#cparse("/super/config/super.config.vm") #cparse("${superIncludes}/super.before-doctype.fhtml") #cparse("${superIncludes}/super.legacy-doctype.fhtml") #cparse("${superIncludes}/super.head-top.fhtml") Nature World Conference on Science #cparse("${superIncludes}/super.head-bottom.fhtml") #cparse("${superIncludes}/super.body-top.fhtml")
to nature home page World Conference on Science
 
home
search

introduction news opinion meetings



Part 2

Table of contents

Forum I - Science: Achievements, Shortcomings and Challenges
1.1 The nature of science
1.2 The universal value of fundamental science
1.3 The scientific approach to complex systems
1.4 International co-operation in science
1.5 The teaching of science

Forum II - Science in Society
2.1 Public perception of science: between acceptance and rejection
2.2 Science for development
2.3 Setting priorities in the new socio-economic context
2.4 Science: the gender issue
2.5 A new social contract for science

[Back to Part 1]

Forum I - Science: Achievements, Shortcomings and Challenges

1.1 The nature of science
Science is a systematic means of obtaining knowledge about the world. One of the basic facts underlying this endeavour is the observation that the world exhibits order. This observation seems to be shared by all cultures. An important step in understanding the order of the natural world consists in systematically describing its phenomena. Of the innumerable and endlessly variable phenomena, some exhibit their own order and can be classified into groups. Modern science, which originated in the 17th century, tends to depict this order in a special way, namely by positing laws of nature. Laws of nature are general regularities that hold between classes of events. Such regularities form the basis of scientific predictions and scientific explanations and are an essential part of scientific theories. Even those theories which scientists refer to as 'models' posit regularities of some sort or another. Thus, one of the core activities of science consists in the classification of events and the discovery of general regularities among such events in order to explain them and to predict their behaviour.

Many scientific explanations are reductionist; that is, an explanation of the behaviour of a system is sought in terms of its components and the laws governing their behaviour. This is why the scientific approach has often been called analytic. In a reductionist research strategy, systems are analysed into their components, their configurations, and their interactions in order to understand the system as a whole. Such strategies are a very powerful systematic means of directing research. Even when reductive explanations are unsuccessful, they usually lead to interesting results and foster some knowledge of the systems in question. Whether or not this reductionist approach can always be completely successful is a controversial questions. There may be systems whose behaviour cannot be fully understood in terms of their component parts in principle (see section 1.3).

Human beings are prone to error, prejudice and superstition. However, human beings are also capable of learning from error, and science employs systematic means to do so. Thus, science is not only systematic in displaying orderly structures of nature, but also in how it establishes knowledge claims and improves their precision. A large variety of critical methods designed to locate and contain different kinds of errors have been developed. As science is a human endeavour, it cannot entirely eliminate error, but it can minimise the probability of errors and evaluate their magnitudes.

In order to produce knowledge, science employs a curious mix of speculative and critical elements. The speculative element is necessary in order to find general regularities because they are not easy to identify and cannot simply be read off the phenomena themselves. To explain observed regularities, science often posits barely observable or even outright unobservable entities. At first, their very existence may even be highly dubious. Well-known examples of such initially speculative entities are atoms and genes. Such entities, and the properties attributed to them, must then demonstrate their existence either by revealing themselves indirectly in a broad variety of different phenomena, or by becoming observable through the development of new means of observation.

Since antiquity, the unaided observation of phenomena has been one means of making discoveries and controlling theoretical knowledge claims. But modern science has invented additional means to these ends, such as increasingly powerful instruments which mediate observation. Moreover, not only can science provide the human senses with artificial aides, but it has also discovered means of observation for which we lack sense organs altogether. A classical example is radio waves, which allow astronomers to explore the depths of outer space. The other principal means of knowledge acquisition is the scientific experiment. While all cultures have used some sort of experimentation, or 'trial and error' procedure, for improving their technologies, the systematic use of experiments for the acquisition of theoretical knowledge is a fundamental novel aspect of modern science. Roughly, experiments can be classified into two classes. First, the so-called 'explorative' experiments discover novel kinds of phenomena or hitherto unknown connections between different phenomena. In this way, they can suggest new avenues for further empirical and theoretical investigation. Second, experiments are used to test specific hypotheses about general regularities. For example, by systematically varying the different factors influencing the behaviour of a given system, and recording the system's response, experimenters are capable of distinguishing between genuine causal relations and mere correlations.

For several centuries, there has been a general belief in the existence of a specific scientific method which ensures the reliability and excellency of scientific knowledge. This idea dates back to some of the pioneers of modern science such as René Descartes and Isaac Newton. They suggested that scientific knowledge could only be gained by adherence to a set of absolutely binding rules, later dubbed 'The Scientific Method'. Since the late 19th century, however, a different picture of the advancement of science has emerged. This new account is mainly based on detailed historical research into scientific research processes. According to this new picture, science proceeds from the knowledge that it has already produced. In particular, eminent solutions to research problems function as exemplars for the identification and solution of other problems. Thus, in many cases, the advancement of science can be seen as a self-amplifying process in which existing knowledge forms the basis for new knowledge. In each particular field, continuous research traditions typically emerge. But the productive potential of some body of knowledge to guide a research tradition may eventually exhaust itself, and fundamental changes may be necessary in order to secure further progress. These changes take place in scientific revolutions when a fundamentally new point of view is generated which can transform the conceptual foundations of a discipline. Most notably, chemistry, biology and physics have seen such revolutionary transformations in the 18th, 19th, and 20th centuries respectively. In this new picture of science, the reliability of scientific knowledge is guaranteed by the precise focus and depth of research. If there is a hidden mismatch between nature and the theories underlying research, this mismatch will reveal itself in the research process. Eventually this will necessitate substantial corrective changes in the theories involved. In this way, the research process itself is capable of detecting and locating errors in the theory-nature fit.

The public character of scientific knowledge and its in-built mechanisms of self-correction and expansion distinguish science from most traditional forms of knowledge. Many cultures have developed highly sophisticated systems of knowledge, especially in astronomy, natural products, medicine, and mathematics, to mention just a few. But traditional knowledge has often been restricted; for instance to the leaders of certain religious elites, making broad dissemination of this knowledge impossible. More importantly, there does not seem to have been a systematic mechanism for effectively checking the reliability of knowledge and securing its growth into new areas. These undoubtedly unique qualities of scientific knowledge should not lead to an uncritical dismissal of traditional knowledge. In some areas, especially in medicine, there are still stocks of traditional knowledge not yet understood or even considered by science, but which nevertheless remain extremely useful, especially with regard to practical application. For many cultures, much traditional and popular knowledge has been and continues to be essential to survival. Furthermore, it should not be forgotten that traditional knowledge has contributed to the very development of modern science and an interaction between the two can be productive for all parties. However this does not warrant the conclusion that we could do without science, as some anti-science movements suggest. All countries can ignore science only at their own peril. It can be an important weapon in our fight against ignorance, poverty, superstition, and diseases and should be recognised as such.

The growth of science over the last 400 years displays an amazing increase in diversity. Even a very rough classification of the sciences specifies several hundred different special disciplines. The advantage of specialisation is obvious: it makes in-depth knowledge of a particular domain of phenomena possible. However, there are also some disadvantages of specialisation. The more the fragmentation into specialities and sub-specialities of science progresses, the more difficult communication between these specialities becomes. And as much research into urgent problems has to draw on resources from different disciplines and sub-disciplines, the fragmentation of science can inhibit progress. Fragmentation also creates communication problems between science and the public.

However, there are two main tendencies in scientific development which counteract the tendency toward ever-increasing diversity. The first tendency is the development of increasingly comprehensive theories - physics and biology provide telling examples. Seemingly disparate fields of science become connected by these very general theories. Thus, over and above the increase in diversity among scientific fields, a conceptually unifying net of theories tends to develop. A second unifying tendency of modern science is the internally driven inter-disciplinary research that leads to a growing overlap in the fundamental sciences such as biology, chemistry, and physics. Especially in research in molecular science, this inter-disciplinary overlap is clearly visible. New knowledge gained about as yet unknown molecular mechanisms will strongly influence the future of health research, the environmental sciences, and research on novel materials, all of which will continue to have a deep impact on society. [Back to Table of Contents]

1.2 The universal value of fundamental science
For the most part, there are two different sources of the problems that science tackles. One source is science itself. Because of its systematic character (see section 1.1), science generates its own questions, both with respect to content and with respect to method. The attempt to explain some phenomena systematically, to establish knowledge claims systematically, and to improve their precision within their domain - each of these generates certain questions that must be addressed. This is the domain of what is called fundamental (or basic) science. Fundamental science can be summed up as the generation of new knowledge. It tackles questions that are generated by the science system itself. The developmental dynamics of fundamental science is thus mainly driven from within science. This process depends on various resources and can be influenced by them (see section 2.3). In particular, new technological resources like better experimental apparatus, measuring instruments with higher resolution, or faster computers can make intrinsically interesting problems accessible which were previously out of reach, thus opening up new frontiers for research.

By contrast, the other important source of scientific problems is the social environment, or some subsystems of it, in which scientific research finds itself. Any given society has many problems to solve, be it on a national or an international level. For many of these attempted solutions, society turns to science. Obviously, some of these problems may be tackled with the scientific resources already at hand. In these cases, the term 'applied science' is fully adequate. By applied science, we simply mean the use of already existing scientific knowledge which is sufficient for solving a given problem. Here, there is often a smooth, but often time consuming and costly, transition between applied science and the commercial development of new products.

At this point, the following questions arise: why should society finance fundamental science? Isn't fundamental science simply a playground for scientists that is otherwise useless for society? Don't we have enough pressing problems that well-educated people like scientists should be trying to tackle without wasting time and resources on fundamental problems largely driven by the scientist's curiosity? Or, to put it more directly, isn't fundamental science a waste of money that even the industrialised countries, let alone the less developed countries, can no longer afford? Aren't those politicians who tend to cut expenses for fundamental science in times of low public budgets on the right track?

In spite of these reservations, there are some compelling reasons why fundamental science is imperative for both the industrialised and less developed countries. First of all, the knowledge needed to solve many of the most pressing problems the world faces does not yet exist. Thus, certain societal needs and desires directly trigger fundamental science. In the 20th century, questions posed to science by society have led to new areas of research and ground-breaking discoveries. And even though the questions are generated from outside of the scientific arena, they do trigger fundamental scientific research. In other words, scientific problems arising as a result of societal needs can lead to research that is focused on the discovery of novel knowledge in a certain domain. The first large-scale example of this was probably the development of the atomic bomb, but research on cancer, nuclear fusion, novel materials and various environmental problems exemplify the same pattern. It is quite likely (and will even be necessary) that the proportion of fundamental research driven by societal needs will increase in the future (see section 2.5). This is because, on a finite Earth with an increasing population, increasingly complex problems are arising which do not fall under the rubric of any particular science. Again, environmental problems provide a host of telling examples. In these circumstances, the know-how to solve these problems cannot simply be drawn from a well-established discipline and then applied to the particular set of problems at issue. Rather, it will be necessary to consult several disciplines, each of which is inadequate by itself. This must include co-operation and contributions from the social sciences. Interdisciplinary co-operation can lead to new solutions, and in this way new interdisciplinary approaches are continually evolving.

Second, apart from this problem-induced form of fundamental research, a case can be made for supporting curiosity-driven fundamental research, in which the application of results to real world problems is not the primary intention, and may not even be foreseeable. First of all, fundamental science simply stimulates the human creative potential. But the real reason why fundamental science is invaluable is found in the countless examples drawn from the history of science in which knowledge produced for its own sake later resulted in socially invaluable technological potential. The isolation of penicillin, for example, was the result of years of basic research into the nature of mould which at first did not appear to have any practical benefits or economic applications at all. Quantum mechanics, to take another example, is a physical theory that was invented because physicists saw various flaws in its predecessor, Newtonian physics, especially because the theory could not explain why normal matter is fairly stable and does not collapse into nothingness. For the lay person, the stability of a stone is a trivial fact of experience which does not need an explanation. It is simply taken for granted. At first glance, scientists who consider this fundamental question seem to be wasting their time. But in fact, this seemingly innocuous and irrelevant question led to innumerable technological applications in material science. Some of these applications have dramatically changed our world, most notably the invention of the transistor. Another striking example of the development of a powerful new technology that arose out of an attempt to answer a basic scientific question is the discovery of restriction enzymes. This discovery opened up the possibility of applying molecular genetic analysis and introducing specific genetic changes to any kind of organisms (see section 1). Genetic engineering is now widely recognised as a key technology for the pharmaceutical and biotechnology industries, as well as an indispensable research tool in almost all of the life sciences. Or consider a whole discipline like botany. Strictly speaking, botany is a fundamental science concerned with investigating the nature of plants. Many aspects of botany, however, have direct importance for human welfare and development. Such fields as forestry and horticulture are closely tied to fundamental botanical studies and others, such as pharmacology and agronomy, still depend on basic botanical knowledge. Thus, in the long run, the possibility of solving problems by scientific means depends on the existence of fundamental theories, methods, and insights that are provided by fundamental scientific research alone. But we should not overlook the fact that the transition of fundamental science to technology is far from automatic. Additional intellectual and institutional means are necessary in order to put the fruits of fundamental research to sustainable practical use.

Thus, with respect to the potential benefits, fundamental research can be seen as a long-term investment. As an investment, fundamental science has the unusual feature that the potential benefits of research are often unforeseeable. A large part of research in fundamental science may never yield any economic returns, but when fundamental science does turn out to be economically rewarding, the benefits can be immense. This element of unpredictability in the economic returns of fundamental science makes goal-directed action and priority-setting complicated and difficult (see section 2.3). The reason for this unpredictability lies in the very nature of fundamental science itself. It arises because fundamental science is mainly aimed at gaining novel knowledge in some field, whereas applied science is oriented towards a goal and tends to work with phenomena which are already well-known. No research foresight strategy can completely eliminate this unpredictability. However, it is clear that this kind of investment must be supported mainly by institutions which have a special responsibility for the long-term prospects of global society. That is why applied science supported by industry is often subsidised by short-term investments directed at very specific product- or service-oriented goals. Market proximity and profitability is affecting research and development financed by the private sector.

Third, scientific knowledge gained through fundamental science is often needed for long-term purposes which have no immediate economic returns. This holds especially for the use of fundamental science in long-term planning, both on a national and an international scale. It is important for governments and inter-governmental panels to be aware of long-term prospects such as climate change or demographic tendencies, which may include a large range of projected consequences in such areas as the health-care system, insurance policies, etc. Again, this holds for all countries, independently of their degree of industrialisation. Hypotheses on these issues must often be based on data collected over long periods of time. This is especially obvious in the case of extrapolation on climatic trends, such as estimates of the prospects of global warming or occurrences of global phenomena like El Niņo, which may in turn have consequences for public health, such as an increase in the rate of diseases like malaria.

Fourth, it is extremely important to note that not all of the benefits of fundamental science come in the form of economically exploitable discoveries. A considerable share of fundamental scientific research is done by young people today, in particular by graduate and post-doctoral students. Some of these young scientists will pursue academic careers. But others will seek employment in industry or in government positions in order to work in research and development, in laboratories providing different kinds of services such as food quality control, medical testing, policy decisions, etc. In other words, society needs scientific experts in all kinds of positions outside fundamental scientific research. But fundamental scientific research is clearly the best form of training for scientific experts. This is due to the fact that it is providing those engaged in it with a sound understanding of state-of-the-art scientific theories, the technical skills needed, as well as the systematic approach required in many areas of the workplace and other societal activities. Thus, fundamental research has an important educational role, which is frequently underestimated. This point will be taken up again in Section 1.5.

Lastly, the cultural aspect of fundamental science should not be overlooked. Science can provide us with an extremely rich picture of our world, from its most minute details right up to the largest objects in the universe. All cultures have developed a desire to know the world in which they live, and science is a particularly strong method of fulfilling this desire. Fundamental science is imperative for every nation, industrialised and industrialising alike. [Back to Table of Contents]

1.3 The scientific approach to complex systems
'Complexity' designates a set of loosely connected scientific ideas having to do with the phenomenon that certain systems exhibit, in spite of being governed by relatively simple laws, a number of unexpected properties. An offshoot of the theory of dynamical systems, complexity has become a subject of considerable scientific interest over the last few decades. Parts of it have become widely known under the fashionable rubric of chaos theory. This shift in focus towards complexity is not so much a consequence of some dramatic new discovery or revolutionary development, since certain essential details have been known for quite some time. Instead, it was primarily the result of advances in computer technology which have allowed scientists to address previously intractable problems. Nevertheless, whether this shift towards complexity constitutes some sort of scientific revolution is controversial.

Complexity research claims a high degree of generality: it is supposed to apply to extremely heterogeneous areas. The unifying idea that binds these heterogeneous areas is that of a complex system. Complex systems typically show different patterns of behaviour than simple systems. Research into complex systems can only meaningfully begin once the simple systems (i.e. parts of complex systems) are more or less understood. There are several areas of science where simple systems were indeed deciphered in the course of the first half of the 20th century. Interest then moved on to more complicated systems. There are several source of such a general idea of a complex system.

The first source was the research into dynamical systems in the context of classical mechanics that started as early as the end of the 19th century. The solar system exemplifies such a dynamical system. The solar system appears to be fairly simple because the planets look as if they would revolve around the sun forever. Yet, mathematical research has shown that it is far from clear that the revolution of planets around the sun will indeed go on forever. For instance, it is quite imaginable that one of the planets gains so much energy from the other planets that it leaves the solar system altogether, leaving the others in states of lower energy such that they circle the sun in lower orbits. The question that results is whether the solar system is stable or not.

A second source of complexity research is computer science and this in a two-fold sense. Computers are the primary tool used for complexity research. For instance, for dynamical systems, exact solutions are hardly ever obtained. This is the reason why the field lay mostly dormant for about half a century. Most of the insights of complexity thinking are gained through the use of computer models of the most diverse kinds. Computer models represent a given situation by abstracting from everything that does not seem relevant to those aspects of the system's behaviour one is interested in studying. Thus, seen as naturalistic representations of a given system, computer models appear as gross misrepresentations. Still, in successful models this does not prevent them from mimicking exactly the relevant dynamical aspects. For instance, in the very active field of artificial life research, most characteristics of the real physiological set-up of actual animals is completely ignored. Only a few traits reminiscent of real animals, like the production of offspring or certain rudimentary forms of locomotion or predation, enter the picture. If they are cleverly chosen, they prove sufficient to mimic certain aspects of the dynamics of a population of real animals. For instance, in a computer model that simulates a population with replication, mutation, and competition among individuals, spontaneous emergence of parasites may occur along with some new phenomena these parasites might generate.

Computer science stimulates complexity research in a second sense. Computations themselves provide a comparatively perspicuous model for the distinction between the simple and the complex, which is at the core of complexity research. Still, it should be recognised that there is no definition of the central notion of complexity that is both wide enough to cover all of the paradigm cases, and narrow enough to exclude trivialities. Further sources of complexity research include cybernetics, information theory, the theories of automata, of autopoiesis, and of molecular self-organisation, as well as systems theory, non-equilibrium thermodynamics, and synergetics.

The area of complexity research emerging from these various sources attempts to be a new, unified way of contemplating nature, human social behaviour, life, and the universe itself. It is an interdisciplinary approach fuelled by sophisticated mathematics, mathematical modelling, and computer simulation. It is inspired by observations that have been made on complex systems in the most diverse fields: meteorology, climate research, ecology, economics, physics, embryology, computer networks, and many more. These systems exhibit behaviour that is strikingly different from that of more simple systems. Typically, the behaviour of complex systems cannot be guessed or calculated on the basis of knowledge of their parts and their composition in the system. In fact, the components of the system interact in a way which severely limits predictability. The limits to predictability come in degrees. Some of these limits can (and will) be overcome by greater computing power and better algorithms. Some limits are of a deeper nature, but could be overcome if we had unlimited computing power and exact calculations. But some limits are of an absolute nature and could not be overcome by any possible means. Thus, complex systems exhibit so-called emergent properties and laws. In other words, they exhibit properties and laws shown only by systems of that degree of complexity that come as a surprise given the knowledge of the system's components and their composition. Complexity research is, therefore, seen by many of its proponents as anti-reductionist (see section 1.1), since new levels with new laws emerge which could not have been predicted by the analytic procedures characteristic of reductionistic research strategies.

One of the key processes responsible for the surprising behaviour of complex systems is self-organisation. This is the emergence of an orderly behaviour of some or all components of the system; in other words, some co-ordination among them. The crucial point is that this co-ordination is not brought about by some force or influence acting on the system as a whole, but by the interaction of the components which leads to this collective effect under certain circumstances. Self-organisation is a paradigm case for the development of order out of disorder. Typically, the emergence of new order happens in systems that are neither too orderly (like crystals), nor too disorderly (like turbulent fluids). Metaphorically speaking, it happens in systems 'on the edge of chaos'.

As a result of self-organisation, complex systems may exhibit spontaneous transitions into new states without apparent macroscopic causes. The reason for this is either that minuscule outside influences may bring about huge effects, or that the system's intrinsic instability drives it in some direction. Of particular interest are complex adaptive systems that occur in various sciences such as economics (for example the economy of a certain region), ecology (the ecosystem of a pond), biology (the immune or nervous system of an organism, the development of an embryo), and artificial intelligence (computer networks), to name just a few. These systems adapt to changes in their environment in an often extremely surprising way. In these cases, the idea of complexity research is that there must be some general common principles governing this sort of adaptive behaviour.

One of the most intriguing features of complexity is the fact that very complex behavioural patterns of a system can be generated by the rather simple mathematical rules the system follows. Many dynamical models begin by replacing the continuous flow of time by a set of equidistant points in time. The system's behaviour is then modelled as a series of discrete states. This series is generated through the repetitive application of a fairly simple rule starting with some initial state. Even if that rule is fairly simple, extremely complex and surprising behavioural patterns which do not seem to be built into the design of the rule can result. Typically, these transformation rules are non-linear. Non-linearity is a precise mathematical concept that can be explained as follows. A system is linear if that system's behaviour can somehow be described by a proportion. For instance, if a system's response to a disturbance is linear, then the response will increase with an increase in the disturbance and it will decrease with a decrease in the disturbance. If a behaviour over time is linearly dependent on its initial conditions, then small changes of the initial conditions will result in small changes of the system's behaviour over time. In non-linear systems, these properties that make the behaviour of linear systems easy to predict do not hold. Small changes may have disproportionally huge effects. The so-called butterfly effect catches this point nicely. Due to the extreme non-linearity of the global weather system, the disturbance caused by a single butterfly in Africa could result in a tornado in North America thirty days later. Thus, the properties of complex systems mentioned above, such as unpredictability, the emergence of new properties and laws, and self-organisation, are all related to the non-linearity of those systems.

Sometimes, highly enthusiastic pronouncements about complexity research are heard. It has been claimed that complexity even offers an entirely new world view which has the potential to settle some large issues; for example, how did the world become so complicated? Why is there so much order and structure in a world with so much instability? Why does innovation seem to thrive at the boundary separating order and disorder? It remains to be seen whether these promises will be fulfilled. [Back to Table of Contents]

1.4 International co-operation in science
Science is a social enterprise dependent on communication and co-operation among scientists. Communication has a two-fold function in science. It is necessary both in order to avoid wasteful duplications of research efforts (this is accomplished through the quick dissemination of research results), and to ensure that systematic criticisms of any claim to scientific knowledge can be made through independent evaluation. Science's specific claim to knowledge includes its objectivity, and objectivity implies intersubjectivity. This means that the validity of the results of scientific research should be independent of factors such as gender, ethnicity, age, and nationality, as well as any other distinguishing characteristics of the researchers involved. Thus, according to the nature of science, there should not be any national barriers which hinder the dissemination of research results and their critical evaluation. Furthermore, because the properties of the systems and objects which are studied by many different fields are of a universal nature, such as properties of matter, principles of life, etc., the world-wide exchange of data, knowledge, and ideas is to the advantage of researchers around the world.

As a matter of fact, few enterprises in the world are as thoroughly internationalised as science. In many scientific disciplines, the leading laboratories and institutions are scattered over different parts of the world, but they exchange personnel, ideas, and research materials. There are innumerable international science organisations, including international field-specific unions of scientists, which unite national science organisations. These international unions share an umbrella organisation, ICSU (International Council for Science), together with official national representatives. Research results are published in a growing number of international journals. In most institutions, the credentials of scientists are judged according to how well their work is represented in such journals. Scientific consensus, where it exists, transcends any national, cultural, or continental borders. Where it doesn't exist, the lack of consensus usually has nothing to do with national mentalities or cultural differences, at least not for the last 50 years or so. But even earlier, there are many known cases of scientists trying to collaborate with colleagues in an enemy state during a state of war. Science allows people from very different cultural backgrounds to communicate and share ideas in the interests of the common good.

Another reason for the international co-operation in science which has steadily continued to develop over the last few decades is simply the sheer size of many large-scale projects. The size and expense of these projects has made it simply quite unfeasible for many nations to maintain scientific research activities in a growing number of fields; unfeasible, that is, unless they enter into co-operative arrangements for the construction and operation of expensive scientific facilities. Perhaps the most well-known example of such co-operation is the planning and construction of the international space station. But right here on Earth, high-energy particle accelerators, the human genome project, and many forms of global environmental research are all examples of enormous projects which require international co-ordination and co-operation.

The most important reasons for international co-operation have been recognised over the last 20 years. The detrimental effects of human activities have become so widespread and intensive that they are affecting the environment on a global scale. Such changes in atmospheric composition, in land cover, and in the oceans, as well as related climate changes and diminishing biodiversity are now beyond dispute and are collected under the rubric 'global environmental change' or 'global change'. Understanding and addressing global change requires a truly international scientific effort of unprecedented co-operation and interdisciplinarity. In response to this challenge, UNESCO and ICSU have sponsored or co-sponsored the World Climate Research Programme (WCRP), DIVERSITAS, a programme on biodiversity science, the International Geosphere-Biosphere Programme: a Study of Global Change (IGBP), and HDP, a programme that addresses the Human Dimensions of Global Environmental Change, to name just a few. Moreover, these programmes co-operate where suitable, especially in the interfaces between the natural and the social sciences. There have also been many international co-operative conventions, mainly as a response to global environmental degradation. These include the Montreal Protocol on Substances that Deplete the Ozone Layer (1987), the Basel Convention on the Transboundary Movement of Hazardous Wastes and Disposal (1992), conventions on Biological Diversity (1992), Climate Change (1992), and Desertification (1994), again to name just a few.

All nations share a vital interest in monitoring global change, and they also have much to contribute to the common understanding of the Earth system. Satellites can scan the globe, but much of the information needed by researchers must be obtained locally from the land or the oceans. Whilst laboratories, data banks, and computers produce impressive analyses, local observations and insights are needed to bring theses analyses to life. Only studies focused on regional and local conditions can adequately assess the real implications of global environmental changes on a global scale. For instance, the Global Change System for Analysis, Research and Training (START), co-sponsored by ICSU and UNESCO (among others), is the international scientific community's response to the need for research on regional environmental change. START promotes interdisciplinary research at a regional level by developing research networks. The purpose of these networks is to assess regional impact and to provide regionally important information. The START initiative also helps build endogenous capacities in the developing regions of the world so that they can participate effectively in the various scientific projects of the global change research programmes. Thus, regional research networks are instrumental in mobilising resources to augment existing scientific capabilities and infrastructure - something of particular importance for developing countries because the developing countries also influence global change and are especially sensitive to it. Further objectives of START are the enhancement of communication between researchers and the strengthening of data and information-system capabilities supporting the regional research networks.

The growing need for international collaboration and the novel characteristics of many of these enterprises raise several policy issues which deserve careful consideration. First of all, it is important to recognise that international co-operation on fundamental scientific research projects has not grown proportionately with the internationalisation of emerging science and technology issues facing the global community today. Nor has international co-operation increased proportionately with the spread of scientific competence and the advent of new information and communication technologies. Given the large incentives and potential benefits that large-scale collaboration can provide, why isn't there more international collaboration in scientific and technological research? How can the potential for free riding be reduced without jeopardising knowledge distribution? Who will take the initiative and who will manage participation in international projects involving sovereign countries?

Because nearly all scientific research is still funded, organised, and implemented at the national level, international co-operation presents a major challenge to the global community. In order to help build coherent efforts out of basically national research projects, the international programmes have adopted a 'value-adding' approach aimed at knitting together the contributions from individual projects in order to address larger issues. This includes building consensus on research priorities and plans, and co-ordinating the use of expensive infrastructure in order to achieve effectiveness and resource efficiency. In addition, this approach supports the creation and development of international, inderdisciplinary research networks, common experimental protocols, standardised methodologies and data. It also supports the comparison of models, the integration and synthesis of global change research, and the timely and appropriate dissemination of knowledge to the policy and resource management sectors.

This last activity is a major challenge facing the international global change programmes. Global change itself is a manifestation of current unsustainable development undertaken by societies around the world. The transition to sustainability must be based on a sound scientific understanding of the global systems and their effects on human beings.

There appear to be enormous opportunities for constructive change in global scientific research management. The increasing use of new information and communication technologies will facilitate information transfer and allow international research networks to develop. There seems little doubt that, faced with global challenges, these enhanced technologies will lead to a growing clamour for reform aimed at increasing integration of resource allocation, research planning, and information distribution. One of the most significant challenges facing the international community in the coming century will be the development of management and policy mechanisms for the co-ordination of the international scientific community in order to assure that the division of labour reflects abilities to contribute to co-operative scientific projects rather than national interests. Of course, a primary aim should be to develop realistic and practical policies to reduce the increasing disparities and disadvantages that already exist.

The new information and communication technologies have modified the techniques of scientific investigation. They have provided researchers with the tools for simulating, controlling, recording, and analysing huge quantities of data. In economic terms, they have helped reduce the costs of scientific research by making data collectively available - data which is often very expensive to gather or produce. ICSU has established a Committee on Data for Science and Technology (CODATA) which is concerned with the organisation, management, quality control, and dissemination of scientific and technical data. Collective data sharing is leading to new institutional configurations and establishing electronic relations between researchers around the globe. In short, science is developing into a test-bed for many complex technical, economic, social, and organisational issues concerning international communication and information distribution.

But the tensions emerging out of these new developments should not be overlooked. Science needs unrestricted access to data world wide. The private sector, however, has a strong interest in protecting data in some areas. But in other areas, it has an equally strong interest in ensuring the free collection of data of various kinds. On the other hand, individuals have a desire and right to protect their privacy. Actions have been taken by the World Intellectual Property Organisation, the European Union, and the United States aimed at introducing new intellectual property laws. Data bases are not covered by copyright because they do not meet the test of creativity in the arrangement of data. Sectors of the information industry, however, believe that a new copyright clause is needed to protect their investment in creating databases and to guard against piracy. Extending these rights could impose serious constraints on science and education, undermining the ability of researchers and educators to access and use scientific data. It would make it more difficult for scientists to compile global or regional data bases, or to re-use and re-combine data for publication or instructional purposes. The trend toward commercialisation of scientific data is a cause of great concern for all developing countries because it counteracts access to scientific knowledge and data which is indispensable.

In addition, because of rising costs, the mainstream scientific literature which is produced in a few countries has increasingly become inaccessible to scientists, students, and even libraries in a growing number of countries. Scientific publishing has become a big business and a commodity only for those countries who can afford it. This increases the importance of efforts aimed at increasing participation in science publishing by all countries. The potential benefits of modern information technologies are not being sufficiently put to the service of the global scientific community. Instead, they are being used to the economic advantage of just a few enterprises. In 1992, ICSU, in co-operation with UNESCO, established the International Network for the Availability of Scientific Publications (INASP), in order to address many of these challenges. Its objectives are to support and strengthen existing programmes for the distribution, publication, exchange and donation of books and journals, to encourage new initiatives to improve the availability of quality scientific literature, and to help establish the sustainable exchange and distribution of scientific publications.

It is difficult for scientists in less developed countries to compete with colleagues in more affluent nations, as they often lack adequate resources. Laboratory equipment, journals, software, and other necessary items tend to carry Western price-tags. The same is true of page charges in many scientific journals. In addition, the dominance of the English language in the international scientific community conveys an advantage to scientists from English-speaking countries and from Europe. All of these factors lead to a marginalisation of the developing world within science, and consequently to a very low representation of developing countries in the leading international journals. To make matters worse, a further consequence is that assessments of scientific productivity solely relying on citation analysis drastically underestimate the research output of developing countries, as the scientometric institutes mainly index international journals. South-South co-operation is one strategy for addressing these problem. For instance, the Third World Network of Scientific Organisations, in collaboration with the Third World Academy of Sciences and with the participation of the Centre for Science and Technology of the Non-aligned and other Developing Countries, has produced many useful publications including Profiles of Institutes for Scientific Exchange and Training in the South. The African Energy Policy Research Network (AFREPREN) is a good example of a successful African network coordinated and run entirely by Africans. The main conclusions to be drawn are that there are still many challenges which need to be met before the ideal of science as truly international can be realised, and international co-operation must be used more effectively, especially to counteract monopolist trends and to ensure a wider and more democratic distribution of the resources and products of scientific activity. [Back to Table of Contents]

1.5 The teaching of science
It is evident that human resources constitute the ultimate basis for the wealth of nations. Whereas capital and natural resources are passive factors of production, human beings are the active agents who accumulate capital, exploit natural resources, build social, economic, and political organisations, and carry forward national development. Furthermore, as the 1987 report of the World Commission on the Environment and Development Our Common Future emphasised, development must become sustainable to ensure that it meets the needs of the present generation without jeopardising the ability of future generations to meet their own needs. But the first prerequisite for sustainable development is education. Only an informed public and a trained workforce can introduce the new sustainable production and consumption pattern which is required. What are the primary skills that people must have in order to carry forward the development of their country with the additional constraint that this development must be sustainable? What knowledge do they have to have? What resources must they be able to use? On what can they rely? UNESCO and ICSU maintain that one of the key components that make the development of these abilities possible is science, and more specifically, on top of a solid general education, a proper science education. What are the grounds for this conviction?

A look at recent history may be helpful. Japan and Germany have been enormously successful in re-building their countries after World War II. Among the developing countries, some have been much more successful than others in developing their economy, especially in Southeast Asia. Clearly, as stated above, it is human beings who drive development. But what do the countries mentioned have in common? Two things stand out. First, all of these countries have placed great emphasis on secondary education with a strong science component. Secondly, research and development activities in these countries were supported by a variety of measures. These factors played an important role in the positive development of these countries as a pool of trained individuals is of prime importance for developmental activity. Fundamental sciences form an essential part of any curriculum for training these individuals.

Why is this so? Why is training in mathematics, physics, chemistry, and biology so helpful? Through this training, students acquire skills that derive from the special nature of science, more specifically, from its systematic nature (see section 1.1). Of course, this training provides students with the ability to address scientific questions as they arise within the realm of science. But perhaps even more importantly, it also provides the intellectual skills needed for addressing some of the general questions the global community is facing today. For most of today's environmental and developmental issues, the sciences are essential for detecting and analysing problems, identifying solutions, and ensuring sound policies and actions. At the same time, the complexity of problems makes interdisciplinarity and integrated approaches, which include contributions from the social science, important methodological tools.

On the one hand, science trains systematic thinking by forcing the articulation of appropriate questions and the identification of the resources needed for their answers. This systematic training is applied to concrete problem-solving tasks. On the other hand, the systematic approach must not petrify the thinking process. Innovative, critical thinking is required as not all solutions to given problems can be arrived at by applying well-established methods. In science, the prime authority is the better argument, not custom, social authority, or convention. However, the need for creative innovation does not imply an anything-goes mentality. If science has developed ways to cope with a particular set of problems successfully, then any would-be innovation must be backed up by very good arguments, otherwise it will be ignored or discarded.

Science students must first learn some mathematics and informatics because these are essential parts of science which serve many functions: theories are frequently articulated in mathematical language. Testing theories and hypotheses involves statistics. Given large quantities of data, filtering out the interesting ones requires mathematical screening methods. Writing computer programmes makes use of algorithms, and much more. Mathematical thinking trains the ability to abstract: to learn to distinguish the unimportant from the relevant in a given context, and to discard the unimportant. In every science problem, a clear articulation of the situation to be dealt with and the questions to be answered is required. What are the relevant aspects of the situation at hand? What are the resources for identifying the relevant? Is it a well-confirmed theory, working hypothesis, or sheer prejudice? What is it that I want to know about the given situation? Is it possible, on the basis of the information about the given situation, to answer these questions? Do I need more information, and if so, what do I have to do in order to get it? Do I need additional observations or experiments, or can I rely on existing information? Where is this information accessible? Am I able, given my resources, to perform the necessary steps, be they informational, experimental, observational, calculational, or theoretical? What theories, hypotheses, or models have to be used to answer the main questions? Are these theoretical assumptions secure enough for the given purpose? How is the relevant information about the given situation to be extracted from the theoretical assumptions to be used?

These are the kinds of questions that students are trained to deal with during their science education. They are also the kinds of questions asked and answered in the process of productive scientific research, which usually begins at the postgraduate level. Yet, there are also other abilities which come into play during active research into partly unknown territory. First of all, in spite of the fact that initial ideas come from individuals, productive research in the natural sciences nowadays mostly occurs in groups. This, then, involves all the social skills necessary to deal with such a situation. Second, in many research endeavours, a narrow scientific background will not suffice because knowledge from many areas must be drawn upon if the problem at hand is to be successfully solved. Thus, perspectives from other disciplines have to be integrated, which necessitates the establishment of interdisciplinary communication. As the cultures of different scientific disciplines are fairly different (subdisciplines may differ considerably - even within a single discipline), productive interaction between people with different backgrounds and outlooks is often required.

Looking back at what has been said about training in the sciences, it is obvious that the abilities gained by science education are essential for people who want to promote development in their countries. In order to be an effective means for development, science education must start at the primary level. To this end, ICSU recently established the Programme on Capacity Building on Science (PCBS) which focuses on primary education, as well as the public understanding of science. Realising the great importance of science education, at the 1995 International Conference on Donor Support to Development Oriented Research in Basic Sciences at Uppsala, Sweden, a declaration was issued emphasising the need for attention to education in basic sciences in developing countries. It contains a set of recommendations for actions by developing countries. It also emphasised the need for capacity building in the fundamental sciences, for supporting research and higher education in the basic sciences, for increasing co-ordination and co-operation, and for improving access to information supporting fundamental sciences. Many universities in both industrialised and less developed countries face particular academic and economic crises, and unflagging efforts are needed world wide to solve these challenges. Most especially, steps must be taken to arrest the deteriorating standards at universities in the Third World. To achieve these ends, a viable interaction between science education and industry is required both nationally and internationally. [Back to Table of Contents]

Forum II - Science in society: towards a new contract

2.1 Public perception of science: between acceptance and rejection
The public perception of science is not something stable. It has constantly changed throughout history, and it continues to fluctuate today. It also varies from culture to culture as well as within different sectors of the same culture. But for most of its lifetime, modern science has been seen as an indisputable vehicle for progress. The progress envisioned was of a material as well as of a spiritual and often even of a political nature. Materially, science was able to solve many problems of physical survival. Spiritually, science brought reason into areas where superstition or religious prejudices had dominated. Furthermore, in the political arena, science as an in principle anti-authoritarian and democratic enterprise has gradually transformed institutions and patterns of thinking dating back to the Middle Ages.

In some of the founding documents of modern science dating from the 17th century, science's promise was clearly expressed. Scholarly freedom was granted under the condition that science not interfere with politics or religious teaching. The enlightenment movement in Europe was deeply influenced by modern science and its achievements. The existence of modern science showed that reason can inform practice in unprecedented ways.

Science has fostered a great number of innovations improving human living conditions. By laying the foundations for the production of goods, contributing to public health, providing energy and new information channels, and through a myriad of other innovations, modern science has lived up to many of the expectations connected with it from its beginnings. So far, mainly the industrialised regions of the North have benefited from modern science and technology. A major challenge for the 21st century is to make these benefits more available to people living in less developed countries in order to fight poverty, disease, and environmental degradation.

Within the last few decades, this thoroughly positive assessment of science has been challenged, mainly in the industrialised countries. While there may be some unfounded accusations, a more critical appraisal of science and technology is indeed justified. The overly progressivist view holds that science can only have benign consequences because it is a product of reason. This naive conviction received its first serious blow with the advent of chemical warfare in World War I. However, this did not yet seriously undermine the public perception of science as something essentially and inevitably good. It would take World War II, especially the horror of the atomic bomb, to shake seriously the public belief in science as something intrinsically good. Despite its indubitable essential use of reason, science is not simply good for humankind by its very nature, but it is an extremely powerful instrument that can be directed toward very different purposes. Science and technology are not good in themselves, but they are when used wisely. Given this perspective, the more radical anti-science tendencies in some wealthier countries can be interpreted as an expression of disappointment. But this disappointment is grounded in an overly optimistic view of science in the first place, namely, that science and all of its effects are good by their very nature, which is simply false.

However, the more sober criticisms of science that have surfaced in recent decades should be viewed as being part of a learning process of how to develop and apply science more carefully and more wisely in the future. Such criticism is important in order to pinpoint actual and potential problems and to promote constructive debate among scientists, policy-makers, and the public.

Public surveys conducted in various countries indicate, on average, a considerable degree of public interest in science, as well as a certain amount of appreciation of scientific achievements and their actual and potential contributions toward improving living conditions. However, moderate scepticism as well as outright hostility towards modern science and/or its technological applications have been expressed for quite a number of different reasons by a wide range of groups. Typical sources of discontent with modern science include religious beliefs which may conflict with scientific theories, unwillingness to accept risks associated with new technologies, concerns about human dignity and animal rights, fear that technological change could veer, or has already veered, out of control, and pacifistic repugnance of military-industrial complexes. Others include various kinds of romanticism about nature and pre-industrial forms of life, or the idea that science and its claims to universality are yet another manifestation of Western cultural imperialism.

Some of these sceptical ideas thrive in the institutions of higher learning themselves, whereas others have the character of grass-roots movements or broader socio-political movements. Some reflect real challenges, whereas others may amount to unfounded accusations. In order to separate the two, so as to address the former and dispel the latter, scientists are advised to engage actively in open discourse not only with policy-makers, but also with the public. Furthermore, the social sciences and humanities could contribute to bridging the existing gaps between scientists and the general public.

Recent surveys of public scientific literacy in several major industrial nations have come to a somewhat sobering conclusion: a considerable fraction of the general public lacks knowledge of even some of the simplest scientific facts, such as that the Earth revolves around the sun, or that antibiotics are ineffective against viruses. With some notable exceptions such as belief in the theory of evolution, the average degree of scientific literacy seems largely independent of culture and of a nation's scientific and economic competitiveness. A surprisingly large number of those surveyed reported a very high interest in science, which seems to be somewhat at odds with the low scientific literacy measured by the surveys. Although these studies do not necessarily reflect how much people really care about science, they do show that science still enjoys high social prestige.

It is frequently assumed that a negative attitude towards the sciences must be based on an insufficient level of scientific literacy. However, such a 'deficit model' of the public understanding of science is rejected by most social scientists studying the interaction of science and society. While improving scientific literacy world wide is a desirable goal regardless of the adequacy of the deficit model (see section 1.5), problems at the science-society interface should not simply be attributed to public ignorance.

Parts of the public have the feeling that science and technology are becoming increasingly powerful, and that lay-people have little impact on how this power is used. For some, science has even come to be seen as infringing on democratic rights. Another reason for contemporary dissatisfaction with science reflects impatience. When there is a promise that AIDS will soon be cured, it is natural that people suffering from the condition are angered by delays. But perhaps the main reason for the decline in public confidence in scientists and technical experts and their respective institutions lies in their past failures to anticipate and control possible negative consequences of science and technology. There are a significant number of widely known examples where unexpected effects threatened the environment or public health. Recall, for instance, the careless use of DDT, the series of thalidomide-induced birth defects, the toxic chemical spills from Bhopal or Seveso, or the nuclear reactor incident from Chernobyl. The human tragedy involved in these non-natural disasters provides a ready explanation for the current crisis in confidence: non-specialists simply make a rational decision in revoking their faith in the experts and institutions who either were not aware of the possible hazards or were ignoring them. What has made things worse is the phenomenon known as 'hired brain' experts, who back powerful lobby interests with convoluted technical reports. While it is clear that risks cannot be completely eliminated and that unforeseen events can always occur, it is of utmost importance for the future to improve transparency and risk control. In addition, public consensus on the acceptable levels of risk for different kinds of technology must be established. This is the only way of restoring public confidence.

One proposed approach is known as 'citizen participation' in technology assessment, an idea which goes back to the 1970s (e.g., the public hearings held by the Experimentation Review Board in Cambridge, Massachusetts). By and large, it has been shown since then that well-informed lay people can come to intelligent and responsible conclusions in science and technology policy, for example, in assessing the safety and ethical soundness of genetic engineering techniques. A highly efficient procedure, which has been developed over the last few years, is known as the 'consensus conference', where a group of lay people work closely together with specialists to reach informed conclusions on the safety of specific technological systems, the soundness of the underlying knowledge, or whatever issues that may affect the public. For example, after Denmark, the Netherlands, and the United Kingdom, a French Conférence des citoyens on genetically modified organisms held in 1998 concluded that more research is necessary in order to assess accurately the ecological safety of releasing genetically modified plants outside of contained facilities. Another possibility are direct plebiscites at the regional or national level. The Swiss electorate, for instance, has recently been called to vote on a proposed constitutional amendment which intended to ban the production, acquisition, and distribution of transgenic animals, and the deliberate release of any genetically engineered organisms. With such democratic procedures, the responsibility for difficult policy decisions can be better distributed. In addition, one of the main causes of public mistrust in science could be eliminated. [Back to Table of Contents]

2.2 Science for development
Fundamental research plays a crucial role in the development of the natural sciences and their application (see section 1.2). Education is equally vital and complementary to research (see section 1.5). International co-operation can help reduce the glaring disparities in science and technology that lie at the heart of developmental problems by promoting the exchange of individuals, resources, and ideas (see section 1.4). By now, science and technology has been universally recognised as an important tool which can help solve a myriad of problems including malnutrition, infectious diseases, water-shortage, environmental degradation, and biodiversity depletion. Developing scientific research facilities can also help sustain economic growth and employment, as well as social equity. Science and science education, based on a solid primary education, have also become a cultural necessity for raising awareness of various aspects of the health and prosperity of the population, and for providing the skills for coping with the bewildering advances made possible by science and technology. A viable strategy aimed at addressing this complicated knot of challenges must proceed on various fronts.

First of all, in many countries, the national research potential in science has to be strengthened and the institutional facilities improved. Any sustainable strategy cannot overlook the need for forcefully promoting science education at the primary level. To this end, ICSU recently established the Programme on Capacity Building in Science (PCBS). In addition, emphasis in university education and post-doctoral training have to be placed on the development of low-cost laboratory equipment and practical work manuals, the introduction of recent discoveries and crucial new concepts into course curricula, and the training of laboratory technicians, etc. Training programmes like those offered at the International Centre for Theoretical Physics at Trieste, Italy (co-sponsored by UNESCO) are a very good example of international co-operation, but local and regional institutions should be strengthened or created, taking special care to gear these towards the specific needs and working conditions of the different countries and local regions. International co-operation has to be broadened and strengthened through collaborative actions with intergovernmental bodies like UNESCO and non-governmental organisations like ICSU.

Recent developments in computer and information technology, biotechnology, and energy use are expected to bring about radical changes to peoples' lives. The combined advances in computer technology and microelectronics seem likely to result in a decrease in the number of commercial outlets for products from less industrialised countries because the latter will be less competitive. Automated manufacturing procedures bring the threat of unemployment to countries where there are adequate supplies of labour, but not of capital. Yet these technologies also offer great benefits to those countries capable of exploiting them; for example, in data processing, communications, manufacturing, and quality control. Applied microbiology and biotechnology offer the possibility of producing a great number of substances and compounds essential to human life and prosperity. Improved fermentation processes with higher yields, improved fertilisation techniques, cheap production of biologically produced fuels for cooking and heating, and biotechnological production of foodstuffs all offer distinct benefits, especially to developing countries. Furthermore, biotechnology is particularly important for those regions which are rich in biodiversity. If biological resources are not utilised locally, there is a risk that they will be exploited by foreign-owned or multinational corporations which may not sufficiently compensate the local regions with economic benefits. Renewable forms of energy like sunlight, which are sometimes more effective and accessible than conventional forms, while at the same time environmentally less destructive, are especially attractive options because they promise savings on energy imports. More generally, humankind should be careful not to insist in making the same mistakes that were made by industrialised countries when they opted for unsustainable modes of production and consumption, and science can be a great partner in the effort to correct these mistakes. Although the neglect of sustainability may appear to produce quicker returns in some cases, in the longer run the consequences will be disastrous for everyone. But in some cases, pressing vital needs may require immediate attention at the expense of sustainability. The international community should assist them in compensating the costs of sustainable policies.

The exploitation of scientific knowledge through the production of marketable goods requires an economic and industrial infrastructure, especially skilled labour, production facilities, and capital. These goods are in short supply in non-industrialised countries. In their quest to expand the manufacturing base and with their relative lack of capital for launching new industries, developing countries have to compete with each other when attracting capital through foreign investment. The availability of skilled workers is a major attraction in the competition for such investment. Thus again, education in general, and science education in particular, are of utmost importance for industrial development, as UNESCO has repeatedly emphasised (see section 1.5). In addition, there is a need to build up the infrastructure necessary for development, and for proper governmental policies to back up this process. Countries in Southeast Asia which provided these necessities have registered phenomenal development in recent decades.

One area in which countries should, in their own vital interests, invest in scientific research concerns the causes and consequences of disasters caused by natural phenomena such as floods, earthquakes, El Niņo, and tsunamis. The consequences of these natural hazards are often specific to particular regions, depending on topographical features, settlement structure, housing, preventive measures, and the like. Some natural disasters must be confronted by the international community, as they are of global proportions, and UNESCO and ICSU play an active role in addressing them. But the local consequences of natural disasters are still a burden primarily for the respective region, and these regions should be as well-prepared as possible to cope with those consequences. Another research area that has region-specific aspects and is of vital interest to many countries is water research. Knowledge of the availability of freshwater resources and water use technologies are of vital importance, especially in vulnerable environments, for example in arid and semi-arid zones. It is also highly relevant to activities aimed at mitigating desertification and rehabilitating degraded land. Sustainable practices for water resource management have to be found in such environments.

The argument that developing countries should not participate in the international research effort in fundamental science because the latter is irrelevant to their most pressing needs and therefore a bad investment of their resources is fallacious. There is an air of condescension to this view which smacks of cultural colonialism. The developing countries do not want to be reduced to mere spectators in the drama of fundamental science, and for good reasons. Research in fundamental science attracts bright young people and provides them with state-of-the-art scientific knowledge and problem-solving skills which cannot be learned from books alone. Thus, fundamental science is an important source of scientific and technical expertise which is of prime importance for a country's transition from a less developed to a more advanced nation (see section 1.2).

One problem which needs to be addressed is a particular form of international migration. International migration is a complex phenomenon and can have many diverse causes. Historically, many nations have benefited from migration. However, when the migration is of highly educated and skilled people who go from poorer to richer countries, there is a special problem - the so-called 'brain drain'. Normally this phenomenon, when it refers to the migration of scientists, is the result of poor working conditions, lack of resources, scarcity of jobs, unstable institutional and governmental support for science and technology, as well as lack of incentives to scientists and science students, etc. Those countries which have fewer scientists per capita and badly need to increase their numbers, are also just the ones that are 'exporting' them to the richer countries. Brain drain, which so severely affects some of the less developed countries, can only be reversed by changing the above mentioned conditions. Of course, international co-operation may help in counteracting or mitigating the negative effects of such migration.

Countries with fewer scientific resources or less scientific capacity need to be better integrated in the information flow within the scientific community. The increasing cost of journals and books limits the accessibility of scientific information in such countries (see section 1.4). Furthermore, many scientists in less developed countries find it difficult to present their research results in international journals, which are published mainly in Europe and North America. Again, international bodies are needed to alleviate the problem. The exchange of professors and researches on temporary assignment, joint research projects, and multi-media teams electronically linked can help alleviate the problem. UNESCO is actively engaged in strengthening existing networks involved in collecting, storing, retrieving, and disseminating information relating to the sciences, as well as sharing data bases, and publishing directories. A special effort should be made to facilitate access, especially for researchers in developing countries, to scientific information through the development of electronic networks. If these challenges can be met, then science can promote a country's development by helping it achieve economic growth, employment, and social equity, as well as effective protection of the environment and prudent and sustainable use of natural resources. [Back to Table of Contents]

2.3 Setting priorities in a new socio-economic context
Organising and financing scientific research is a key challenge facing governments all over the world. The growth of science has been strained as the rising costs of fundamental scientific research confront limited national budgets. In order to overcome this bottleneck, universities are seeking tighter collaboration with private-sector industry. The number of patents filed by universities and university-industry collaborations has greatly increased over the last few years. Thus, we are witnessing an increasing commercialisation of scientific knowledge. This is at loggerheads with some values associated with science, as outlined below.

Moving science closer to the market does undoubtedly promote more efficient and more effective mechanisms for advancing commercial technologies. Bringing the production of scientific knowledge closer to the market helps put knowledge to practical use. Conversely, coupling research to commercial interests provides powerful incentives for generating new knowledge. But could an increase in private-sector funding offset a decrease in public-sector funding? After all, private companies in major industrialised countries spend considerable portions of their income on research, as do governments, if they find it profitable. The US aerospace industry is an impressive example, as are pharmaceutical companies all around the world. However, only a small fraction of private R&D expenditure goes into research in fundamental science. Most of it concentrates on specific product development with foreseeable profits.

Furthermore, as industry continues to finance an increasing portion of scientific research, the public character of scientific knowledge is under a growing threat. In fact, fundamental scientific research is carried out under the open principle: new knowledge is disclosed and disseminated quickly and completely. Economists explain that this principle of open science provides private incentives to generate public goods. In order for scientific knowledge to be effectively used both in the generation of new knowledge and in its application to problem-solving on a global and local scale, the fruits of fundamental scientific research must not only be accessible, but also widely distributed and quickly disseminated. This is essential both for the development of new knowledge and for the application of existing knowledge to practical problems. But this stands in direct conflict with the current socio-economic trend towards the commercialisation of knowledge. Commercial constraints generate significant restrictions on the disclosure and dissemination of scientific knowledge and may even threaten the autonomy of fundamental science. Market forces generate the need for either the explicit protection of knowledge through patents or for secrecy in order to provide incentives for investing in knowledge production. These tendencies increasingly confront the traditional notion that fundamental discoveries supported by public funds should be designated as public goods.

In this context, it is important to recognise that new information and communication technologies (see section 1.4) may have a paradoxical effect on unrestricted knowledge distribution. The Internet is widely believed to promote an increase in the free flow of information. While it is indeed a powerful means of knowledge distribution, it may create incentives for companies to withhold information or raise the costs of access to it in order to extract the full value and recognition of a scientific discovery.

Thus, while closer co-operation between science and industry contributes to the public good by promoting increasingly effective mechanisms for technical applications, traditional academic norms, such as the commitment to the free flow of information and the full public disclosure of research results, threaten to deteriorate. This is an important trade-off for society which generates important policy issues: should we halt the contraction of the institutional space of open research? How can we protect the autonomy of science without inhibiting the benefits of closer ties between science and industry? How can fundamental research be managed in a context favouring market orientation?

The commercialisation of scientific knowledge in a policy context which promotes short-term benefits at the cost of long-term projects also threatens fundamental research funding and could inhibit international co-operation on projects which require attention on a global scale. As is well-known, we are living in a world in which the present value of future benefits is largely under-rated. Historically high levels of real rates of interest since the 1990s reflect a social preference for current consumption at the expense of investment in the future. Current corporate trends actually promote short-term thinking which can erode support for a company's fundamental scientific research programme. The emphasis on quick returns has made it difficult for companies to finance long-term investments. As public research projects adapt themselves to commercial needs, publicly funded fundamental research is following suit and priorities are also moving to short-term projects. This reallocation often occurs at the expense of curiosity driven research with no foreseeable results. But such research has led to significant and important break-throughs in the past as illustrated in the introduction. Because the science-society contract has been renegotiated, and the deal is now much tighter for the scientific party, difficult policy issues are now pressing. They often force hard choices on resource allocation - independently of a country's level of industrialisation. Effective strategic planning and priority setting strategies are increasingly necessary in order for the efficient allocation of increasingly limited funds. But the results have not always been successful. For example, global spending on malaria research by both the public and private sectors has sharply declined just as an increasing resistance to existing drugs makes the need for research and development greater than ever. Clearly, greater international coordination will be needed in order to fund increasingly global challenges.

But, cost-benefit analysis cannot provide an adequate basis for decision-making concerning fundamental research funding for two reasons. First, as mentioned earlier, the benefits of fundamental research are often unforeseeable. Second, policy decisions made today could have considerable effects on the welfare of future generations, which are not accounted for in a cost-benefit analysis. In the coming century, we are faced with the pressing need for new approaches which take inter-generation equity into account. Global science policy must address questions such as what kinds of mechanisms can serve the interests of future generations. Another problem concerns how access to public information (such as human genome data) can be maintained. Little consensus has been reached about how best to meet these challenges and others like them. There is a pressing need to develop new approaches.

One possible avenue for generating funding for fundamental scientific research which is often over-looked, and which could perhaps avoid some of the tension generated by commercial interests, is the use of charity organisations. There is some evidence suggesting that charitable motives are a very effective means of raising money for scientific research. For example, in France, the telethon for research on myopathic diseases, which takes place annually, now raises more money than the department of life sciences of the CNRS (Centre national pour la recherche scientifique). Part of the success of such operations can be attributed to the participation of popular public figures such as movie stars and sport heros.

But given the inherent limitations on these various sources of funding (industry, charities), as we face the 21st century, science management and science policy will have to invent and implement new mechanisms for generating an investment in fundamental science and the healthy distribution of scientific knowledge. [Back to Table of Contents]

2.4 Science: the gender issue
Science, by its very essence, is an enterprise which should be neutral with respect to the particulars of those who practise it. The reality is that, on average, women are under-represented in the sciences. The degree of this under-representation varies somewhat from discipline to discipline as well as from country to country. The skew in gender ratio towards males increases as one moves from the undergraduate to the graduate and post-doctoral levels. It is largest in the senior positions. To some extent, in many countries the under-representation of women in the sciences mirrors the general dominance of males in most occupations enjoying a high social prestige. For instance, in many professions outside of science, the percentage of woman decreases as the position increases hierarchically. But in addition, there are specific social and institutional mechanisms responsible for the under-representation of women in science that must be identified and eliminated wherever possible. The basis for this postulate is two-fold. First, this asymmetry violates the now widely accepted principle of gender equality. Second, by not giving equal opportunities to women, science does not make adequate use of its most precious resource, namely, human talent. Women may bring different perspectives into play, observe neglected areas, generate different ideas, place different emphases, make different value judgements, and so on. Science can neglect this variety only to its own disadvantage and to the disadvantage of the global community. Science is dependent on the flow of new ideas. Creativity and the generation of new perspectives is a notoriously scarce resource which can have a notoriously deep impact.

What are the main social causes of gender inequality in science? What are the mechanisms that erode the comparatively high percentage of female undergraduate students as one moves up the ladder to graduate students, junior researchers, and senior researchers? A whole array of factors which inhibit women in their pursuit of a scientific career have been identified by recent social science research.

First, as early as the secondary school level, co-education does not seem to have only beneficial effects in science classes. Typically, male students dominate the class-room leaving little space for different learning and communication patterns practised by some female students. Discouragement and lower self-confidence may result.

Second, false stereotypes about women's purportedly lower abilities to understand or practise science are widespread. They tend to amplify all the adverse factors, and in addition, create the false impression that no discriminatory social factors and circumstances are causally relevant. For instance, lower average grades of female students in high school science find an easy explanation in a supposed asymmetry in the distribution of abilities, but this dubious explanation leaves the relevant questions about social mechanisms unasked.

Third, gender-related cultural values are too often still in place, discouraging young women from pursuing a professional career in science. If the pursuit of a scientific career is seen as 'masculine', then it does not come as a surprise that many women are reluctant to choose science as a profession.

Fourth, recent studies indicate that in many places it is still the case that women need more scientific credentials than their male colleagues to have an equal chance of securing a post or obtaining funding for a research project. In other words, women are still actively discriminated against. Evaluations of scientific research conducted by governments or independent institutes frequently fail to take women into account due to the lack of gender-specific indicators. This causes a lack of transparency and makes it impossible to address these problems systematically and to take appropriate actions to correct the inequalities.

Fifth, there are certain features of the academic career system which make it difficult to harmonise family lives with research requirements. Science is, in many areas, highly competitive. Creative scientific work is almost always done under time pressure because there is always the danger that someone else will be first, and in scientific research and publication 'the winner takes all'. Thus, part-time research which allows for a synchronously balanced time investment in raising children and in doing science is much more the exception than the rule - both for men and for women. But also diachronically, allocation of time between science and family needs meets serious difficulties. Research typically moves so quickly that after even a small interruption, one may not even understand the very questions that are currently being asked. Furthermore, due to a widespread lack of day-care facilities at scientific institutions, the parallel pursuit of science and parenthood meets with serious difficulties.

Lastly, the high relocation rate required of young scientists cannot easily be met by those responsible for young children. Thus, as long as women are more strongly involved in child-raising than men, these factors will strongly hinder women from pursuing a scientific career, and may force especially women into making an inhuman choice between science and parenthood.

Another important issue concerns the marginal role of women in science policy in a wide sense. This is, of course, due to their lower presence in both the scientific and administrative systems. Setting research priorities, allocating grants, evaluating research activities, assessing the safety of technological systems, etc. are activities mostly done by men. Thus, women are not sufficiently involved in decisions which affect them to the same or even to a greater extent than men. By virtue of their different historical, cultural, and social position, and also because of different interests, women often have a different vision of how to put science to use for the benefit of society, e.g., in biotechnology. Again, there are two aspects to this question, the aspect of gender equality, and the loss of talent necessary for future development. In the future, full use must be made of women's competence, experience, and potential.

Finally, further analysis of the different specific effects science and technology have on women and men is necessary, both with respect to its positive and negative consequences. Policy decisions on resource allocation can motivate gender issues. For example, breast cancer and birth-control pills raise issues about gender equality in resource allocation. While this is most obvious in the field of biotechnology, it may be relevant in other fields as well. This holds for more and less industrialised countries alike. It is particularly pressing in regions of the world where women have virtually no influence on policy decisions, and yet bear the main burden in the daily functioning of the community.

Paradoxically, in spite of the existing asymmetry in science with respect to gender, the sciences may increasingly be powerful factors for the advancement of women. In a world which is increasingly becoming knowledge-based, the neglect of a huge source of talent is plainly dysfunctional, even if considerations of fairness were left aside. In addition, with the further dissemination of information technologies, many workplaces will no longer depend on a particular location, thus allowing for new patterns of co-existence between private and professional life. [Back to Table of Contents]

2.5 A new social contract for science
The idea of a social contract for science is a way of describing the relationship between science and society. To express this relationship in such terms, the mutual benefits must be identified. Over the last few decades, governments have funded universities and other research institutions without providing many directives about how this money should be spent. In return, these institutions delivered exploitable knowledge which benefited society in the form of contributions to economic growth, public health, national prestige, and national security. During the Cold War, this last contribution was seen as especially important, as is reflected by the large portion of the investment in science devoted to military-related research by industrialised countries. For society, science was the only eligible partner for such a contract; no other institution, existing or imagined, had a similar potential for meeting society's needs. The traditional science-society contract was predicated on the assumption that market forces cannot guarantee the optimal allocation of resources to research. Market forces tend to direct the flow of investment to areas where short-term returns on investment through marketable goods are expected (see section 2.3). But the advantages society expected from science were of a different nature. National security and better public health, for example, are not marketable goods.

For a variety of reasons, this traditional contract needs to be replaced by a new one which meets the needs of society in the 21st century. With the end of the Cold War, national security has assumed a lower priority in many countries. The economic context in which science operates has changed and is expected to be different in the future. As outlined in section 2.3, an increasing portion of scientific research will be undertaken by the private sector and by university-industry collaborations. To some extent, this is turning scientific knowledge from a public into a private good. Most importantly, science could help meet some urgent societal needs, including sustainable development and global environmental issues. The urgency of these problems matches that of the national security needs during the Cold War. As in the old contract, science can deliver goods which no other institution can provide.

What are the main features that this new contract should have in order to make the best use of science in the 21st century? First, the new social contract for science should acknowledge that investment in science is, among other things, a matter of intergenerational equity. Just as we now harvest the fruits of scientific advances made by previous generations, future generations, too, will want to stand 'on the shoulders of giants'. The contract should protect the commitment to fundamental science and its associated freedom of research as a benefit to all humankind, present and future.

Second, the contract must also take into consideration that in the 21st century, the boundary between fundamental and applied science will become increasingly blurred in many areas. The reason for this tendency is that increasingly, the discovery and understanding of new phenomena come hand in hand with the applications or developments made possible by them. This trend can already be clearly seen in such areas of research as the human genome, cancer, as well as in biotechnology and nuclear fusion. This development raises a host of challenges, especially concerning the effective distribution of investment and revenue between the private and public sectors.

Third, the new contract should recognise that science can operate most efficiently if important scientific information is allowed to spread rapidly and internationally. At the same time, the contract should concede that the cost of weakening the openness of science may sometimes be compensated by more effective mechanisms for advancing commercial technology. In other words, the possibility of delaying the disclosure of certain kinds of scientific knowledge in order to produce competitive marketable goods provides a commercial incentive for research and development. Thus, mechanisms will have to be established which regulate under which circumstances the non-disclosure of scientific findings is acceptable, especially when these findings substantially profit from publicly funded scientific institutions. Furthermore, the problem of intellectual property rights will have to be discussed anew, given the powerful new means of electronic storage and dissemination of information.

Fourth, due attention must be given to the fact that human civilisation is a major environmental force on our planet. If the current trend continues, the results will be disastrous. The global climate system is likely to be profoundly changed. Rising sea levels and the destruction of the planet's ozone shield add to the severity of the situation. The current practice of using the energy trapped in fossil fuels is not sustainable at the present rate. While carbon dioxide emissions could immediately be reduced through energy conservation, 'cleaner' forms of energy would be invaluable. Science has already made the use of new forms of energy possible, such as photoelectric or nuclear energy. Although the latter is controversial, the examples provide reasons to believe that further progress can and will be made in energy technology. Another related challenge that the international community faces is the depletion of biodiversity on a global scale due to habitat fragmentation and destruction. Science should also provide more effective advice on conservation policy. New means for utilising biological resources could be developed which would create powerful incentives for their conservation.

Fifth, in order to confront these global challenges, the contract should promote stronger interaction between scientific disciplines, and interdisciplinary co-operation which includes the social and human sciences. This holds especially for the articulation of relevant research questions: approaches that neglect the human dimension of a complex problem tend to produce answers that are irrelevant for its solution. Scientists will have to improve their skills for carrying out problem-oriented, instead of discipline-oriented, research. This point also reiterates the blurring between fundamental and applied research (see above, number 2). Governmental, intergovernmental, and non-governmental agencies will have to co-operate at a higher intensity then hitherto practised, especially on global and long-term projects. In order to confront many of the most pressing challenges, new international research networks will have to be created and existing ones will have to be strengthened.

Sixth, the contract should contain a strong commitment to increasing the representation of women at all levels of the scientific community. The mechanisms discouraging young women from pursuing a scientific career and discriminatory practices in scientific institutions need to be identified and eliminated. Statistical censuses which take into account gender as a relevant parameter are urgently required in order to monitor attempts at increasing the representation of women at all levels. Furthermore, decision-making in science and technology policy, which frequently affect women and men differently, must involve more women than is presently the case.

Seventh, the contract should recognise that scientists have a special responsibility. In the traditional contract, the scientist's responsibility almost exclusively concerned the scientific quality of their work. In return for public funds, science had to deliver reliable knowledge, irrespective of its potential use. The latter belonged to the responsibility of those applying this knowledge to practical ends. Science was viewed as being 'value-free', that is, disconnected from evaluations of its applications. In the new contract, scientists are committed to an even stronger form of responsibility. It should be noted at the outset, however, that it is very difficult to determine where the scientist's responsibility ends. Due to the moral issues that sometimes arise directly from their research, scientific responsibility now includes a new ethical dimension. It is part of a scientist's responsibility to keep the public well-informed, both about potential advances, imminent risks, long term effects, and potential dangers of their work. As initially only the scientists may be aware of these various aspects, they should exercise good judgement, wisdom, and humility. They should refrain from the arrogant assumption that their scientific competence extends to issues involving social norms and values. They should be committed to the peaceful, productive use of scientific knowledge. Scientists should not be the sole arbiters of the value of their work and its consequences for society. The ethical issues which science and technology generate should not be decided by market forces either: they should be decided by informed citizen participation, based upon the best available knowledge.

Eighth, it is a vital part of this responsibility that scientists communicate their knowledge to the public in order to increase the public understanding of science, to inform policy decisions, and to make new findings accessible to those who might need them. More efficient bridges need to be built between policy, management, and science, as well as between the public and private sectors. A concomitant requirement exists for the training of interdisciplinary scientists who have special competence in working at the policy-science, management-science, and public-science interfaces. These will be needed to improve risk-benefit assessments, health and safety standards, efficient allocation of resources, and for targeting potentially fruitful local investment opportunities. These goals would be helped by appropriate university curricula and by a more flexible reward system for professional scientists.

Ninth, the contract should acknowledge the need to bridge the widening knowledge gap in order to promote socio-economic development in less developed countries by strengthening their scientific research and teaching capacities. Strengthening the research capacities must include a science policy that is adapted to the respective local situation. For example, science policy models that are effective in a highly industrialised country may be completely inadequate in a less industrialised one. Strengthening the teaching capacities includes improving basic science education at the primary level, as well as in higher-education. Science education should also be recognized as a useful resource outside of the laboratory, and increasing scientific literacy is necessary to optimise informed public policies. The main goal must be to distribute the technological benefits of science more equally around the world.

Finally, the new social contract for science should commit the scientific community to addressing the most urgent needs of society in proportion to their importance. As the development of the atomic bomb demonstrates, scientists can respond to urgent societal needs quickly. One of the most urgent needs in the 21st century will consist in the development of clean technology and the sustainable use and management of natural resources, in order to improve the living conditions prevailing in most countries around the world. Scientific knowledge should play a greater role in addressing some of the most pressing global challenges such as poverty, environment, health, and food and water security. To meet these challenges, scientific research will be needed more than ever. Another continuing challenge for scientists must be the battle against infectious diseases, such as malaria and AIDS.

In conclusion, the new contract should promote political, economic, and social co-operation in an effort to direct scientific knowledge and technologies toward the benefit of humankind. A firm commitment to scientific research and education by all nations, predicated on the new social contract for science outlined above, will be a necessary prerequisite for achieving real human and social development in the 21st century. It will promote human rights and human dignity. It will bring together skilled and dedicated people, and represent humanity in all of its diversity. It will promote the creative exchange of ideas toward a more peaceful world. It will be a continuing project drawing from across the globe working towards our common future.

[Back to Part 1]



introductionnewsopinioncontact us


Macmillan MagazinesNature © Macmillan Publishers Ltd 1999 Registered No. 785998 England.
#cparse("${superIncludes}/super.body-bottom.fhtml")