go into a recent suggestion of MacKinnons. He proposes to draw certain realist conclusions from the intertheoretical character of some relations from mathematical physics.65 In this context he refers to the fact that for instance certain symmetry principles remain valid after theory changes. Thus the form of physical laws is independent of where, when and in which direction we place the reference system in relation to which we formulate these laws.66 This fact leads, for closed systems, to conservation laws for linear momentum, energy and angular momentum, laws which apply both in the classical

152

theories and in quantum field theory. MacKinnon now proposes to interpret this continuity realistically:
The basic invariance laws considered are accepted as true. Yet they are not true of physical reality in the sense intended when true is defined in terms of satisfaction. Thus, the assertion that a law is invariant under time reversal does not mean that the law applies to systems that go backwards in time. Pace Feynman, there are no such systems. These invariance laws are general constraints, or laws about the laws of physics.67

This seems an interesting suggestion certainly deserving closer examination. Still, I do not think we should expect too much from it with respect to the possibility to arrive at more ambitious realist interpretations of natural science. For in the first place it concerns very formal, structural characteristics of physical theories. As MacKinnon indicates himself it does certainly not concern a realism in which notions such as picturing or reflection play a part. Thus the property energy in the classical law of conservation of energy clearly differs from that in the quantum theoretical law, as a consequence of the different internal ontologies of those two theories.68 Secondly, the question remains whether the presupposition of this kind of realism is tenable: is it really true that all these conservation laws are intertheoretically valid? At first sight this viewpoint raises a number of questions, such as: what about the alternative Galilei versus Lorentz invariance; and what about reversible versus irreversible theories; to what extent can we say that the specific symmetries from quantum field theory represent an intertheoretical continuity? Only a systematic investigation into these problems can lead to a more definitive evaluation of this, in itself interesting, suggestion.

5.3

The measurement problem in quantum mechanics

In the last section of this chapter I will discuss the measurement problem in quantum mechanics. I shall approach this problem from the view on the experimental natural sciences as presented above. On that basis the measurement problem is interpreted as a problem concerning the relation between different modes of description (for a certain kind of object): the quantum mechanical, the classical and the daily mode. For

153

its discussion I make use of the terminology introduced above, viz. of the conceptual discontinuities and of the formal and numerical correspondences between modes of description. The notions of material realization, co-reference and referential realism are also applied to this problem. Consequently the problems surrounding the internal ontology of quantum mechanics how this theory views the world will not be raised. I think the measurement problem is a separate problem, standing apart from the question of the ontological interpretation of quantum mechanics. This approach implies, too, that the measurement problem is dealt with from a rather general philosophical angle, so that the more technical physical aspects will necessarily be left aside. I first treat the problem in a general fashion, in daily terms, and then, somewhat more technically, in quantum mechanical terms.69 In a measuring process we want to establish an (at least roughly complete) correlation between certain properties of the measuring apparatus M and of the object system O about which we want to find out something. Suppose we measure with the help of M the probabilistic quantum mechanical quantity A of O and that the end result of the measurement is given by the values of the position G of a pointer on the measuring apparatus M. If the experiment is to be successful such a correlation between M and O should emerge in the measuring process so that from reading the position value G = gi we can conclude with near certainty that A has value ai. Since the measuring apparatus is a macroscopic system, two descriptions are basically possible: one in quantum mechanical terms and one in classical (or even daily) terms. In quantum mechanics we describe the total system O+M with a state vector or a density operator. Now the point is that there are cases of measurements where we must conclude from the quantum mechanical description of O+M that the pointer of M is in a superposition of states by the end of the measuring process, states which correspond with different and macroscopically clearly distinguishable values gi of G. In other words: one can imagine measurements on O+M where interferences between macroscopically clearly distinct positions of the pointer play a part. But from the classical (or daily) description we expect this: since A is a probabilistic quantity, which may take different values ai (i=1,2,...,n), we do not know in advance which value the pointer position G will have by the end of the measuring process. We do know, however, (because this is a general feature of classical quantities) that in each individual experiment at any moment pointer G will take one and only

154

one of the positions gi (i=0,1,...,n). (In practice we will of course meet with measurement inaccuracies when establishing these values, but that is a different problem, independent of the quantum mechanical measurement problem discussed here). By simply looking at M at the end of the measuring process we can find out which values the pointer position G takes in the various experiments. The measurement problem is now that, classically, the pointer cannot be at one and the same moment in a state which corresponds to different values gi, whereas the quantum mechanical description seems to imply exactly this. This measurement problem presented in these general observations can be formulated in more detail in quantum mechanical terms. Suppose that before the measurement O is in state ci|ai> (|ai> are the eigenstates of operator A, with corresponding eigen-values ai) and that M is in the eigenstate |g0> of G corresponding to the rest state of the pointer. To obtain a macroscopically observable measurement result (a pointer position G = gi), the effect of the microscopic object O on the measuring apparatus M has to be amplified to a macroscopic scale. The force accompanying these amplification processes often results in the disappearance of O as an identifiable microscopic object in the measuring apparatus. By the end of such a (so-called non-ideal) measuring process O will certainly be in a different state than at the beginning; in fact we have not the foggiest idea about the final state of O in these cases. That is why we describe the action of the evolution operator (U) as follows: U ( ci|ai>|g0>) = ci|i>|gi> =|f> Since we do not know the state of O at the end of the measuring process, we regard the ci|i> only as coefficients in the development of |f> to |gi> (the |i>are in general not orthogonal). As required, a correlation has emerged between the values of G and those of A during the measuring process on the ground of evolution U. So: the quantum mechanical state of O+M after the measurement interaction is |f>, or in terms of density operators f, with: f =|f> <f| = ( ci|i>|gi>)(ci*<i|<gi|) (1)

This product essentially contains the so-called interference terms (where i i), corresponding to different pointer positions gi and gi.

155

From a classical viewpoint final state (1) seems to raise problems: classically no interference can occur. For that reason the final state of O+M should much rather be

f , with:
(2)

f = |ci||i>|gi> <gi|<i|
In this equation the interference terms are clearly absent and thus

corresponds with a state in which G does have fixed values gi at the end of the measuring process. Each of these occurs with a chance of |ci|, which is shown when a great number of identical experiments are repeated. Now the problem is that we can certainly postulate as a general requirement that

f is the adequate final state from the classical point

of view, but the question remains which state, (1) or (2), is empirically correct. We can after all imagine measurements on the combined O+M system for which (1) and (2) predict different results. This is especially the case when we measure an observable of the type L = 1 F on O+M for which [F, G] 0, that is for which F and G are incompatible observables (their operators do not commute).70 In other words: in general (1) and (2) are empirically not equivalent. If we now stick to the correctness of the purely quantum mechanical analysis (1), we must conclude that (2), stating that the position of the pointer at the end of the measuring process has welldefined values gi, is incorrect (despite the fact that this statement seems evidently true on the basis of classical and daily insights!). It is this contrast between the final states (1) and (2) which I take as a starting point for the following discussion of the measurement problem in quantum mechanics. I do not deny that there are some objections that might be raised against the measurement problem as it stands in the above presentation. To give some examples: (1) and (2) are formulated in terms of state functions, so that the entire argumentation and conclusion depends crucially on the specific interpretation of these state functions. One may for instance try to find a way out of the problem by postulating that |f> should not be applied to individual systems, but only to ensembles of systems.71 Or one could object that even (2) offers no solution because the decomposition of

f is not unique, so that we still cannot conclude from

(2) that M is really in state |gi> at the end of the measuring process.72 Finally one may try to solve the problem with the help of the insight that

156

the used interaction model is physically much too simple and should be replaced by more realistic models.73 In this section I shall, however, hardly or not touch on these kinds of arguments and the counter-arguments evoked by them. For I think that many of these arguments assume unjustly that the occurrence of a measurement problem should be blamed on the general nature of the argumentation followed up to now. And in order to solve, or rather, to avoid this problem they then turn to more balanced physical accounts and/or philosophical interpretations with respect to the quantum mechanical formalism. What I do intend is first of all to examine closely the philosophical background of the measurement problem as framed above. It will turn out that on the basis of the view on the experimental natural sciences advocated in this study the previous general argumentation, which led to the contrast between (1) and (2), has to be rejected as incorrect, so that the measurement problem vanishes. At the same time this implies that the more specialist arguments mentioned are superfluous for a solution of the measurement problem in quantum mechanics whatever interesting observations they otherwise may have pointed out. In the next three sections I shall analyze three kinds of philosophical argument on the basis of which one often formulates the contrast between (1) and (2) and experiences this contrast as a problem: 1. The pointer is an ordinary observable object and therefore has at any moment a fixed position. Maybe this is never exactly known, but in any case no interference between macroscopically clearly distinguishable positions is possible. 2. The pointer is a classical physical object and, on the basis of this theory, has a well-defined position. In this case the problem is that the classical and the quantum mechanical descriptions are both valid, even though they differ radically from a conceptual point of view. 3. Concerning experimental results (such as the values of the quantity pointer position unambiguous communication should be possible. This requires the pointer position to have a fixed value at the end of the measuring process.

5.3.1

Measurement problem and realism

Thus, the main aim in this section 5.3 is to analyze and evaluate the philosophical views from which follow the arguments which are given to show the existence of a quantum mechanical measurement problem. The

157

first argument is supplied by, among others, dEspagnat. He postulates that state (1), in any case at first sight,
also seems to have the meaning that apparatus M is simultaneously in a superposition of states that correspond to different positions gn of the pointer. Such a meaning, however, comes into direct conflict with what appears to us as a part of our most immediate experience, which is that macroscopic objects cannot occupy different positions at the same time. It seems, therefore, inherently absurd.74

But this indeed only holds true at first sight, because dEspagnat of course knows very well that: (a) for all measurements actually carried out on O+M the difference between (1) and (2) is either zero or neglibly small. (Note that in principle every observation of the pointer plus the measuring apparatus in which O has vanished can be regarded as such a measurement on O+M, as long as O was in the required superposition before the measurement interaction). (b) measurements on O+M which highlight the difference between (1) and (2) (such as L) have not yet been carried out and cannot in fact be carried out without destroying measuring apparatus M as a macroscopic object. We would, after all, have to measure on M a quantum mechanical observable F which is incompatible with G. However, for dEspagnat this is not a sufficient reason to conclude that (as yet) there is no measurement problem. He wishes to view scientific theories realistically and not just as instruments for explaining or predicting experimental results.75 His counter-argument against (a) is that if a scientific theory describes reality, we can draw conclusions with respect to this reality from this theory, independently of the possible practical measurability or unmeasurability of those conclusions. The fact that, from a practical point of view, no difference has been proven between (1) and (2) up to now is insufficient ground within such a realist interpretation to avoid the measurement problem. And dEspagnats counter-argument against (b) is that in order to know that a scientific theory describes reality it suffices if all predictions have come true up till now.76 Realists do not need to postpone their claims until all possible predictions (which are in principle of an unlimited number) have been in fact tested. The only thing (b) implies is that realist claims are always fallible: they can be refuted by future falsifications. If quantum mechanics, in contrast with our most immediate experience, predicts as the correct final state (1) and not (2), then we cannot solve this conflict by pointing out the unmeasurability or the practival negligibility of the differences between (1) and (2), or the fact

158

that this specific prediction of quantum mechanics has not yet been tested. Such an argumentation is, according to dEspagnat, contrary to a realist point of view. However: by all means not every form of realism leads in this way to a measurement problem. DEspagnat adheres to a specific kind of realism, which he calls unirealism.77 This unirealism assumes that reality coincides with physical reality and physical reality in its turn with quantum mechanical reality. Quantum mechanics is taken as the fundamental theory which in principle is capable of describing all of reality. But for certain macroscopic phenomena (such as pointer positions) we have at our disposal another valid description based on our most direct experience and which also forms an essential part of classical theories. DEspagnat wants to hold on both to the validity of this aspect of the classical description and also to the limitless applicability of the quantum mechanical description on the ground of his unirealism. Even though the difference between (1) and (2) may be unmeasurable or practically negligible, the conceptual differences in the interpretations of (1) and (2) continue to exist. Therefore dEspagnats measurement problem is really the fact that these two descriptions are conceptually discontinuous. This entails that its solution would require a conceptual continuity between both modes of description. It will be clear that unirealism plays a crucial role in dEspagnats argumentation on the basis of which he concludes that there does exist a quantum mechanical measurement problem. However, unirealism is, I think, not very plausible as a philosophical thesis. For a start this position does not comply with the no-privilege principle, because of its fixation on the present quantum theory as the fundamental theory. A second objection is unirealisms exaggerated pretension to universality. This claim does not just hold that quantum mechanical processes may have specific consequences on a macroscopic level which cannot be explained by classical theories (this is quite likely, see for instance superconductivity). Over and above this unirealism postulates that various conceptual characteristics of the quantum description are applicable to all phenomena, in some absolute sense. This is why dEspagnats argumentation implies that other descriptions, if their adequacy seems undeniable for certain domains, should also possess these same characteristics. That such a conceptual continuity is an illusion in the case of quantum and classical theory, I hope to have shown in the previous section. For this reason the measurement problem only presents to the unirealist one tiny little

159

problem in a sea of larger (and in my opinion) for him or her unsolvable problems, which crop up as a consequence of the occurrence of conceptual discontinuities between theories. On these grounds unirealism seems to be an untenable philosophical position. Its pretension to universality is a long way from being validated. On the other hand, the realism defended above also contains certain universal claims. If a theory T has been successfully tested for a number of phenomena of a certain kind, the claim is that T applies to all phenomena of that kind. I have interpreted this realistically in the sense that the terms of T involved refer to elements in an independently existing domain of phenomena. If we postulate that a term refers, then we also express by this a universal claim (for all phenomena of that domain) via the requirement of reproducibility. But the question how large the total domain is for which the theory can be interpreted realistically is in my eyes of an empirical, a practical nature. A fortiori it is impossible to establish beforehand, as unirealism does, that there is (or should be) one fundamental theory which describes all of reality correctly. On the basis of all this I would conclude that in general (1), or a more detailed version of it, is the right quantum mechanical description of the measuring process and that therefore no transition from (1) to (2) takes place; that there is a clear contrast with description (2), which is based on characteristics of theories which are conceptually incompatible with quantum mechanics; but, that this only leads to a measurement problem from an (implausible) unirealist point of view. Apart from unirealism, dEspagnats statement that a realist cannot appeal to the unmeasurability or negligibility of the difference between (1) and (2) is not acceptable just like that. What dEspagnat generally means by this he formulated clearly:
Although we are ourselves embedded in reality, reality is independent of us in the senses that it existed before the advent of life and consciousness and that it would exist even if no human being existed.78

This I consider to be a consistent realist assumption. After this, dEspagnat concentrates on the part of this sentence after the comma. The start of this quotation, however, implies that it certainly is not simply non-realist to take the limits and impossibilities of human perception into account. The fact that we do not see interferences between observables of the measuring apparatus M may very well have physical causes, which have to do with specific properties of the apparatus and/or our senses.

160

Suppose we have a good quantum mechanical description of the operation and possibilities of the measuring apparatus and of the senseorgans. If we could then show with the help of this description that, on the basis of the specific macroscopic character of M and in the light of the limited possibilities of our sense-organs, measurements on M which could prove the difference between (1) and (2) are physically impossible, then I think this should be entirely acceptable for any realist. This conclusion does not conflict with dEspagnats realist assumption quoted above. It follows after all from the fact that people are embedded in reality. The only remaining (measurement) problem is the showing of correspondence between the above-mentioned hypothetical, quantum mechanical description of measuring apparatus and sense-organs and its classical description. This correspondence problem is therefore the subject of the next section.

5.3.2

The measurement problem as a problem of correspondence

The second argument assumes the measuring apparatus M to be always a macroscopic system, which therefore can be described by classical theories. Here the position of the pointer is an ordinary classical quantity which has a well-determined, fixed value at any moment. On the other hand it is definitely legitimate to try to describe M (and the interaction with O) quantum mechanically. The situation is thus that we have two theories which both, at least in principle, should be considered applicable to one and the same domain of phenomena, viz. certain properties of measuring apparatus such as pointer positions. For this situation we can therefore raise the question following the observations in 5.2 whether the two theories correspond with each other in that domain. First of all section 5.2 established that in general no conceptual correspondence (or continuity) exists between quantum theory and the classical theories. Thus it is not surprising that in the special instance discussed here this is not the case either. The conceptual framework of quantum mechanics is radically different from that of the classical theories and this does not change when we apply quantum mechanics to macroscopic systems. Conceptually we are stuck with the quantum mechanical superposition (1) to describe O+M after the measurement interaction. Within the philosophical view on science expounded here

161

this fact does not raise any immediate problems as we saw in the previous section. Next we can examine whether there is numerical and formal correspondence and co-reference of terms from the theories involved in this case. We will then have to leave the extremely schematic description (1) and start on a more physical quantum theoretical description which takes into account in particular the specific properties of the macroscopic measuring apparatus M. Thus the macroscopic character of M implies that we are dealing with a complex system of a multitude of particles. Several attempts have in the meantime been made to tackle the measurement problem along these lines.79 For the numerical correspondence the point is to show that in the domain concerned (that of the values of pointer positions and comparable quantities) the quantum mechanical and classical descriptions yield the same numerical results. Therefore it should be first demonstrated that physical aspects (of the operation) of measuring apparatus M allow O+M to be described at the end of the measuring process by (2) or by a comparable mixture. In other words: because of the specific character of the phenomena in the domain involved, quantum theoretical descriptions (1) and (2) are empirically equivalent in this domain. The demonstrating of a numerical correspondence then implies that the numerical values of the pointer position following from quantum mechanical description (2) and of the one yielded by a classical description of M are identical.80 Within these discussions of the measurement problem there is no real transition from a state of form (1) to one of form (2). We only observe that, on the basis of specific characteristics of the domain of application, (1) and (2) are empirically equivalent and we try to give physical reasons for the existence of this equivalence.81 With formal correspondence the point is to define certain macrostates within the quantum mechanical description, which are each builtup out of the micro-states of the constitutive microscopic particles of the measuring apparatus M and which we can each relate via a one-to-one correspondence to different classical states of M, e.g. to states of the channel detectors each representing a well-determined pointerposition.82 The establishing of numerical and formal correspondence is due to Ms complexity a rather difficult physical problem, so that philosophers cannot achieve very much here in the first place. Only once there are acceptable theoretical-physical descriptions which allow the assigning of formal and numerical correspondences, we can check as philosophers, following the lines charted in 5.2.5, whether this is a case

162

of co-reference of terms from the quantum mechanical and the classical description of macroscopic systems such as measuring apparatus. When we treat this problem in such a way it does not raise as a specific measurement problem, leaving aside the problematic of correspondence special philosophical difficulties. I would like to add two observations to what I have said thus far in this section. Firstly, correspondence is, in my opinion, always correspondence in a certain domain. This implies that there may be domains for which only the quantum mechanical description (1) is valid. An example is the domain in which observables of the form L = 1 F, with [F, G] 0, play a part. On the other hand there is of course nothing new in the remark that for a domain in which the observables F and G both play a part there is no numerical correspondence between the quantum and the classical theory. After all, quantum mechanics was developed precisely because the classical theories were unable to deal with this kind of incompatible observables. A second and last remark: the measurement problem as a correspondence problem, as explained here, has as yet very little to do with measuring or experimenting. The measuring process is included only to serve the purpose of being able to describe O+M after the interaction with a superposition. If we could by some means or other, bring M on its own into a superposition, then every reference to the measuring process would be superfluous, while the correspondence problem would continue to exist.

5.3.3

Quantum mechanical measuring process and communication

It is especially Bohr who stresses the importance of communication in his interpretation of quantum mechanics. In this section I will explain the implications of Bohrs view with respect to the measurement problem dealt with here, and then criticize it on the basis of my own viewpoints. It does not seem necessary to describe Bohrs complete philosophy in great detail.83 I will restrict myself to a brief listing of those aspects important to my immediate purpose. First and foremost Bohr postulates that the activities and results of the experimental sciences (including quantum mechanics) should always be objective, meaning by this intersubjectively communicable. For him this implies that they should be described in daily terms (if necessary complemented with relevant parts from classical physics):

163

The decisive point is to recognize that the description of the experimental arrangement and the recordings of the observations must be given in plain language, suitably refined by the usual physical terminology. This is a simple logical demand, since by the word experiment we can only mean a procedure regarding which we are able to communicate to others what we have done and what we have learnt.84

A result which would meet this requirement of intersubjective communicability is for example: At time t we measured the momentum of object O with measuring apparatus M (under such and such circumstances) and found value pi. In classical discussions we can, however, always translate this unambiguously into: Object O has at time t momentum pi. Such an objectivist idealization, in which we abstract from the measuring process and restrict the description to objects and their properties, is quite common in classical physics and does not immediately raise any problems. Specific to quantum mechanics, according to Bohr, is that it does not allow of such an objectivist idealization: quantum mechanical objects in themselves do not have well-defined properties such as position or momentum. We can only speak of Os properties (such as its momentum) in a meaningful way when they are related to a well-determined measuring arrangement M (e.g. a momentum meter). Thus in Bohrs view the quantum phenomena obtain a feature of wholeness: the object, interaction and measuring apparatus are inextricably linked. For this complex of object, interaction and measuring apparatus Bohr uses the term phenomenon. This viewpoint has direct implications for Bohrs interpretation of the quantum mechanical formalism. We can describe object O with a state vector |> or with a density operator , but in Bohrs opinion such a description with the help of the formalism remains formal. The only function of the formalism is to make predictions of measurement results obtained under specific conditions that are described in daily or classical language:
The entire formalism is to be considered as a tool for deriving predictions, of definite or statistical character, as regards information obtainable under experimental conditions described in classical terms ...85

In this view it is therefore senseless to look for an objectivist interpretation of |> or , independent of any specific experimental arrangement. On the basis of this, what can be said about the measurement problem as I sketched it above? First of all it ought to be mentioned that

164

Bohr has never stated anything explicitly about the specific problem of the transition from (1) to (2). We shall therefore ourselves have to develop consistently Bohrs viewpoint with respect to this problem. Firstly: it is certainly not in conflict with Bohrs philosophy to try to describe a macroscopic system such as O+M with the help of quantum mechanics. We must, however, in order to obtain a complete phenomenon (in Bohrs sense), choose a new measuring apparatus M which is described in daily and/or classical terms. (O+M then becomes the quantum mechanical object system O). Bohrs solution of the measurement problem is actually quite simple. State (1) is a quantum mechanical description of the O+M system. If we would consider (1) by itself, we would get certain cases of M being in a superposition of pointer positions gi. The position G of the pointer would not be well-defined in that case. According to Bohr, the statement pointer G is at the end of the measuring process in a superposition of pointer positions gi is not fit for an unambiguous communication. Therefore, Bohr says, we should turn the argumentation around. Intersubjective communicability is a priori. For quantum mechanics this means that O+M and M cannot be objectively separated and that state vector |> has only a formal meaning. Within the formalism (1), or a more exact elaboration of (1), supplies the correct description of the final state of the quantum mechanical measuring process. But precisely because we cannot ascribe any physical reality to this formal description it cannot follow from (1) that M is really in a superposition. We can only meaningfully ascribe physical reality to a phenomenon in this case to O+M.86 The specifically chosen experimental conditions determine which properties of the object involved (viz. O) we can meaningfully call real and which we cannot. If we choose a G-meter for M (e.g. a camera which takes a photo of the pointers position at the end of the measuring process), then G is a welldefined property of O with well-determined values for every single system O. In other words: because of this specific choice of M interferences between pointer positions cannot play a part in these experiments. For quantum phenomena of this specific kind, (2) does and (1) does not supply the adequate description of O+M. The transition of description (1) to description (2), the disappearance of interference terms, is in Bohrs opinion a formal step justified by the specific choice of the experimental conditions necessary to carry out a measurement with G-meter M. Hence, in Bohrs view there is a direct link between the requirement of intersubjective communicability, the daily (and possibly

165

classical) description of the experimental conditions and observations of M, and the formal step in the theoretical description of the quantum system O+M, from (1) to (2). On the other side we may certainly try to measure, following Bohr, the quantity L = 1 F, with [F, G] 0, on O+M. The formal description of O+M is then again (1). In order to have a complete phenomenon we must choose another M, namely an L-meter. But since [F, G] 0 the pointer positions G of M are not well-defined. We do not get a (formal) transition to (2), but to a state in which terms occur which formally correspond with interferences between pointer positions in complete accordance with Gs not being well-defined. From a practical point of view the fact that in this case the quantity pointer position is not well-defined, means that in order to be able to measure L on O+M we have to violate this system in fact so much that as an unambiguously identifiable macroscopic system with unambiguous pointer positions it will be destroyed. No wonder that in such a case we cannot make the demand of daily or classical interpretability applicable to the pointer of O+M. To elucidate this a bit more I shall briefly digress into the literature. It has often proved difficult to apply Bohrs way of thinking consistently. Thus Daneri, Loinger and Prosperi, for example assert that their measurement theory is in harmony with Bohrs. But even though one could reconstruct their approach as such, they themselves sow the seeds of doubt about this statement. The major problem is that they do not distinguish systematically enough between measuring apparatus M as sub-system of the object O = O+M to be described quantum mechanically, and measuring apparatus M (in Bohrs sense the one and only) with which we carry out measurements on O+M. They postulate that:
A satisfactory theory of the measuring process must start from a characterization of the macroscopic properties of a large body. Such properties must have an objective character; it is therefore necessary that no interference terms appear in the relations concerning them.87

In other words: the properties of a system such as M have to be objective in the sense of intersubjectively communicable while at the same time its quantum mechanical description should not contain any interference terms. However, it is correctly Bohrian that in the case under consideration M (and not M) should possess intersubjectively communicable properties (to be described in daily terms) and that it is

166

the choice of a specific M which causes the formal disappearance of the interference terms. It is true that Daneri, Loinger and Prosperi in other passages come closer to Bohrs viewpoint. They postulate explicitly, for example, that the disappearance of interference terms in (1) rests on the fact that with apparatus M one can only measure on M observables which are compatible with pointer position G.88 This proposition resembles closely Bohrs viewpoint that it is the specific choice of M as a G-meter which determines the formal transition of (1) to (2). What is certainly not the case is Rosenfelds assertion89, that this transition, in Daneris, Loingers and Prosperis theory, is caused by irreversible, ergodic (and thus physical) processes in the system M. Whether this, completely unBohrian, misinterpretation by Rosenfeld is the result of the confusing presentation by Daneri, Loinger and Prosperi, or of the inadequacy of his own interpretation of their theory90, I leave aside here. So far this exposition of Bohrs solution of the measurement problem. In the remaining part of this section I shall comment upon this solution with the help of the insights obtained in this work. To guarantee an unequivocal communication Bohr requires the experimental conditions and observation results to be ultimately phrased in plain language (plus the probably needed classical terminology). At first sight there would seem to be an overlap with the description of the material realization I proposed. Still, the differences are bigger than the similarities.91 The main difference concerns the role played by this daily description. For Bohr the choice of the experimental conditions, described in daily terms, decides what can be called physically real on a theoretical level. Only within the context of a phenomenon can we meaningfully speak about the position or momentum of a particle. Thus there is a more or less direct relation between the requirement of communicability, the description of experiments in plain terms and the internal ontology of quantum mechanics.92 In contrast with this, the invariance of the description of the material realization only serves to achieve knowledge of the reference of terms to (in themselves independently existing) elements in domains of reality. Application of the criterion of reference tells us whether the entities etc. which the terms concerned refer to, do or do not factually exist, which is something very different from the possibility to be at all able to talk meaningfully about the physical reality of these entities.

167

To give an example: for Bohr it is only meaningful to assign a position to an electron in relation to an actually existing position meter. But, given this position meter one can regard the electrons position as a classical property. Bohrs internal ontology for quantum mechanics is really still the classical one, except that in the quantum mechanical domain one can only speak significantly about classical properties in relation to well-determined measurement arrangements. But in the realist view as proposed in this study, we might for example conceive of the position of a particle as a well-defined probabilistic quantity, independent of experimental conditions. As I observed in section 4.2 we can also carry out reproducible experiments with respect to chance processes, so that we can apply the criterion of reference to probabilistic quantities too. We then get a completely different ontology, one which does not attempt to connect as much as possible with the classical. Next, one can try to make such an internal ontology plausible with the help of all sorts of arguments, including theoretical ones. The form a plausible internal ontology for quantum mechanics should take is a complex and much debated problem, which is, however, not under discussion here.93 I have, however, argued in general that a realistic interpretation of this kind of internal ontology can only claim reference, and not picturing or reflection. This means that even a plausible quantum mechanical ontology does not picture reality as it really is. A second difference between Bohrs daily and classical language and the description of the material realization involves the relation between daily and classical language. For Bohr these two languages go together quite naturally. He seems to assume that in our entire language of daily experience we speak about objects of which the properties have fixed and identifiable values at all moments, and on the basis of this he restricts intersubjective communication to communication about (or in terms of) these daily and classical objects with their properties. In contrast with this the description of the material realization is formulated only in daily language. In other words: I do not specify to what extent this daily language corresponds with the classical, as far as ontological characteristics are concerned. In my opinion there are in part correspondences and in part differences (the daily language is not one monolithic unity!), so that it is certainly wrong to restrict verbal communication to that part of the daily language showing correspondences with the classical.94 If we call a person aggressive or a social structure capitalist, we do not primarily allude to things and events but rather to dispositions or tendencies. Still, unambiguous communication

168

is possible in these matters. Speculating a bit further I would like to suggest that in this way we can perhaps view the strangeness of the quantum mechanical ontology differently. This ontology is especially strange when contrasted with classical objects and their properties, but it might be closer to the ontology of the previously-mentioned other parts of the daily language. A comparative study of such ontologies could contribute to the designing of a plausible internal ontology for quantum mechanics. Concluding I would like to briefly summarize the results of the discussions set down in this section. Firstly dEspagnats philosophical argumentation for the existence of a measurement problem: this argumentation does not prove convincing, because (a) its unirealist premiss does not agree with the no-privilege principle and (b) the argumentation as a whole presupposes an unwarrantable degree of conceptual continuity. Secondly, we can regard the measurement problem in terms of correspondence within a certain domain. A physical problem is that, at least at the present, the necessary quantum mechanical theories about complex macroscopic systems, such as measurement apparatus, have not yet been sufficiently developed. Once this is the case, we can philosophically interpret this situation following the lines set out in this work. We can then try to formulate explicit formal and numerical correspondences and accurately reconstruct between which theoretical terms and in which domain of phenomena co-reference exists. Finally, it proves possible to avoid a number of problematic aspects of Bohrs interpretation of the measurement problem by interpreting the role of the daily language differently via the notion of the description of the material realization. In particular, this view does not hold, as Bohrs does, that the internal ontology is, as it were, merely a derivative of specific aspects of this description. The link between the theoretical description and the description of the material realization is that both are interpretations of the same material realization, as was explained in chapter 4. Only through the notion of reproducibility are the two different modes of description connected with each other. My conclusion is that the view of the experimental natural sciences expounded in this study makes it possible to achieve a satisfactory solution of the measurement problem in quantum mechanics, in so far it is a philosophical problem.

169

Conclusion

Concluding this study I would like to look back at the whole of the discussions presented. I am guided by two matters. The first one is that of how the two parts are connected. I shall deal with this by explaining the major similarities and differences between Habermas views and the view of the experimental natural sciences proposed in part II. The second one is that of placing this view in a wider context, viz. that of the natural sciences as a social-cultural phenomenon. I shall touch on implications and possible elaborations of this view with respect to this wider context. First of all the comparison with Habermas. The main similarities are, in my eyes, these: a) Quite justly Habermas stresses the special place in the practice of the natural sciences taken by instrumental and experimental action. He focuses particularly on those actions undertaken with the purpose of producing reproducible results. Likewise, notions such as acting and reproducibility play an important role in my view. These form an aspect of the experimental natural sciences which usually is ignored in modern philosophy of science. b) However, in the natural sciences there is not only instrumental action, theoretical argumentations, too, are presented. Habermas takes this fact into account in his epistemology by distinguishing between the objectivity of experience and the truth of propositions. I regard these distinctions as most sensible and have made grateful use of them as a starting-point for the discussions of material realization versus theoretical description of experiments, and for those of reference of theoretical terms versus truth of propositions. Habermas also realized that experimental action is discursively interpreted right from the beginning. But as he observes himself, a systematic explanation of the relation between objective experience and true propositions, and of the part played by experimentation in this, is, up to the present, lacking in his philosophy. The discussion of the epistemic role of the (description of the) material realization with respect

170

to theoretical propositions can be taken, to some extent, as an attempt to fill this void in Habermas. c) Habermas refuses to regard science as an autonomous, isolated activity. He points to its roots in society and demands philosophical space for criticism of science and society. The best known example is his criticism of science and technology as ideology. I share these fundamental starting-points with Habermas, even though my concrete elaboration differs from his in some respects (see below). Besides these similarities there are, however, also some obvious and fundamental differences. a) First of all there is the issue of universality. Although Habermas has abandoned in his later work attempts to find a quasi-transcendental foundation, he maintains the universality of his philosophical claims: objective experience is for him always and only experience in terms of things and events; and the ideal speech situation is a foundation which is anticipated in every communication. In contrast with this, my claims are clearly less universalistic. The material realization of experiments should be carried out with a group of laypersons who all speak, more or less, the same language of material realization. But this does not yet establish what kind of language it is. In principle it is also possible to apply the criterion of reference with the aid of laypersons who do not experience or interpret (the results of) their actions in terms of things and events. Even greater is the difference with respect to the problem of truth. I think the notion of an ideal speech situation as an operative fiction is itself a fiction. Habermas is in fact still looking for the same thing as are the metaphysical realist advocates of the correspondence theory, viz. not truth, but the ultimate Truth, thereby making truth once again transcendent and unattainable. In my view there are in the practice of the experimental natural sciences many propositions which are justly called true, even though this judgement will always remain open to revision. It is philosophys task to make explicit this concept of truth in its applicability to paradigmatic cases. The result of the explication given in this study is that it is not so much truth which is transcendent but reality. It is the latter which transcends all our, untrue and true, linguistic conceptualizations. b) Despite the fact that I have made use of his general distinction between objectivity and truth, its further elaboration differs markedly from that of Habermas. This is mainly due to my realist viewpoint. For Habermas, both fundamental and theoretical cognitive schemata play a role in the constituting of object-domains. Consequently, what he is

171

looking for is the mechanism of the transformation of objective experience in discursively useful experimental data. In my opinion the exact nature of this transformation is philosophically less important. For in fact all transformations of the theoretical description into the description of the material realization are adequate, as long as they enable us to determine reproducibility and coreference. In this question we must clearly distinguish ontology from epistemology. Scientific object-domains are theoretically interpreted right from the start. These interpretations fix the internal-theoretical ontologies. The daily description of the material realization, however, plays an exclusively epistemic part, namely in the acquiring of knowledge of the reference and co-reference of theoretical terms. Thus the realist distinction between ontology and epistemology implies a severe restricting of the philosophical relevance of the problem of transformation. c) As said, I share Habermas critical intention with respect to science and society. However, we both work out that intention quite differently. A sensible and possibly effective criticism of science as a social practice needs, among other things, an adequate philosophical conception of science. Since Habermas philosophical views with respect to the experimental natural sciences proved to be unsatisfactory in a number of ways, I have in the foregoing primarily focused on these. The aspect of social criticism of science has received little explicit attention so far. This is why I shall treat it separately to some extent in the following. Again obvious differences will crop up. The second question I have asked myself in this conclusion is: what does the philosophical view on the experimental natural sciences presented in part II produce from a social-cultural perspective? To answer this we should pay attention not only to the conclusion of the discussion, the realistic realism, but also, primarily, to the whole philosophical position which makes the attaining of this conclusion possible. Realism forms only one aspect of this view. This is, in the first instance, of internal-philosophical importance. It offers a satisfactory explanation of the continuity which is necessary to make the conceptual discontinuities factually occurring in the development of science understandable. I think that we can deduce from the occurrence and the function of trans-theoretical existence-claims in scientific practice one of the few truly plausible arguments for a (limited form of) scientific realism. The realistic realism proposed here gives via the criterion of (co-)reference a precise explanation of what these existence-claims do or

172

do not imply. Moreover, it is exactly in the aspect of reproducibility of material realization that a crucial similarity between experimental natural science and scientified technology lies. By interpreting this aspect realistically I would like to stress that, if we want to successfully use science for technological projects, we are always stuck with the fact that we will have to materially realize these projects in a reproducible manner. Further on I shall come back to this issue. Finally there is the question how and to what extent the kind of arguments which were used here in relation to scientific realism can be extended to take issue with those philosophers (e.g. Quine) who base themselves on discontinuities between whole cultures which have been revealed particularly by anthropologists. In an interesting article Doyal and Harris introduced a notion of constitutive activities, which seems congenial to the idea of material realization. For science they describe this concept as follows:
it is the non-arbitrary shared practice of the scientists (e.g., what they succeed in making and/or manipulating when they conduct an experiment) which mediates between the disparities of belief and/or conceptual understanding with which they carry on their disagreement.1

For the various cultures within the tradition of the experimental natural sciences this shared practice does indeed offer the possibility to refute idealism, as we saw above. The question remains whether the same holds true for cultures differing much more radically from each other. Doyal and Harris give for instance eating, sleeping, playing and healing as examples of constitutive activities which are to some extent in and from themselves understandable, independently of their cultural-specific interpretations.2 But especially for healing and playing one might very well doubt this. This is, however, not the place to go into this matter. I will conclude that this problem of intercultural continuity and discontinuity is a much more difficult one than the analogous problem within the natural sciences. An approach la Doyal and Harris may be fruitful, but the conclusions they have presented up till now are generally premature. A second central aspect of the philosophical view defended here is conceptual relativism and the ensuing refutation of convergent realism. Which is to say that the conceptualizations from the experimental natural sciences are no (approximate) representation of reality as it is in itself. The positive side to this is that the verifiable truth of propositions from the natural sciences can also be analyzed with the help

173

of the notion of plausibility, that plausibility resulting from an interplay between contextual, cognitive and social factors. This view offers the possibility of a fertile exchange between philosophical discussions (about plausibility and thus about truth) and the research in the sociology of knowledge which is swiftly growing in size and significance. I think the social importance of such a viewpoint is primarily generally ideological.3 It relativizes the distinction between scientific and other (political, literary, religious) conceptualizations of reality. This could lead to a weakening of reductionist and scientistic ideologies, which is an advantage in an era in which science is still deployed rightly and wrongly as the only remedy for all kinds of social troubles. In other words: in this situation conceptual relativism can broaden the scope for a non-scientific approach to problems.4 Thirdly, there is the aspect of material realization and, more in general, of experimental natural science as a concrete production process. Because this aspect has, through scientified technology, profoundly influenced and changed society, man and nature, I think it high time that it be given an important place in philosophical discussions of the natural sciences. In the context of the exposition about realism I primarily made use of the idea of a description (in more or less constant daily terms) of the material realization. But at the same time it was stressed that this description of the material realization process implies a certain abstraction. In fact such a material realization is always part of a social context and thus has some essentially social aspects: the material realization is always also a social realization. The question which aspect predominates, the material one or the social one, seems to me to be unanswerable.5 This social realization of the experimental natural sciences can be analyzed in more detail in various ways. For a start we can focus on, what I call, the extrinsic factors in the social realization of closed experimental systems. One could for instance think of the hierarchical and sexist power relations which all too often characterize this social realization in modern scientific practice. Or one might think of the industrialization, militarization and commercialization of the practice. These extrinsic factors of the social realization of experiments should be subjected to criticism and change. They are theoretically analyzed in the so-called science, technology and society studies.6 There are also intrinsic social factors active in the production of closed experimental systems. As we have seen, theoretical knowledge is necessary in order to materially realize such systems. This is nevertheless not purely an internal-scientific question. For in order to achieve

174

closedness we also have to succeed in implementing the control required by theoretical science in society. Experimental projects, for instance, should be physically and socially planned in such a way that they are protected from the interference of playing children, students who feel antagonistic towards science, hooligans etc. For the experimental natural sciences this would mostly require social control on a relatively small scale.7 This analysis becomes interesting when we extend it to technological projects, which in principle should also yield stable, reproducible results. Compared with experimenting there are two related, but most important differences in the case of a scientified technology. In the first place, the social control needed to obtain stable and reproducible technological results often permeates an entire society. This is quite obviously the case in the use of experimental results from medical science.8 Considering the strongly technologized character of our present society, much detailed research is required in order to unravel this specific entanglement of knowledge and power. On the basis of this we can ask the question who or which authorities exert this social control, which social ideologies play a part in this, and also whether we wish to comply with this control and, if not, how we can resist it. The latter brings to the fore a second difference in degree between experiment and scientified technology. Generally speaking, it will prove quite problematic whether the social control, needed for the production of stable and reproducible technological systems can in fact be realized. This holds true particularly for present-day society, in which we are often concerned with the social realization of large-scale technological systems over longer periods of time. That is why we must ask ourselves whether there are risks involved in the production of (possibly) unstable or non-reproducible technological systems, and if so, whether we do or do not want to live with them.9 These discussions about possible further developments of the view on the experimental natural sciences proposed in this study are obviously tentative and preliminary. Nevertheless, I hope that they have fulfilled their purpose, namely that of pointing out that studies in philosophy of science should not only be judged by their internal philosophical merits, but that they should also explicitly take into account, and be evaluated by what they have to say about, the historical and social place of the phenomenon of science.10

175

Notes

Introduction
1. 2. Marx, second thesis on Feuerbach. See Marx (1959), 243. Please note: the philosophical claims in this study generally refer to the experimental natural sciences. For stylistic reasons I also use a shortened form when misunderstanding seems impossible. See e.g. Koyr (1968) and Kuhn (1977a). An important different attempt can be found in Hacking (1983). I shall briefly discuss his views on realism and experimentation in chapter 4. Habermas (1982), 274. However, he did not sit around doing nothing: see his theory of communicative action, Habermas (1981).

3. 4. 5.

Chapter 1
1. The purpose of this book is certainly not to give an exegesis of the development of Habermas writing. The main focus is on a number of philosophical problems relating to the experimental natural sciences. I treat these partially in connection with Habermas work. See Habermas (1978b), 89; and Habermas (1981), Volume I, 1617 and Volume II, 586 593. It remains to be seen whether Habermas himself escapes this fundamentalism entirely. Cf. McCarthy (1978), 110. Habermas (1979a), 24. See also section 1.4.1. In Habermas (1978a), ch. 4, he develops this criticism of positivism on the basis of an analysis of the philosophies of Comte and Mach. One could, however, doubt whether his criticism is also completely adequate with respect to the modern version, i.e. logical positivism. Habermas (1978a), vii. See, e.g., Habermas (1978b), 308311. For the definitions of the terms instrumental, strategic and communicative action, see 1.3.1. See, e.g., Habermas (1974), 713. See Habermas (1971c). Cf. also Marcuse (1968). See Habermas (1982), 242250. See for this especially Habermas (1981). Habermas (1971c), 9192. See also Habermas (1979b), 117120. See McCarthy (1978), 2330, for a discussion of a number of problems Habermas early definition of the term strategic action has to face. Habermas (1971c), 92. See, e.g., Habermas (1978b).

2.

3. 4. 5.

6. 7. 8. 9. 10. 11. 12. 13. 14. 15.

176

16. McCarthy (1978), 2628. In Habermas (1979b), 117120, Habermas himself follows this rephrasing by McCarthy. 17. One might very well doubt whether this is a satisfactory solution of the problem. See for this the Conclusion. 18. See Habermas (1978c), 360380. Cf. also McCarthy (1978), 91125. 19. Habermas (1973), 214. Cf. also Habermas (1971a), 114115. 20. See Habermas (1979a). For accounts of the distinctions only indicated here, see McCarthy (1978), 272291; or Korthals, Kunneman (1979), 112122. 21. In contrast with this the claim to genuineness can only be validated within the actionframe, and not discursively; while the claim to comprehensibility can be justified by an actual, instead of a grounded, consensus. See Habermas (1973), 221222. In his (1981), Volume I, 3545, Habermas introduced with respect to these earlier views some additions, nuances, and alterations. However, these are, within the context of the present study, of lesser importance. 22. Habermas (1978c), 375. 23. Habermas (1971b), 207. 24. Habermas (1978c), 368. Cf. also Habermas (1971b), 206213. 25. In Habermas (1971b), 207209, he also adds the arithmetical categories (numerals) which make quantification possible. In this work he also postulates that causal expressions have a somewhat different status, because they do not refer directly, as e.g. here and now do. 26. See for instance Piaget (1973). 27. See Habermas (1978c), 363 and 365. Note that identification of objectivity with (a certain kind of) intersubjectivity is not unusual. We also find this in such different philosophers as Popper and Bohr (see Popper (1959), 44; and Bohr (1958), 67). Nevertheless this identification implies a severe restriction of the objectivity problem. In the next chapter I shall indicate an alternative approach, to be developed subsequently in part II of this study. 28. See Habermas (1971a), 127. For the probable meaning of quasi, see section 2.2.2. 29. Habermas (1978a), 127. 30. Occasionally Habermas does not just speak of things and events, but also of deep structures which manifest themselves in terms of things and events (Habermas (1974), 20; and Habermas (1978c), 370). But this in itself important distinction does not in fact play a well-defined role in Habermas epistemology and its addition is rather gratuitous. 31. E.g. Habermas (1978a), 68. 32. Habermas (1978a), 130 (see also Habermas (1978c), 374375). Also cf. Habermas (1982), 242 and further. See also McCarthy (1978), 117. Here I am merely concerned with Habermas intention. I do not mean to say that by this he also solved this problem satisfactorily. 33. See e.g. Bhme (1976). 34. Habermas (1978a), 128. 35. Habermas (1978a), 126. 36. Habermas (1978c), 383. 37. See Habermas (1978a), 124; Habermas (1978c), 369. 38. Habermas (1973), 211212. 39. Habermas (1973), 211. 40. Habermas (1973), 215218. 41. Habermas (1973), 216. See also Habermas (1971a), 123124. 42. Habermas (1973), 212 (and 218). 43. Habermas (1981), Volume I, 65. 44. Habermas (1974), 18.

177

45. Note that the point is the linking, not the identification, of truth and validity of argumentations. This interpretation of validity in Habermas as validity of a discursive argumentation seems to me quite plausible for truth. Still, there is one difficult side to it, in that the claim to genuineness (whereby Habermas also speaks of validity) is not linked up with such a discursive argumentation. 46. Habermas (1973), 218. 47. Cf., for instance, Wittgenstein (1974, 71, 73, 232 and 239) about the role of doubting. 48. Habermas (1973), 239240. Cf. also Habermas (1971a), 129. 49. Habermas (1973), 240. 50. I use the translation adequacy (for Angemessenheit), because this term carries the double connotation of something being adequate for something. Habermas himself also points out the material aspect of his theory of truth although he does not work it out in great detail when he proposes to conceive of the cognitive schemata in the sense of Piaget (and also of a materialist theory of knowledge, which regards social work in society as a synthesis). (Habermas (1973), 246). 51. See Hempel, Oppenheim (1948) for a more detailed formulation of this explanatory scheme. 52. Habermas (1978b), 308. 53. Habermas (1973), 243244. Cf. also McCarthy (1978), 405406, note 22. 54. Cf., for the distinction between epistemological and metaphysical positivism, Radder (1983), 172173. 55. Habermas (1973), 246. 56. See e.g. Piaget (1973). 57. See for a more detailed elaboration of this example, although in a different context, Koningsveld (1973), 87100. 58. This is also one of the key points in Wittgensteins (1974); see e.g. 125, 164, 310, 315, 341. 59. It is this indirect relatedness to reality which according to Habermas makes it possible for true empirical-analytic propositions to be technically applied to the domain of things and events. 60. See Kuhn (1970a). 61. Habermas (1978c), 375. 62. See Habermas (1973), 252254. 63. Habermas (1973), 255. For a brief and lucid summary see Kunneman (1980), especially 230231. 64. Habermas formulations with respect to notions such as meaning, criterion and condition for truth are not always clear (cf. Puntel (1978), 158159). I interpret criterion here quite minimally as merely a sufficient condition. This is because in chapter 2 I shall mainly criticize the sufficiency of Habermas criterion of truth. Still there is something to say for an interpretation of criterion as both a sufficient and a necessary condition. Since this problem is less relevant for my purposes, I leave it aside. 65. See Keuth (1979), 384385. 66. See Puntel (1978), 159. 67. Habermas (1973), 255. 68. See Habermas (1973), 219. 69. I shall paraphrase these conditions (in my own words and focused on the theoretical discourse), so as not to have to represent Habermas theory and classification of speech acts entirely. Cf. further Habermas (1973), 255 ff. 70. Habermas (1973), 256. 71. Habermas (1973), 256. 72. Habermas (1973), 257259. See also Habermas (1971a), 139141. Cf. also Habermas (1982), 277.

178

73. 74. 75. 76. 77. 78.

See also Kunnemann (1979) and (1980); and Boon (1980). Habermas (1971a), 128. Beckermann (1972), 6771. Habermas (1978c), 360, refers to Apels Peirce-interpretation. See also Apel (1980). Habermas (1978c), 366 and 375. Habermas (1978c), 366.

Chapter 2
1. 2. 3. Habermas (1978c), 367369. See for instance Shapin (1979), in particular 4350. He no longer speaks about (quasi-)transcendental philosophy, but about horizontal and vertical reconstructions of universal competences for following rules. See Habermas (1971b), 174175; Habermas (1978c), 378379. Cf. also McCarthy (1982). See Habermas (1981), Volume I, 72113. Differentiating between these worlds is one aspect of rationalization, the other is the institutionalizing of discourses, which allows systematic criticism of validity claims. See Habermas (1981), Volume I, 7685. Cf. Haleber (1983). I shall however come back to this later on, when developing my own viewpoints in part II. Certainly, as far as instrumental action is concerned, the difference between these horizontal reconstructions and his earlier transcendental philosophy is quite minimal. That is why I maintain the older terminology on this point for the sake of clarity. See Habermas (1978c), 368369 and 377378. Also cf. McCarthy (1978), 296, who argues that Habermas follows the analytic Kant reception in this issue.

4.

5. 6. 7. 8.

9.

10. Neither in these recent discussions is there any agreement about the correct (re-) interpretation of transcendental philosophical notions. To give some examples: the interpretation of the condition of possibility in terms of sufficient conditions is proposed by Baum and Wilkerson (Baum (1979), 7, and Wilkerson, quoted in Bennett (1979), 5556); while for Bhaskar and Bennett the crucial point of transcendental argumentation lies in the formulation of necessary conditions (Bhaskar (1979), 67, and Bennett (1979), 50 and 5556). 11. Rorty (1979), 82; Bhaskar (1979), 68 and 153158. As opposed to Rorty, Bhaskar also points out the positive side to transcendental philosophical arguments: even though such a conceptual analysis only makes explicit what already was given implicitly, still its realization is a non-trivial question, for which a critical discussion about the exact form of the arguments is both possible and desirable. See also Bhaskar (1978), 257. 12. Habermas (1973), 216. Cf. also Habermas (1971a), 123124. See also section 1.5.1. 13. See for instance Bhaskar (1978), 3645. Cf. on this point also Lobkowicz (1969), particularly 266268. 14. Note that in Habermas facts are, paradoxically enough, not something in the world like things and events are! See Habermas (1973), 215. 15. See especially section 4.3, but also compare section 2.3.1. 16. Habermas (1971a), 127. 17. Janich (1978), 912 and 23. Cf. also Hacking (1981). 18. Habermas (1982), 275. 19. See Hamlyn (1970), 112117; and Puntel (1978), 35. 20. Habermas, too, postulates explicitly that his theory of truth does not concern this third question. See Habermas (1973), 239; and Habermas (1982), 273.

179

21. Habermas (1982), 273. Cf. also McCarthy (1978), 300303. 22. Here I follow the view of criterion which Habermas advocates in his (1973). In Habermas (1982), 273, he uses the term differently, viz. more in the sense of a decision procedure to verify the truth of specific propositions. 23. See the quotation at the beginning of section 1.5.2. 24. An example of the confusion of concepts, which often reigns supreme in discussions of these problems, is a debate between Kunneman and Boon. In Kunneman (1979), 177, we read that Habermas defines truth as a claim to truth attached to propositions about which a consensus has been reached in a discourse, on valid grounds. This shows a confusion between definition of truth and criterion of truth. Boon criticizes Kunnemans and Habermas views, but in his argumentation he simply equates criterion of truth and specific testing procedure (Boon (1980), 141142). In the end this is quite justly rejected by Kunneman (1980), 236, where he says that grounded consensus is a criterion of truth and does not serve as a vehicle for scientific progress. It is, however, strange that Kunneman writes this down just like that, making no reference to his earlier proposed interpretation of a consensus achieved on valid grounds as a definition of truth. 25. See Radder (1978), 3336. The same argument can be found in Keuth (1979), 380381. 26. Habermas (1973), 249. 27. In a variation on Habermas (1973), 218: A grounded claim is not yet by any means a claim founded on experience. 28. Cf. Keuth (1979), 381. 29. Keuths criticism of Habermas (Keuth (1979), 384385), is not right on this issue, for he all but ignores the role of the communicative condition in the criterion of truth, and, moreover, does not bear in mind consistently enough the epistemological character of Habermas entire theory of truth. 30. It is quite likely that Habermas himself does not feel too happy about this point, given his statement that the evidential dimension of the concept of truth is badly in need of further clarification (Habermas (1982), 275). 31. Habermas (1973), 232233. Already in his earlier work Habermas sees the justification of general propositions via induction as a central problem for the theory of science. See Habermas (1978), chs. 5 and 6. 32. Puntel (1978), 163164, stresses this. I do, however, feel that he pays far too little attention to what I have called the material aspect of Habermas theory of truth, viz. the proven adequacy of the theoretical language system. 33. Cf., for what follows, Habermas (1973), 244250. 34. However, also cf. point 5 of this section, where I shall go into the relation between these two kinds of cognitive schemata. 35. Habermas (1978c), 366. 36. It is certainly not the case that for Habermas a theoretical language system also contains theoretical laws apart from the theoretical language. For first of all he argues that the cognitive development, guaranteeing the adequacy of the theoretical language system, precedes the separate argumentations for or against specific theoretical laws (see Habermas (1973), 249). Secondly, if the laws were included in the theoretical language system, the plausibility of laws would follow by definition from the adequacy of this language system. But instead of this Habermas keeps saying that this plausibility, via an inductive step which should be pragmatically interpreted, rests on the adequacy of the theoretical language system (see 1.5.2). 37. In section 4.2 I shall go into the notion of plausibility in greater detail. 38. This is also Boons criticism (Boon (1980), 142143). 39. Kunneman (1980), 231235.

180

40. Kunneman (1980), 237239, defends Habermas in this question by referring to the possibility of institutional measures within the group of scientists. As said, this in itself problematic suggestion only touches one aspect of Habermas definition of the ideal speech situation. 41. Habermas (1978c) , 366. 42. Cf. the observations about this in section 2.2.1. See further Habermas (1981). 43. Habermas (1982), 275. 44. See Habermas (1978c), 366. 45. Habermas (1973), 246. 46. Hesse (1982), 103, my italics. In his Reply Habermas does not oppose this interpretation (cf. Habermas (1982), 275). 47. These restrictions are, however, clearly less severe than for example in Kant. See for an example section 5.3.3, where I criticize Bohrs views, which are comparable to Habermas on this point. Compare Habermas viewpoint in the quotation of note 35. 48. In a more exhaustive criticism one could also ask questions about the exact role of consensus in the criterion of truth, or about the reasons for the restriction to the deductive-nomological scheme Habermas uses in the explanation of his concept validity of argumentations.

Chapter 3
1.

2.

3.

4. 5. 6. 7.

8. 9. 10.

11. 12. 13. 14.

See Janich (1978), 919. Note that this is a somewhat elliptical formulation. In fact p, p1, etc. always represent theoretical descriptions of episodes or situations, i.e. propositions. To make the text better readable and to avoid being unnecessarily pedantic I shall, however, call p and p1 from time to time situations or episodes (except for cases in which this might lead to misunderstandings). This concept of reproducibility plays a major role in the discussion. What I shall say about it now is very sketchy. In what follows I shall go further into it, on various occasions. I therefore share Habermas opinion on this issue. See the quotation of note 29 in section 1.4.1. See for a more extensive analysis of this example Koningsveld (1973), 87100. See in this context also Barnes (1983). Cf. also Kuhn (1977b). See Von Wright (1971), 3482. On the one hand this means that a number of criticisms made against Von Wright (e.g. Tuomela (1978), 3144; Bhaskar (1978), 54) are not applicable to my discussions; and, on the other hand, that my own criticism of Von Wright only applies to his claims with respect to experimentation in the natural sciences. Von Wright (1971), 54. Von Wright (1971), 63. For Von Wrights reply to the question of the possibility of knowledge of non-manipulable systems, see Von Wright (1971), 7071. Reasonably, for we can never be absolutely sure whether a0 would not have changed into a of itself. We assume this as reasonable on the basis of earlier experiences. Von Wright argues that this counterfactual aspect (that certain things would not happen unless we do them, and that others would unless we prevent them) is an essential property of the concept of action. Von Wright (1971), 69 and 72. Von Wright (1971), 64 and 82. Bhaskar (1978), 256. For the definition see Bhaskar (1978), 70. For the criticism see Jongeling (1981), 2940.

181

15. As the following example shows, these conditions may also be necessary for the experimental result itself, and not only, as Von Wright postulates, for the carrying out of the experiment. See Von Wright (1971), 63. 16. Von Wright (1971), 186. 17. I do want to add that the essence of my criticism is not so much that apart from sufficient also necessary conditions should be involved in the discussions: Von Wright can also do this by a slight (formal) alteration in his definition of closedness. The main point of concern is that not only the occurrence of initial situation p generally requires active human interference, but that this also holds true, among others, for situation q. Hence, an important purpose of the entire analysis is to determine the exact meaning of the phrase that in closed systems not only p, but also q is artificially produced by people. 18. This means: sufficiently general within the context of this study. In other contexts, particularly those in which the social aspects of the producing of closed systems are prominent, we should also include the influence of the system on its surroundings. For this see Radder (1986). 19. Quoted in Derksen (1980), 160. 20. This is primarily an elaboration of Radder (1978), 3842. 21. Note: I use this example of Newtonian and Einsteinian mass, here and in the next chapter, only to indicate the philosophical problem (see further, e.g., Kuhn (1970a), 101 102; and Field (1973), 463473). A more detailed example is worked out in chapter 5, in which I discuss the relation between the classical physical theories and quantum mechanics. 22. This research has not been published, but it is reported rather comprehensively in Lynch, Livingston, Garfinkel (1983), especially 225229. 23. Lynch, Livingston, Garfinkel (1983), 225. 24. Lynch, Livingston, Garfinkel (1983), 228. 25. Cf. Lynch, Livingston, Garfinkel (1983), 209212; and Polanyi (1967), ch. 1. 26. Also compare the distinction Collins introduces between the algorithmic and the enculturational model for the replication of experiments. Collins (1975), 206208. 27. Lynch, Livingston, Garfinkel (1983), 228, my italics. 28. See e.g. LvyLeblond (1976), 147151; Latour, Woolgar (1979), esp. ch. 2; and Rip (1982), 225226 and 229231. 29. Latour, Woolgar (1979), 127. 30. There are of course exceptions. See e.g. Hacking (1981) and (1983). 31. See also Van Lieshout, Mol (1982), esp. 3061. They regard science as a conglomerate of talk (the verbal, the theoretical), things (the material) and activities (the social), whereby these aspects are not hierarchically related but mutually structure each other. A recent example of ignoring the material can be found in De Vries (1982). When discussing Laboratory Life he leaves out almost completely the notion of materialization to which Latour and Woolgar attach so much importance, restricting himself to discussing their concept credibility. 32. Lynch, Livingston, Garfinkel (1983), 210. 33. At least: in our culture. See further section 4.4. 34. Bhaskar (1982), 342. See for the following also: Bhaskar (1978), 3336, 5356 and 255 262; and Bhaskar (1979), 514. For a summary of Bhaskars work see Radder (1980). 35. See Von Wright (1971), 7374. 36. Bhaskar (1982), 343. Bhaskars second argument thus also implies a criticism of Habermas viewpoint with respect to technological application: while for Habermas technology follows naturally from the experiment (as anticipated technology), there is a fundamental difference according to Bhaskar: in contrast with experimentation technology generally takes place in basically open systems, where different kinds of

182

37. 38. 39. 40.

mechanisms operate simultaneously and contrarily. It would carry us too far to subject this important difference between Habermas and Bhaskar to a closer examination. Cf. also Radder (1986). See also e.g. Benton (1981) and Gibson (1983). See Bhaskar (1978), 3656. Cf. Bhaskar (1978), ch. 3. See also section 4.5.

Chapter 4
1.

2. 3. 4. 5.

6. 7. 8. 9. 10.

11.

12. 13. 14. 15. 16. 17. 18.

19. 20. 21. 22.

It will be understood that also in the criterion to be proposed here, we are not concerned with offering specific methods to verify concrete propositions from the natural sciences. Cf. the distinctions at the beginning of section 2.3. Universal propositions will be evaluated in terms of plausibility; see further on, under in the sixth place. See e.g. Harvey (1981). See e.g. Panofsky, Phillips (1962), 276277. It follows from the criterion that a proposition C which was accepted in the first place as verified true can generally prove to be not verifiably true after all in four ways: a. The material realization of the experiment is not reproducible. b. The required general laws are not plausible. c. The logical inference rules have not been correctly applied. d. C has been inferred from other propositions which, after all, prove to be not verifiably true. In fact, social and cognitive aspects are always interwoven in the practice of scientific research. See e.g. Bhaskar (1978), 184. However, for such an independent testing new auxiliary hypotheses are usually required. Cf. e.g. Woolgar (1981), 385386. See Jammer (1966), 5661; and Kuhn (1978). Hendry (1980), 66. See Jammer (1966), 161165. A complete acceptance (which was, however, accompanied by conceptual changes) only became possible after the emergence of modern quantum mechanics in the years 1925 and 1926. See Hendry (1980), 78. In general one cannot characterize the use of models simply as cognitive. (See e.g. Van Lieshout (1982), 6162). Here the entanglement of social and cognitive aspects, which I referred to in note 6, becomes apparent. See Lakatos (1970), 140154. See Derksen (1980), 324. See Derksen (1980), 321. See Radder (1982a), 102. See Derksen (1980), 324 and 326. Cf. also Shapere (1980), 9193. Shapin (1979), 41. In Shapin (1982) he offers an extensive survey of the (mainly English) studies in this field. For a general introduction of this strong programme, see Van Lieshout (1982). For a survey of the various theoretical approaches in the field of sociology of scientific knowledge, see Knorr-Cetina, Mulkay (1983). Harvey (1981). For a good introduction to the physical problem, see dEspagnat (1979). Harvey (1981), 108112. Meanwhile this test has been carried out with a positive result for quantum mechanics. See Aspect, Dalibard, Roger (1982). Harvey (1981), 124.

183

23. See e.g. Kwa, Van Lieshout, Mol, Pekelharing, Radder (1982). Cf. also De Vries (1982). 24. For a somewhat more extensive exposition of this criticism of the claims of this kind of micro-sociological studies of science, see Kwa, Van Lieshout, Mol, Pekelharing, Radder (1982), the even pages. 25. Cf. Bhaskar (1978), especially ch. I. 26. Cooke (1978), 320 (my italics). Another example: I distinguish sharply between what is the case and how we judge what is the case. This reflects a strong commitment to realism. I take scientific theories to consist, in the main, of true or false descriptions of reality irrespective of our views of the matter. (Devitt (1979), 30). As if our (!) theories are no part of our views of the matter. 27. Leplin (1979), 276. 28. This point seems debatable. One might object that there is the rest mass which is invariant, or that in the later Lorentz theory the mass of the electron was also variable (see McCormmach (1970), esp. 4752). The simplest answer to this is to argue that as far as mass is concerned the conceptual break proves to be earlier, namely between Webers theory and later theories (see Leplin (1979), 271 and further). On the other side one could point out the problems which the identifying of Lorentz mass with either rest mass or relativistic mass raises (see Field (1973), 463473). 29. Georgi (1981), 45. 30. Some aspects of the example above (point 1-4) are treated in more detail in chapter 5. 31. Cf. for instance Boyd (1973) and Newton-Smith (1980). 32. In the next section I shall work out the relation between domain and material realization. 33. What the relation between the theoretical languages and the description of the material realization is, I shall explain in 4.4, under point (2). 34. In principle it also seems possible to maintain differences in degree between theoretical and observational propositions (cf. Newton-Smith (1980), 9495). If we, however, use the thesis of underdetermination in discussions about realism, such a pragmatic solution is insufficient. See the examples in section 5.2.4. 35. Two examples of this can be found in Newton-Smith (1980), 105110. Another wellknown example comes from Poincar. See for an account and discussion of the latter, Sklar (1976), 88146. 36. See the example analyzed above. 37. For a clear exposition of this view see Laudan (1981), esp. 2022. Also compare Hardin, Rosenberg (1982). 38. Cf. with respect to this distinction between learning and translating also Kuhn (1970b), 266271; and Hacking (1982), 5661. 39. For example Laudan (1981) contains much relevant material. 40. Hesse (1980), 143144. 41. My evaluation follows more or less the same lines as Hesses (1980), viixiv. 42. Putnam (1978), 24. Cf. also Doorman (1983). 43. Apparently Putnam realized this himself, considering his recent change to an internal realism. See Realism and Reason, in Putnam (1978), 123140. For a survey see Doorman (1983). Please note: in this chapter I shall regularly refer to various views of Putnam. However, I do not intend to describe his development or even sketch it. See for this, for example, Putnam (1983), viixviii. 44. As the next chapter will show the explanation of T2 is better than T1 in terms of T1 is a limiting case of T2, if the latter is exactly analyzed, will in general not fall under (a) just like that. This is contrary to what advocates of such an explanation suggest. (See e.g. Putnam (1978), 1922). The problem is that they do give all kinds of examples, but never work out one example in detail. 45. Laudan (1981), 40.

184

46. Cf. also Laudan (1981) for a detailed criticism of convergent realism. 47. Cf. also Barnes (1983). For a brief survey of different forms of relativism see Hollis, Lukes (1982), 112. 48. In the next sections I shall examine what the function is of speaking about e (electron) or m (mass) as such. 49. Cf. also Hacking (1982), 58 and further, who argues that the crucial point of the problem is not in truth but in truth or falsity. 50. See Newton-Smith (1982), 106108. 51. See e.g. Shapere (1980), Pekelharing (1982), Barnes (1983). 52. Cf. the discussions in section 3.4. 53. Note: the point is not so much this A and her specific theoretical description. It suffices for the reproducibility of the material realization that there is some experimenter who, when repeating the experiment, will arrive again and again at the same theoretical description. 54. For the sake of simplicity I speak here of theory, singular. In fact various theories always come into play in the theoretical description of experimental situations, as we saw in the example in 3.2. If the material realization is reproducible then all terms from the theoretical description of all experimental episodes refer. The distinctions between the various required theories (e.g. between auxiliary theories and theories to be tested) are not relevant with respect to the reference of terms. 55. See Collins (1975). Cf. also the controversy about locality represented in 4.2. 56. Note that in a similar way, via the criterion of verification, the truth of propositions always relates to a domain. For closer examinations of this concept of domain see Shapere (1977) and Nickles (1977). 57. Cf. Fine (1975), 2930. See for an example section 5.3.1. 58. Cf. Pekelharing (1982), 3940. 59. Extensive discussions of historical and philosophical aspects of correspondence can be found in the next chapter, in particular in section 5.2. 60. In this way we can certainly not relate all E- and N-equations. Cf. Bunge (1970), 288-291. 61. Cf. also Zahar (1983), 246248. 62. Although I can agree generally with many of Hesses viewpoints (see note 41), there is a radical difference on this point. Hesse does aim at a (kind of) translation of one theoretical language into the other, through a set of true (referring) observation sentences (see Hesse (1980), 146). In my opinion this difference can be found at some other points as well, but it would carry me too far astray to treat this explicitly in this context. 63. The transition to the theory or relativity is for example less radical than that to quantum theory. But to repeat: the main reason why the example of Newtonian and Einsteinian mass is used in this general philosophical chapter is that this has an easy appeal for pointing out these problems. By this I do not mean to say that it is the best case for supporting my views. 64. Putnam (1975a), 200; cf. also Putnam (1978), 1925. 65. Hardin, Rosenberg (1982), 608 (my italics). I do want to add that these authors present this as only one of the possible views on reference. 66. See Derksen (1980), 282286. 67. See for instance Putnam (1975b), 236; Putnam (1978), 2122; and Leplin (1979), 281 282. Cf. Hesse (1980), 153155. 68. See e.g. Putnam (1975a) and (1975b). His theory is akin to Kripkes (see Kripke (1980)). For a survey see Van Eyck (1980). 69. On this point I agree with a number of critics of Putnam. See e.g. Derksen (1980), 283 284; Leplin (1979), 278281; Nola (1980), 524526. 70. This is against Nola (1980), 515517.

185

71. See e.g. Derksen (1980), 282283; Nola (1980), 515 and further. 72. Accepting for the moment that these are corresponding terms. As said, the formulating of formal correspondences in specific cases requires new detailed research. The reason for this is that the way in which terms can formally correspond with each other may differ in general from case to case. However, an examination of the formal correspondence between phlogiston and oxygen would carry me too far from the main line of the argument. 73. See Hesse (1980), 155156. 74. Again accepting that these terms formally correspond with each other in this domain. See note 72. 75. Cf. Field (1973). Cf. also the observation under note 60. 76. We can only observe the larger pluriformity from the point of view of the modern theory; e.g. on the basis of the referential refinement (see for this term Derksen (1980), 285 286) of phlogiston to sometimes oxygen and sometimes hydrogen. 77. Putnam (1978), 25. I have left out Putnams italics. 78. I will speak about his more recent internal realism in the next section. 79. See Laudan (1981), 3236. 80. To specify this a bit more I shall make two observations. First of all, we will have to restrict the propositions C to empirical claims. In other words: the reality conditions do not apply of course to logically true propositions or to propositions with logically true parts. Next we should be careful when interpreting the negation, by clearly differentiating between C has no truth value and C is false. We can then conceive of the referring of the terms of C as a presupposition of Cs possessing a truth value. Two reasons support such a view. On the one hand it links up with what was said in section 4.3 about conceptual relativism with respect to truth. On the other hand, if we identify not(C is true) with C is false, we are left with contradictions. (From not(all descriptive terms of C refer) then follows both C is false and C is true, because C and notC contain the same descriptive terms). 81. Note that these conditions are restricted to singular propositions. As was explained in section 3.2, singular propositions in the experimental natural sciences always refer to a theoretically defined domain of reproducible phenomena. This is the reason why, through this reproducibility, universal laws and theories also play a role in the reference and truth of singular propositions. From these universal propositions we can, in my eyes, only demand that they be plausible. As said, I focus in this study on the problem of reference. Wherever the argumentation requires the notion of plausibility, I have tried to use it in an intuitively clear and acceptable way. 82. Heisenberg (1948), 334. 83. Hacking (1983), 265. 84. Hacking (1983), 8687. 85. See e.g. the article Realism and Reason, in Putnam (1978), 123140. Cf. also Putnam (1983), viixviii. 86. Putnam (1978), 137. 87. See, e.g., Doorman (1983), 109117. 88. Cf. Putnam (1978), 2122; and Doorman (1983), 114116. 89. Doorman (1983), 116. 90. Koyr (1968), 90. One should, however, keep in mind that this fusion of theoretical and experimental tradition did not emerge everywhere at the same time and its development was not equally rapid in all fields. See Kuhn (1977a), about the rise of the Baconian sciences. Apart from this, the question of the macro-social background of the rise of this experimental tradition remains important. Cf. for instance Keulartz (1981), part I, 29-36.

186

Chapter 5
See the discussion about conceptual discontinuity in section 4.3. This is clearly a different approach from Meyer-Abichs, who tries to give an (unequivocal) interpretation of all Bohr has said about correspondence and the correspondence principle. See Meyer-Abich (1965), esp. ch. II. 3. Bohr (1913). 4. See Hund (1974), 65. For the events leading up to Bohrs atomic theory, see Heilbron, Kuhn (1969), 211-266. 5. See Bohr (1913), 874875; the assumptions 4 and 5 mentioned there I leave aside for the present. 6. See for this development Heilbron, Kuhn (1969), 266277. 7. See Bohr (1914). This is the translated text of a lecture given in December 1913. 8. Bohr (1914), 295. 9. See Bohr (1913), 1314. 10. Bohr (1913), 14. So this should really be completed with the correspondence between
1.

2.

n, and

n vn for > 1.

11. 12. 13. 14. 15. 16. 17. 18.

19.

20.

21.

22. 23. 24. 25. 26. 27. 28.

Bohr (1914), 287288 and 295; Bohr (1913), 12. See for instance Bohr (1914), 301. See Bohr (1913), 1415. See also Heilbron, Kuhn (1969), 277280. See for this and the following: Jammer (1966), 89109. For a systematic explanation, see e.g. Kramers (1919), 145. See Bohr (1918), 125130. For a lucid representation of this proof, see Tomonaga (1962), 142146. Bohr (1918), 110111 and 130131. Kramers (1919), 48. He made this correspondence plausible by referring to the fact that, with respect to the intensity distributions, the radiation laws of Planck and of RayleighJeans for low frequencies also correspond with each other. Kramers (1919), 61 and 72. Please note: for the sake of brevity I will not deal any further with correspondence regarding the polarizations. It is remarkable, though, that here in some cases an exact correspondence occurs, also for small quantum numbers. See for this Bohr (1918), 110111 and 130133; Jammer (1966), 113114; and Tomonaga (1962), 147150. A letter from Bohr to Rutherford, 27 December 1917, in Bohr (1976), 682683; about the same time that the first results of Kramers work became known. See this same letter and also Radder (1982c), 226. Compare Epsteins reaction to Bohrs On the Quantum Theory of Line Spectra: It really seems that the discrepancy between the quantum mechanical and classical approach is not at all as large as was assumed until now. (Letter from Epstein to Bohr, 14 May 1918, quoted in Bohr (1976), 637). Bohr (1918), 110. Cf. also 130131. Kramers (1919), 4750. See equation (14) in section 5.2.3. Kramers (1919), 52. Bohr (1976), 577; see also Kramers (1923), 552. Cf. also Tomonaga (1962), 159162. See Kramers (1919), 48 and 50. Here I do not consider the rather opaque use of the correspondence principle in Bohrs so-called second atomic theory (in the years 19201923). See for this: Kragh (1979), esp. 155159, and also Kramers (1923), 554559; and Jammer (1966), 133155.
For different applications, see note 28 and the references mentioned there.

29. This holds good at least for the more systematic applications of the correspondence principle. 30. See e.g. Bohr (1923), 472473.

187

31. Bohr, Kramers, Slater (1924). 32. Jammer (1966), 184. See also Radder (1982c), 242 and 252, note 65. 33. This probabilistic aspect of the virtual field model has been too scantily treated in the historical literature, in my opinion. See Radder (1982c), 242243. 34. Quoted in Van der Waerden (1967), 20. 35. Born (1924). 36. See Serwer (1977), 214219; and Hendry (1982), 204209. 37. Hendry (1982), 207. 38. Heisenberg (1925). For extensive discussions of this see Van der Waerden (1967), esp. 2535; and MacKinnon (1977), esp. 161188. 39. Quoted in Serwer (1977), 242. See also Paulis letter to Bohr of 21 February 1924, quoted in Serwer (1977), 230. 40. Meyer-Abich does try to give such a reconstruction of Bohrs view, on the basis of a lot of material. See Meyer-Abich (1965), 7293. The final result of this (8687) is, however, both too vague (and thus too little elucidating) and unsatisfactory, because it is still formulated in terms of the harmonic components of the motion of the electrons. 41. This against Meyer-Abich (1965), 85. Also see Paulis views as quoted in Serwer (1977), 242. 42. See Kramers (1924), 201; and Heisenberg (1925), 262. 43. Kramers (1925), 328; cf. also Kramers (1924), 201. 44. See Forman (1971) and Radder (1983). 45. See Born (1924). 46. For claritys sake I describe Borns equation in my own notation here. 47. See Born (1924), 192194. 48. Born (1924), 190. 49. Letter of Bohr to Oseen, January 1926. Quoted in Rosenfeld, Rdinger (1967), 73. 50. An example is the substitution of Poisson brackets by commutators. See e.g. Messiah (1969), 317318. 51. See Hendry (1980). 52. See e.g. Tomonaga (1962), 136138. 53. See its calculation in Pauli (1926). 54. See Lakatos (1970), 147148. 55. Cf. also Hanson (1972), 149157. He speaks of logical discontinuity. 56. The question may arise whether the correspondence language is a real language. My answer to this is that here I do not use the term language in a strict sense, as is not uncommon in philosophy. Another possibility could be to speak of a correspondence terminology. 57. It is certainly not the case that this and other intertheoretical relations always connect the complete formalisms of the theories concerned. Not only in Borns equation (15), but also for instance in the Ehrenfest theorem, we are dealing with a relation which only applies to certain cases for certain kinds of functions. Cf. Bunge (1970), 288293. As for the Ehrenfest theorem see e.g. Messiah, who writes: It is generally not correct to state that the mean values <qi> and <pi> follow the laws of Classical Mechanics. (Messiah (1969), 217). Thus Nickles rendition of the theorem that any relation that appears in classical mechanics must be valid as a relation between quantum theoretical expectation values, is clearly wrong (Nickles (1973), 194). 58. Note that this numerical correspondence between the Fourier components and the transition amplitudes is established indirectly (in contrast with that between the classical and quantum frequencies), for instance on the basis of experiments about the dispersion of light by atoms. See in this context the clarification of the criterion of reference at the beginning of section 4.4. 59. Cf. Kripkes conception of proper names, in Kripke (1980), esp. 4760. 60. See Kuhn (1970a), 101102.

188

61. 62. 63. 64. 65. 66. 67.

68. 69.

70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80.

MacKinnon (1982), 446. MacKinnon (1982), 448. MacKinnon (1982), 448449. MacKinnon (1982), 448 (my italics). MacKinnon (1982), 457460. Note that the examples of intertheoretical relations he gives are different from Borns relation discussed here. Cf. also Wigner (1970), articles 15. Interestingly, MacKinnon relates these geometrical invariances to measurement and measurability (MacKinnon (1982), 453). MacKinnon (1982), 459. Post, too, points out the importance of such trans-theoretical symmetries and invariances. However, he does not use these in an argumentation for realism, but in a plea for a rational heuristics. See Post (1971), 226228. One can express this difference in terms of possessed versus latent observables. See e.g. Park (1968), 210212. I certainly do not give an extensive explanation of the measurement problem. See for this the literature, e.g. dEspagnat (1976); Fransen (1981); Park (1968); Jammer (1974). A complete understanding of the entire argumentation with all its details certainly requires technical-physical knowledge. Here I have aimed at presenting it in such a way that the outline of the argument will also be understandable to the reader who is not schooled in physics. See e.g. Fransen (1981), 3334. See e.g. Ballentine (1970). See e.g. Park (1968), 215; Belinfante (1975), 5862; and Fransen (1981), 1012. In 5.3.2 I shall come back to this objection. For a discussion of various possibilities see Fransen (1981) and dEspagnat (1976). DEspagnat (1976), 162, my italics. See dEspagnat (1976), 163165. Cf. what dEspagnat (1979) says about induction. DEspagnat (1976), 239. DEspagnat (1976), 233. See e.g. Daneri, Loinger, Prosperi (1962); Machida, Namiki (1980); for surveys see Fransen (1981), 4180; dEspagnat (1976), 161226; Jammer (1974), 471521. Note that this does not imply that (2) now possesses the properties of a classical ensemble, in the sense that it would be built up from sub-ensembles, to every element of which could (as far as M is concerned) be unambiguously assigned state |gi>. After all, the decomposition of f is in fact not unique; there are always other observables of M which are not dispersion-free on the sub-ensembles with pointer positions gi. See Park (1968), 212217; and Fransen (1981), 1012. However, I do not think all this raises any unsurmountable problems for the measurement problem as it has been discussed here. For, in the context of this problem, it is not necessary to show that every M is in a unique state |gi>. It suffices to show for the domain in question that a physically acceptable description exists which assigns the correct numerical value to classical quantities such as the pointer position for any apparatus M at any moment. For the latter a description such as (2) is adequate. Cf. the remark at the end of section 5.3.1. Note: in practice we will not always succeed in supplying physical reasons for the equivalence of (1) and (2). I think that Machida and Namiki (1980) have come closest to achieving this end. Daneri, Loinger, Prosperi (1962) have assumed this equivalence but for non-physical reasons (see 5.3.3). See e.g. Daneri, Loinger, Prosperi (1962), 299. See for this e.g. Honner (1982); Stapp (1972); Scheibe (1973), 949; Radder (1979). Bohr (1963), 3. Bohr (1948), 314. See Bohr (1935), 138139.

81.

82. 83. 84. 85. 86.

189

87. Daneri, Loinger, Prosperi (1962), 305; see also 298. 88. This is explicitly in conflict with Bubs statement that Daneri, Loinger and Prosperi pretend to use only purely quantum mechanical arguments. See Bub (1968), 505 and 515. Apart from that Bubs criticism seems to agree with mine. 89. See Rosenfeld (1965), 225, 229 and 230. 90. See Fransen (1981), 5556. 91. In Radder (1979) my viewpoint with respect to the measurement problem was closer to Bohrs. 92. Honner (1982) interprets this relation as transcendental. And indeed there does exist a clear analogy with the way Habermas establishes a relationship between intersubjectivity on the basis of instrumental controllability, the description of sensorial experience with the help of fundamental cognitive schemata, and the limits they impose on the basic predicates of internal-theoretical ontologies. Cf. section 2.3.2, point 5. My criticism of Bohr is also analogous to that of Habermas. 93. Nowadays this discussion concerns most of all the role played by the notions of nonlocality or non-separability in the theorems of Einstein, Podolski and Rosen and Bell. See e.g. Selleri, Tarozzi (1981); and dEspagnat (1976) and (1979). 94. Habermas would say: to the language of things and events. See section 1.4.1. But it is even to be doubted whether things and events form an adequate ontology for the classical theories. See Bhaskar (1978) and the discussion in section 3.4.

Conclusion
1.

2. 3.

Doyal, Harris (1983), 75. See Doyal, Harris (1983), 67. There is also the more specific question whether conceptual relativism opens perspectives for conscious, normative choices between concrete alternatives within (instead of for) science. This problem is intriguing but too complex to briefly discuss it here. See e.g. Kunneman (1979) and Keulartz (1981). 4. For claritys sake: by this I, of course, do not mean to say that non-scientific approaches are always preferable to scientific ones or that any approach which is non-scientific is per se recommendable. See for a closer examination of the problem of regular and alternative medical treatment, Keulartz, Kwa, Radder (1985). 5. This implies also a rejection of Habermas distinction between instrumental and communicative action as inadequate as far as this does not refer to aspects of concrete actions to be distinguished analytically, but to empirical characterizations of those actions themselves. See the end of section 1.3.1. Cf. on this point Bader (1983), esp. 331. 6. See e.g. Ravetz (1971) or Boers (1981), esp. ch. III. 7. Cf. the scaling-down scaling-up model in Latour (1983). 8. See e.g. Keulartz, Kwa, Radder (1985). 9. See for a further elaboration of these questions Radder (1986). 10. Cf. the criticism of Derksen in Radder (1982b), 8889.

190

Bibliography

Apel, K.-O. (1980): C.S. Peirce and the Post-Tarskian Problem of an Adequate Explication of the Meaning of Truth: Towards a Transcendental-Pragmatic Theory of Truth, Part I, The Monist, 63, 386407. Aspect, A., Dalibard, J., Roger, G. (1982): Experimental Test of Bells Inequalities Using Time-Varying Analyzers, Physical Review Letters, 49, 18041807. Bader, V.-M. (1983): Schmerzlose Entkopplung von System und Lebenswelt?, Kennis en Methode, 7, 329355. Ballentine, L.E. (1970): The Statistical Interpretation of Quantum Mechanics, Reviews of Modern Physics, 42, 358381. Barnes, B. (1983): On the Conventional Character of Knowledge and Cognition, in: Knorr-Cetina, Mulkay (1983), 1951. Baum, M. (1979): Transcendental Proofs in the Critique of Pure Reason, in: Bieri, Horstmann, Krger (1979), 326. Beckermann, A. (1972): Die realistischen Voraussetzungen der Konsenstheorie von J. Habermas, Zeitschrift fr allgemeine Wissenschaftstheorie, 3, 6380. Belinfante, F.J. (1975): Measurements and Time Reversal in Objective Quantum Theory, Pergamon Press, Oxford. Bennett, J. (1979): Analytic Transcendental Arguments, in: Bieri, Horstmann, Krger (1979), 4564. Benton, T. (1981): Realism and Social Science. Some Comments on Roy Bhaskars The Possibility of Naturalism, Radical Philosophy, nr. 27, 1321. Bhaskar, R. (1978): A Realist Theory of Science, Harvester Press, Sussex. Bhaskar, R. (1979): The Possibility of Naturalism. A Philosophical Critique of the Contemporary Human Sciences, Harvester Press, Sussex. Bhaskar, R. (1982): Realism in the Natural Sciences, in: L.J. Cohen et al. (eds.), Logic, Methodology and Philosophy of Science VI, NorthHolland Publishing Company, Amsterdam, 337354.

191

Bieri, P., Horstmann, R.-P., Krger, L., eds. (1979): Transcendental Arguments and Science, Reidel, Dordrecht. Boers, C.(1981): Wetenschap, techniek en samenleving, Boom, Meppel. Bhme, G., ed. (1979): Protophysik, Suhrkamp, Frankfurt. Bohr, N. (1913): On the Constitution of Atoms and Molecules, Part I, II, III, Philosophical Magazine, 26, 125, 476502, 857875. Reprinted in Bohr (1981). Bohr, N. (1914): On the Spectrum of Hydrogen, in Bohr (1981), 283 301. Bohr, N. (1918): On the Quantum Theory of Line Spectra, reprinted in Van der Waerden (1967), 95136. Bohr, N. (1923): On the Application of the Quantum Theory to Atomic Structure. I. The Fundamental Postulates of the Quantum Theory, reprinted in Bohr (1976), 458499. Bohr, N. (1935): Can Quantum-Mechanical Description of Physical Reality Be Considered Complete? reprinted in: S. Toulmin (ed.), Physical Reality, Harper and Row, New York, 130142. Bohr, N. (1948): On the Notions of Causality and Complementarity, Dialectica, 2, 312319. Bohr, N. (1958): Atomic Physics and Human Knowledge, Wesley, New York. Bohr, N. (1963): Quantum Physics and Philosophy Causality and Complementarity, in: N. Bohr, Essays 19581962 on Atomic Physics and Human Knowledge, Interscience, New York, 17. Bohr, N. (1976): Collected Works, Volume 3, NorthHolland Publishing Company, Amsterdam. Bohr, N. (1981): Collected Works, Volume 2, NorthHolland Publishing Company, Amsterdam. Bohr, N., Kramers, H.A., Slater, J.C. (1924): The Quantum Theory of Radiation, reprinted in Van der Waerden (1967), 159176. Boon, L. (1980): Moet de wetenschap rationeel zijn?, Kennis en Methode, 4, 140151. Born, M. (1924): Quantum Mechanics, translated and reprinted in Van der Waerden (1967), 181198. Boyd, R.N. (1973): Realism, Underdetermination, and a Causal Theory of Evidence, Nos, 7, 112. Bub, J. (1968): The Daneri-Loinger-Prosperi Quantum Theory of Measurement, Nuovo Cimento, 57B, 503520. Bunge, M. (1970): Problems Concerning Intertheory Relations, in: P. Weingartner, G. Zecha (eds.), Induction, Physics and Ethics, Reidel, Dordrecht, 285325.

192

Collins, H.M. (1975): The Seven Sexes: A Study in the Sociology of a Phenomenon, or the Replication of Experiments in Physics, Sociology, 9, 205224. Cooke, R.M. (1978): Realism and Content, Kennis en Methode, 2, 319 340. Daneri, A., Loinger, A., Prosperi, G.M., (1962): Quantum Theory of Measurement and Ergodicity Conditions, Nuclear Physics, 33, 297 319. Derksen, A.A. (1980): Rationaliteit en wetenschap, Van Gorcum, Assen. Devitt, M. (1979): Against Incommensurability, Australasian Journal of Philosophy, 57, 2950. Doorman, S.J. (1983): Nieuw Realisme? Kennis en Methode, 7, 98120. Doyal, L., Harris, R. (1983): The Practical Foundations of Human Understanding, New Left Review, nr. 139, 5978. DEspagnat, B. (1976): Conceptual Foundations of Quantum Mechanics, Benjamin, London. DEspagnat, B. (1979): The Quantum Theory and Reality, Scientific American, November 1979, 128140. Eyck, J. van (1980): Refererende termen, Kennis en Methode, 4, 206 228. Field, H. (1973): Theory Change and the Indeterminacy of Reference, The Journal of Philosophy, 70, 462481. Fine, A. (1975): How to Compare Theories: Reference and Change, Nos, 9, 1732. Forman, P. (1971): Weimar Culture, Causality and Quantum Theory, 19181927: Adaptation by German Physicists and Mathematicians to a Hostile Intellectual Environment, Historical Studies in the Physical Sciences, 3, 1115. Fransen, M. (1981): Het meetprobleem in de quantummechanika, Masters Thesis, University of Amsterdam. Georgi, H. (1981): A Unified Theory of Elementary Particles and Forces, Scientific American, April 1981, 4055. Gibson, Q. (1983): Tendencies, Philosophy of Science, 50, 296308. Habermas, J. (1971a): Vorbereitende Bemerkungen zu einer Theorie der kommunikativen Kompetenz, in: J. Habermas and N. Luhmann (eds.), Theorie der Gesellschaft oder Sozialtechnologie Was leistet die Systemforschung?, Suhrkamp, Frankfurt, 101141. Habermas, J. (1971b): Theorie der Gesellschaft oder Sozialtechnologie? Eine Auseinandersetzung mit Niklas Luhmann, in: J. Habermas and N. Luhmann (eds.), Theorie der Gesellschaft oder Sozial-

193

technologie Was leistet die Systemforschung?, Suhrkamp, Frankfurt, 142190. Habermas, J. (1971c): Technology and Science as Ideology, in: J. Habermas, Towards a Rational Society, Heinemann, London, 81-122. Habermas, J. (1973): Wahrheitstheorien, in: H. Fahrenbach (ed.), Wirklichkeit und Reflexion. Festschrift fr W. Schulz, Neske, Pfullingen, 211265. Habermas, J. (1974): Theory and Practice, Heinemann, London. Habermas, J. (1978a): Knowledge and Human Interests, 2nd edition, Heinemann, London. Habermas, J. (1978b): Knowledge and Human Interests: A General Perspective, in: Habermas (1978a), 301317. Habermas, J. (1978c): A Postscript to Knowledge and Human Interests, in: Habermas (1978a), 351380. Habermas, J. (1979a): What is Universal Pragmatics?, in: J. Habermas, Communication and the Evolution of Society, Beacon Press, Boston, 168. Habermas, J. (1979b): Historical Materialism and the Development of Normative Structures, in: J. Habermas, Communication and the Evolution of Society, Beacon Press, Boston, 95129. Habermas, J. (1981): Theorie des kommunikativen Handelns, Suhrkamp, Frankfurt (2 Volumes). Habermas, J. (1982): A Reply to my Critics, in: Thomson, Held (1982), 219283. Hacking, I. (1981): Spekulation, Berechnung und die Erschaffung von Phnomenen, in: H.P. Duerr (ed.), Versuchungen. Aufstze zur Philosophie Paul Feyerabends (second volume), Suhrkamp, Frankfurt, 126158. Hacking, I. (1982): Language, Truth and Reason, in: Hollis, Lukes (1982), 4866. Hacking, I. (1983): Representing and Intervening, Cambridge University Press, Cambridge. Haleber, R. (1983): Het etnocentrisme van de westerse sociologie. Habermas en het evolutionaire moderniseringsparadigma, Krisis, nr. 13, 2750. Hamlyn, D.W. (1970): The Theory of Knowledge, MacMillan Press, London. Hanson, N.R. (1972): Patterns of Discovery, Cambridge University Press, London. Hardin, C.L., Rosenberg, A. (1982): In Defense of Convergent Realism, Philosophy of Science, 49, 604615.

194

Harvey, B. (1981): Plausibility and the Evaluation of Knowledge: A Case-Study of Experimental Quantum Mechanics, Social Studies of Science, 11, 95-130. Heilbron, J.L., Kuhn, T.S. (1969): The Genesis of the Bohr Atom, Historical Studies in the Physical Sciences, 1, 211290. Heisenberg, W. (1925): Quantum-theoretical Reinterpretation of Kinematic and Mechanical Relations, translated and reprinted in Van der Waerden (1967), 261276. Heisenberg, W. (1948): Der Begriff abgeschlossene Theorie in der modernen Naturwissenschaft, Dialectica, 2, 331336. Hempel, C.G., Oppenheim, P. (1948): Studies in the Logic of Explanation, Philosophy of Science, 15, 135175. Hendry, J. (1980): The Development of Attitudes to the Wave-Particle Duality of Light and Quantum Theory, 19001920, Annals of Science, 37, 5979. Hendry, J. (1982): Bohr-Kramers-Slater: A Virtual Theory of Virtual Oscillators and Its Role in the History of Quantum Mechanics, Centaurus, 25, 189211. Hesse, M. (1980): Revolutions and Reconstructions in the Philosophy of Science, Harvester Press, Brighton. Hesse, M. (1982): Science and Objectivity, in: Thomson, Held (1982), 98-115. Hollis, M., Lukes, S., eds. (1982): Rationality and Relativism, Blackwell, Oxford. Honner, J. (1982): The Transcendental Philosophy of Niels Bohr, Studies in History and Philosophy of Science, 13, 129. Hund, F. (1974): The History of Quantum Theory, Harrap, London. Jammer, M. (1966): The Conceptual Development of Quantum Mechanics, McGraw-Hill, New York. Jammer, M. (1974): The Philosophy of Quantum Mechanics, Wiley, New York. Janich, P. (1978): Physics Natural Science or Technology? in: W. Krohn, E.T. Layton, P. Weingart (eds.), The Dynamics of Science and Technology, Reidel, Dordrecht, 327. Jongeling, B. (1981): Bhaskar en de regelmatigheidsfilosofie, Research Report, Free University, Amsterdam. Keulartz, J. (1981): De Starnbergers over alternatieven in de wetenschap, Part I, Krisis, nr. 4, 2344 and Part II, Krisis, nr. 6, 83109. Keulartz, J., Kwa, C.-L., Radder, H. (1985): Scientific and Social Problems and Perspectives of Alternative Medicine, Radical Philosophy, nr. 41, 29. Keuth, H. (1979): Erkenntnis oder Entscheidung? Die Konsenstheorien der Wahrheit und der Richtigkeit von Jrgen Habermas, Zeitschrift fr allgemeine Wissenschaftstheorie, 10, 375393.

195

Knorr-Cetina, K.D., Mulkay, M., eds. (1983): Science Observed, Sage, London. Koningsveld, H. (1973): Empirical Laws, Regularity and Necessity, Veenman, Wageningen. Korthals, M., Kunneman, H. (1979): Arbeid en interaktie, Coutinho, Muiderberg. Koyr, A. (1968): Metaphysics and Measurement, Chapman & Hall, London. Kragh, H. (1979): Niels Bohrs Second Atomic Theory, Historical Studies in the Physical Sciences, 10, 123186. Kramers, H.A. (1919): Intensities of Spectral Lines, in: H.A. Kramers, Collected Scientific Papers, North-Holland Publishing Company, Amsterdam, 1108. Kramers, H.A. (1923): Das Korrespondenzprinzip und der Schalenbau des Atoms, Die Naturwissenschaften, 550559. Kramers, H.A. (1924): The Quantum Theory of Dispersion, reprinted in Van der Waerden (1967), 199201. Kramers, H.A. (1925): On the Behaviour of Atoms in an Electromagnetic Wave Field, in: H.A. Kramers, Collected Scientific Papers, North-Holland Publishing Company, Amsterdam, 321331. Kripke, S.A. (1980): Naming and Necessity, Blackwell, Oxford. Kuhn, T.S. (1970a): The Structure of Scientific Revolutions, University of Chicago Press, Chicago. Kuhn, T.S. (1970b): Reflections on my Critics, in: I. Lakatos, A. Musgrave (eds.), Criticism and the Growth of Knowledge, Cambridge University Press, Cambridge, 231278. Kuhn, T.S. (1977a): Mathematical versus Experimental Traditions in the Development of Physical Science, in: T.S. Kuhn, The Essential Tension, University of Chicago Press, Chicago, 3165. Kuhn, T.S. (1977b): The Function of Measurement in Modern Physical Science, in: T.S. Kuhn, The Essential Tension, University of Chicago Press, Chicago, 178224. Kuhn, T.S. (1978): Black Body Theory and the Quantum Discontinuity, 18941912, Clarendon Press, Oxford. Kunneman, H. (1979): Cognitieve en normatieve rationaliteit, Kennis en Methode, 3, 173198. Kunneman, H. (1980): Rationaliteit en consensus, Kennis en Methode, 4, 229245. Kwa, C.-L., Lieshout, P. van, Mol, A., Pekelharing, P., Radder, H. (1982): Onderhandelen over het laboratoriumleven, Krisis, nr. 8, 8896. Lakatos, I. (1970): Falsification and the Methodology of Scientific Research Programmes, in: I. Lakatos, A. Musgrave (eds.), Criticism and the Growth of Knowledge, Cambridge University Press, Cambridge, 91195.

196

Latour, B. (1983): Give Me a Laboratory and I will Raise the World, in: KnorrCetina, Mulkay (1983), 141170. Latour, B., Woolgar, S., (1979): Laboratory Life. The Social Construction of Scientific Facts, Sage Publications, Beverly Hills. Laudan, L. (1981): A Confutation of Convergent Realism, Philosophy of Science, 48, 1949. Leplin, J. (1979): Reference and Scientific Realism, Studies in History and Philosophy of Science, 10, 265284. Lvy-Leblond, J.-M. (1976): Ideology of/in Contemporary Physics, in: H. Rose, S. Rose (eds.), The Radicalisation of Science. Ideology of/in the Natural Sciences, MacMillan Press, London, 136175. Lieshout, P. van (1982): Het strong programme: de oplossing van wetenschapsfilosofie in wetenschapssociologie, Krisis, nr. 7, 5471. Lieshout, P. van, Mol, A. (1982): Aanzetten tot een kritiese wetenschapstheorie, Masters Thesis, University of Utrecht. Lobkowicz, N. (1969): Interesse und Objektivitt, Philosophische Rundschau, 16, 249273. Lynch, M., Livingston, E., Garfinkel, H. (1983): Temporal Order in Laboratory Work, in: Knorr-Cetina, Mulkay (1983), 205238. Machida, S., Namiki, M. (1980): Theory of Measurement in Quantum Mechanics, Progress of Theoretical Physics, 63, Part I: 14571473; Part II: 18331847. MacKinnon, E. (1977): Heisenberg, Models, and the Rise of Matrix Mechanics, Historical Studies in the Physical Sciences, 8, 137188. MacKinnon, E. (1982): The Truth of Scientific Claims, Philosophy of Science, 49, 437462. Marcuse, H. (1968): One Dimensional Man, Sphere Books, London. Marx, K. (1959): Theses on Feuerbach, in: K. Marx, F. Engels, Basic Writings on Politics and Philosophy (ed. L.S. Feuer), Anchor Books, New York, 243245. McCarthy, T. (1978): The Critical Theory of Jrgen Habermas, The MIT Press, Cambridge Massachusetts. McCarthy, T. (1982): Rationality and Relativism: Habermas Overcoming of Hermeneutics, in: Thomson, Held (1982), 5778. McCormmach, R. (1970): Einstein, Lorentz, and the Electron Theory, Historical Studies in the Physical Sciences, 2, 4187. Messiah, A. (1969): Quantum Mechanics, North-Holland Publishing Company, Amsterdam (2 Volumes). Meyer-Abich, K. (1965): Korrespondenz, Individualitt und Komplementaritt, Franz Steiner Verlag, Wiesbaden.

197

Newton-Smith, W. (1980): The Underdetermination of Theory by Data, in: R. Hilpinen (ed.), Rationality in Science, Reidel, Dordrecht, 91 110. Newton-Smith, W. (1982): Relativism and the Possibility of Interpretation, in: Hollis, Lukes (1982), 106122. Nickles, T. (1973): Two Concepts of Intertheoretic Reduction, The Journal of Philosophy, 70, 181201. Nickles, T. (1977): Heuristics and Justification in Scientific Research: Comments on Shapere, in: F. Suppe (ed.), The Structure of Scientific Theories, University of Illinois Press, Urbana, 571589. Nola, R. (1980): Fixing the Reference of Theoretical Terms, Philosophy of Science, 47, 505531. Panofsky, W.K.H., Phillips, M. (1962): Classical Electricity and Magnetism, Addison-Wesley, Reading, Massachusetts. Park, J.L. (1968): Quantum Theoretical Concepts of Measurement, Philosophy of Science, 35, Part I: 205231; Part II: 389411. Pauli, W. (1926): On the Hydrogen Spectrum from the Standpoint of the New Quantum Mechanics, translated and reprinted in Van der Waerden (1967), 387415. Pekelharing, P. (1982): Cyklopiese filosofie, Krisis, nr. 9, 2852. Piaget, J. (1973): Psychologie en kennisleer, Het Spectrum, Utrecht. Polanyi, M. (1967): The Tacit Dimension, Routledge and Kegan Paul, London. Popper, K.R. (1959): The Logic of Scientific Discovery, Harper and Row, New York. Post, H.R. (1971): Correspondence, Invariance and Heuristics, Studies in History and Philosophy of Science, 2, 213255. Puntel, L.B. (1978): Wahrheitstheorien in der neueren Philosophie, Wissenschaftliche Buchgesellschaft, Darmstadt. Putnam, H. (1975a): Explanation and Reference, in: H. Putnam, Mind, Language and Reality. Philosophical Papers, Vol. 2, Cambridge University Press, Cambridge, 196214. Putnam, H. (1975b): The Meaning of Meaning, in: H. Putnam, Mind, Language and Reality. Philosophical Papers, Vol. 2, Cambridge University Press, Cambridge, 215271. Putnam, H. (1978): Meaning and the Moral Sciences, Routledge and Kegan Paul, Boston. Putnam, H. (1983): Realism and Reason. Philosophical Papers, Vol. 3, Cambridge University Press, Cambridge.

198

Radder, H. (1978): Objektiviteit en waarheid in de kwantummechanika. In konfrontatie met Bohr en Habermas, Masters Thesis, University of Amsterdam. Radder, H. (1979): Bohrs filosofie van de quantummechanica. Analyse en kritiek, Kennis en methode, 3, 411432. Radder, H. (1980): Realisme in natuur- en menswetenschappen. Een bespreking van het werk van Roy Bhaskar, Krisis, nr. 2, 6181. Radder, H. (1982a): An Immanent Criticism of Lakatos Account of the Degenerating Phase of Bohrs Atomic Theory, Zeitschrift fr allgemeine Wissenschaftstheorie, 13, 99109. Radder, H. (1982b): Rationaliteit en natuurwetenschap in de angelsaksiese wetenschapstheorie, Krisis, nr. 7, 8689. Radder, H. (1982c): Between Bohrs Atomic Theory and Heisenbergs Matrix Mechanics. A study of the Role of the Dutch Physicist H.A. Kramers, Janus, 69, 223252. Radder, H. (1983): Kramers and the Forman Theses, History of Science, 21, 165182. Radder, H. (1986): Experiment, Technology and the Intrinsic Connection between Knowledge and Power, Social Studies of Science, 16, 663683. Ravetz, J.R. (1971): Scientific Knowledge and its Social Problems, Oxford University Press, Oxford. Rip, A. (1982): The Development of Restrictedness in the Sciences, in: N. Elias, H. Martins, R. Whitley (eds.), Scientific Establishments and Hierarchies, Reidel, Dordrecht, 219238. Rorty, R. (1979): Transcendental Arguments, Self-Reference and Pragmatism, in: Bieri, Horstmann, Krger (1979), 77103. Rosenfeld, L. (1965): The Measuring Process in Quantum Mechanics, Supplement of the Progress of Theoretical Physics, 222231. Rosenfeld, L., Rdinger, E. (1967): The Decisive Years 19111918, in: S. Rozental (ed.), Niels Bohr. His Life and Work as Seen by his Friends and Colleagues, North-Holland Publishing Company, Amsterdam, 3873. Scheibe, E. (1973): The Logical Analysis of Quantum Mechanics, Pergamon Press, Oxford. Serwer, D. (1977): Unmechanischer Zwang: Pauli, Heisenberg, and the Rejection of the Mechanical Atom, 19231925, Historical Studies in the Physical Sciences, 8, 189256. Selleri, F., Tarozzi, G. (1981): Quantum Mechanics, Reality and Separability, Rivista del Nuovo Cimento, 4, 153.

199

Shapere, D. (1977): Scientific Theories and their Domains, in: F. Suppe (ed.), The Structure of Scientific Theories, University of Illinois Press, Urbana, 518565. Shapere, D. (1980): The Character of Scientific Change, in: T. Nickles (ed.), Scientific Discovery, Logic, and Rationality, Reidel, Dordrecht, 61101. Shapin, S. (1979): Homo Phrenologicus: Anthropological Perspectives on an Historical Problem, in: B. Barnes, S. Shapin (eds.), Natural Order. Studies of Scientific Culture, Sage Publications, Beverly Hills, 4171. Shapin, S. (1982): History of Science and its Sociological Reconstructions, History of Science, 20, 157211. Sklar, L. (1976): Space, Time and Space-Time, University of California Press, Berkeley. Stapp, H.P. (1972): The Copenhagen Interpretation, American Journal of Physics, 40, 10981116. Thomson, J.B., Held, D., eds. (1982): Habermas: Critical Debates, The MIT Press, Cambridge Massachusetts. Tomonaga, S.-I. (1962): Quantum Mechanics, Vol. I, North-Holland Publishing Company, Amsterdam. Tuomela, R. (1978): Erklren und verstehen menschlichen Verhaltens, in: K.-O. Apel, J. Manninen, R. Tuomela (eds.), Neue Versuche ber erklren und verstehen, Suhrkamp, Frankfurt, 3058. Vries, G. de (1982): De ontwikkeling van wetenschappelijke kennis, sociologisch beschouwd, Kennis en Methode, 6, 190220. Waerden, B.L. van der (1967): Sources of Quantum Mechanics, NorthHolland Publishing Company, Amsterdam. Wigner, E.P. (1970): Symmetries and Reflections, The MIT Press, Cambridge Massachusetts. Wittgenstein, L. (1974): ber Gewissheit/On Certainty, Blackwell, Oxford. Woolgar, S. (1981): Interests and Explanation in the Social Study of Science, Social Studies of Science, 11, 365394. Wright, G.H. von (1971): Explanation and Understanding, Routledge and Kegan Paul, London. Zahar, E. (1983): Logic of Discovery or Psychology of Invention?, British Journal for the Philosophy of Science, 34, 243261.

200

Index of names

Apel, K.-O. 179 Aspect, A. 88, 183 Austin, J.L. 15 Bader, V.-M. 190 Ballentine, L.E. 189 Balmer, J.J. 127, 128, 130, 145 Barnes, B. 181, 185 Baum, M. 179 Beckermann, A. 35, 179 Bell, J.S. 190 Belinfante, F.J. 189 Bennett, J. 179 Benton, T. 182 Berthollet, C.L. 119 Bhaskar, R. 5, 41, 65, 67, 77 79, 103, 179, 181184, 190 Biot, J.B. 119 Boers, C. 190 Bhme, G. 177 Bohr, N. 86, 92, 99, 110, 125 151, 163169, 177, 181, 187190 Boon, L. 179, 180 Born, M. 126, 136139, 140 144, 149, 188189 Boyd, R.N. 97, 184 Boyle, R. 53, 85 Bub, J. 190 Bunge, M. 185, 188 Collins, H.M. 182, 185 Comte, A. 176 Compton, A.H. 86 Cooke, R.M. 90, 184

Dalibard, J. 183 Daneri, A. 166, 167, 189, 190 Derksen, A.A. 86, 110, 135, 182, 183, 185, 186, 190 Devitt, M. 184 Dingler, H. 21 Dirac, P.A.M. 2, 93, 96, 97, 116, 120, 121, 151 Doorman, S.J. 121, 184, 186 Doppler, C. 86 Doyal, L. 173, 190 Duhem, P. 69 Ehrenfest, P. 188 Einstein, A. 85, 86, 95, 118, 127, 132, 145, 190 Epstein, P.S. 131, 187 dEspagnat, B. 158161, 169, 183, 189, 190 Eyck, J. van 185 Feuerbach, L. 176 Feyerabend, P. 74, 151 Feynman, R.P. 153 Field, H. 113, 182, 184, 186 Fine, A. 185 Forman, P. 140, 188 Fowler, A. 148 Franck, J. 146, 147 Fransen, M. 189, 190 Galilei, G. 153 Garfinkel, H. 182 Georgi, H. 184 Gibson, Q. 183 Goodman, N. 120

201

Habermas, J. 159, 65, 76, 78, 80, 81, 83, 90, 106, 115, 118, 119, 170172, 176183, 190 Hacking, I. 119, 120, 176, 179, 182, 184186 Haleber, R. 179 Hamlyn, D.W. 179 Hanson, N.R. 74, 151, 188 Hardin, C.L. 110, 184, 185 Harris, R. 173, 190 Harvey, B. 87, 88, 183 Heilbron, J.L. 187 Heisenberg, W. 118, 126, 136140, 146, 186, 188 Hempel, C.G. 28, 178 Hendry, J. 138, 183, 188 Hertz, G. 146, 147 Hesse, M. 56, 98, 181, 184 186 Hollis, M. 185 Holt, R.A. 84, 87 Honner, J. 189, 190 Hund, F. 187 Jammer, M. 137, 183, 187 189 Janich, P. 43, 59, 63, 179, 181 Jeans, J.H. 129, 187 Jongeling, B. 181 Kant, I. 8, 9, 18, 35, 40, 179, 181 Kepler, J. 127, 128, 131, 143 Keulartz, J. 186, 190 Keuth, H. 32, 178, 180 Knorr-Cetina, K.D. 183 Koningsveld, H. 178, 181 Korthals, M. 177 Koyr, A. 122, 176, 186 Kragh, H. 187 Kramers, H.A. 126, 132134, 136138, 140, 141, 187, 188 Kripke, S. 185, 188 Kuhn, T.S. 74, 150, 151, 176, 178, 181184, 186188

Kunneman, H. 53, 177181, 190 Kwa, C.-L. 184, 190 Lakatos, I. 74, 86, 183, 188 Laplace, P.S. de 119 Latour, B. 182, 190 Laudan, L. 100, 115, 184 186 Lavoisier, A. 119 Leplin, J. 135, 184, 185 Lvy-Leblond, J.-M. 182 Lieshout, P. van 182184 Livingston, E. 182 Lobkowicz, N. 179 Loinger, A. 166, 167, 189, 190 Lorentz, H.A. 2, 9193, 96, 97, 109, 110, 116, 120, 129, 151, 153, 184, Lorenzen, P. 21 Lukes, S. 185 Lynch, M. 182 Mach, E. 176 Machida, S. 189 MacKinnon, E. 151153, 188, 189 Marcuse, H. 13, 176 Marx, K. 9, 13, 176 McCarthy, T. 13, 176180 McCormmach, R. 184 McMullin, E. 86 Messiah, A. 188 Meyer-Abich, K. 187, 188 Miller, D.C. 84 Mol, A. 182, 184 Mulkay, M. 183 Namiki, M. 189 Newton, I. 85, 118 Newton-Smith, W. 97, 101, 184, 185 Nickles, T. 185, 188 Nola, R. 135, 185, 186 Oseen, C.W. 188 Oppenheim, P. 28, 178

202

Panofsky, W.K.H. 183 Park, J.L. 189 Paschen, F. 133 Pauli, W. 136, 139, 140, 188 Peirce, C.S. 35, 39, 97, 179 Pekelharing, P. 184, 185 Phillips, M. 183 Piaget, J. 18, 19, 28, 29, 39, 177, 178 Pickering, E.C. 148 Pipkin, F.M. 87, 88 Planck, M. 85, 127129, 187 Podolski, B. 190 Poincar, H. 92, 184 Polanyi, M. 182 Popper, K.R. 74, 97, 177 Post, H.R. 189 Priestley, J. 112114, 116, 118, 119 Prosperi, G. 166, 167, 189, 190 Puntel, L.B. 32, 178180 Putnam, H. 6, 9799, 110 112, 115, 119121, 135, 184186 Quine, W.V. 173 Ramsey, F.P. 23 Ravetz, J.R. 190 Rayleigh, Lord 129, 187 Rip, A. 182 Roger, G. 183 Rorty, R. 41, 179 Rosen, N. 190 Rosenberg, A. 110, 184, 185 Rosenfeld, L. 167, 188, 190 Rutherford, E. 99, 127, 187 Rdinger, E. 188

Rydberg, J.R. 128 Scheibe, E. 189 Schndelbach, H. 22 Schrecker, F. 7174 Schwarzschild, K. 130, 131 Searle, J. 15 Sellars, W. 97 Selleri, F. 190 Serwer, D. 188 Shapere, D. 183, 185 Shapin, S. 87, 179, 183 Sklar, L. 184 Slater, J.C. 137, 188 Sommerfeld, A. 130, 131 Stapp, H.P. 189 Stark, J. 131, 133, 148 Strawson, P. 23 Tarozzi, G. 190 Thales 112, 116, 118 Tomonaga, S.-I. 187, 188 Toulmin, S. 27 Tuomela, R. 181 Vleck, J.H. van 136, 137 Vries, G. de 182, 184 Waerden, B.L. van der 188 Weber, W.E. 184 Wigner, E.P. 189 Wilkerson, T.E. 179 Wittgenstein, L. 178 Woolgar, S. 182, 183 Wright, G.H. von 5, 6365, 67, 68, 78, 181, 182 Zahar, E. 185 Zeeman, P. 91, 138, 148

203

Вам также может понравиться