Sie sind auf Seite 1von 116

Lecture Note

Philosophy of science
Page | 1 BY
Alloy S Ihuah PhD

 The philosophy of science, a sub-branch of epistemology. It is the branch of


philosophy that studies the philosophical assumptions, foundations, and implications
of science, including the natural sciences such as physics, chemistry, and biology, the
social sciences such as psychology, history, and sociology, and sometimes—
especially beginning about the second decade of the twentieth century—the formal
sciences, such as logic, mathematics, set theory, and proof theory. In this last respect,
the philosophy of science is often closely related to philosophy of language,
philosophy of mathematics, and to formal systems of logic and formal languages.
Questions Addressed by Philosophy of Science
Philosophy of science investigates and seeks to explain such questions as:
 What is science? Is there one thing that constitutes science, or are there many
different kinds or fields of inquiry that are different but are nevertheless called
sciences?
 Does or can science lead to certainty?
 How is genuine or true science to be distinguished—demarcated, to use the usual
philosopher's term—from non-science or pseudo-science? Or is this impossible, and,
if so, what does this do for the claims that some things are pseudosciences?
 What is the nature of scientific statements, concepts, and conclusions; how are they
are created; and how are they justified (if justification is indeed possible)?
 Is there any such thing as a scientific method? If there is, what are the types of
reasoning used to arrive at conclusions and the formulation of it, and is there any limit
to this method or methods?
 Is the growth of science cumulative or revolutionary?
 For a new scientific theory, can one say it is “nearer to the truth,” and, if so, how?
Does science make progress, in some sense of that term, or does it merely change? If
it does make progress, how is progress determined and measured?
 What means should be used for determining the acceptability, validity, or truthfulness
of statements in science, i.e. is objectivity possible, and how can it be achieved?
 How does science explain, predict and, through technology, harness nature?
 What are the implications of scientific methods and models for the larger society,
including for the sciences themselves?
 What is the relationship, if any, between science and religion and science and ethics,
or are these completely separate?

Those questions may always have existed in some form, but they came to the fore in Western
philosophy after the coming of what has been called the scientific revolution, and they
became especially central and much-discussed in the twentieth century, when philosophy of
science became a self-conscious and highly investigated discipline.
It must be noted that, despite what some scientists or other people may say or think, all
science is philosophy-embedded. Philosopher Daniel Dennett (1995:37) has written, “There
is no such thing as philosophy-free science; there is only science whose philosophical
baggage is taken on board without examination.”
Towards a Definition:
Etymologically, the word science is derived from the Latin word scientia, ie “knowledge”.
Understood as human activity thus said to be “knowledge” is a human undertaking to learn
about the world around us through a special method called “scientific method”. We shall
some back to this latter. But suffice it to say, however, that there is no one and only definition
Page | 2 of science. This is partly because the standpoints from which authors look at science differ. In
his book “What is Science?” Normal Campbell writes that science can be looked at from two
aspects: firstly; science is a body of knowledge and a method of obtaining it. Secondly,
science is a pure intellectual study, and so in this regard akin to painting, sculpture or
literature rather than the technical arts. Understood in this light, science aim only at satisfying
the needs of the mind and not those of the body. It appeals to nothing but the disinterested
curiosity of mankind.
We may say perhaps that though limited in scope the second aspect of science is closely
linked with the first. Both project science as a whole body of knowledge, logically
interconnected and directed at achieving some desired goal; spiritual or material. Arguably,
such an endeavour requires systematic coherence, objective and standardized method as its
import ingredients. This conception of science may have informed Amadi’s definitions of
science, that it could also mean (i) knowledge especially of fact or principles gained by
systematic study. (ii) a particular branch of knowledge especially one dealing with body of
facts or truth systematically arrange and showing the operations of general laws as the
science of mathematics. (iii) Systematized knowledge especially of the law and facts of the
physical or material world. He thus sums up science as,
The pursuit of knowledge and understanding of the natural and social dimensions of our
world of observation, formulating descriptive systems by controlled experiments to determine
the degree to which these system represent the phenomena world etc.
Understood as such, science is concerned with both man and his way as much as everything
that is foreign and external to man. It is a branch of pure learning which is concerned with the
properties of the external world of nature which business is to find accurately those properties
are, to interpret them, and to make them intelligible to man; the intellectual satisfaction at
which its aims would be secured completely if this external world could be reduced to order
and be shown to be directed by principles which are in harmony with our intellectual and
moral desires. As an intellectual endeavour, science is both natural philosophy as well as
moral philosophy which learning arose ultimately from man’s desire to understand the world.
Perhaps it is this understanding of science that the complex adjectival form of the world
“science”, namely “scientific “i.e. knowledge- making has today come to be accepted as the
real province of science which in the early beginnings was the original enterprise of natural
philosophy. No wonder therefore that science has come to be accepted as “the making of
knowledge i.e. research instead of knowledge as such. Thus, science described as such is the
systematic process of making knowledge; of building knowledge.
Thus far, science can be provisionally defined as the process, or the group of interrelated
process though which we can acquire modern and over –changing knowledge of the natural
world which encompasses inanimate nature, life, human nature and human society. It thus
means that the result of science can never be static; changing and gain. Perhaps it is this
quality of science that Ogbinaka (1998:178) writes that the intellectual frontiers of science
have ever been expanding with very little of its contents being dropped. This has been done
to such an extent that the intellectual results of ‘yester science’ look crude and naïve in the
face of today’s science. Quoting from the Encyclopedia Britannica vol. 6, he argues further
that this conception of science has provided f very strong for the following concepts of
science;
(i) Science can be taken to be a mood in which the World is considered. Being a mood, we
should accommodate its changing states. Just as no man is always in the same mood, and no
man of science remains permanently in the same scientific mood.
(ii) Science is always developing. It is not a static body of knowledge.
(iii) Science is more of the making of “knowledge” (i.e. in contradictions with a claim that it
Page | 3 is knowledge itself’, so it is close to be called a research; a method employed in pursuit of a
goal which involves “the acquisition of systematic generalized knowledge concerning the
natural world; knowledge which helps man to understand nature, to predict natural events and
to control force.”
This again involves the use of previously accumulated knowledge to construct general
theories or systems from which testable hypotheses can be derived, and the testing of such
hypotheses via quantified observation under controlled conditions.
Thus far we may ask whether such a conception of science, argued above adequate for the
analysis of the impact of science oh human development, but in particular African
development. We may argue that such a conception of science reduced to a “method”
employed in pursuit of a goal is inadequate on the following ground.
(i) As the acquisition of “Systematic” generalized knowledge concerning the natural world,
science is made to be a scarce commodity reserved only for the west to the exclusion of the
developing world. But this is clearly fallacious, for science id a widely distributed
commodity, found in every culture tradition.
(ii) Science as a whole is a process, which transcends particular scientists, research teams,
and institutes. Hence, to argue that scientific goals encompass outcomes toward which
movement occurs is to miss the point. Put in proper perspective, a “goal” as usually
understood in an outcome, which people strive or more generally toward which the internal
functioning of a system is directed. Suffice it to say then that the meaning of a statement
attributing a “goal” to such a process would require clarification. No doubt, science produces
certain outcomes and some of these outcomes are goals of individual scientists and research
teams; but it does not necessary follow that science must be defined in terms of movement
towards the goal. As rightly confirmed by Richter; (1972:14)
It is entirely possible that the most significant aspect of science involve movement, over a
long time span, in direction which have not been intended or recognized by scientists
generally, and which have emerge accidentally, even if there he has also been movement in
directions which may be identified as corresponding to a “goal” of science
(iii) Even if science is defined as a process of moving toward a goal, it does not follows that
science thereby becomes equivalent to a “method”. Rightly defined, a method is a process
employed deliberately in pursuit of goals. It refers to the specification of steps, which must be
taken in a given order, to achieve a given end. As a function in scientific inquiry, “methods
are used within scientific inquires. “However, the concept of method cannot reasonably be
applied to some important types of event through which the findings of different inquiries are
interpreted and integrated by the scientific community as a whole. This is because the nature
of the steps and the details of their specification depend on the end sough and the variety of
ways of achieving it.
We may thus argue here that the concepts of goals and method used as a quality of any
scientific endeavour can only be recognised as appropriately applicable at relatively
microscopic levels. As Maurice Richter (ibid.) conclude on this matter that, when we seek
instead to analyse science macroscopically, taking into account not merely what happens
within particular research projects but also the integration of findings of many such projects
in different disciplines over centuries, the concepts of goal and method appear to lose their
relevance. The method of science therefore vary according to whether its end is taken to be
the conquest of nature of the discovery of truth and in the light different theories about the
relation between those ends and man’s primitive condition of impotence and ignorance.
The conception of science as a social institution, science as an occupation and lastly science
as a profession are also inadequate insofar as they imply comparatively stable relationship
between science and society, with science performing certain functions or services on a
Page | 4 relatively consistent basis. The way ahead here is the conception of science as a cultural
process with alternative avoids the difficulties and shortcoming of the above conceptions.
Science as a cultural process is associated with a distinct view of nature as operating
according to general laws, which remain largely hidden under ordinary observation
circumstances but which can be uncovered through systematically controlled observation and
experimentation as for example Isaac Newton’s one set of propose law; the law of
gravitation, the principle of calculus and the compound nature of light; and his three laws of
motion (Richter 1972:16). Such scientific laws, as they are many, provided an
overwhelmingly impressive demonstration of potentialities of this approach, thus reasonably
clearly differentiated from such related phenomena as philosophy region, technology and
magic among others.
This conception of science argues a position that the role of scientists became differentiated
from other roles and scientists came to communicate with each other systematically in ways,
which mark the beginning of what we call the scientific community. Captured with vision,
Maurice Richter distinguished three perspectives on science, perspectives which are mutually
compatible and capable of beginning integrated into single more comprehensive perspectives,
but which differ in the insights, which they provide. These include
(i) Science as cultural counterpart of individual cognitive development
(ii) Science as an out growth of traditional cultural knowledge
(iii)Science as a cognitive form of cultural development.
Taking together, the aspect of looking at science helps to provide a richer understanding of
the meaning and place of science in human development than any one or two of them alone.
According to this conception therefore, science is a transition in which cultural knowledge
systems with certain characteristics are displaced by other knowledge system sharing these
same characteristics but which also differ in certain in certain respect from those which they
displace. As a process, the direction of scientific development is similar to that of individual
cognitive development which starting is traditional cultural knowledge. Maurice Richter
(p.58) adumbrates the point further thus;
The structure of scientific development is similar to that of the evolution process in general
and the process of cultural evolutionary process in general and the process of cultural
evolution in particular science is an extension of cognitive development from the individual
to the cultural level. And a developmental outgrowth of traditional cultural knowledge, and
a specialized cognitive variant and extension of cultural evolution.
Through this cognitive process, long established systems in the scientist area are displaced,
though unintentionally but which new and genuine contribution is made to science and or
opening up the possibility of further inquires which will expose serious defects and or
frontiers in the system. Linked with this conception is, the understanding that, science as a
cultural process encapsulate the study of the maker of culture; man and his environment. By
extension therefore science could be said to be study of the mastery of man’s material
environment. This nature of science in terms of its employments in solving man’s
environmental or materials problems is what could be said to be the systematically ordered
knowledge of man and his environment. This most obviously explains why philosophers
describe man is an explorer who finds himself on a strange island, without knowledge of his
origin, his mission and the nature of his environment. But which scientific knowledge
provide contours and all the available routes for his journey in life. Rightly understood, the
future of man, and in general cosmic harmony is to a large extent dependent on man’s
creative mind.
This creative mind of man is what could be likened to a cultural counterpart of the individual
cognitive development which involves a process of learning to correct for certain limitations
of man’s observational capacities. This properly speaking is science for us which procedure is
Page | 5 in line with man’s ontology. Our capacity to observe the world is limited by selectivity of
our perceptions and by selective focusing of our attention on some aspects of the
environment to the neglect of other aspects. We are also limited by our restricted locations in
time and space. Richter (p. 59) argues in the spirit of Thomas Kuhns paradigm shift that,
through scientific process of selection on the basis of simplicity and predictive capacity, new
and alternative systems replace the old ones.

A scientist may take a genuine contribution to science in a way, which strengthens an already

established system, e.g. by demonstrating its applicability to a new class of phenomena, but in doing
so he, may very well be contributing to the ultimate displacement of that system by opening up the
possibility of further inquiries which will expose serious defects
in the syste
The Nature and Structure of Philosophy of Science
Scientists are unbiased observers who use the scientific method to conclusively confirm and
conclusively falsify various theories. These experts have no preconceptions in gathering the data
and logically derive theories from these objective observations. One great strength of science is
that it’s self-correcting, because scientists readily abandon theories when they are shown to be
irrational. Although such eminent views of science have been accepted by many people, they
are almost completely untrue. Data can neither conclusively confirm nor conclusively falsify
theories, there really is no such thing as the scientific method, data become somewhat subjective
in practice, and scientists have displayed a surprisingly fierce loyalty to their theories. There
have been many misconceptions of what science is and is not. I’ll discuss why these
misconstruals are inaccurate later, but first I’d like to begin by talking about some of the basics
of what science is.
Science is a project whose goal is to obtain knowledge of the natural world. The philosophy of
science is a discipline that deals with the system of science itself. It examines science’s
structure, components, techniques, assumptions, limitations, and so forth.
The Basic Structure of Science
To properly understand the contemporary philosophy of science, it is necessary to examine
some basic components of science. The components of science are data, theories, and what is
sometimes called shaping principles.
Data are the collections of information about physical processes. Sometimes collecting and
finding data to support theories can be rather laborious. This is because the specific details of
data that come into play can make science such a tricky business that some scientists, when
talking to laymen, sometime leave them out. Also, it is easy to fit a theory in with the data if the
data are vague and overgeneralized. It usually becomes more difficult to fit the theory with
specific data, especially since the details make it more likely for the theory to become less
plausible. Even so, data are important parts of theories and of science.
Theories come in roughly two forms. Contrary to what some might think, a theory in the
scientific sense does not have anything to do with whether or not it is supported by the evidence,
contradicted by the evidence, well liked among scientists, and so forth. It only has to do with its
structure and the way it functions. That is, just because a theory is a scientific theory does not
mean that the scientific community currently accepts it. There are many theories that, though
technically scientific, have been rejected because the scientific evidence is strongly against it.
Phenomenological theories are empirical generalizations of data. They merely describe the
recurring processes of nature and do not refer to their causes or mechanisms. Phenomenological
theories are also called scientific laws, physical laws, and natural laws. Newton’s third law is
one example. It says that every action has an equal and opposite reaction. Explanatory theories
attempt to explain the observations rather than generalize them. Whereas laws are descriptions
of empirical regularities, explanatory theories are conceptual constructions to explain why the
data exist. For example, atomic theory explains why we see certain observations. The same
could be said with DNA and relativity. Explanatory theories are particularly helpful in such
cases where the entities (like atoms, DNA, and so forth) cannot be directly observed.
But before we proceed, let us come to terms with the meanings of the terms “Fact” “Law” and
theory as they relate to our discussion.
Fact: Denotes actual state –of affairs. It is that which corresponds to a proposition; it is what
makes a proposition true. In order words, fact is synonymous with the proposition. Therefore, a
proposition can be said to be true if it described the state -of- affairs, that is, saying or stating of
what is as it is or of what is not as it is not.
(ii) Law: In relation to science, the law is used in terms of the law of nature or better still,
scientific law. Hey describe how nature works and do not describe anything. For
instance, Kepler’s law of planatory notion only describes how planets actually do
move: it does not prescribe how planet should move, and the penalties they would
suffer if they fail. In a sense, scientific laws describe certain uniformities in nature. It
is true, universal, empirical proposition. It applies to all members of a class without
exception; say “All iron rust when exposed to oxygen” is a true universal proposition
different from a statement of fact like “this piece of iron rusts”.
(iii) Theory: It is a unified system of laws or hypotheses with explanatory force
constructed to explain the laws of nature. This distinguished it from scientific law,
which are discovered. One most distinguished feature of scientific theory is that it
normally contains a number of terms that denote an unobserved entity. Take for
instance “there are protons and electrons”.
While it is an ocular fact that the trio are distinct features in scientific endeavour they all the
same have a systematic relationship. Uduigwomen (pp. 69-70) succinctly states the relationship
thus
While particular facts lead the scientist to discover scientific laws, the need to unify and explain
scientific law lead the scientist to construct or devise theories. Thus, while scientific laws unify
and explain facts theories are devised to unify and explain both laws and facts

Scientific Research and Theory Formation


Research in science is a tedious assignment. Its requirements are patience, ingenuity, interest,
availability of technical equipment and the needs for carefulness. Each research or experiment
we embark on in science takes us a long way. Aristotle was right when he stated that each
experiment we embark on adds a little to our knowledge of nature.
The starting point of very experiment is in the mind, it starts at any point in time we are curious
or ask question about something. Our research essentially aims to solve a particular problem.
And the process of this we should be able to ask questions on what see, read, about why and
how things happen and under what circumstances they will happen again. This kind of attitude
helps one to think.
In selecting a problem to research on a lot of care need to be taken. We have to ensure that the
subject matter of the problem we selected is very exciting to us. After selecting a problem it has
to be narrowed down so that it will have a focus. Before the actual research we need to lay down
our groundwork, assemble all the materials and practice any skill needed for the research. We
also need to find out what other researchers have done in the particular area we want to work
on. This step is necessary as it provides one with the benefits of failures or successes of earlier
experiments. It tells what pitfalls may lay ahead and what errors are to be avoided. It gives
information on the techniques and skill on how to solve our problems.
When are research/experiment starts one should be able to give answers problems as they arise.
One is expected to guess at answers – and this is not unscientific - as this kind of guess is called
hypothesis. It is an educated guess based on knowledge, background and experience.
In our experiment of research there is the need for a control like checking temperature, the
inflow of Air, measurement and such other checks that will work positively for the expected
result of the experiment.
The collection, arrangement and the analysis of data constitute the other important phases in our
research/experiment. At this level there is the need for one to keep complete and accurate
records of everything you do and everything that happens. There is the need to avoid at this
level certain psychological factors that can distort one’s power of observation. In
research/experiment there is the temptation to see only these things that favour ones hypothesis
and avoid things that disfavour it. This is a wrong attitude in scientific investigation. The data
we collected has to be properly arranged using a table or a chart as the situation dictates. The
raw data we have collected has to be put in the way that it will be informative.
Our conclusion should be the outcome of the analysis derived from our data. Our personal
opinion should not come in: when it does it has to be stated clearly. If there is anything
unexpected we find in our research it has to be stated if the experience/research fails it has to be
stated. It can be stated that a given experiment has failed to answer a number of questions
posited. In real practice scientists are more eager to advertise their successes than failures. A
good experiment or research work must be reproducible. This informs why all the steps taken in
an experiment have to be stated.
Theories in Science
As an empirical discipline science requires theories. When a crucial experiment give answer to a
proposition than the knowledge derives from such research be comes valuable for tackling
problems in science. And such body of knowledge is usually seen as a theory. Thus in empirical
sciences theories are common features, and they are methodologically formulated to embody
one’s idea. Theories in science are supposed to have strong foundation as the world of science is
a world of competition between rival theories. Hence the logic of scientific research is seen as
theory of theories and scientific theories are seen as universal statements and like it is the case
in linguistics representation they are system of signs or symbols. They “are nets cast to catch
what we call ‘the world’: to rationalize, to explain and to master it. Thus, “A good theory is that
which has passed through a crucial test – that takes us closer to the truth. It is a fact that we may
not get all the truth we require in a given problem, but we can get close to it. A theory should be
able to open up new ideas, be able to make novel predictions, as we have in Einstein theory of
relativity or Lavoisier thesis of combustion.
A theory in science is not supposed to be ever lasting. At any point in time a theory is unable to
answer a number of questions passed to it and within the area it covers, such a theory loses its
validity. And scientists should not have so much anxiety to protect their theories as to ensure
that they have focus. This explains the need for a critical attitude in science. There is no point to
“demand and search for infallible intellectual authorities but (we) should instead try to build a
philosophical programme for counteracting intellectual error”. The concern of philosophers and
scientists should be on how our intellectual life and institution should be arranged to the extent
that it could open up our beliefs, conjectures, policies, sources of ideas, traditional practices,
among other, whether justifiable or not to optimum critical evaluation so as to eliminate
intellectual errors. Theories in science are not absolute truth, they are not supposed to be, and
they are attempts to understand mysteries and hidden laws of nature.

Shaping Principles
Shaping principles are non-empirical factors and assumptions that form the basis of science and
go into selecting a “good” theory. Why are they necessary? Can’t theories be selected solely on
the basis of empirical data? Surprisingly, the answer is no. Why not? Describing some mistaken
views of science come in handy for explaining the answer. It is evident that theories and data by
themselves are insufficient for science to work, and thus other factors are needed for science to
operate. This group of factors in the nature of science is that of shaping principles, which can be
used to select theories and form the foundations of science. Many assumptions are made in
science. One example is the uniformity of nature. That is, the belief that natural processes
operate in a fairly consistent manner. This shaping principle is the basis for the idea of natural
laws. For example, Newton’s laws are said to apply throughout the universe. This is believed
even though scientists have not actually tested the laws everywhere in the universe. Natural
laws could not exist in science without assuming the uniformity of nature. Other assumptions
made for science to operate include that there exists an external objective reality, that our senses
are generally reliable, and so forth.
Another set of shaping principles evaluates the empirical evidence to select theories. Because of
the underdetermination of theories, there is always an infinite number of competing theories that
can accommodate any given set of empirical data. Since these competing theories are
empirically indistinguishable from each other, if science is to pick out a theory from among
these numerous competitors and claim that it is correct, then such a selection must be based on
nonempirical principles (whether they be philosophical, personal, societal, or whatever). The
law of parsimony is one of them. This principle of logic states that, if all other aspects are equal,
the simplest theory is preferred over other theories involving additional factors. This is also
called Ockham’s razor (sometimes spelled as Occam's razor). The law of parsimony is often
used because a theory conforming to this principle fits the data more easily. This principle
especially applies to theories with ad hoc hypotheses. The lower the number of ad hoc
hypotheses a scientific theory has, the better. Other principles include (but are not limited to)
empirical adequacy (covering the pertinent data in some suitable way), self-consistency,
fruitfulness (giving rise to other understandings and having stimulated pioneering investigations
and advancements), and explanatory power. Another key principle is how well a theory ties in
with other scientific theories and concepts that are rational to believe. It is only when these
kinds of shaping principles interact with data can science then provide rational support for a
theory over its competitors.
However, there are a few exceptions to the idea that there is no conclusive proof in science.
Logic is the closest we can get to rigorous proof and falsification. For example, suppose our
friend Bob has this theory: hairless men have no hair. By the rules of logic, Bob’s theory must
be true. Of course, Bob’s theory is a tautology (needless repetition of an idea, in this case it’s
the repeated concept of hairless men), and tautologies are typically not very helpful. Sadly, not
very many helpful theories can be thoroughly proved by logic, and logic disproving a scientific
theory is almost never used because seldom does a scientist propose a theory that is logically
impossible. Most of the time science relies on other shaping principles to pick theories.
It becomes easier to understand these principles when they are put into action. In the “moon is
made of cheese” example, we can reject it because of the law of parsimony. It uses an ad hoc
hypothesis, whereas the theory of the moon being like a rock does not. Often times, of course,
more than one shaping principle becomes applicable. For example, suppose Bob’s computer is
malfunctioning. One theory he has is that an invisible gremlin has caused such problems, and
another is that a computer virus has invaded his machine through his modem, computer
programming, and some fairly complex electronic systems in his computer as well as on the
Internet. The gremlin theory is simpler, and thus it would seem to appeal to the law of
parsimony. Yet the gremlin theory hardly seems empirically adequate in this case. This is
because other considerations need to be taken into account. Another fact to consider here is that
the computer virus theory ties in with electronic concepts that are supported by evidence,
whereas the gremlin theory does not. Because so many shaping principles are used and because
they can often conflict with each other, we should be careful about justifying how much the
evidence supports a theory.
Unfortunately, there are still limitations involved in scientific practice and shaping principles do
not solve the entire problem, even in the basic foundational beliefs of science. Take the
uniformity of nature, for example. We believe nature is consistent enough so that the
experimental data (such as from testing physical laws) obtained from two years ago on Earth
will essentially be the same if the experiments were to be conducted in identical conditions on
Mars next week. But there really is no logical principle to tell us that physical laws will hold
true in places where we haven’t tested them (even if that place is the future).[36] A similar sort
of problem arises when we choose between empirically identical theories. When using shaping
principles to select a theory, we must have some philosophical basis for believing that nature’s
preferences are similar to ours. And for many of these principles there is no logical rule to imply
their reliability. For example, in picking out a theory from among it’s empirically
indistinguishable competitors (and when all other factors are held constant), the notion that
reality favors simple theories over complex ones is nevertheless a philosophical principle.
Although these indicators of theoretical truth are necessary for science to work, they are
significantly indirect, circumstantial, highly fallible, and are still unable to prove/disprove
theories. While science may be the best we can do, the limitations should still be recognized.
On top of that, there is no known clear-cut method that tells us to what degree the evidence
confirms a scientific theory, despite attempts at finding one. This becomes problematic when
scientists must decide on what theory to accept as the most rational one. Scientists intuitively
feel how rational scientific theories are, rather than having a precise logical method for such
judgments. These intuitive feelings result from shaping principles. The interactions of shaping
principles in the minds of scientists are so complex and so numerous that we may never come
up with a rigorously logical system to select theories. Most of the shaping principles are
frequently unspoken and sometimes scientists themselves do not know they are using them.
Although some shaping principles are based on logic, others are not always so sensible and
objective. Scientists (and regular human beings) are also affected by cultural, social, and
personal beliefs. Indeed, such factors have been significant influences in scientific revolutions.
This is because many activities in science, such constructing theories, involve numerous aspects
of oneself. In the case of making theories, the theories themselves are creative inventions that
come from the minds of scientists. Science is a human activity, and what affects scientists will
have an effect on science. One may think that having such unscientific factors affect theory
judgments is bad for science. That may very well be true, but unfortunately there is no known
way to separate the helpful principles (explanatory power etc.) from the unfavorable ones
(personal biases etc.) in the subconscious minds of scientists that make these theory judgments.
Because every human being has their own unique set of shaping principles, different scientists
(and regular human beings) can look at the exact same set of data and disagree about which
theory most rationally explains the observations. Rather than the traditional view that science is
to be protected from biases and other imperfections of people, it turns out that science is
inescapably infected with humanness.

Tapestry
It would seem that there is a delicate tapestry in interpreting the data. It is uncommon for a
theory to be tested in isolation because of the Duhem-Quine problem. Because we often rely on
background assumptions to derive predictions for a theory, and because those background
theories depend on other auxiliary theories and principles for their empirical expectations etc., it
would seem to follow that the collection of theories combined with their shaping and
background principles thus make up an explanatory matrix, or conceptual grid, in which to fit
the data. Modifications to the explanatory matrix can be made in attempts to get a better fit, but
because of the interwoven nature of the tapestry, often times one cannot supplant aspects of the
grid without changing things in some way elsewhere. So it’s possible that the need would arise
for an entire conceptual system to be replaced. Additionally, the nature of science can make it
difficult, if not impossible, to empirically test an individual theory completely independent of
this matrix. However, it is also quite possible for nature to teach us some things in carrying out
our investigation. That is, the interaction between the explanatory matrix and the data can be a
sort of two-way process. As we uncover more data, we can learn better ways to shape the grid
and how to go about it.

( a ) Elements of Physical Science.


The physical science deals with all non-living aspects of man’s environment. The physical
science include such subject areas as:
Astronomy, Chemistry, Geo-science, Physics, Mathematics
Astronomy: Is science that is concerned with the study of the heavens. It deals with the
relationships, movements, compositions, sizes, and distances of the heavenly bodies within and
beyond the solar system.
Through careful observations with the telescope it has been possible to plot the orbit of planets
and comets. An instrument called the spectroscope has provided the means for determining the
temperatures and compositions of distant stars.
Astronomy of the research typology branches into:
Astrometry, Astrophysics, Celestial mechanics, Cosmology, Radar astronomy, Radio
astronomy, Astronautics.
Astrometry is concerned with measurement of the size, distance and motion of heavenly
bodies, Astronautics is the science that enables man to navigate in outer space by means of
instruments which can calculate speed and position relative to the movements of the heavenly
bodies, Astrophysics is a combined science which studies the radiation from celestial objects
which information astrophysicists can derive certain of the physical and chemical properties of
heavenly bodies, including such factors as composition, temperature and motion, Celestial
Mechanics on the other hand investigates the motion of bodies in space as they are influenced
by gravitational attraction; this branch is unimportant in determining the weight and speed of
earth satellites, Cosmology is concerned with the origin, structure and possible evolution of the
entire universe. The assumption here is that the appearance of the observable universe can
provide information about the rest of the universe, Radar Astronomy is the study which involves
the sending of radar signals from each to some nearly heavenly bodies such as the moon, to gain
information about them by means of the radar echoes, Radar Astronomy is the study of radar
waves received on the earth from Distant Stars and Nebulae, In the area of applied astronomy
we have celestial, navigation which is a way of determining one’s location on earth by
measuring the position of two or three stars and by learning where those stars are above the
earth of that particular time.
Chemistry: Is the study of matter and its behaviour when combined with other material. Matter
is investigation to find out what it is made of and how it changes.
Chemistry is important in improving standards of living, as fro e.g to make stronger metals to
improve soil, and to destroy bacterial. It also make possible the development of substances such
as rubber, nylon, and plastics from other completely different materials, Thus through analysis
i.e. taking substances apart and putting them together in different desired combinations, and
synthesis i.e. by combining the parts the chemist may produce a completely new substance or an
improved substance similar to the old one, Knowledge of chemistry is important in a great
number of occupations, such as medicine, agriculture, and engineering to mention but a few, It
helps man to understand his world and change the chemical substances in it to his essential
advantage in varying areas (above), Some common terms in chemistry include among others the
following:
Atom – The smallest element which retains the elements properties: they are indestructible and
cannot be broken down into ordinary chemical means, Elements – Substance composed entirely
of atoms of one kind. It cannot be broken down chemically, Energy – A measure of capacity to
do work, Mass – The quantity of matter is an object or body, Matter – Anything that has mass
and takes up space, Mixture – Substance composed of two or more elements or compounds each
of which retains its individual properties, Molecule – The smallest part of a compound which
retains that compounds properties. The atom in a molecule can be separated by chemical means,
Physical Change – A change in for (solid, liquid, or gas) but not in chemical composition,
Properties – The characteristics of a substance the physical properties (mass, melting point
hardness etc) can be measured and expressed in numbers. The chemical properties (valence,
ionization etc) determine how a substance reacts chemically.
Reaction for Chemical Change – A change in the chemical composition of a substance.
Branches Of Chemistry:
Biochemistry: Explores the chemical changes, investigates chemical changes and products of
living organism e.g metabolism and digestion.
Inorganic Chemistry: Confined to analysing the chemical elements and their compounds that do
not contain carbon i.e. their properties preparation and their uses are investigated. It is further
subdivided into (a) analytical chemistry (b) qualitative analysis (c) quantitative analysis.
Analytic – Study of chemical compound to determine their composition and structure,
Qualitative – To determine the kinds of elements that are present in the compound or nature,
Quantitative – Confined to discovering the quantity of each element. Present in a compound or
mixture.
Organic Chemistry: Deals with chemical compounds, which contain the element carbon. It
includes identifying and purifying compounds found in nature, their synthetic preparation and a
systematic study of compound of unknown structure directed towards future synthesis.
Physical Chemistry: Is a combined science that studies the laws, which control the physical and
chemical properties of matter. It is divided into:
Electrochemistry, which deals with, the changes produced by electric current e.g electroplating,
Nuclear chemistry, which is the study of the effects of bombardment of atomic nuclei subatomic
particles, Photochemistry is concerned with the chemical changes produced by light,
Thermochemistry is the measurement of temperature changes in the chemical system.
In the area of applied chemistry, science fund practical relevance in the human environment in
the following areas of human endeavours:
Agriculture:
The conversion of farm products to industrial use, testing of soil and preparation of chemical
fertilizers, Chemical Engineering: Concerned with developing manufacturing processes, which
involve chemical changes e.g, manufacture of steel and rubber, Control Chemistry: Is the
specialty of analysing samples of materials produced or processed by industry to be sure they
are what they are said to be, Metallurgy: Deals with recovery of metals, from one and with their
purification. Another branch of metallurgy is concerned with the making of metal alloys for
specific purposes, Petroleum Chemistry: Is confined to the manufacture of products from crude
oil, Pharmaceutical Chemistry: Deals with the preparation of drugs for medical use. This
involves the extraction of medical chemical (elements or compounds) from barks and herbs etc,
Synthetic Chemistry: Is concerned with the production of new chemical materials such as
cellophane. It also develops substitute for natural products e.g textile, fibres and rubber, Textile
Chemistry: Is the specialty of testing fibres of synthetic or natural materials for their ability to
take dyes.
Geo-Science.
This branch of the physical science of a combined nature, which includes many aspects of
physics and chemistry as they are, related to the study of earth science i.e. its atmosphere its
surface, and its internal structure. This is done through careful measurements of the earth’s
magnetism, gravity and size.
As a research science, it comprises geochemistry, geology and geophysics.
Geochemistry: Is the study of the chemical composition of the earth and its atmosphere. It
includes a study of past and present chemical changes.
Geology: Deals with the earth’s structure, including successive physical changes of the past and
present. Subdivision under this branch include:
Economic Geology – the study of valuable deposits of minerals like oil, gas, coal, cement, salt
etc.
Geognosy, which is concerned with the materials, which make up the earths and with their
arrangements.
Mineralogy deals specifically with the identification and classification of minerals.
Pedology – the study of soils, including their composition origin and classification.
Petrography – systematic study of the structure, texture and mineral composition of rocks.
Petrology is concerned with the rocks, which make up the crust of the earth, including their
descriptive and classification.
Historical geology deals with the origin and evolution of the earth.
Geodynamics is the study of mechanical processes involved in the formation of the earth.
Geomorphology deals with the origin of the topographic features of the earth.
Stratigraphy is concerned with the order and arrangement of rocks.
Paleontology deals with fossils of now extinct forms of plants and animals, and with their
relationship to the rocks and strata in which they are found.
Geophysics: This is the combined study of physics and mathematics as they are related to the
earth; including its oceans and atmosphere. Subdivisions under this branch include: Cosmic
physics which concerns the electrical particles, which travel through space and strike the earth,
Geodesy, which is the study, which measures the earth to determine its size and shape,
Geomagnatology the study of the earth’s magnetic field, Glaciology, which deals with the
characteristics of, ice on the earths surface and the changes it produces, Hydrology – the stud of
water of the earth, its properties and distribution, Isostasy – the study of the earth’s interior,
Meteorology concerns the atmosphere and its relation to whether and climate, Oceanography
deals with the ocean, including the bottom and shore. It involves relationship between animals,
plants and their environment, Seismology is the study of earthquakes, Volcanology concerns the
forces, which produces volcanic eruptions on the earth’s surface.
Geoscience is applied in the following areas of human endeavour.
Coal technology, which involves the processing of coal for industrial purposes other than direct
combustion. It is then used in the generation of heat and energy.
Engineering geology, which investigates the ground upon which construction is done. This
helps to save both material and human resources from untimely destructions.
Meteorology involves the collection of information regarding whether conditions for the
purpose of providing forecasts.
Mining geology, which helps in the location of mineral deposits.
Petroleum geology, which deals with the location of oil deposits.
Physics
Is the study of forms of energy such as sound, heat and light? It is concerned with the nature and
sources of energy and with investigations of the ways in which one form of energy can be
changed to another. It also involves devising instruments to measure various forms of energy as
for example the investigation and control of nuclear energy.
Physics researchers in the areas of: Biophysics, which as a combined science is the study of
energy as related to the living process.
Electricity and magnetism, which deals with the effects of electricity, particularly its
electromagnetic effects as utilised in the operation of motors and electric generator.
Electronics concerns the stud and control of electrons, especially in relation to vacuum tubes
and transistors.
Heat – this area of physics deals with the nature of temperature changes in substance;
subdivisions under this area includes: Cryogenics – the study of the behaviour of substances at
extremely low temperature.
High temperature – the behaviour of materials of extremely high temperatures such as those
achieved by means of solar furnaces.
Thermoanalysis – the study of the effects of heat on the chemical or physical properties of
materials.
Thermodynamics – the study of heat as it is produced by the molecules.
Light – branch of physics which deals with the physical characteristics of radiant energy as it
affects sight. A major sub-area here is optic which studies visible light – rays. It has also come
to include all phenomena of electromagnetic waves of wave lengths less than those of
microwaves, yet greater than those of x-rays.
Mechanics – concerns the effects of forces acting upon bodies in motion or at test. Sub-areas
under this branch of physics include:
Dynamics – deals with the forces which produces motion or which change station.
Aerodynamics – the study of fluid mechanics as it is related to motion between a fluid, air and a
solid.
Hydrodynamics – this area of physics is concerned with liquids in motion.
Kinematics – the study of motion apart from its effects upon bodies.
Kinetics – deals with the changes in motion as they are caused by forces not in equilibrium.
Statics concerns balanced forces or bodies at rest.
Aerostatics – the study of fluids at rest.
Hydrostatics – the study of forces that are produced by liquids at rest.
Nuclear physics – the study of particles found in the nuclei of atoms, together with the energy
effects produced when the nuclear particles are disturbed by external forces.
Solid state physics deals with the properties and structures of solid materials. It is subdivided
into: Crystallography – the study of crystals. Acoustics – the transmission of sound through
different materials. Refraction, reflection, interference and absorption are studied.
Ultrasonics – concerns high frequency vibrations, far beyond the range of hearing.
Physics, in its applied form translates its entire material nature to man’s essential needs and
desires of good living. The broad science, which does this, is called engineering, which, in a
broad sense deals with the materials and the power that are useful to man.
Subdivision under this area of human endeavour includes:
Acoustic Engineering – the specialty of controlling sound.
Architectural Engineering – the decision and construction of buildings.
Civil Engineering – the design and construction of public property such as bridges, dams,
reservoirs and transportation systems.
Electronic Engineering – design and use of electron tubes and transistors.
Metallurgical Engineering – the extraction of metals from ores and with their purification and
alloying properties.
Public Health Engineering – this involves water supply, sewage, disposal and such projects as
rodent, fly and insect control.

Mathematics
Mathematics is the oldest of the sciences. It began with man’s need to count objects and to
measure distances. It is the most exact of all the sciences; proper use of its methods can provide
only one correct answer to a specific problem. It is the language used by all the other
sciences/all things (Onigbinde 1999:258)
It is divided into:
i. Algebra: The study of relationships between numbers as they are presented by symbols,
though multiplication, division, raising to a power, and extracting a root.
ii. Arithmetic: The science of computation by the use of numbers, through addition,
subtraction, multiplication and division.
iii. Calculus: The system used to figure the rate of change of a function. It is further
subdivided into:
Differential calculus which deals with the rate of change of variables and is a means of finding
tangents of curves.
Integral calculus which concerns 1 belt with the limiting values of differentials and is a means
of determining length, volume and area.
Geometry:
This branch deals with the measurement and relationships of line angles. Subdivisions of this
branch includes:
Analytic geometry – application of algebraic to geometry.
Descriptive geometry – used to solve problems dealing with the space relationships to geometric
forms which comprise an object.
Differential geometry – application of calculus to the study of flat surfaces and curves.
Elliptic geometry – deals with symmetrical forms that are not true cycles.
Plane geometry – concerns the study of polygons, triangles and cycles that can be drawn with a
ruler and compass. It is restricted to magnitudes of two dimensions in a single plane.
Trigonometry – geometric treatment of triangles. Spherical trigonometry involves triangles
inscribed in cycles.
Solid geometry – deals with figures of three dimensions such as cubes, cylinders and spheres.
Topology – geometry of distorted geometric forms.
iv. Statistics:
The collection of numerical facts, together with the processes of tabulation and interpretation. It
is a science of reaching conclusions from materials that are variable and of predicting result in
forms of probability. It is divided into:
Descriptive Statistics – which concerns itself with collecting and tabulating of data and
summarization of processed data.
Mathematical Statistics – this area deals with mathematical proofs used in statistical methods.
Mathematics finds practical relevance in the following areas of human learning.
Actuary studies which involves the calculation of risks and the establishment of premiums and
dividends for insurance companies.
Accounting – the analysis of the financial records used in business industry and the preparation
of statements and reports based on these records.
Similarly, mathematics finds practical relevance in the field of engineering.
In statistics as well, practical application is very much in use. This involves collecting,
tabulating and analysing data to discover relationships between variable happenings so as to
predict the probable outcomes under known conditions e.g. population census as it affects both
birth and death registrations. This help nations to plan their economics, social and infrastructure.
Descartes thus argued that nothing was more sorely needed than precisely this extension of
mathematical reason into wider fields (Collins J. 1967:3). Mathematics thus acts as a key and
primary science. It excludes the false and probable, but also proves its own conclusions by the
most certain and universally accepted demonstrations.
Thus, as the study of symbols, terms and mathematical methods, mathematics as a branch of
science is the key of science which practical application in human relationships is no doubt
indispensable, especially for the purpose of establishing consistency. In its applied nature,
mathematics find expression in literally every area of science (unlike the character of the other
sciences) in the service of all mankind in a practical way. Such is what informed Descartes to
generalize the use of the mathematical method in understanding the operator of nature. Thus,
theories for him, argues Onuobia should be trimmed down to those susceptible mathematical
development (1991:15) perhaps such was what Pythagoras means when he says “numbers
furnished a conceptual model of the universe” (Ibid p. 9).
a)
b) The Nature of the Physical Science
Philosophy is closely linked to science (and technology). All scientific and technological
innovations (and inventions) have conceptual basis.
Man being a rational creature, seeks rational answers to his needs through his rational faculties
by devising systematically techniques modifying and utilizing his environment methodically
and creatively. This as it were brings in philosophy in science and technology, Understood as
the knowledge of the ultimate or basic principles or causes of all things and of knowledge of
ultimate cause of all being, philosophy investigates nature (cosmology). Natural phenomena in
the direction of utility i.e. in order to utilise its resources and forces for human needs, This
involves the explanation of the phenomena of nature from the aspect of their existence and
traces them back to the conditions of their possibility, There are three major aspects of
cosmology and each of them provides man with fundamental knowledge of nature that helps
him to generate techniques and skill to tap the resources and forces of nature.
There is the aspect which inquires into the problems concerning mans knowledge of nature – the
epistemology of nature. This controls both pre-scientific and scientific knowledge of nature.
There is the aspect of cosmology that investigates the natural phenomena and the basic
categories and concepts of the physical science like motion, time, space, power, energy, matter,
weight, life, change etc. – the metaphysics of nature.
This aspect of cosmology is entangled with technology because its investigation of the natural
phenomena and the basic concepts of the physical science in forms far adequately in the
formulation of techniques and skills to confront nature methodically. For instance, it is from the
metaphysical understanding of the principle of motion or change, power or force or energy
matter or mass, space or time etc that man is able to conceive, fashion and establish techniques
about the proper utilisation of motion, power and energy, space and time, improvement of life
etc. this is a clear crossing from philosophy to science, and technology.
Metaphysics of nature, lays effective foundation for techniques (science) “it tries to understand
them as best as it can by reducing them to the ontological conditions of their possibility which
are implicit in the concrete world of nature and by grasping the metaphysical essence of
corporeal existence” (Junks p. 270) the ruling method here is aposteroristic.
Natural philosophy (cosmology) controls the demand of the metaphysics of nature and man’s
actual knowledge of concrete nature in order to arrive at concepts which give an answer to the
question about the natures of things and which show the proper relationship between natural
reality, man and the whole of reality including the Absolute.
Generally, then natural philosophy gives man the fundamental knowledge of nature, answering
the question about the nature of things, in order to be able to generate workable science and
technology to tap resources and forces of nature. It also controls, directs and orients man
towards the proper telos (end) and nature with reference to the whole of reality, the good of man
himself and in reference to the ultimate being: the Absolute.
It may be argued here that there is a rational relationship between the principles that are the
immediate objects of philosophical reflection, the theoretical (scientific) principles and the
practical principles that ‘hardens’ into techniques which we can call technological principles.
Both are drawn from ontological ratio which is the base of every given reality.
Basically every reality draws its meaning from three levels of reason (ratio) namely:
a. Ontological Ratio, Methodological Ratio,
Logical Ratio
1) Ontological Ratio has to do with the basic principles which forms the foundational
meaning of each “given reality”. unraveled in its natural state; and this defines the
amplitude which nature has in every given reality that opens for man’s approach e.g.
There is a ratio in a virgin land that presents it as a cultivatable its natural amplitude
for cultivation
2) Methodological Ratio has to do with the principle which forms the meaning which the
human reason in complimentarity with the ontological season gives as practical
principles workable for man’s appraisal land utilization of the given.
3) Logical Ratio is the deposit of the methodological principles in scientific form
coordinated for further application.
4) The basic question here is that of rationality in science and technology. This to us is a
philosophico-anthropological question. It is the question of how well man uses his
rationality in the exploration of the natural environment or investigation of the
natural phenomena to the telos of man; respecting both the material and spiritual
destiny of man.
The details required by this, postulation may be reserved fro later discussion but suffice it to say
that no better deal could be possible here without venturing into the basic issue of the methods
of the physical sciences which study is architectonic to exposing how well the physical sciences
could be used in healing humanity or how worse it could be in the disservice of humanity – to
kill humanity (attack and destroy man). But such also, can be discussed after a historical
development of science is ventured into.

Scientific Methods
The primary attempts of the philosophy of science are to elucidate the various elements
involved in the process of scientific inquiry. This, though not universally patterned follows the
observational procedures, pattern of argument, method of representation and calculation,
metaphysical presuppositions. Additionally, philosophy of science sets out to evaluate the
grounds to the validity of scientific inquiry from the points of view of formal logic practical
methodology and metaphysics.
Largely the scientific world harbours men with very different professional backgrounds and
interests. This to a large extent has resulted to a barrage of different methods of arriving as a
scientific truth.
It is however noted that, the scientific method is historically traced to Francis Bacon who in his
Novum Organ on (New instrument) showed how scientific method could be applied to the crafts
and experimental facts. That “we should start scientific enquiry with empirical facts”.
Following from this Descartes moderated Bacon’s position insisting rather that “we should start
with general principles which are provided by mathematics. This suggests to us that scientific
method proceeds by deduction rather than induction.
More recently dialectical materialism has been suggested by Karl Marx as the most accessible
methodology. Such for him is the point of departure for any science matters he argues exist in
eternity and alters its modes of existence and occurrence continually. Marx upheld that
All phenomenons are interconnected and this fact is discoverable by scientific method.
(i) The origin of a phenomenon gives the key to its understanding.
(ii) Each natural process is inherently limited because it develops its own
contradiction, which is usually overcome to produce new process.
It suffices to say therefore that the acquisition of scientific knowledge is multi-methodological.
According to him dialectical methodology is the triadic interpretation of things and process viz:
thesis – anti thesis – synthesis or universal particular – individual.
Generally, the most acknowledged characteristics which clearly distinguish scientific
knowledge from other sorts of knowledge come about though the following me thods;
(i) Hypothesis
This is usually a proposal, a statement taken as a starting point of an explanation or reasoning.
Used as a method of scientific inquiry it is a principle of investigation, which is not itself proved
as true. Its acceptance is therefore tentative until proved. Researchers go into the creation or
postulation of hypothesis in cases where an indefinite wait for proved theories constitute great
hindrance or handicap to further investigation.
It is therefore a formal position taken for granted and tentatively used as a means of explanation.
The logical positivist’s on the other hand identified hypothesis and proposition as two
dimensions of verifications. A hypothesis, they say is a proposed law connecting various
observations or much different immediate expression. A hypothesis is potentially testable or
verifiable and when it is eventually verified to become a law, each record of an observation is a
proposition. A hypothesis consists of many propositions, which are many verification of it.
Hypotheses are confirmed only by an appeal to the verifying propositions. This method of
basing general statements or laws on accumulated verifying propositions is known as induction
and is seen by the logical positivist as the haillinare of science.
(ii) Observation
This has to do with the general behaviour of natural phenomena. Apart from mathematics all the
disciplines under the classification of physical science begin with the observation of facts.
Galileo was the first in the history of science to start studying or observing the “behaviour of
falling bodies. From this observation or study, more facts were collected about moving bodies
which refuted earlier theories (or hypothesis) of motion such as that of Aristotle which asserted
that “the rate of fall was prepositional to the weight to the body” in question.
The argued position here is that observation as a method of science seeks to legitimize or reject
a proposal which was argued above is but a starting point of an explanation it is an attempt in a
practical way at proving or disproving a stated position (tentative, yet to be proved position).
Thus progress in science has been largely assumed by the observation of things as regards how
they as a rule behave. This methodology is also referred to as position methodology of science
which spirit is expressed in the attempt to study facts by observing the constant relations
between things and formulating the laws of science simply as the laws of constant relations
between phenomena (Stumpf 1971:343) it was this spirit that prompted Galiled to seek to
understand the involvement and relations of stars without probing into the their physical
constitution. The same spirit dragged Newton to seek to describe phenomena of the physical
without probing into the intricate issue of the essential nature of things. Similarly Fourier
discovered the mathematical laws of the diffusion of heart, without asking of questions
bordering on the essential nature of heat, and curvier formulated some laws about the structure
of living things without delving into the intricate question of the nature of life.
This method is clearly positivistic tic by its insistence on observation as a basis for verifiability.
Hence the logical positivists adopted the verification principle according to which a proposition
(hypothesis) is meaningful if it can be verified directly or is capable of being verified in future
experience.
The use of this method helped to serve as a criterion of demarcation between science of non-
science or Psenso-science and so enhance the unity science of all metaphysical notions. e.g
Epistemological Metaphysical or Ethical assertions are all debunked as meaningless. They do
not live up to condition of empirical verifiability such is why charmers says science is a
structure built out of facts” (1982:1).
It must be said that this method of inquiry is itself inductive. Statements (singular and general)
so inferred from are products of observation or experience, when one infers a general statement
from singular statements he does so provided (when) the number of observation are repeated
under a wide variety of conditions and provided no acceptance observation statement conflicts
with the Derived Universal law or theory, it is legitimate to infer a universal law or theory from
a limited list of statements. For example, concerning heated metals, we can legitimately draw
the universal law that “All metals expand when heated” i.e. generalizing from a finite number of
observation statements to an universal law or theory (we shall explain leads that if a wile variety
of conditions and if all these observed. As without exception possessed the property B then all
as have the property B.
Analogously this method scientific knowledge to a building resting on the secure basis provided
by observation. Chalmer corroborates this thinking succinctly when he says
As the number of facts established by observation and experiments groups and as the fact
become more retuned and esoteric due to improvement in our observational and experimental
skills, so more and more laws and theories of over more generality and scope are constructed
by careful inductive reasoning. The growth of science is continuous, over onward and upward,
as the fund of observation data is increased (p.5)
It is to be said then that scientific method begins with observation, moves to theory and
produces a generalization (or Universal statement with a predictive ability. If the generalization
is a good one, it is (or will be) considered a law of nature. Here understood, science produces
objective results which can be confirmed by anyone who want to go out and repeat the original
tests.
But this simple view of scientific method though widespread even among practicing scientist),
is unsatisfactory in a number of ways. e.g. The assumption that our knowledge and expectations
do not e=affect our observations, and that it is possible to make observation in a completely
unprejudiced way is an inaccurate description of what observation is all about. Our
understanding of psychology/perception argues for us here that seeing something is just having
an image of on your retina. As the philosopher N.R Hanson put it; “There is more to seeing than
meet the eyeball (Nigel Warburton, 1999:115) our knowledge and our expectation of what we
are likely to see affect what we actually do see. Our professional background affect what we see
(e.g. ordinary person seeing telephone cable sees a chaotic tangle of coloured wires. but a
telephone engineer sees patterns of connections order and purpose or trained physicist looking a
an electron microscope and someone from a pre-scientific culture looking of the same
equipment the physicist would understand the inter-relationship between the different parts of
the instrument and would appreciate how to lose it and what could be done with it.
Secondly the simple view of science neglects the nature of observation statements. The
language the scientist uses to move these observation built into it, there is no such thing as a
completely neutral observation statement. Observation statement are theory laden – even an
everyday statement as:
“He touched the base wise and gave himself an electric shock”
Assumes that there is such a thing as electricity and that it can be harmful. By using the word
electric, the speaker presupposes a whole theory about the causes of the harm experienced by
the persons touching the wire.
INDUCTIVISM
The notion of induction was first introduced into public vocabulary by Aristotle as a method of
learning distinct from demonstration. Aristotle stamps his authority in the scientific community
as a pioneer of scientific methodology. In this prior analysis or summative induction, Aristotle
talks of induction as a kind of syllogism in which we reach a universal conclusion from an
exhaustive survey of the cases it covers. In his posterior or intuitive induction, he talk of
induction as the establishment of a universal truth by consideration of an instance or instance
which reveal to though the necessity of the connection asserted. In his estimation therefore
science is knowledge of causes. Thus. Francis Bacon. David Hum, and J.S Mill all assumed that
the business of the empirical scientist was to establish universal propositions about causal
connection though, they differed from Aristotle in the account that they gave of causes.
Inductivism is then methodology of arriving at scientific fact or laws, and theories. Science is
not based on speculative imagination, but on what we observe. This most obviously instigated
the thinking of the textbooks of nature and the treaties of philosophers like Aristotle and Bacon
who both argued that the most reliable source of knowledge is experience for science is a
structure built on facts.
It is pertinent to note that the means to this end is what is called inductionism; the use of which
scientist establish general laws derived rigorously from a great number of particular observation
of facts or experience. This is the principal concern of a scientist. Bacon thus acknowledges that
it is the business of the scientist to discovered the forms of phenomena; a necessary and
sufficient condition for scientific knowledge. He says, ‘we should start scientific inquiry with
empirical facts”. This method involves moving from a limited number of observation statement
to the justification of universal statement. Chalmer (1982:1) restates the principle of induction
thus:
If a large number of A’s have being observed under a wide variety of conditions, and if all the
observed A’s without exception possessed the property of B. then all A’s have the property of B.
The inductivist thinks of scientific knowledge as a building resting on the secure basis provided
by observation. This method proceeds through rigorous arguments to establish a truth. If the
premises are known to be true, then it is probable that the conclusion will be true perhaps the
logic of inductionism is more correctly stated by Chalmer thus:
As the number of facts established by observation and experiment grows, as the facts become
more refine and esoteric due to improvements in our observation and experimental skills, so
more and more laws and theories of over more generality and scope are constructed by
careful inductive reasoning.
This methodology referred to as empirio-criticism is concerned for all or most of the time with
generalizations from experience, through the process of scientific experimentation i.e what is
not empirically verifiable does not qualify as true knowledge.

Mistaken Beliefs of the Scientific Method


Many students (including me) were brought up with a somewhat eminent view of science, or at
least a fairly eminent view of science as it should be done. As I have found however, the status
of science which most of us were taught may have been a bit misleading. Some ideas of what
“the scientific method” is have also been erroneous. This is perhaps because scientists
themselves tend to be ignorant of the philosophy of science. Changes have been made in history
about what science is and how it should be done.
In the early years of science, the system of acquiring knowledge was viewed as completely
objective, rational, and empirical. This traditional view of science held that scientific theories
and laws were to be conclusively confirmed or conclusively falsified based on objective data.
This was supposed to be done through “the scientific method.” Apparently some sort of method
was necessary because humans seemed to have a variety of tendencies and feelings that were
not very trustworthy, including biases, feelings, intuitions, and so forth. These kinds of things
had to be prevented from infecting science so that knowledge could be reliably obtained.
Rigorous and precise procedure (“the scientific method”) was to be followed so that such
imperfections of humanity would not hinder the process of discovering nature.
Baconian inductivism in the early seventeenth century was at one point considered to be the
scientific method. The basic idea at the time was this: collect numerous observations (as much
as humanly possible) being unaffected by any prior prejudice or theoretical preconceptions
while gathering the data, inductively infer theories from those data (by generalizing the data into
physical laws), and collect more data to modify or reject the hypothesis if needed. In many
instances, this concept seemed to work. One can collect numerous observations of physical
processes and experiments to derive natural laws, such the conservation of mass-energy. Alas,
Baconian inductivism is an inaccurate picture of scientific method. When using inductivism to
arrive at natural laws, certain theoretical preconceptions are absolutely vital. To generalize the
data into physical laws, the individual must assume that the laws apply for physical processes
not observed. This results in several assumptions being held, such as the uniform operation of
nature. Even if we put aside the fact that inductive logic is invariably based on such
postulations, there is another problem. Science deals with concepts and explanatory theories that
cannot be directly observed, including atomic theory and the theory of gravity. Many other
theories include unobservable concepts like forces, fields, and subatomic particles. There is no
known rigorous inductive logic that can infer those theories and concepts solely from the data
they explain. If inductivism is the correct scientific method, then such theories cannot be
legitimate science. As if these difficulties weren’t enough, inductivism has other major technical
problems that have led to its demise.
Sir Isaac Newton developed hypothetico-deductivism in the late 1600s (though the method was
actually named at a later date). Essentially, one starts with a hypothesis (a hypothesis is
basically a provisional theory) and then deduces what we would expect to find in the empirical
world as a result of that hypothesis, hence the name hypothetico-deductivism. Here the idea was
to quarantine human irrationality. One could make a theory for any or no reason. The sources of
theories would be irrelevant in hypothetico-deductivism since the theories could be tested
against the empirical world and be confirmed or refuted that way. A theory did not become a
good theory by its origins, but because of the hypothetico-deductive method of verification.
Inductivism, recall, could not work because empirical data cannot be the sole source of a theory.
Some scientists and philosophers of science who rejected inductivism embraced hypothetico-
deductivism. A significant reason is that it allowed ideas like atomic theory to be legitimate
science whereas they would not be in inductivism.
Unfortunately, hypothetico-deductivism also has problems. The philosophy that rigorous proof
is necessary for good science has serious problems even if we assume that sense experience,
memory, and testimony are all generally reliable. For one thing, we cannot be sure that we have
examined all the germane data. There is always the opportunity for future observations to topple
even the most established of theories. For example, there is always the possibility that an
observation could conflict with any known scientific law. This is what caused Newtonian
mechanics to be cut down to size. Rather than being a total account for the nature and dynamics
of the universe, Einstein, Heisenberg, and other physicists demonstrated that the realm of
Newtonian mechanics is much more restricted than what was once thought. Unrevealed data can
also contradict the predictions of any explanatory theory as well. Every theory has an infinite
number of expected empirical outcomes, and we are incapable of testing all of those
expectations. So even though a theory can be confirmed to some extent by empirical data, it can
never be conclusively confirmed. Apart from this, hypothetico-deductivism’s method of
verification has this sort of structure where T is a theory and D a set of data that we would
expect if the theory were true:
1. If T then D.
2. D.
3. Therefore, T.
This is not a logically valid argument. Indeed, an argument of this sort of structure is called the
fallacy of affirming the conclusion. Have T = “An invisible unicorn from Mars flew into the sky
to cause rain,” and D = “It is raining.” Logically, the first premise must be correct (If T is true,
then D would be true). Suppose the second premise is correct. It is raining. Even so, the
conclusion doesn’t logically follow. Why doesn’t it work? Because there could be other
possibilities for D other than T. That is, more than one theory could exist to explain the data.
And this is indeed the case. In this example, it could simply be natural weather patterns, not a
flying invisible unicorn from Mars, that caused the rain. In science or anywhere else, any given
body of data (no matter how large) will always be agreeable with an unlimited number of
alternative theories. Invariably there are many theories that explain the exact same data, and at
least some of the theories will contradict each other. This fact is sometimes expressed as data
underdetermining theories, or is simply referred to as the underdetermination of theories.
Because such competing theories are consistent with the same set of data, all of these theories
are empirically identical. This means that empirical data by itself cannot exclusively confirm
one theory from among its empirically indistinguishable competitors. Some of these empirically
indistinguishable theories may be elegantly simple and others may be outrageously complex,
but multiple alternatives exist for any set of data. There are examples of this problem in the real
world. In one such instance, Tyco Brahe and Copernicus each had a competing theory of the
solar system. It can be shown mathematically that every bit of data that is predicted by one
theory would be predicted by the other theory. We may not always be able to think of
alternative theories, but this only has to do with problems of human imagination in constructing
such theories, not the logic of the circumstances. Of course, the underdetermination of theories
also poses yet another problem for Baconian inductivism. (Explanatory theories cannot be
inferred from data alone if there are always numerous alternatives that explain that same set of
data.) As a result of the underdetermination of theories and the risk of undiscovered,
contradictory empirical evidence, a scientific theory cannot be conclusively proven merely
through the data. Even if we take out the notion of conclusive proof from hypothetico-
deductivism, it seems that this idea of the scientific method dreadfully oversimplifies how
science works. No rational scientist would accept the flying invisible unicorn from mars theory
simply because it passed the empirical confirmation test in the above example, for instance.
Popperian falsification is another belief of what the scientific method is. Karl Popper, regarded
by many as one of the finest and most influential philosophers of science of the twentieth
century, realized the flaws of inductivism and rejected it. Popper recognized that one could not
record everything observed, because that is simply not feasible. Some sort of selection is
needed, and thus observation is always selective. That being true, Popper believed that a
hypothesis had to be created first for scientific investigation to begin. Otherwise there would be
no other way to tell which data are germane. Since theories must be created first in order to
decide what observations were relevant, such theoretical preconceptions would be essential to
doing science (contrary to Baconian inductivism). This was one of the reasons he believed
inductivism is unworkable. He also denied the concept of conclusive proof and instead stressed
the idea that falsifiability is the necessary criterion for a theory to be legitimate science. In other
words, if a theory cannot be falsified through some conceivable observation, then such a theory
is not genuine science. The necessity for a scientific theory to be conclusively falsifiable is
known as the demarcation criterion. This idea seemed reasonable enough, since scientific
theories can make predictions. Popperian falsification suggested that if a prediction does not
come true, then the scientific theory must be false. Popper’s idea of the scientific method was
for scientists to test scientific theories in experiments where the outcome could potentially
falsify the theory, especially in experiments where the theory would most likely collapse.
Science still had some of its traditional quality in that it could make definite progress by
conclusively eliminating theories.
Yet, like inductivism, Popper’s ideas are not entirely successful either. (Consequently, some
regard Popper’s contribution to the philosophy of science to be overrated.) Popper was certainly
correct that data are selective, but they need not have a theory to guide the selection (though that
is often the case). For instance, one can record data and apply assumptions to the data to form a
theory, as is sometimes the case with scientific laws. (Note that since assumptions need to be
accepted for the theory to be created, this would not be an example of inductivism without
assumptions in action.) The demarcation criterion is even more flawed. Surprisingly, the
problem is that it is impossible to conclusively falsify theories by empirical data. One reason is
that theories by themselves are incapable of making predictions. Instead, the empirical
consequences of a theory invariably rest on background assumptions (also called auxiliary
assumptions) from which to derive predictions and even to obtain data.
Suppose we have a particle theory that says if we process a certain particle in a particular way,
we will get specified values on various measurements.
1. All theories (the particular electrical, atomic, particle, etc. models that are used) involved
in deriving the prediction are correct;
2. The specific version of those theories and models (from #1) from which the predictions
are derived from are correct (for example, belief in atoms have been widely accepted for
quite some time now, but the precise details and models of the exact composition,
components etc. have significantly varied.);
3. The prediction derived from those theories and specific versions of those models is
mathematically or logically correct; and
4. Some other things we’ll skip.
Note that most of the items depend on scientific theories. But scientific theories, remember,
cannot be conclusively proven. The dependence on background assumptions to make
predictions is sometimes called the Duhem-Quine problem. There are real life examples of this
problem. To “disprove” the idea that the earth was moving, some people noted that birds did not
get thrown off into the sky whenever they let go of a tree branch. That data is no longer
accepted as empirical evidence that the earth is not moving because we have adopted a different
background system of physics that allows us to make different predictions. So if a theory’s
prediction does not come true, one can claim that the theory is correct and that at least one of the
auxiliary assumptions is wrong.
Besides using auxiliary assumptions to make predictions, such assumptions are necessary to
find out if the predictions come true. Suppose that in order to test our particle theory in the real
world we must use a certain particle accelerator in a particular way. To experimentally test this,
we must adhere to the following statements:
1. All of the theories and models (particle, electronic, engineering) used in what we believe
happens inside this accelerator are correct (including the specifics);
2. All theories (electronics and so forth) on how the detector works are correct (including
the specifics of the models involved);
3. Both the detection devices and the accelerator are operating as designed;
4. Both of the above devices are being used properly (including the assumption that the
readings are recorded correctly); and Some other things we’ll skip again.
Notice that several of the items are again dependent on scientific theories, which cannot be
rigorously proven. Suppose the prediction does not come true and we observe that, “this particle
did not have the specified properties that it should’ve had.” That observation would be heavily
dependent on theories. Although it is possible that our theory could be wrong, it is also possible
that instead one or more of the assumptions listed are wrong. Often, the terminology used to
describe experimental results in addition to the measurements and instruments used in testing
theories make up another set of background assumptions. The dependence on such postulations
for obtaining data is described as observations being theory-laden. In this example, we have to
assume these kinds of assumptions (including #1, #2, #3, and #4 on the list above) to accept the
observation of what properties the particle produced. A completely theoretically neutral
language for recording data is not always possible. Suppose Bob’s prediction comes true. There
is still the possibility of the background assumptions being wrong. Consequently, theories can
neither be conclusively proven nor conclusively falsified by empirical data.
Also, it is possible to salvage a troubled theory or make arguments against a well-supported
theory. This can be done because one can alter auxiliary assumptions to produce different
predictions or change the meaning of theory-laden observations. For example, suppose I
proposed the theory that the moon is made of cheese. To refute this theory, many people would
point out that astronauts have gone up there and found out that it is more like a rock than a huge
piece of cheese. I could counter that argument by saying something like, “the moon with its
great age would naturally accumulate massive quantities of rocks and other particles from space.
Under that layer of space debris, however, is the cheese.” This type of argument that explains
away such evidence is called an ad hoc hypothesis, especially if the theory-saving device lacks
further significant evidence to support itself. Of course, it is possible to rationally discard this
absurd theory, but the point is one cannot do this merely by pointing to the data. When the right
ad hoc hypotheses are made, the theory of the moon being made of cheese becomes empirically
identical to the moon being rock-like. This sort of thing is not limited to ridiculous theories
about the moon’s composition. It’s possible to modify virtually any theory so that it’s consistent
with whatever data that might come up.
Despite the fact that Karl Popper was not completely successful, he did make some useful
contributions. He pointed out that data are selective and subject to human choice (and thus
demonstrated that data are not quite as objective as once thought). He also showed the flaws of
inductivism and why a theory cannot originate exclusively from empirical data.
So it does seem that, if the only way to evaluate theories is in terms of empirical predictions,
science is in trouble. In testing theories, scientists use auxiliary assumptions for which they have
rational reason for being true, even though the assumptions and theories are not conclusively
proven. Yet, given the underdetermination of theories, we can’t just pick a theory and justify it
solely by the data. We can’t even justify a particular theory as probable by the empirical
evidence since there are an infinite number of other theories that can explain the exact same set
of data. How can science function?
Shaping Principles
It is evident that theories and data by themselves are insufficient for science to work, and thus
other factors are needed for science to operate. This group of factors in the nature of science is
that of shaping principles, which can be used to select theories and form the foundations of
science. Many assumptions are made in science. One example is the uniformity of nature. That
is, the belief that natural processes operate in a fairly consistent manner. This shaping principle
is the basis for the idea of natural laws. For example, Newton’s laws are said to apply
throughout the universe. This is believed even though scientists have not actually tested the laws
everywhere in the universe. Natural laws could not exist in science without assuming the
uniformity of nature. Other assumptions made for science to operate include that there exists an
external objective reality, that our senses are generally reliable, and so forth.
Another set of shaping principles evaluates the empirical evidence to select theories. Because of
the underdetermination of theories, there is always an infinite number of competing theories that
can accommodate any given set of empirical data. Since these competing theories are
empirically indistinguishable from each other, if science is to pick out a theory from among
these numerous competitors and claim that it is correct, then such a selection must be based on
nonempirical principles (whether they be philosophical, personal, societal, or whatever). The
law of parsimony is one of them. This principle of logic states that, if all other aspects are equal,
the simplest theory is preferred over other theories involving additional factors. This is also
called Ockham’s razor (sometimes spelled as Occam's razor). The law of parsimony is often
used because a theory conforming to this principle fits the data more easily. This principle
especially applies to theories with ad hoc hypotheses. The lower the number of ad hoc
hypotheses a scientific theory has, the better. Other principles include (but are not limited to)
empirical adequacy (covering the pertinent data in some suitable way), self-consistency,
fruitfulness (giving rise to other understandings and having stimulated pioneering investigations
and advancements), and explanatory power. Another key principle is how well a theory ties in
with other scientific theories and concepts that are rational to believe. It is only when these
kinds of shaping principles interact with data can science then provide rational support for a
theory over its competitors.
However, there are a few exceptions to the idea that there is no conclusive proof in science.
Logic is the closest we can get to rigorous proof and falsification. For example, suppose our
friend Bob has this theory: hairless men have no hair. By the rules of logic, Bob’s theory must
be true. Of course, Bob’s theory is a tautology (needless repetition of an idea, in this case it’s
the repeated concept of hairless men), and tautologies are typically not very helpful. Sadly, not
very many helpful theories can be thoroughly proved by logic, and logic disproving a scientific
theory is almost never used because seldom does a scientist propose a theory that is logically
impossible. Most of the time science relies on other shaping principles to pick theories.
It becomes easier to understand these principles when they are put into action. In the “moon is
made of cheese” example, we can reject it because of the law of parsimony. It uses an ad hoc
hypothesis, whereas the theory of the moon being like a rock does not. Often times, of course,
more than one shaping principle becomes applicable. For example, suppose Bob’s computer is
malfunctioning. One theory he has is that an invisible gremlin has caused such problems, and
another is that a computer virus has invaded his machine through his modem, computer
programming, and some fairly complex electronic systems in his computer as well as on the
Internet. The gremlin theory is simpler, and thus it would seem to appeal to the law of
parsimony. Yet the gremlin theory hardly seems empirically adequate in this case. This is
because other considerations need to be taken into account. Another fact to consider here is that
the computer virus theory ties in with electronic concepts that are supported by evidence,
whereas the gremlin theory does not. Because so many shaping principles are used and because
they can often conflict with each other, we should be careful about justifying how much the
evidence supports a theory.
Unfortunately, there are still limitations involved in scientific practice and shaping principles do
not solve the entire problem, even in the basic foundational beliefs of science. Take the
uniformity of nature, for example. We believe nature is consistent enough so that the
experimental data (such as from testing physical laws) obtained from two years ago on Earth
will essentially be the same if the experiments were to be conducted in identical conditions on
Mars next week. But there really is no logical principle to tell us that physical laws will hold
true in places where we haven’t tested them (even if that place is the future).[36] A similar sort
of problem arises when we choose between empirically identical theories. When using shaping
principles to select a theory, we must have some philosophical basis for believing that nature’s
preferences are similar to ours. And for many of these principles there is no logical rule to imply
their reliability. For example, in picking out a theory from among it’s empirically
indistinguishable competitors (and when all other factors are held constant), the notion that
reality favors simple theories over complex ones is nevertheless a philosophical principle.
Although these indicators of theoretical truth are necessary for science to work, they are
significantly indirect, circumstantial, highly fallible, and are still unable to prove/disprove
theories. While science may be the best we can do, the limitations should still be recognized.
On top of that, there is no known clear-cut method that tells us to what degree the evidence
confirms a scientific theory, despite attempts at finding one. This becomes problematic when
scientists must decide on what theory to accept as the most rational one. Scientists intuitively
feel how rational scientific theories are, rather than having a precise logical method for such
judgments. These intuitive feelings result from shaping principles. The interactions of shaping
principles in the minds of scientists are so complex and so numerous that we may never come
up with a rigorously logical system to select theories. Most of the shaping principles are
frequently unspoken and sometimes scientists themselves do not know they are using them.
Although some shaping principles are based on logic, others are not always so sensible and
objective. Scientists (and regular human beings) are also affected by cultural, social, and
personal beliefs. Indeed, such factors have been significant influences in scientific revolutions.
This is because many activities in science, such constructing theories, involve numerous aspects
of oneself. In the case of making theories, the theories themselves are creative inventions that
come from the minds of scientists. Science is a human activity, and what affects scientists will
have an effect on science. One may think that having such unscientific factors affect theory
judgments is bad for science. That may very well be true, but unfortunately there is no known
way to separate the helpful principles (explanatory power etc.) from the unfavorable ones
(personal biases etc.) in the subconscious minds of scientists that make these theory judgments.
Because every human being has their own unique set of shaping principles, different scientists
(and regular human beings) can look at the exact same set of data and disagree about which
theory most rationally explains the observations. Rather than the traditional view that science is
to be protected from biases and other imperfections of people, it turns out that science is
inescapably infected with humanness.

Tapestry
It would seem that there is a delicate tapestry in interpreting the data. It is uncommon for a
theory to be tested in isolation because of the Duhem-Quine problem. Because we often rely on
background assumptions to derive predictions for a theory, and because those background
theories depend on other auxiliary theories and principles for their empirical expectations etc., it
would seem to follow that the collection of theories combined with their shaping and
background principles thus make up an explanatory matrix, or conceptual grid, in which to fit
the data. Modifications to the explanatory matrix can be made in attempts to get a better fit, but
because of the interwoven nature of the tapestry, often times one cannot supplant aspects of the
grid without changing things in some way elsewhere. So it’s possible that the need would arise
for an entire conceptual system to be replaced. Additionally, the nature of science can make it
difficult, if not impossible, to empirically test an individual theory completely independent of
this matrix. However, it is also quite possible for nature to teach us some things in carrying out
our investigation. That is, the interaction between the explanatory matrix and the data can be a
sort of two-way process. As we uncover more data, we can learn better ways to shape the grid
and how to go about it.
Limitations of Science as a Result of Scientists
Some have pictured the scientist as a completely objective individual who is free of bias and
preconceptions, and who is willing to quickly abandon even the most well accepted theory if it
were shown to be scientifically inadequate. This belief is not close to the truth. The reality is
that scientists are humans, and humans are fallible beings. They have weaknesses just like the
rest of us. For one thing, a bias towards favored theories is actually built into all scientific
research. (Recall the necessity of background assumptions to make predictions and test
theories.)
A related imperfection, and to many a startling one, is a shaping principle called tenacity (also
referred to as belief-perseverance by psychologists). Scientists throughout history have shown a
surprisingly severe loyalty to their theories, even with theories that are in trouble with the
evidence. Furthermore, this sort of tenacity persists in scientists for rather long periods of time.
Why is this the case? The reasons become clear when one considers what scientists do in their
field of work. When people put enormous amounts of effort into something over great lengths of
time, as scientists often do with their theories, they have a tendency to become attached to it.
Scientists in such cases have an inclination to want the theory to be true and it becomes
psychologically more difficult for them to reject it as false, even if they are presented with
strong evidence against the theory. The satisfaction of destroying a theory one has arduously
worked for can be small compared to watching the theory become successful. Furthermore, the
reluctance to give up long-held beliefs is part of human nature, and scientists are not immune to
it. Not many of us would renounce the idea that two plus two equals four even if we were
presented with a mathematical proof disproving that idea. Consequently, a scientist whose
career and livelihood are invested in a scientific theory will probably not give it up effortlessly.
Needless to say, not everyone has been aware of this, including scientists. How is it then that
new theories emerge in science? Nobel prize winning physicist Max Planck has said, “A new
scientific truth does not triumph by convincing its opponents and making them see the light, but
rather because its opponents eventually die, and a new generation grows up that is familiar with
it.”
However, tenacity is not necessarily a bad thing. Ironically, belief-perseverance is one of the
reasons science has advanced as far as it has. This is because scientific theories are not perfect,
and the only way to make real progress with a theory is to be committed to it. Virtually every
scientific theory has some sort of problems with the scientific evidence; which are sometimes
explained away by ad hoc hypotheses, at times there is some waiting for the problems to be
eventually solved, sometimes the problems are unnoticed, at times they are simply ignored, and
from time to time a theory is kept because there is no better alternative. If science abandoned
every theory that had contradictory evidence, science would barely have any theories at all.
Furthermore, if a theory’s problems are eventually solved, then we have tenacity to thank for
preventing the premature abandonment of the theory. Besides that, consider this hypothetical
case. Suppose a scientist who possesses no tenacity writes a paper for a scientific journal and
points out all the ways a concept or experiment might be flawed. Such a paper is likely to be
rejected. Part of the responsibility of a scientist is to provide the most favorable case for his
theories and leaving the criticism of the theories to other scientists. Belief-perseverance helps
accomplish this and thus can work well when science deals with theories that only a few
scientific workers really care about. When significant tenacity to an accepted theory is only
limited to a single scientist or a small group of scientists, the theory can easily be weeded out.
So some amount of tenacity is reasonable, and is part of what makes science function.
Nevertheless, tenacity can become a major problem when the majority of scientists fervently
accept a scientific theory that does not have enough rational support. Naturally, there is an
extent where the amount of tenacity becomes excessive and it’s time to abandon the theory in
favor of a different theory that has more evidence behind it. Unfortunately, there is no clear-cut
agreeable procedure to decide when such scientific concepts should be discarded. Feelings and
other shaping principles play a part in deciding when that time should come, and scientists can
sometimes disagree reasonably on that issue.
Another imperfection is that of observation. Because scientists are human, we cannot obtain
completely objective observations even if there could be total theoretical neutrality. One time it
was believed (because of direct observation) by Thomas Huxley that he discovered a being
halfway between a living organism and a dead one. Many other scientists made observations
that came to support that view. Later, however, it was discovered to be purely mineral. Over a
hundred independent observations corroborated Rene Blondlot’s concept of N-rays, but later it
was discovered that there were no such things as N-rays. These are, of course, extreme cases,
but it does demonstrate that data are not totally uncontaminated by humans. In practice, data are
somewhat subjective. This is because shaping principles influence the data we perceive, and
also because of the tendency for the mind to unconsciously fill in patterns based on these
notions. Such human contamination is called internal theoretical orientation of data. As a result,
totally objective data cannot be obtained.
Besides honest confusion of data, there is also deliberate distortion. Often times the scientist
who commits the fraud thinks he knows the answer. Some people may have justified faking the
data by thinking they were just speeding up the process. Some examples include that of Cyril
Burt; a psychologist who forged data on identical twins to support the idea that intelligence was
inherited. It is possible this was done because finding thirty-three identical twins who were
separated at birth would be a bit tricky. A more famous case would be that of Piltdown man, an
alleged missing link in human evolution. This is also an example of internal theoretical
orientation of data, because the fraud was an obvious one and yet persisted for over forty years.
Of course, these things do not happen all the time, but it should be noted that scientists are not
perfectly moral beings either, and sometimes this can have a debilitating effect on science.
Religion, Philosophy and Science
The notion that religion and science have constantly been at war is not without foundation. It is
true that there have been some religious people who have disagreed with the scientific
community (e.g. Biblical creationists). It is also true that many religious people once held views
contrary to what is now accepted (such as the Catholic Church accepting geocentricism).
However, these sorts of events should not be overgeneralized. While many attempts have been
made to show that religion is unhealthy for science (particularly in the 19th century),
contemporary historians see that work as more propaganda than legitimate history.
Even so, some believe that religion and science are utterly incompatible. Actually, that view is
relatively recent. It dates back not to Galileo, but to the liberal theologians of the Enlightenment.
(Incidentally, Galileo was not actually branded a heretic, the sentence he received was for
disobeying orders.) Not every educated person believes that science is against religion. There
are a growing number of people who believe otherwise, and that have rational support for the
idea that theology and science cannot be totally separated. Many scientists (including Newton,
Faraday, and even Galileo) have been deeply religious. To add to that, some scientists have
actually implanted their religion into their scientific work, including Newton, Boyle, Maxwell,
Pasteur, and others. Clearly, religion and science are not always bitter enemies.
Also, the evidence suggests that religion (and more specifically the theistic philosophy that
stemmed from the Christian worldview) was a significant factor in the birth of modern science,
at least partly because it provided some unique philosophical principles that science requires.
Why, for instance, would a rational investigation of nature be successful? Because a rationally
orderly God created the universe. (Nature consistently operating in mathematical patterns would
especially be confirmative for this belief.) According to the Christian religion of that time and
area, the universe is orderly, this orderly world can be known, and there is a motive to discover
this order. Indeed, many of the founders of modern science were Christians trying to
demonstrate that humanity lived in an orderly universe. Why should the investigation of nature
be empirical? Because God could have created an orderly universe in more than one way. This
sort of mindset is rather different from classical atheism (which was even accepted in the 16th
century), which holds to the metaphysical view of a universe dominated by chance events. This
philosophy hardly implied an orderly universe.
Conclusion
So what exactly is the scientific method? Although scientists certainly do something in their
field of work, there really is no such thing as the scientific method. This is true for a number of
reasons. First, the majority opinion in the scientific community is often wrong. Someone not
going along with what the majority does can produce something scientifically useful, and this
has been done many times. This is the position advanced by Paul Fayerabend in his Anarchistic
method of anything goes. Second, science has many specialized fields, and scientists in those
fields require certain craft skills unique in that field to conduct experiments. Such experiments
do not involve precise rules that give detailed instructions on what to do at each step. What may
appear to be misconduct to an outsider may actually be quite valid scientific practice in that
field. Furthermore, rapid progress in science will be more likely if scientists do not follow a
single standardized method. Individual scientists have numerous ways of making theories and
evaluating them, which explains why there can be disagreements among scientists. The different
shaping principles that interact with data can produce different results with each scientific
worker, including on how scientists should approach things. Sometimes these disconformities
help to produce useful scientific revolutions. At times revolutions in science happen in large
part because these kinds of shaping principles that are accepted by the majority change over
time. Great changes in shaping principles create another reason why there has never been a
single scientific method used by all scientists. Although there are some general objectives to
achieve in science (e.g. finding scientific theories that are rationally supported), there are a
number of ways to go about this, and not every scientist shares the same method.
It does seem that science contains various imperfections and some serious limitations on
certainty. Many have pointed out the existence of technology as a sign that we are on the right
track. But just because technology works doesn’t necessarily mean that our theories of why it
works are correct. Often, the reliability of technology depends more upon empirical
regularities, rather than explanatory concepts. For example, candles and light bulbs have
worked and will continue to work even though our theories of why they work have changed over
time (light as particles, waves, or some combination of the two; the rejection of the phlogiston
theory of heat, etc.). The underdetermination of theories applies to explaining the effectiveness
of technology just like any other data. Some have believed that science has been successful in
acquiring knowledge, yet there really is no way of verifying this. Data are incapable of
conclusively proving theories, and we can’t exactly read an omniscient “book of truth” to see
how often our theories have been correct. Historically speaking, almost every theory in science
eventually becomes discarded as wrong. Consequently, there have been so many false starts in
science that it would be rather incredible if we were the ones who are finally on the right track.
It would be especially amazing considering that the theories that we’ve already discarded have
not even been conclusively falsified by the data. Even so, this is not to say science isn’t worth
having around. On the contrary, science provides significant benefits for humanity. For one
thing, science has helped us to alleviate the struggle to survive. Whether or not we are on the
right track, it seems clear that science is conducive for useful technology. Various aspects of
science can be used for the needs of people, understanding ourselves and even our place in the
universe. Although there is a very real possibility of being wrong, we can increase our chances
of being right through further accumulation of data. Despite all its imperfections and limitations,
science may very well be the best tool we have for discovering nature.

The Ancient Greek Period


Although existence of philosophy of science as a separate and self conscious branch of
philosophy occurred relatively recently—within the last two or three centuries in Western
thought—its beginnings go back to the beginning of philosophy altogether. The first epoch
occurred in ancient Greek philosophy, in what is now known as the Pre-Socratic period,
encompassing those philosophers now known as the Pre-Socratics: Thales, Anaximander,
Anaximenes, Xenophanes, Pythagoras, Parmenides, Heraclitus, Anaxagoras, Empedocles,
Democritus, and Protagoras, and going on to include the heights of Greek philosophy with
Socrates, Plato, and Aristotle. The main emphasis of the Pre-Socratics was determining the
basic elements of the universe: Proposals included water, air, fire, the boundless, numbers,
atoms, and being itself. An attempt to find the ordering principle of existence—or at least an
ordering principle—was undertaken; mind and mechanical causation were suggested. Later
on in the Pre-Socratic period, with the arrival of the Sophists, attention shifted from nature to
man; we could call this a movement away from natural science to social science.
Another central feature of the Pre-Socratic epoch was a movement away from a deocentric
(God or gods-based) explanation of things to a naturalistic (nature-centered) one. One could
say, for a salient example, that the first verse of Genesis in the Bible, "In the beginning God
created the heavens and the earth," (Genesis 1:1), offered a deocentric account of the origin
of everything, and implied, tacitly if not explicitly, that study of nature, or natural science,
would need to begin with or at least make central reference and connection to theology. The
Pre-Socratics, however, moved away from that view, and since their time there has been
tension and open—and sometimes heated and hostile—disagreement among philosophers,
scientists, and theologians and religious-minded people about whether there is or should be a
role for or connection between science and religion and between philosophy and theology.
Socrates and Plato were more known for their interests in ethics and politics than in natural
philosophy, but they did make important and lasting contributions here too. One of those was
a focus on definition, or drawing boundaries around things or concepts.
Socrates seems to have thought that to know is to be able to define. Later on, especially in the
twentieth century, what is known as the demarcation problem became prominent in
philosophy of science; this was the problem of drawing a hard or bright line between science
and non-science.
The Rationalists and the Growth of Scientific Knowledge

Introduction
Professor S. B. Oluwole (1991:40) once argued that, “one of the fundamental
problems of the philosophic enterprise today is that philosophers themselves are not fully
agreed on the definition of the main tasks, goals, and the challenges of philosophy”. Suffice
it to say that, the suggestion by R.J. Hirst (1968:8) that, “philosophy is the rational
investigation of certain fundamental problems about the nature of man and the world he lives
in” gives us cause to argue that philosophy attempts to provide rational solutions to such
problems. How this is done or can be done, is what we shall soon undertake; most obviously
by going back to the traditional claim that the philosopher possesses a special intuition which
enables him to draw peculiarly philosophical conclusions from everyday experience. In
venturing into this undertaking, we are not unmindful of our mission; to argue out a necessary
link between philosophy (rationalism, idealism) and technology (the art of doing things) by
which implication we shall conclude that scientific technology involves the application of
reason to techniques.
Understanding Science:
Science means “knowledge”, used in a wide sense, science is the systematic study of
anything according to laid down intrinsic principles (i.e scientific methods). Thus, any study
carried out using this model is a scientific study. The knowledge derived from such a study is
scientific knowledge. In the narrow sense of the word, science is restricted to the positive or
empirical sciences such as physics, chemistry, biology among other areas of the physical
sciences. Our work adopts its wide sense of usage in the tradition of Aristotle through the
medieval to contemporary scholars who as it were used the term to mean episteme i.e
theoretical knowledge
Etymologically, the word science is derived from the Latin word scientia, i.e.
“knowledge”. Understood as a human activity, science seen as “knowledge” is a human
undertaking to learn about the world around us through a special method called “scientific
method”. Suffice it to say, however, that, there is no univocal definition of science. This is
partly because the standpoint from which authors look at science differs. In his book What is
Science? (1952) Norman Campbell writes that science can be looked at from two aspects:
firstly; science is a body of knowledge and a method of obtaining it. Secondly, science is a
pure intellectual study, and so in this regard akin to painting, sculpture or literature rather
than the technical arts. Understood in this light, science aims only at satisfying the needs of
the mind and not those of the body. It appeals to nothing but the disinterested curiosity of
mankind.
We may say perhaps that though limited in scope, the second definition of science is
closely linked with the first. Both project science as a whole body of knowledge, logically
interconnected and directed at achieving some desired goal; spiritual or material. Such an
endeavour requires systematic coherence, objectivity and standardised method as its
important ingredients. This conception of science may have informed Amadi’s definitions of
science, that it could also mean (i) knowledge, especially of facts or principles gained by
systematic study. (ii) a particular branch of knowledge especially one dealing with body of
facts or truth systematically arranged and showing the operations of general laws as the
science of mathematics. (iii) systematised knowledge especially of the laws and facts of the
physical or material world. He thus sums up science as,
the pursuit of knowledge and understanding of the natural and social dimensions of our
world of observations, formulating descriptive systems by controlled experiments to
determine the degree to which these systems represent the phenomenal world etc. (Amadi,
1991:185).

Understood as such, science is concerned with both man and his ways as much as
everything that is foreign and external to man. It is a branch of pure learning which is
concerned with the properties of the external world of nature; its business is to find out
accurately what those properties are, to interpret them, and to make them intelligible to man.
The intellectual satisfaction at which its aims would be secured completely if this external
world could be reduced to order and be shown to be directed by principles which are in
harmony with our intellectual and moral desires. As an intellectual endeavour, science arose
ultimately from man’s desire to understand the world. Perhaps it is this understanding of
science that the complex adjectival form of the world “science”, namely “scientific” i.e.
knowledge – making has today come to be accepted as the real province of science which in
the early beginnings was the original enterprise of natural philosophy. No wonder, therefore,
that science has come to be accepted as “the making of knowledge i.e. research instead of
knowledge as such. Thus, science described as such is the systematic process of making
knowledge; of building knowledge.
It is this quality of science that Ogbinaka (1998:178) writes that the intellectual
frontiers of science have ever been expanding, with very little of its contents being dropped.
Quoting from the Encyclopaedia Britannica vol. 6, he argues further that, this conception of
science has provided very strong basis for the following concepts of science;
(a) Science can be taken to be a mood in which the world is considered. Being a mood, we
should accommodate its changing states. Just as no man is always in the same mood, and no
man of science remains permanently in the same scientific mood.
(b) Science is always developing. It is not a static body of knowledge.
(c) Science is more of the making of “knowledge” (i.e. in contradistinction with a claim that it is
“knowledge itself”, so it is close to be called a research; a method employed in pursuit of a
goal which involves “the acquisition of systematic generalised knowledge concerning the
natural world; knowledge which helps man to understand nature, to predict natural events and
to control natural forces.”
This again involves the use of previously accumulated knowledge to construct general
theories or systems from which testable hypotheses can be derived, and the testing of such
hypotheses is carried out quantified observations under controlled conditions.
One may ask whether such a conception of science, as argued above is adequate for
the analysis of the impact of science on human development, but in particular African
development. We argue that such a conception of science reduced to a “method” employed
in pursuit of a goal is inadequate on the following grounds:
(a) As the acquisition of “systematic” generalised knowledge concerning the natural world,
science is made to be a scarce commodity reserved only for the west to the exclusion of the
developing world. But this is clearly fallacious, for science is a widely distributed
commodity, found in every culture and tradition.
(b) Science as a whole is a process, which transcends particular scientists, research teams, and
institutes. Hence, to argue that scientific goals encompass outcomes toward which movement
occurs is to miss the point. Put in proper perspective, a “goal” as usually understood is an
outcome toward which people strive, or more generally toward which the internal functioning
of a system is directed. Suffice it to say, then, that the meaning of a statement attributing a
“goal” to such a process would require clarification. No doubt, science produces certain
outcomes, and some of these outcomes are goals of individual scientists and research teams;
but it does not necessarily follow that science must be defined in terms of movement towards
that goal. As rightly confirmed by Richter:
It is entirely possible that the most significant aspects of science involve movement, over a
long time span, in directions which have not been intended or recognised by scientists
generally, and which have emerged accidentally, even if there has also been movement in
directions, which may be identified as corresponding to a “goal” of science (Richter,
1972:4).
(c) Even if science is defined as a process of moving toward a goal, it does not follow that
science thereby becomes equivalent to a “method”. Rightly defined, a method is a process
employed deliberately in pursuit of a goal. It refers to the specification of steps, which must
be taken in a given order, to achieve a given end. As a function in scientific inquiry,
“methods are used within scientific inquiries. “However, the concept of method cannot
reasonably be applied to some important types of events through which the findings of
different inquiries are interpreted and integrated by the scientific community as a whole.”
This is because the nature of the steps and the details of their specification depend on the end
sought and the variety of ways of achieving it.
We may thus argue here that the concepts of goal and method used as a quality of any
scientific endeavour can only be recognised as applicable at relatively microscopic levels. As
Maurice Richter (ibid.) concludes on this matter, that, when we seek instead to analyse
science macroscopically, taking into account not merely what happens within particular
research projects but also the integration of findings of many such projects in different
disciplines over centuries, the concepts of goal and method appear to lose their relevance.
The method of science therefore vary according to whether its end is taken to be the conquest
of nature or the discovery of truth and in the light different theories about the relation
between those ends and man’s primitive condition of impotence and ignorance.
The conceptions of science as a social institution, as an occupation and, lastly, as a
profession are also inadequate insofar as they imply a comparatively stable relationship
between science and society, with science performing certain functions or services on a
relatively consistent basis. The way ahead here is the conception of science as a cultural
process which alternative avoids the difficulties and shortcomings of the above conceptions.
Science as a cultural process is associated with a distinct view of nature as operating
according to general laws which remain largely hidden under ordinary observational
circumstances but which can be uncovered through systematically controlled observation and
experimentation as for example Isaac Newton’s one set of proposed laws; the law of
gravitation; the principle of calculus and the compound nature of light; and his three laws of
motion (Richter 1972:16). Such scientific laws, as they are many, provided an
overwhelmingly impressive demonstration of potentialities of this approach, thus reasonably
and clearly differentiated from such related phenomena as philosophy, religion, technology
and magic, among others.

Understood as such, science involves observational procedures, patterns of arguments,


methods of representation and calculations, and the evaluation of the grounds of their validity
from the point of view of formal logic, practical methodology and metaphysics. Certain basic
characteristics distinguish scientific knowledge from other sorts of knowledge. They include,
among others, observation of facts, collection of data, experimentation, and research; a sort of
a self-contained logical relationship empirically demonstrated as valid under given
conditions.
We may index our clarification by arguing further that science has its aim (locus) and
end (telos) in man, and must be so defined not only in terms of its relevance to the society in
general but to man in particular. Hence the locus and telos of science can be more properly
understood when we analyse its “two-factor theory” i.e. the realist theory of Galileo, and the
instrumentalist theory of Bellarmine. According to the realist theory, science aims at telling
the literal truth about the world. It aims at knowledge of how the world really is, in its
intrinsic sense. The instrumentalist on the other hand sees science as aiming to provide
useful aid to thought about everyday world, which is in turn directed toward organizing and
improving our lives within this world. On the other hand, the intrinsic value of science
undertakes the expanded grasp of the interrelatedness of all reality and, thus, situates science
in the context of its inherent value to humanity. Hence science is for the ultimate good of
man the whole man. St. Thomas Aquinas very lucidly captures the true scientific endeavour
thus: “Any culture or society that does not submit the sciences to the critical leadership of
philosophy (science of the ultimate good conduct) heads to confusion and low rationality”
(Nwoko, 1992:12).
While not arguing against the instrumental value of science, we hasten to say that this
form of scientific endeavour is a one sided intellectual knowledge which in itself distorts
one’s view of life and exaggerates scientific form and method to the detriment of its solid
content. In itself, such a scientific world-view is based on individualism, dry rationality and
therefore inadequate for human development. Science in its solid form and content involves
a moral issue. It is tailored towards enlightenment and mastery of nature for the service of
mankind. But its unguided endeavour (instrumental value) can lead us to ruin; hence Sophie
Oluwole emphatically warns that “if we allow ourselves to be carried away by the
innumerable physical successes of science, we may invite our own doom too” (Oluwole,
1991:21).
The Philosophical Character in Science
Philosophy is asking questions and questioning answers, it is a search for self knowledge in
an attempt to transcend man’s environment
We shall begin our investigation of the nature of the philosophical enterprise by
attempting a clarification of three characteristics of philosophy as follows:
 The epistemological concerns of philosophy.
 The metaphysical concerns of philosophy.
 Rational dialogue; of questions and answers, and the re-evaluation.
Epistemological Character
Philosophy is a critical analysis of all human knowledge with the view to establishing
its scope and limits. As a critical analysis, philosophy examines, estimates, and judges the
claims made on behalf of different points of view regarding principles, concepts and
decisions etc. This means that philosophy is very much concerned with the evaluation of
human knowledge and the quest for truth with the view to establishing its scope and limits,
even though philosophers are divided on the extent of human knowledge.

Metaphysical Character
Philosophy is a body of knowledge methodically acquired and ordered, which
undertakes to give the fundamental explanation of all things”. Understood as a “body of
natural knowledge, philosophy is here distinguished from theology and ranks it as dealing
with rational knowledge. Furthermore, philosophy is seen as a “methodically acquired and
ordered knowledge”. This view of philosophy put it as proceeding from observation and
experience, to a reasoned explanation of both of them. Lastly, philosophy understood as an
exercise that “undertakes to give the fundamental explanation of all things”, distinguishes
philosophy as a natural form of inquiry and puts it above other everyday activities. It is
fundamental and foundation to scientific endeavours. It is a search for ultimate reality; the
source of all things in their ultimate causes as known through the natural light of reason.
Rational Dialogue
This is a rational process of asking questions and questioning answers until we come
to answers that are unquestionable and questions that are unanswerable. It is a conscious and
rigorous pursuit of truth without which there are no answers, but which answers, when found,
are further subjected to critical scrutiny so as to obtain clarity, change or reject beliefs or
positions formerly held to tenaciously. Such an approach, Bertrand Russell (1959: ii) says, is
a scientific spirit of a prudent man who according to him will not claim that his present
beliefs are wholly true, though he may console himself with the thought that his earlier
beliefs were perhaps not wholly false.
Basic to the above conceptions of philosophy is the common-thread notion of
philosophy as a process of generating ideas, which are further processed and put into practice
through the art of doing things – techniques. Thus understood, philosophy is meta-science –
conceiving rational ideas which, when translated into science, is conceived as a means of
getting to know the world. For Russell therefore, the question whether objective truth belongs
to human thinking is not a question of theory, but a practical question. The truth i.e. the
reality and power of thought, must be demonstrated in practice. The contest as to the reality
or non-reality of a thought, which is isolated from practice, is a purely scholastic question.
Karl Marx voiced a similar view when he says philosophers have only interpreted the world
in various ways, but the real task is to alter it (Egner and Dennon 1959:636).
Philosophy is both fundamental and foundational to science and technology. Being
an endeavour whose knowledge encompasses the whole, and which seeks to explain the
interconnecting link between things and events of the universe, philosophy (which is
complete knowledge) thus guards, and guides both science and technology rationally toward
a people-centred development. This, is the real enterprise of philosophy in science, and
technology.
We may, therefore outline the main tasks or goals of philosophy in relation to science
and technology, starting with a discussion of the very basic descriptions of philosophy as:
 A Rational Basis to Life
Socrates stated long ago that “the un-examined life is not worth living”. Similarly,
Thoreau maintained that “to be a philosopher is not merely to have subtle thoughts… but to
love wisdom so as to live according to its dictates, a life of simplicity, independence,
magnanimity and trust”. Thus philosophy is not a sterile discussion of abstract notions lying
outside experience but a resolute and rational attempt at understanding life in all its
immensity, variety but above all, its totality. For life has meaning; to find its meaning is what
man’s activities is all about. It is a rational basis to life.
 A Method of Reflective Thinking and Reasoned Inquiry
The second description of philosophy which has implication for science and
technology is its conception as a method of reflective thinking and reasoned enquiry.
Described as a reflective endeavour, philosophy proceeds by way of argument and criticism,
and not by experimental verification. It supposes experience and experiment, but goes
beyond the empirical (experience) while it reflects on it.
Secondly, philosophy is general in its method. That is, it inquires into the general
nature of things or the meaning of general concepts e.g. knowledge, value etc.
Thirdly, philosophy is definitional. It concerns itself with typical questions with a
view to discovering the essence or definition or at least the description of concepts and things
e.g. What is progress? What constitutes development? What constitute a good science, or
technology? are usually a request for definition.
Fourthly, philosophical method is reflective. By this we mean philosophy is
concerned with the meaning and relations between various concepts. It presents a way of
“seeing” the world; a distinctive approach or insight on things.
 An Attempt to Gain a View of the Whole
Philosophy is an attempt to gain a view of the whole. This is an invocation of the
traditional view of philosophy; as a search – an activity through which man reaches the
unknown, into that which is hidden from him, but which he already has at least some initial
notion of what he is looking for. The drive behind the search is exactly the desire to verify
and elucidate our knowledge of reality. Thus, philosophy does not introduce to man a new
world of knowledge, but to a new knowledge of a world he or she knew.
The point at issue here is that the philosophical search goes far beyond the values and
events of everyday life. It opens a new horizon, which though strange, is all the same
fascinating, revealing and beneficial. It is a search for reality.
 Analysis of Language and the Clarification of Concepts

Philosophy is also described as the logical analysis of language and the clarification
of the meanings of concepts. G.E. Moore (1903) writes in his book Principia Ethica that: “in
philosophical studies, the difficulties and disagreement of which its history is full, are mainly
due to a very simple cause; namely to attempt to answer questions without first discovering
precisely what questions it is which you desire to answer” (Moore, 1906:30). Perhaps a
careful study of how language is actually used, taught and developed in everyday discourse
can illuminate, and even transform or dissolve, time-honoured philosophical problems.
These problems are seen as arising, often if not invariably, because thinkers, misled by
superficial grammatical similarities or their own fondness for uniformity have ignored
relevant differences in the function of terms and hence misused them. The result being that
they have drawn wrong conclusions based on a misunderstanding of the function of language.
We may thus argue, like Witgenstein, that whenever there exists a perennial and irresolvable
dispute concerning so-called philosophical problems, it is language that has gone on holiday.
 A Group of Problems as Well as Theories about the Solutions
As an endeavour that concentrates on the wide range of problems, philosophy
attempts through analysis and criticism to raise theories that find solutions to these problems.
Thus character of philosophy oiled the great machine of the early centuries to engineer to the
fore, our present day scientific endeavours. It is fashionable today, to ask “what (if anything)
to expect from today’s philosopher”. Such was the title of a remarkable article that appeared
in the Time Magazine on January 7th, 1966. A similar title appears in one of the chapters of
Kwasi Wiredu’s recent book, Philosophy and an African Culture (1980). The chapter in
question is “What can philosophy do for Africa?” Such titles, provocative though, seem to
suggest that there is some doubt as to what philosophers (or philosophy) have (has) to offer to
the world today. There may even be the suspicion that it has nothing to offer. Perhaps such
question are not asked or may not be asked of engineers, doctors or bankers because it
appears to be quite obvious that they have something to offer and what they have to offer is
obvious to the society.
Philosophical Science
Now, the philosophical enterprise, in science and technology yesterday and today has
ever remained the same. That is, the meta-scientific enterprise, which made philosophy
fundamental and foundational to all human endeavours. Perhaps a retrospection of three
centuries ago argues for the reinvention of the philosophical spirit of the antiquity. Although,
at that point in time, there was confusion in the air, the importance of philosophy was not
seriously in question. We discover in particular that philosophy was held in high esteem,
even though some of its practitioners may have been severely criticised. In antiquity,
philosophy was understood as the science of life which enables people “to think well, to
reason practically – neither too much nor too little because it is the means of living well as
circumstances permit us to live and to see clearly into the causes of things, to analyse our
own acts and motives, and to try to understand those of others.
Admittedly, such rational behaviour of man provided a base for what came to be later
delineated as science which found practical reality in technical application in the quest to
fulfil man’s essential needs. This, according to Abraham Kaplan, is the essential business of
philosophy, which he says, “is to articulate principles by which man can live; not just as a
scientist, citizen, religious Being or whatever, but as the whole man that he is” (Kaplan,
1961:4). It means for us, then, that philosophy is a technical discipline, that intertwines all
areas of human knowledge, and which in pursuit, helps men to think more clearly and more
truly about themselves and the world in which they inhabit. Drawing inspiration from John
Wild in this direction of thought, K. C. Anyanwu argues that “any man who thinks, speaks,
and lives necessarily needs philosophy, and true philosophy, is a therapy for the common
intellect of common men, and true therapy does not try to destroy what it is trying to heal or
perfect” (Anyanwu, 1983:40).
Informed by this thinking, Plotinus (d. 270 A.D) says philosophy was a dear delight.
On the other hand, Plato would place philosophers over the affairs of men. He died because
of this conviction. Protagoras of Abdera (d. 411 B.C.) made the human mind the measure of
all things and philosophising itself was the greatest activity of that mind. Marcus Aurelius (d.
180 A.D.) loved philosophy more than his throne. Boethius (d. 524 A.D) consoled himself
and even wrote a book while in prison entitled The Consolation of Philosophy (Green, 1962).
Little wonder then that the Time Magazine article chronicles what at best remains the
outstanding legacy of philosophy:
The world has both favoured and feared the philosophers’ answers. Thomas Aquinas was a
saint, Aristotle was tutor to Alexander the Great, and Voltaire was a confidant of Kings. But
Socrates was put to death, and Giordano Bruno was burned at the stake. (Green, 1962:8)
Thus, the power of philosophy, the burning desire to ask questions and question
answers, found proper relevance in the 6th century B.C. when the men from the coast of Asia
Minor began asking questions that had never been asked. They began to ask what the world
was made of, and how it originated. Such questions, which are scientific in content, elicited
scientific answers, which activities today can heal, or kill. Suffice it to say however, that the
question about the origins and nature of things were questions that might be resolved by
rational discussion. Precisely, such questions, philosophical in nature though but scientific in
content assisted not only to enrich man’s consciousness, but ultimately led to man’s control
over the forces of nature. Essentially, such questions led to the emergence of science as
presently understood. Perhaps we may reserve such discussion for the later part of this thesis
but argue here briefly that the search for the ultimate reality jointly carried out by the natural
philosophers (c 600. c 300 B.C.) started not only for information but also for understanding.
Thus, the method of examination and critical analysis took the centre stage. This means that,
appeals to tradition and authority were replaced with appeals before the court of human
reason.
Such was philosophy in relation to science in Greek antiquity that philosophy (then
called philosophy of nature) and science (then called natural science) were one and the same
discipline- Philosophy of Nature. It is on record that Aristotle considered as a single science
what is now called philosophy of nature, cosmology, chemistry and biology. V. E. Smith in
the New Catholic Encyclopaedia (NCE) (Vol. II:317), acknowledges succinctly that, “such a
unified view of philosophy and science survives in the title of Isaac Newton’s work, The
Mathematical Principles of Natural Philosophy (1687), and more than a century later in John
Dalton’s A New System of Chemical Philosophy (3 v. 1808-10-27). This is in addition to
another title under the caption, Experimental Philosophy, which philosophies, are today
considered as science, a term that, with the foundation of the British Academy of Science in
1831, came into vogue to designate modern physics, chemistry, biology, and related
disciplines (ibid.). Notwithstanding the subject matter of natural science in present day
organised society, which province is the study of the material world, philosophy still
functions as a lubricant to all human learning.
Though still taken as an endeavour which object is to explain material realities in
terms of the four causes: matter, form, agent and end, the physical sciences are still largely
carried out under the rules of reasoning long formulated by philosophy, of asking questions
and questioning answers, of examination and critical analysis of the basic constitution of the
particular beings that enter the world of human experience. In particular, what is the nature
of the physical universe? How are scientific laws established and validated? etc are some of
the interrogatives which provoke scientific research. Understood in this light, philosophy,
thus, becomes the science of all things in their ultimate causes as known through the natural
light of reason.
Such is what is called science in the Aristotelian – Thomistic tradition, which
designate a type of perfect knowing (scire simpliciter). Knowledge of any object, argues
Aristotle, is obtained when one knows its cause, when one knows that cause is what makes
the object be what it is, and when one knows that the object could not be otherwise than it is
(NCE Vol 12:1190). Following from Aristotle’s reduction, St. Thomas Aquinas similarly
taught that science is knowledge of something through its proper cause. It is a purely
intellectual act as opposed to sense knowledge; mediate intellectual knowledge as opposed to
immediate knowledge of concepts and first principles in so far as it is acquired through the
prior knowledge of principles or causes (ibid.).
It may be said, thus, that science in the spirit of Newton and his successors was left as
the only legitimate body of speculative knowledge concerning existing things. Science stood
alone as a study of things, whereas philosophy, with respect to science, was purely critical
and epistemological, thus acting as a catalyst to the scientific endeavours of our twentieth
century world. Such new philosophical currents which were put in motion to raise science
and technology to its present height include among others, idealism, positivism, realism, and
pragmatism. We shall single out positivism, pragmatism and realism for discussion in this
chapter.
 Positivism
This philosophical current was begun by Auguste Comte. Unlike Aristotle whose
concern was with causes or origins of things, Comte on the other hand was concerned with
their invariable relations of succession and resemblance. Thus, scientific methodology is
according to this school, apparently descriptive as opposed to being explanatory. No wonder
then that Spencer, a later positivist, assigned to philosophy the role of synthesizing scientific
results. However, most positivists conceived the main burden of speculative philosophy as
one of accounting for the apparent necessity and universality in the laws discovered by the
sciences (Smith 1967:317).
Before then Kant had argued that valid knowledge come through phenomena, The
phenomenal world, he argues, could not give rise to the universality and necessity found in
physical laws, and that such universality and necessity had therefore to come from a priori
structures in the human mind (ibid.). Perhaps this attack motivated Comte who posited three
stages in the development of the human mind:
(a) a theological stage, wherein the world is explained by an appeal to supernatural deities.
(b) a metaphysical stage wherein things are explained by abstract essences
(c) and a positivistic stage, wherein reality is accounted for by sciences like that of Sir Isaac
Newton.
This historical syllogism argues out properly a position of a self-contained logical
relationship empirically verified as valid under given conditions, in contrast to the dogmas of
theology and unobservable facts of metaphysics. As argued by Galileo in his Dialogues
Concerning Two New Sciences,
Anyone may invent any arbitrary type of motion and discuss its properties, but we have
decided to consider the phenomenon of bodies falling with acceleration such as actually
occurs in nature… And this at least, after repeated efforts we trust we have succeeded in
doing. In this belief we are confirmed mainly by the consideration that experimental results
are seen to agree with and exactly correspond with those properties which have been, one
after another, demonstrated by us (Smith, 1967:160)
Such is the positivists’ contribution to science that Ernst Mach regarded scientific
laws as economies of thought that make it psychologically easier for man to study nature.
Argued on this score, Philosophy is fundamental and foundational to science and technology.
It is the method of scientific engagement.
 Pragmatism
This philosophical outlook owes its origin to Charles Sanders Peirce, who held that
ideas could be made clear only by looking to their effects. Somewhat almost like Aristotle’s
postulations, but in complete contrast to positivism, Peirce regarded man’s first questions
about nature as being “the most general and abstract ones”. For him therefore philosophy
comes before science, meaning then that rationalism breeds science, and complements
science.
Perhaps it is this complimentary role of philosophy that made William James and
John Dewey, compatriots of Peirce in the pragmatists’ camp to step down distinctions of any
importance between philosophy and science. They insisted, however, that experience extends
beyond the phenomena of Kant or the sense data of British empiricism. There is personal
experience, religious experience, experience of values etc. Such an enlargement of the
Kantian and positivist notion of experience, while important in itself, prepared the way for
philosophies of science like Whitehead’s. If we have only empirical facts, only individual
things, only one thing after another, then we can find no general laws but only summaries of
events, lists of observed regularities. As Whitehead writes “we must not ascribe, we must not
expect, one step beyond our direct knowledge. The (empiricist) has no foothold on which to
rely for specification beyond the region of direct observation. There is no probability beyond
the region of direct observation” (cf. Lewis 1962:126). John Stuart Mill, himself an
empiricist, failed in his attempt to justify the validity of inductive logic, which arrives at
general laws from particular instances. Such scientific mind may have influenced other
scholars like David Hume and Immanuel Kant who in their wisdom argue forcefully that a
regular succession is not a sufficient reason for believing in the inevitability of that
succession continuing, it only accounts for our expectation.
 Realism
Scholars under this philosophical current are common in their opposition to
positivism and to idealism. Prominent among these scholars are Emile Meyerson, Henri
Bergson, and Alfred North Whitehead. Meyerson is of the view that there was ontology in all
science, as shown by the scientist’s commitment to the existence of abiding identities in a
changing world. Another philosopher of science, Henri Bergson, maintained that science as
such presents a geometricised, hence, static view of a world in motion, and that motion can
be grasped only by an intuition that lies beyond the techniques of science. Whitehead on his
part proposed that the scientist in advance of his science commits himself to “half truths”, to
which the philosopher must examine (ibid. p. 319).
This commentary on science by these philosophers of science argues for us an
interesting point in the relationship between science and philosophy, that what today is
referred to as the physical science or the modern sciences known by the general name
empiriology from which we have empirioschematic and empiriometric was one and the same.
In particular, it is argued in this same spirit that empirioschematic knowledge i.e. science
which uses qualitative models, is not a distinct science, but a continuation of the philosophy
of nature – (cosmology and the science of nature).
Generally taken to be an intuitive endeavour, philosophy assists science to investigate
into the constitution of the physical universe. Perhaps we may say here that, both philosophy
and the physical sciences are two complementary old friends, which acting in concert could
make nature a completion. Such a thought is more cryptically captured in Smith’s The
General Science of Nature (1958) that, since the 19th century, efforts to construct a priori
philosophies of nature such as idealism, or to deny a philosophy of nature, as with positivism,
important 20th century western philosophers seem to have rediscovered the need for a
realistic evaluation of nature, one that considers mobile being at a level more general than the
specialised natural sciences and at a level more natural than mathematical physics. Such is
the philosophy we seek to advance in this thesis; a philosophy of human sustainable
development. That, though dualistic in matter and form, there is directionality in the cosmos,
which Whitehead describes as the causality of the end; by which all our physical causes are in
more or less a conscious way.
Furthermore, a better understanding of the essential connection between philosophy
and science can be located in Bertrand Russell’s conception of philosophy as a field of study
which seeks to explain the relationship that man shares with the universe. He posits:
Philosophy as I shall understand the word, is something intermediate between theology and
science. Like theology, it consists of speculation on matters as to which definite knowledge
has so far been unascertainable; but like science it appeals to human reason rather than to
authority… All definite knowledge, so I should contend - belong to science; all dogma as to
what surpass definite knowledge belongs to theology. But between theology and science,
there is a No Man’s land exposed to attack from both sides; this no man’s land is philosophy
(Russell 1962:35).
The relationship between philosophy and science is in its joint appeal to human
reason rather than to authority. While “a better philosopher” is not made through knowing
mere scientific facts, he all the same learns from its principles and methods and general
conceptions. Philosophy, Russell argues then, should be piecemeal and provisional like
science. He stated the essential relationship of the two disciplines thus:
Philosophical knowledge… does not differ essentially from scientific knowledge. There is no
special source of wisdom which is open to philosophy but not to science and the result
obtained by philosophy are not radically different from those obtained from science. The
essential characteristic of philosophy, which makes it a study distinct from science is
criticism (Russell, 1927:2-3).
The implication here is that both philosophy and science are in agreement on the
question of method, which, according to Russell, is the logical analytic method according to
which objective knowledge is possible. This kind of knowledge gives unity and system to the
body of the sciences, the kind of knowledge which results from a critical examination of
grounds of our convictions, prejudices and beliefs. Philosophy fulfils an intellectual role for
the sciences, and as soon as definite knowledge concerning any subject becomes possible,
this subject ceases to be called philosophy and becomes a separate science (Ogbinaka,
2000:38).
It, thus, appears that science and philosophy are two distinctive, unrelated disciplines.
Russell himself had seemingly implied this when he distinguished science as “what you
know” from philosophy as “what you don’t know”. But such conception of the two
disciplines, dangerous though, has turned out to be valuable on epistemic grounds. This is
because, the dialectical spirit of philosophy, its critical comparative analysis and synthesis
argues for further and better interior reconstruction of fundamental scientific facts, the end of
which is the institution of exact scientific knowledge. Such is what should be the
epistemological attitude of the philosopher who should be critical of scientific knowledge not
from a point of view which is ultimately different from that of science, but from a point of
view concerned with the harmony of the whole body of special sciences.
True, science has been able to solve certain problems that philosophy could not solve
but to argue from this and assert that the philosopher has no business in the conclusions of the
scientist is fundamentally fallacious. For even after science has solved certain problems,
there remains the need to understand, interpret and evaluate the body of facts accumulated by
science. The human quest for a rational and fulfilling destiny, thus, imposes on philosophy
the double role of clarifying and analysing scientific concepts and theories which aim at
making their scientific usage clear, and secondly, functioning as a second order discipline
that attempts to answer the following questions: What characteristics distinguish scientific
inquiry from other types of investigation? What procedures should scientists follow in
investigating nature? What conditions must be fulfilled for a scientific explanation to be
correct? What is the cognitive status of scientific laws and principles?
Such interrogatives while not taking away the function of the scientist as the one who
judges one theory to be superior to another, all the same stamps the philosopher’s feet as one
who evaluates the criteria of acceptability implied in the judgement of the scientist. Most
obviously, the two disciplines are indispensable to each other. Hence, the convinced words
of John Losee:
The scientist who is ignorant of precedents in the evaluation of theories is not likely to do an
adequate job of evaluation himself. And the philosopher of science who is ignorant of
scientific practice is not likely to make perceptive pronouncements on scientific method.
(Losee, 1972:2)
This informing drive of the men from the coast of Asia Minor in the 6th century
marked the advent of modern science, to which we shall now turn – the evolutions in science.
We shall endeavour to discuss this subject as a philosophic-scientific endeavour to unravel
the secrets of being, a vision of a single, undivided universe, of unrealised potentialities of the
human mind and heart, of an ideal order lurking behind the manifold appearances of things.
It thus stands to reason that, the philosophic spirit of interrogation enriches the human
mind to question traditions and dogmas in search of verification. Such is when it is said that,
the more mature a science the richer and better tested theories it includes – the more accurate
projections it can make and the more it can learn from both successes and failures in
projecting. Such is the spirit that defines the endeavour of the rationalists to whom science is
more or less a method of knowledge acquisition or knowledge making.
For the rationalists therefore, philosophical science cover such areas as the
methodology of scientific research, the status of scientific theories and the factors that could
enhance the development of science. The four most notable ones include: Raymond Karl
Popper, Thomas Kuhn, Paul Feyerabend and Imre Lakatos. Popper hangs on the falsification,
Kuhn on what he called Normal Science. Lakatos stresses on essence of hard core’ theory in
science whereas Feyerabend did not hide his resentment to methodology in science. We will
state briefly the positions of these philosophers on the issues already raised.
Popper on Falsification
Popper conceives science as a dynamic discipline. It is an open system that is characterised
by inputs of new ideas and abandonment of obsolete ones. There is no limitation to the
acquisition of scientific knowledge. Thus, progress in science consists in the constant
evolution of new ideas, testability and refutation of old and porous theories. And it (science)
continues to be exciting as long as it remain an open system that welcomes criticism. In
essence “science is essentially critical that consists of bold conjectures, controlled by
criticism and … may … be described as revolutionary”. This does not mean that Popper was
entirely opposed to any kind of dogmatism. A dogmatic scientist has significant part to play.
“He did not advocate that we should give in easily, for if we do so we shall fail to notice
where the real power of our theories lies”.
A scientist, Popper stresses is faced the problem of getting close to the truth. The line with
this tradition, he is confronted too with the task of how to falsify theories or replace falsified
one or replace adhoc hypothesis, he also attempt to see the possibility of a merger of two
already existing theories.
The totality of the endeavours here is meant to source a very effective way of surmounting a
giving obstacle in quest of truth. If a scientist eventually achieves success in producing a
theory that serve as a solution to the problem at stake then, his achievements would have
been immense. Nonetheless, this is not the end to the challenge of a scientific enquiry; there
are still much more requirement that are meant to be tackled.
The point is that we have at hand a problem to be tacked and there is a new theory meant for
settling the problem. There are a number of requirements this new theory must have to meet.
The new theories proceed from some simple, new and powerful, unifying idea about some
connection or relation (such as gravitation attraction) between hitherto unconnected things
(such as planets apples) or facts (such as inertia and gravitational mass).
The second requirement of out new theory is that it should be capable of facing independent
testability. This is to say that, in addition to explaining all the explicanda, the new theory
must have new and testable consequence (especially novel consequences); it must also lead to
the prediction of phenomena that so far have not been observed.
The imperative of our requirements as stated above is that in its absence, our new theory
might be adhoc, as it is always possible to produce a theory to fit any given set of explicanda.
The first two requirements are meant to check the range of our choice among the possible
solutions - many of which not relevant to the issue at hand.
We would have made some progress whatever are outcome of our new tests if the second
requirement it met. This is because its testability is enhanced than the previous theory, as it
capable of explaining all the explicanda of the previous theory, in addition to giving rise to
new tests.
Another significant impact of our second requirement is that we are assured that our new
theory, will to some extent fruitful as an instrument of experiments lead to the refutation of
the theory, we will have made an addition to our factual knowledge resulting of the theory,
we will have made an addition to our factual knowledge resulting at least from the
unexpected outcome of the new experiments. Besides new problems will emerge from them,
which are to be solved by new explanatory theories.
However, satisfactory and thorough the above two requirements might be the need to add the
third one is very pressing. This is the requirement that the theory should pass some new and
severe tests. The difference between the third requirement and the previous two ones is that in
the third requirement its fulfillment can only be established by testing the new theory
empirically. This is a material requirement, a requirement of empirical success. And this third
requirement more crucial than the previous two as in indispensable in the decision as to
whether the theory in question is to be accepted at all as a member within the scientific
structure and also if it is an interesting and promising theory.
It is noteworthy, however, that it does happen at time that some of the most interesting
theories are refuted at the first step of its testing.
This shows that not even the greatest physicist can anticipate the secrets of nature his
inspiration can only be guesses, and it is no fault of his, or of his theory, if is refuted.
Newton’s theory is spite of its influence and general acceptability was after all refuted. It is
only by such happening in the science arena that we keep on the tempo of sourcing for new
theories in addition to improving on the existing ones. A theory can be new theories in
addition to improving on the existing ones. A theory can be refuted at any point in time – 10
months or 10 years after its promulgation: it is all simply a matter of historical accident. If a
theory is met be sudden death, Popper believers that the scientist that propounded the theory
should be remembered for various reasons, foremost, the scientist should be remembered for
making available to us in theory that unearth to us new and perhaps still, unexplained
experimental facts coupled with new problem, and for the services it has by doing rendered to
the progress of science during its successful but short life.
What this suggests is that even it a theory is incapable of meeting up to the stipulation of our
third requirements, it can still make contribution t o scientific growth. Nonetheless the third
requirement cannot be dispensed with. Scientists, as matter of fact, aim more than just
contributing importantly to the growth of science. They aim to solve problems with their
theories. Which is to say that “if the progress of science is to continue and its rationality not
to decline, we need not only successful refutations, but also positive successes”.
We should also aim always to produce theories that entail new prediction moreover
predictions of new effects, new testable consequences as embedded in the new theory that
could be seen in many significant well to novel. Instances of such predictions obey Kepler’s
law; of that light, despite its zero mass, would prove to be subject to gravitational attraction
that is Einstein’s eclipse effect. The progress of science centers essentially on the predictions
of this nature and a possible corroboration for them, this is necessary in view of the fact that
scientific theories are meant to achieve conquest of the unknown, a new success in prediction
what had never been thought of before. When a predicted theory is corroborated a reasonable
advancement is made in science.
An unbroken sequence of refuted theories would soon leave us bewildered and helpless: we
should have no clue about the parts of each of these theories or of our background
knowledge – to which have might tentatively attribute the failure of that theory.
As earlier noted, science will be dormant and lose its empirical character if we are unable to
obtain refutations. In the same vein science would stagnate and lose its empirical character if
we are unable to obtain corroboration of new predictions that is if we are only able to produce
theories that meet out first two requirements, and fails the third. If for instance, we are to
produce an unbroken sequence of explanatory theories, each of which would explain all the
explicanda in its field, the experiments that refuted its predecessors inclusive, each would
also be independently testable by predicted effects, yet each would be at one refuted when
these predictions were put to the test. Thus each would satisfy first two requirements.
Kuhn on Normal Science
What a scientist does, according to Kuhn in his daily enterprise amount to what he calls
“normal science’. Any scientist that does not pre-occupy himself with spurious scientific
research must meet up to the requirement of ‘normal science’. To Kuhn “normal science
means the research that is firmly rooted upon one or more past scientific achievements. Such
achievement must be acknowledged by a given scientific community for a given point in time
and it is such that provide the basis for further scientific enterprises. This kind of
achievements, argues, are today recounted though not often in their original form, by science
textbooks elementary and advanced. Such books expound the body of accepted theories
illustrate many of all of its successful applications and compare these application with
exemplary observation and experiment.
Some accepted instance of actual scientific practice according to Kuhn includes law, theory,
application and instrumentation given models from which emanate particular coherent
tradition of scientific research. A scientist possesses two essential attributes. Their
achievement was sufficiently unprecedented to attract an enduring group of adherents away
from competing modes of scientific activity. Simultaneously, it was sufficiently to resolve.
Achievements that share these characteristics Kuhn calls “paradigms” a term which he say
relates these closely to normal science’. These constitute the tradition that the historian
describes under such rubrics as ‘Polemics astronomy’ or ‘Copernican’ or Aristotelian
dynamics’ or Newtonian corpuscular optics’ or wave optics.
The study of paradigms including many that are far more specialized than those already
stated above is what fundamentally prepares the students for membership in a given scientific
community with which he will later practice. Men whose research is based on shared
paradigms are committed to the same rules and standards for scientific practice. That and the
apparent consensus it produces are prerequisite for normal science i.e. for the genesis and
continuation of a particular tradition”. The ability to be acquainted with a paradigm in
addition to the more esoteric type of research it permits signify maturity and advancement in
the development of any scientific field, e.g. the wave theory that is embraced by almost all
the practitioners of optical science. In the 18th century however, the paradigm for this field
was provided by Newton’s optics which taught that light was materials corpuscles “The
transformations of the paradigms of physical optics are scientific revolutions, and the
successive transition from one paradigm, to another via revolution is the usual development
pattern of mature science”.
In biology, for instance, part of its study center on the study of heredity which is a most
universally received paradigms. The essence of having paradigms in the view of Kuhn is
essentially because history reveals that the road to a firm research consensus is extraordinary
arduous. Where there are no paradigms all the facts that relate to the development of a given
science may seem equally relevant.
As one theoretical system gradually receives general acceptance, a paradigm is established.
Kuhn argues thus, that during normal science, rather than attempt to falsify theories, scientists
engage in puzzle-solving activity. All paradigms, he says, have inherent anomalies which, if
and when they become so large that cannot be contained by the ad hoc hypothesis lead to loss
of faith by members of the scientific community. This according to Kuhn may result in the
articulation of several other alternative theoretical structures or paradigms. If, and when the
new paradigm achieves general acceptance, these occurs what Kuhn calls Paradigm Shift.
Lakatos on Hard Core Theory
Lakatos enunciated what he terms ‘hard-core’ in his methodology. This resembles to a
reasonable extent Kuhn’s paradigm shift, for scientists involved in research programme are
bound to agree with the fundamental tenets. The difference however is that in Lakatos thesis
there are several ‘hard-cores’ competing at any state in history. And
The hard-core is a set of statements, which are protected from refutation by the attempt to
modify other assumption when apparent refutation arises.
It consists of a family of theoretical assertions and for a theory to be part of scientific
research programme (S.R.P) it must share those assumptions. There is also the negative
heuristic to the programme. And this “is a methodological principle stipulating that the
components in the ‘hard-core’ not to be abandoned in the face of anomalies”. For instance,
the Newtonian gravitational theory, the three laws of dynamics and the law of universal
gravitation are seen by Lakatos as constituting the ‘hard-core’. The essence of the appeal to
the negative heuristic is to ensure that anomalies emanating from the application of the theory
are not taken as refuting these postulates. The heat produced by anomalies is to be cushioned
by the modification of auxiliary hypothesis, observational hypothesis, or hypothesis stating
initial conditions. As to what is to be done in the face of anomalies, the guidance is provided
by the positive heuristic of the programme. And this comprises of a partly articulated set of
suggestions or hints as to how to change develop the refutable variants of the research
programme, how to modify, sophisticate, and the ‘refutable’ protective belt.
When successive modification by the ‘positive heuristic’ is unable to produce any new
independent tests, which are passed, the research programme is ‘degenerating’. If there is
marketable success in the research programme in guiding us to new discoveries, then it is
progressive; in the long run progressive are chosen over degenerating research programme.
Feyerabend on Method in Science
Paul Feyerabend was anti-methodology. He was opposed to any kind of methodology in
science and he believes he did not recommend any science as a versatile and dynamic
enterprise does not need any methodology. A methodology for science is like a chain tied
round it which in essence is only a disaster to its advancement. And progress he argues, is
very crucial as far as science is concerned. He argues the conclusion that, the only principle
that does not inhibit scientific progress is, anything goes,
He believes that there is no single rule if we should go by historical research, however
plausible and however firmly grounded in epistemology that is not violated at some time or
the other. Such violations certainly, he argues cannot be attributed to accidental events, just
as they cannot be seen as results of insufficient knowledge or of in attention that might have
been avoided. Rather than see it in any of the stated ways, they are to be accepted as
imperatives for progress. He emphasizes that one of the most salient point that is visible in
the history of philosophy and science is the fact that,
That event and development such as the invention of a atomism in antiquity, the Copernican
theory (Kinetic Theory emergence of the wave theory of light occurred only because some
thinker either decided not to be bound by certain ‘obvious’ methodological rules or because
they unwittingly broke them.
The practice of science is in consonant with this pragmatic attitude and it is only when
practitioners of science conform to this requirement that progress in science could be assured.
No matter how fundamental or crucial rule in science may be, there are always circumstances
when it is most available to ignore the rule and adopt the opposite, he asserts.
Aristotle’s Philosophy of Science
Aristotle produced the first great philosophy of science, although his work in this field is
widely denigrated today. Among other problems, his discussions about science were only
qualitative, not quantitative, and he had little appreciation for mathematics. By the modern
definition of the term, Aristotelian philosophy was not science, as this worldview did not
attempt to probe how the world actually worked through experiment and empirical test.
Rather, based on what one's senses told one, Aristotelian philosophy then depended upon the
assumption that the human mind could elucidate all the laws of the universe, based on simple
observation (without experimentation) through reason alone. In contrast, today the term
science refers to the position that thinking alone often leads people astray, and therefore one
must compare one's ideas to the actual world through experimentation; only then can one see
if one's ideas are based in reality.
One of the reasons for Aristotle’s conclusions was that he held that physics was about
changing objects with a reality of their own, whereas mathematics was about unchanging
objects without a reality of their own. In this philosophy, he could not imagine that there was
a relationship between them.
Aristotle presented a doctrine of their being four “causes” of things, but the word cause
(Greek: αἰτἱα, aitia) is not used in the modern sense of “cause and effect,” under which
causes are events or states of affairs. Rather, the four causes—material, formal, efficient, and
final causes—are like different ways of explaining something; the material and formal causes
are internal to the thing and separable only in thought; the efficient and final causes are
external.
The material cause is the material that makes up an object, for example, "the bronze and
silver ... are causes of the statue and the bowl." The formal cause is the blueprint or the idea
commonly held of what an object should be. Aristotle says, "The form is the account (and the
genera of the account) of the essence (for instance, the cause of an octave is the ratio two to
one, and in general number), and the parts that are in the account." The efficient cause is the
person who makes an object, or “unmoved movers” (gods) who move nature. For example,
“a father is a cause of his child; and in general the producer is a cause of the product and the
initiator of the change is a cause.” This is closest to the modern definition of “cause.” The
final cause or telos is the purpose or end that something is supposed to serve. This includes
“all the intermediate steps that are for the end ... for example, slimming, purging, drugs, or
instruments are for health; all of these are for the end, though they differ in that some are
activities while others are instruments.” An example of an artifact that has all four causes
would be a table, which has material causes (wood and nails), a formal cause (the blueprint,
or a generally agreed idea of what tables are), an efficient cause (the carpenter), and a final
cause (using it to dine on).
Aristotle argues that natural objects such as an "individual man" have all four causes. The
material cause of an individual man would be the flesh and bone that make up an individual
man. The formal cause would be the blueprint of man, which is used as a guide to create an
individual man and to keep him in a certain state called man. The efficient cause of an
individual man would be the father of that man, or in the case of all men an “unmoved
mover” who breathed (anima: breath) into the soul (anima: soul) of man. The final cause of
man would be as Aristotle stated, “Now we take the human’s function to be a certain kind of
life, and take this life to be the soul’s activity and actions that express reason. Hence the
excellent man’s function is to do this finely and well. Each function is completed well when
its completion expresses the proper virtue. Therefore the human good turns out to be the
souls’ activity that expresses virtue.”
Aristotle also investigated movement and gravity. He did not know about the principle of
inertia, and held that movement must always be caused by something.
The ancient Greeks, and especially Aristotle, noted the difference between natural objects and
artificial ones or artifacts: Natural objects have self-movement, or the principle of their
motion within them, whereas artificial things have their existence and motion through human
agency. An elephant or an oak tree, for example, are natural objects that move from
principles within them, while an automobile or a bicycle are artifacts that move by principles
external to them and dependent on human agency.
Kant's Philosophy of Science
First published Tue Oct 21, 2003; substantive revision Tue Nov 13, 2007
Kant's philosophy of science has received attention from several different audiences and for a
variety of reasons. It is of interest to contemporary philosophers of science primarily because
of the way in which Kant attempts to articulate a philosophical framework that places
substantive conditions on our scientific knowledge of the world while still respecting the
autonomy and diverse claims of particular sciences. More specifically, Kant develops a
philosophy of science that departs from (i) broadly empiricist views — such as David
Lewis's, according to which purely contingent events in space and time (along with
considerations of simplicity, etc.) determine what the laws of nature ultimately are — and (ii)
certain necessitarian views — such as David Armstrong's, according to which the laws of
nature consist of necessitation relations between universals, which place constraints on what
events occur in space and time. Kant does so by holding that (i) scientific laws do involve
necessity, but that (ii) this necessity is based not on (purely metaphysical and hence
inaccessible) relations between universals, but rather on certain subjective, a priori conditions
under which we can experience objects in space and time.
Kant's scientific writings are also of interest to historians of modern philosophy, historians of
science, and historians of philosophy of science. Historians of modern philosophy are
especially interested in determining how Kant's views on science might complement or
clarify his distinctive metaphysical and epistemological doctrines (e.g., as expressed in the
Critique of Pure Reason). Historians of science reflect on the way in which Kant's position
fits in with the views of other natural philosophers of the period, such as Newton and Leibniz,
including his novel account of the formation of the solar system according to Newtonian
principles. Historians of philosophy of science investigate, among other things, Kant's work
in the conceptual foundations of physics — in particular, his matter theory (e.g., the infinite
divisibility of matter, attractive and repulsive forces, inertia, atoms and the void) and his
dynamical account of the laws of mechanics.
Because physics was Kant's primary (though not exclusive) focus over the course of his
lengthy career, his views on physics during his pre-Critical (1746-1770), Critical (1781-
1790), and Post-Critical periods (after 1790) will be discussed in separate sections.
Subsections will be devoted to each of the chapters of Kant's most influential work in
philosophy of science, the Metaphysical Foundations of Natural Science (1786). Kant's basic
positions on other sciences, including psychology, chemistry, and history, will be presented
thereafter.
1. Physics: The Pre-Critical Period
Kant's early pre-Critical publications (1746-1756) are devoted primarily to solving a variety
of broadly cosmological problems and to developing an increasingly comprehensive
metaphysics that would account for the matter theory that is required by the solutions to these
problems. Kant's first publication, Thoughts on the True Estimation of Living Forces (1746),
explicitly attempts to solve the vis viva controversy, which had been hotly contested ever
since Leibniz's attack on Descartes' laws of motion in the Acta Eruditorum in 1686. While
Kant attempts to occupy an intermediary position between the Cartesian and Leibnizian
positions by maintaining that both mv and mv² could be conserved in different contexts, what
is of particular note is how his solution in Parts II and III rests on the conception of force
developed in Part I. According to this conception, force is understood in terms of the activity
of substances, an activity that Kant then uses to explain how the motions of bodies are
generated, to solve the mind-body problem, and to account for both the possibility of other,
actually existing worlds and the three-dimensionality of space.
Kant develops his account of the nature of substance in greater detail in A New Elucidation of
the First Principles of Metaphysical Cognition (1755). While the first two sections of this
work undertake revisions of Wolff's principles of non-contradiction and sufficient reason, the
third section argues for two substantive principles that are alleged to follow from the
principle of sufficient (or rather, following Crusius, determining) reason, namely the
principles of succession and coexistence. The main thrust of the principle of succession is
directed against Leibnizian pre-established harmony, arguing that only causal connections
between substances can bring about changes in their states. Kant's position appears to be
designed to account primarily for changes of bodily states (with changes in mental states
being parasitic upon them, as was explicitly asserted in the True Estimation). For he
maintains that mutual changes of state require mutual interaction, where it is clear that
changes in motion are precisely the kind of mutual change that he has in mind (since one
body cannot move closer to another without the other body moving closer to it). The principle
of coexistence then argues that harmonious causal interaction between otherwise isolated,
independently existing substances is possible only by means of God's coordination (just as
Leibniz thought was required for harmonious relations between the states of such
substances).
Kant's Physical Monadology (1756) takes as its task the reconciliation of the infinite
divisibility of space, as maintained in geometry, with the simplicity of substances, which
Kant believes is required in metaphysics. As was the case with his earlier works, the essential
feature of his reconciliation lies in the way in which his matter theory is supported by his
metaphysical views. Specifically, Kant asserts that simple substances fill space not by means
of their mere existence, but rather in virtue of their spheres of activity. As a result, any
division of the relevant spheres of activity does not compromise the simplicity of the
substances themselves, since the spatial properties of substances (including the infinite
divisibility of space) arise from the interaction between their activities rather than from their
intrinsic features. In the course of the Physical Monadology, Kant also argues for the
necessity of attractive and repulsive forces and attributes a significant role to the force of
inertia. Kant's acceptance of such Newtonian principles represents an important change of
position over the True Estimation, where Kant rejects the principle of inertia and pursues a
dynamical theory much more in line with Leibniz's views.
In addition to these works, which bridge the gap, as it were, between physics and
metaphysics, during this period Kant is interested in a series of specific issues in cosmology
and empirical physics. For example, Kant writes several short exclusively scientific essays
between 1754 to 1757, including “Brief Outline of Certain Meditations on Fire,”
“Investigation of the question of whether the Earth has suffered changes in its axial
Rotation,” “The Question of the Aging of the Earth, considered physically” as well as three
papers on earthquakes. Of much greater significance is his Universal Natural History and
Theory of the Heavens (1755), which represents an important contribution to science as such.
For in it Kant explains how one can explain the formation of the solar system from an initial
state, in which matter is dispersed like a cloud, solely by means of the interaction of attractive
and repulsive forces. In 1796, Laplace, unaware of Kant's argument, would develop a very
similar derivation, with the result that the view is now typically referred to as the Kant-
Laplace nebular hypothesis.
Later in his pre-Critical period (1763-1770), Kant attempts to build a comprehensive
metaphysical account on the basis of the framework that he had established in his first works.
Thus, in his The Only Possible Basis for a Demonstration of the Existence of God (1763) he
attempts to extend his reasoning to fundamental issues in both philosophical theology and
teleology, presenting, for the first time, his now famous criticisms of the three traditional
arguments for the existence of God, while developing a new theistic proof, based on the idea
that God is necessary as a real ground of the possibilities of things. After reading Hume's
Enquiry concerning Human Understanding in German translation sometime after 1755, Kant
distinguishes between real and logical grounds/opposition in his Attempt to Introduce the
Concept of Negative Magnitudes into Philosophy (1763) in order to avoid Hume's objection
that there is no logical contradiction in the existence of one thing not following the existence
of another. But in this work he is also interested in exploring the notion of a real
ground/opposition further by applying it more widely, e.g., to bodies, mental states, etc. Also
relevant is Kant's Concerning the Ultimate Foundation of the Distinction of the Directions in
Space (1768) which modifies his earlier account of space insofar as he seems to hold that
certain spatial properties may not be able to be explained entirely on the basis of the
interaction between fundamental substances. In his so-called Inaugural Dissertation (1770),
Kant continues to develop a more comprehensive philosophical system, which would
encompass the principles of both the sensible and the intelligible world, and in so doing
modifies his account of space and time even further. Over the course of the next ten years,
during which he published almost nothing, Kant would revise his views more systematically,
with the publication of the Critique of Pure Reason in 1781 representing the first major step
in his “critical turn.”
Adickes (1924), Harman (1982), Friedman (1992), Laywine (1993), Schönfeld (2000),
Kuehn (2001), Lefevre & Wunderlich (2000), and Watkins (1997, 2001, 2003) have
emphasized the importance of scientific issues in the development of Kant's thought during
his pre-Critical period, as he reacted to Leibniz, Newton and other, more immediate
predecessors (such as Christian Wolff, Christian August Crusius, Leonard Euler, Pierre Louis
Moreau de Maupertuis, and Martin Knutzen).

2. Physics: The Critical Period (Metaphysical Foundations of Natural Science)


Though Kant discusses issues relevant to physics in various works throughout the Critical
period (esp. the Critique of Pure Reason), his views on this topic are developed most
explicitly in the Metaphysical Foundations of Natural Science (1786), which consists of a
preface and four chapters.
2.1 Preface
In the Preface to the Metaphysical Foundations Kant (i) analyzes the concepts of nature and
science so as to establish what conditions must be met for a body of knowledge to constitute
natural science in the proper sense, (ii) explains why science so understood requires “a pure
part” (4:469) and what criteria would have to be satisfied for such a pure part to exist, (iii)
argues that chemistry and psychology cannot at present meet these criteria, and (iv) describes
what procedure should be followed to satisfy these criteria and thus to provide the “pure part”
that science proper requires.
The feature of Kant's conception of natural science proper that is most immediately striking is
how restrictive it is. It requires that cognition (i) be systematically ordered (ii) according to
rational principles and (iii) be known a priori with apodictic certainty, i.e., with
“consciousness of their necessity” (4:468). Because properly scientific cognition must satisfy
these strict conditions, it requires “a pure part on which the apodictic certainty that reason
seeks can be based” (4:469). But since Kant identifies pure rational cognition that is
generated from concepts with metaphysics, it follows that science proper requires a
metaphysics of nature. He then specifies that such a metaphysics of nature could consist in
either a “transcendental part,” which discusses the laws that make possible the concept of a
nature in general — “even without relation to any determinate object of experience” (4:469)
— or a “special metaphysical” part, which concerns a “particular nature of this or that kind of
things” for which an empirical concept is given.
Kant's very conception of natural science proper thus immediately gives rise to several
systematically important questions. First, if the “transcendental part” of the metaphysics of
nature can be identified with the results of the Critique of Pure Reason, then the
Metaphysical Foundations is a work in special metaphysics. But what exactly is a special
metaphysics? In particular, what particular natures or kinds of things could be its object? And
how precisely can an empirical concept of such things be given without compromising the
necessity required of the pure part of natural science? Second, how is the special metaphysics
provided by the Metaphysical Foundations supposed to be related to the transcendental part
of the metaphysics of nature that was established in the Critique of Pure Reason? Does the
former presuppose the principles of the latter or are they logically independent, but still
related to each other in some other way?
First, Kant suggests that in special metaphysics the principles of the transcendental part “are
applied to the two species of objects of our senses” (4:470). Thus, the particular kinds of
things that could be investigated in a special metaphysics are (i) the objects of outer sense,
i.e., matter, and (ii) the objects of inner sense, i.e., thinking beings, which would thus result in
a doctrine of body and a doctrine of soul. Kant then argues that because “the possibility of
determinate natural things cannot be cognized from their mere concepts … it is still required
that the intuition corresponding to the concept be given a priori, that is, that the concept be
constructed” (4:470), which is a task that requires mathematics. This is Kant's justification for
his famous claim that “in any special doctrine of nature there can be only as much proper
science as there is mathematics therein” (4:470). This argument suggests that the necessity
required of the pure part of natural science derives from the necessity of the rules by which
the mathematical construction of determinate things must proceed.
Kant then uses the claim that science proper requires the construction of the concept of the
object in a priori intuition to exclude the possibility that chemistry and psychology, at least as
they were practiced at that time, could count as science proper. In the case of chemistry, the
problem is that “no law of the approach or withdrawal of the parts of matter can be specified
according to which … their motions and all the consequences thereof can be made intuitive
and presented a priori in space (a demand that will only with great difficulty ever be
fulfilled)” (4:471). Since its principles are “merely empirical,” it can, at best, be a “systematic
art” (ibid.). The case of psychology is more complex, since Kant provides (at least) two
separate reasons in the Preface for denying it the status of natural science proper. First, Kant
claims that mathematics is inapplicable to the phenomena of inner sense and their laws,
though he grants that the law of continuity (discussed, e.g., at A207-209/B253-255 and
A228-229/B281 in the Critique of Pure Reason) ought to apply to changes in our
representations as well. He downplays the significance of this application of the law of
continuity, however, by noting that time has only one dimension, which does not provide
enough material to extend our cognition significantly. Second, Kant also complains that
empirical psychology cannot separate and recombine the phenomena of inner sense at will;
rather, our inner observations can be separated “only by mere division in thought” (4:471).
Kant's fuller views on chemistry and psychology will be discussed further below.
Second, in explaining how mathematics can be applied to bodies Kant asserts that “principles
for the construction of the concepts that belong to the possibility of matter in general must
first be introduced. Therefore a complete analysis of the concept of a matter in general [must
be provided in which it] makes use of no particular experiences, but only that which it finds
in the isolated (although intrinsically empirical) concept itself, in relation to the pure
intuitions in space and time, and in accordance with laws that already essentially attach to the
concept of nature in general” (4:472). Kant then explains that this means that the concept of
matter must be determined according to the Critique of Pure Reason's categories of quantity,
quality, relation, and modality (4:474-476). Further, Kant holds that “a new determination”
(4:476) must be added to the concept of matter in each chapter of the Metaphysical
Foundations. This suggests not only that the principles argued for in the Metaphysical
Foundations are to be developed “in accordance with” the principles defended in the Critique
of Pure Reason, but also that both the concept of matter and the Metaphysical Foundations
itself is structured according to the Critique of Pure Reason's table of categories.
Unfortunately, these points of clarification do not resolve all of the issues that are
immediately raised by Kant's pronouncements about what is required for natural science
proper. One further issue that is relevant here concerns the concept of matter that is at the
heart of the Metaphysical Foundations. Kant introduces it in the Critique of Pure Reason
(A847-848/B875-876) as the concept of something that is impenetrable, extended, and inert.
Yet, in the beginning of the Preface of the Metaphysical Foundations, he describes it as
whatever is an object of outer sense, and later he argues that the “basic determination of
something that is to be an object of the outer senses had to be motion, because only thereby
can these senses be affected” (4:476). Whatever weight one accords Kant's justification of the
connection between matter, outer sense, and motion, one faces a dilemma. If the concept of
matter, most fundamentally, is simply the concept of any object of outer sense, then how is it
still empirical in any genuine sense (and what has become of the structural difference Kant
draws between the Critique of Pure Reason and the Metaphysical Foundations)? If, by
contrast, impenetrability, extension, and movability are deemed the basic traits of the concept
of matter, then how can one know a priori that any object we might encounter in outer sense
must behave in accordance with the laws that would govern matter so defined?
Moreover, even if one can find an appropriately nuanced sense in which the concept of matter
is empirical while still allowing for an appropriate kind of necessity, questions can still be
posed about the “new determinations” that are to be added to that concept in each chapter of
the Metaphysical Foundations. For example, what is the justification for each specific
determination that is added when one thinks of matter as having a quantity, a quality, etc.?
Also, what is the relationship between each new determination of matter and the various
claims that Kant makes in each chapter of the Metaphysical Foundations? In particular, when
Kant explicitly invokes principles for constructing concepts belonging to the possibility of
matter, is his idea that these principles are required insofar as they make experience of the
relevant “new determination” of matter possible (so that Kant would be developing a
transcendental argument in the Metaphysical Foundations similar in many respects to the
Critique of Pure Reason)? Answers to these questions depend on how one interprets the
arguments Kant develops throughout the Metaphysical Foundations.
The conception of science that Kant presents in the Preface has been the focus of
considerable attention over the past several decades. In the German literature, the issues
raised above have been discussed at length by Plaass (1965), Schäfer (1966), Hoppe (1969),
Gloy (1976), and Cramer (1985). Pollok (2001) has recently produced a detailed and
comprehensive textual commentary on the Metaphysical Foundations. Important work has
also been done in the English literature by Walker (1974), Brittan (1978), Buchdahl (1968,
1969, and 1986), Butts (1986), and Watkins (1998a). Friedman (1992, 2001, and 2002) has
been especially influential on these issues as well.
2.2 Phoronomy
The first chapter of the Metaphysical Foundations, the Phoronomy, considers the quantity of
motion of matter and how it is to be constructed in intuition a priori (so as to produce the kind
of rules that are necessary for our experience of matter in motion). Since extension and
impenetrability are not directly relevant to how different magnitudes (or degrees) of motion
can be represented, Kant restricts his discussion in this chapter to matter considered as a
point. Since the motion of a point in space can be represented straightforwardly, the main
issue is how to represent the composition of two different motions. Kant's primary claim in
this chapter is that due to the relativity of space (i.e., the fact that every motion can be viewed
arbitrarily as either the motion of a body in a space at rest, or as a body in a state of rest in a
space which is in motion in the opposite direction with the same velocity) “the composition
of two motions of one and the same point can only be thought in such a way that one of them
is represented in absolute space, and, instead of the other, a motion of the relative space with
the same speed occurring in the opposite direction is represented as the same as the latter”
(4:490). The proof of this Theorem considers the three possible cases for the composition of
two motions: (i) The two motions are in the same direction; (ii) the two motions are in
opposite directions; (iii) the two motions enclose an angle. Kant then shows how one can
construct a priori in intuition a single motion out of the two motions described in cases (i)-
(iii).
Very little has been written directly on Kant's Phoronomy. (Kant's philosophy of
mathematics, by contrast, has received considerable attention.) The most extensive
discussions of the Phoronomy are by Palter (1972) and Pollok (2001).
2.3 Dynamics
The second chapter of the Metaphysical Foundations, the Dynamics, considers how it is
possible to experience matter as filling a determinate region in space. Propositions 1-4 are
devoted to exhibiting the nature and necessity of repulsive forces. In Proposition 1 Kant
argues that repulsive force is required for matter to fill space, since solidity, understood by
“Lambert and others” as the property matter would have by means of “its mere existence”
(4:497), cannot truly explain how one matter resists another matter's attempt to penetrate it.
Kant then specifies several central features of repulsive forces in Propositions 2 and 3.
Repulsive forces admit of degrees to infinity, since one must always be able to think of a
slightly greater or lesser force, and although matter can be compressed to infinity, it can
never be penetrated, since that would require an infinite compressing force, which is
impossible.
In Proposition 4 Kant draws an important consequence from his characterization of repulsive
forces, namely that matter is infinitely divisible (4:503). What is especially striking about this
point is that it represents a significant departure from his own earlier Physical Monadology,
where he had accepted attractive and repulsive forces, but denied the infinite divisibility of
what ultimately constitutes matter, namely physical points or monads. It is true that part of
Kant's rationale for his change of position on this point stems from the “critical turn”
undertaken in the Critique of Pure Reason (and in its Second Antinomy in particular). For
once one recognizes that both space and spatial properties such as divisibility are not
properties of things in themselves but rather only appearances, one can reject the proposition
that seems to necessitate the acceptance of simple substances, namely the idea that simple
substances must precede the wholes they compose (4:506). However, Kant's proof also seems
to depend, in its details, not merely on the idea that every space is filled by means of some
repulsive force or other, but on the stronger claim that every space is divisible into smaller
spaces that are filled by different repulsive forces.
Propositions 5-8 are all devoted to attractive force. In Proposition 5 Kant argues that matter
must have an attractive force in order to fill space. Kant's argument is that if there were only
repulsive forces, then matter “would disperse itself to infinity” (4:508) since neither space nor
other matter could limit it. Proposition 6 argues that both attractive and repulsive forces must
be considered essential to matter. That is, attractive forces alone are not sufficient to account
for matter filling a space, since if matter consisted solely of attractive forces, there would be
no force to counteract the attractive force being exercised and the universe would collapse
into a single point. Proposition 7 then specifies how attractive forces are to be understood,
namely as the immediate action of matter on other matter through empty space (and therefore
at a distance). Kant thus directly confronts the metaphysical question of how to understand
attraction that Newton attempted to avoid by positing it merely mathematically. As Kant
interprets the situation, Newton “abstracts from all hypotheses purporting to answer the
question as to the cause of the universal attraction of matter … [since] this question is
physical or metaphysical, but not mathematical” (4:515). In response to the “most common
objection to immediate action at a distance,” namely “that a matter cannot act immediately
where it is not” (4:513), Kant argues that action at a distance is no more problematic than
action by means of physical contact (whether it be by collision or pressure), since in both
cases a body is simply acting outside itself. Proposition 8 concludes by arguing that attractive
forces act immediately to infinity and by adding a “preliminary suggestion” (4:518) as to how
one might be able to construct the concept of cohesion (which Kant understands as attraction
that is restricted to contact).
In the General Remark to Dynamics Kant addresses two main issues. First, Kant considers
how it is that the specific varieties of matter (e.g., water as different from mercury) might be
reduced, at least in principle, to the fundamental forces of attraction and repulsion. The
second issue concerns the fundamental distinction between the “mathematical-mechanical”
and the “metaphysical-dynamical mode of explanation”. The former mode of explanation,
which is associated with the postulation of atoms and the void, employs nothing more than
the shapes and motions of fundamental particles and empty interstices interspersed among
them. It contrasts with the metaphysical-dynamical mode, which employs fundamental
moving forces (e.g., attraction and repulsion) in its explanations. Kant grants that the
mathematical-mechanical mode has an advantage over the metaphysical-dynamical mode,
since its fundamental posits can be represented (indeed, “verified” (4:525)) mathematically,
whereas he repeatedly admits that the possibility of fundamental forces can never be
comprehended, i.e., their possibility can never be rendered certain. However, Kant thinks that
this advantage is outweighed by two disadvantages. First, by presupposing absolute
impenetrability, the mathematical-mechanical mode of explanation accepts an “empty
concept” at its foundation. Second, by giving up all forces that would be inherent in matter,
such a mode of explanation provides the imagination with more freedom “than is truly
consistent with the caution of philosophy” (4:525).
Given that the bulk of Kant's matter theory is presented in the Dynamics, it is unsurprising
that it has received the greatest amount of attention in the literature. Of particular note are
discussions by Buchdahl (1968, 1969), Brittan (1978), Kitcher (1983), Butts (1986), Carrier
(1990), Friedman (1990), Malzkorn (1998), Warren (2001), Pollok (2002), and Engelhard
(2005).
2.4 Mechanics
The third chapter of Kant's Metaphysical Foundations, the Mechanics, concerns how it is
possible to experience matter as having a moving force, that is, how one matter
communicates its motion to another by means of its moving force. Kant begins, in
Proposition 1, by clarifying how the quantity of matter is to be estimated before stating, in
Propositions 2-4, three Laws of Mechanics.
After first defining the quantity of matter and the quantity of motion (or, in contemporary
terms, impulse, i=mv), Kant asserts that the quantity of matter, in comparison with every
other matter, can be estimated only by the quantity of motion at a given speed (4:537). Kant's
proof proceeds by way of elimination. The quantity of matter, which is the aggregate of the
movable in a determinate space, cannot be estimated by counting the number of parts it has,
since, as was established in the Dynamics, every matter is infinitely divisible. Nor can one
estimate the quantity of matter merely by considering its volume, since different matters can
have different specific densities. As a result, the only universally applicable way of
estimating the quantity of matter is to hold the velocity of matter constant.
In Proposition 2, Kant states his First Law of Mechanics: the total quantity of matter remains
the same throughout all changes in matter (4:541). His proof seems to rely (i) on the principle
of the First Analogy of Experience that no substance arises or perishes throughout any change
in nature and (ii) on the identification of what in matter must be substantial. On this latter
point, Kant quickly assumes that the ultimate subject of all accidents inhering in matter must
be the movable in space, and that its quantity is the aggregate of the movable in space. In his
remark to this proposition, Kant explicitly notes that there is a crucial difference between
spatial and non-spatial substances, since the latter, unlike the former, could gradually fade
away by degrees. (Kant cites the possibility of consciousness as a concrete example.) Kant
uses this difference to argue that since the quantity of matter consists in a plurality of real
things external to each other that cannot fade away (as consciousness might), the only way to
decrease its quantity is by division.
Kant's Second Law of Mechanics, stated in Proposition 3, is that every change in matter has
an external cause. (Immediately after this principle, Kant adds in parentheses a version of the
law of inertia that is much closer to Newton's: “every body persists in its state of rest or
motion, in the same direction, and with the same speed, it is not compelled by an external
cause to leave this state” (4:543). Since Kant's Second Law of Mechanics is not identical to
Newton's law of inertia, it would require argument to show that, and by means of what
additional assumptions, the former entails the latter.) The proof of the main principle depends
on the Second Analogy of Experience (which asserts that all changes occur in accordance
with the law of cause and effect and thus entails that every change in matter has a cause) as
well as on the further assumption that matter has no internal grounds of determinations (such
as thinking and desiring), but rather only external relations in space. In his remark to this
proposition, which clarifies this “law of inertia,” Kant explains that inertia is to be contrasted
with life or the ability of a substance to determine itself to act from an internal principle.
Thus, a body's inertia “does not mean a positive striving to conserve its state” (4:544), but
rather what it does not do, its lifelessness.
Kant also asserts that the very possibility of natural science proper depends on the law of
inertia, since the rejection of it would be hylozoism, “the death of all natural philosophy”
(4:544). In a later remark in the Mechanics, Kant explicitly objects that “the terminology of
inertial force (vis inertiae) must be entirely banished from natural science, not only because it
carries with it a contradiction in terms, nor even because the law of inertia (lifelessness)
might thereby be easily confused with the law of reaction in every communicated motion, but
primarily because the mistaken idea of those who are not properly acquainted with the
mechanical laws is thereby maintained and even strengthened” (4:550). Kant goes on to point
out that if inertia were to entail an active force of resistance, then it would be possible that
when one moving body hits another, the moving body has to apply part of its motion solely to
overcome the inertia of the one at rest and might not have any motion left over, as it were, to
set the body at rest into motion, which is contrary to experience (and Proposition 2).
Kant's Third Law of Mechanics, expressed in Proposition 4, asserts the equality of action and
reaction in the communication of motion. Kant formulates a version of the Third Analogy of
Experience (according to which all external action in the world is interaction) and suggests
that the main point at issue in mechanics is establishing that mutual action is necessarily
reaction. Kant's argument for this law is based on the following line of thought: (i) if all
changes of matter are changes of motion; (ii) if all changes of motion are reciprocal and equal
(since one body cannot move closer to/farther away from another body without the second
body moving closer to/farther away from the first body and by exactly the same amount); and
(iii) if every change of matter has an external cause (a proposition that was established as the
Second Law of Mechanics), then the cause of the change of motion of the one body entails an
equal and opposite cause of a change of motion of the other body or, in short, action must be
equal to reaction.
In Remark 1, Kant then shows how his position differs from that of other authors. Newton
“by no means dared to prove this law a priori, and therefore appealed rather to experience”
(4:449). Kepler likewise derived it from experience, though he went further, conceiving of it
in terms of a special force of inertia. Certain unnamed “transfusionists” (presumably Locke
and perhaps Descartes or Rohault) attempted to deny the law altogether by suggesting that
motion could simply be transferred from one body to another in the communication of
motion, a view Kant rejects on the grounds that explaining the communication of motion in
terms of the transfer of motion is no explanation at all and also amounts to admitting that
accidents could be literally transferred from one substance to another.
Kant's Laws of Mechanics have been discussed widely in the secondary literature. One can
point to discussions by Palter (1972), Duncan (1984), Friedman (1989, 1992, and 1995),
Brittan (1995), Westphal (1995), Carrier (2001), and Watkins (1997 and 1998b). Friedman's
interpretation (1992) deserves special mention. According to Friedman, in light of Kant's
rejection of Newton's absolute space and time, he must develop some way of providing an
appropriate meaning for absolute motion, i.e., for the real, as opposed to merely apparent,
relative motion of bodies. To this end, he “views the laws of motion as definitive or
constitutive of the spatio-temporal framework of Newtonian theory, and this, in the end, is
why they count as a priori for him” (p. 143). More specifically: “We need to presuppose the
immediacy and universality of gravitational attraction in order to develop a rigorous method
for comparing the masses of the primary bodies in the solar system. We need such a method,
in turn, in order rigorously to determine the center of mass of the solar system. This, in turn,
is necessary for rigorously determining a privileged frame of reference and thus for giving
objective meaning, in experience, to the distinction between true and apparent (absolute and
relative) motion. This, finally, is necessary if matter, as the movable in space (Definition 1 of
the Phoronomy: 480.5-10), is to be itself possible as an object of experience. Hence an
essential — that is, immediate and universal — attraction is necessary to matter as an object
of experience. It follows, for Kant, that the immediacy and universality of gravitational
attraction must be viewed — like the laws of motion themselves — as in an important sense a
priori. These two properties cannot be straightforwardly obtained from our experience of
matter and its motions — by some sort of inductive argument, say — for they are necessarily
presupposed in making an objective experience of matter and its motions possible in the first
place” (pp. 157-158).
2.5 Phenomenology
The final chapter of the Metaphysical Foundations, the Phenomenology, focuses on how the
motion of matter can be experienced modally, that is, in terms of it being possibly, actually,
or necessarily in motion. Its three Propositions specify (in accordance, Kant suggests, with
the results of the three previous chapters) that (i) rectilinear motion is a merely possible
predicate of matter, (ii) circular motion is an actual predicate of matter, and (iii) the equal and
opposite motion of one matter with respect to another is a necessary motion of that matter. In
the General Remark to the Phenomenology, Kant discusses the status of absolute space,
which had been presupposed by the possible, actual, and necessary motions of matter at issue
in the three main propositions, and explains that since it is not itself an object of experience,
it must be represented by means of an idea of reason (in Kant's technical sense of “idea”,
namely as a concept for which a corresponding object could never be given to us in intuition).
Though we can never know absolute space, it none the less functions as a regulative principle
that guides us in our scientific practice by forcing us to look for further conditions for the
conditioned objects we meet with in experience. Kant's view that ideas of reason can function
as regulative principles is developed in the Appendix to the Transcendental Dialectic in the
Critique of Pure Reason.
There is relatively little secondary literature that discusses Kant's Phenomenology. Carrier
(1992) provides a detailed interpretation of Kant's conception of absolute space. Friedman
(1992, Chapter 3, but also 1995) argues that the Phenomenology should be read “as
attempting to turn Newton's argument of Book III of Principia on its head” (1992, p. 142).
3. Physics: The Post-Critical Period (Opus postumum)
Kant's interest in physics continued after the publication of the Metaphysical Foundations, in
fact, until the very end of his productive years. Though Kant never completed a manuscript
that could be put forward as a publication, the various notes, sketches, and drafts on topics in
physics that he was working on intensively during this time (especially after 1796) were
gathered together over a century after his death and published as his so-called Opus
postumum.
Despite the fragmentary nature of the Opus postumum, Kant makes it clear that it is designed
to fill an important gap in his system. Just as the Metaphysical Foundations had attempted to
connect the transcendental principles of the Critique of Pure Reason and the principles that
explain how matter is possible, the Opus postumum undertakes the task of effecting a
transition from the special metaphysics of nature contained in the Metaphysical Foundations
to physics itself. However, Kant does not clarify adequately what systematic principles would
guide this transition project. On the one hand, in a note that stems from a period shortly after
the publication of the Metaphysical Foundations, Kant suggests that one could “follow the
clue given by the categories and bring into play the moving forces of matter according to
their quantity, quality, relation, and modality” (21:311), a procedure that could be similar to
that of the Metaphysical Foundations. On the other hand, if the Metaphysical Foundations
already presupposes an empirical concept (namely matter), the transition to be carried out in
the Opus postumum cannot be understood as moving from something non-empirical to
something empirical. As Kant struggles with the problems that result from trying to account
for now much more specific features of matter, it is unclear that (or how) the categories are
supposed to be of help in structuring Kant's argument. Thus, the precise argumentative
structure of the Opus postumum (i.e., its relationship to Kant's other works and its
fundamental presuppositions) remains problematic.
Whatever its form, the content of the Opus postumum includes reflections on a series of
important topics in physics. Three clusters are particularly noteworthy. (1) Kant develops
more detailed views about a number of outstanding issues concerning matter theory that he
had discussed (often in a tentative way) in the Metaphysical Foundations, such as fluidity,
rigidity, cohesion, and the quantity of matter. (2) Kant argues for the existence of an all-
encompassing ether. This might seem to be a natural development, since the Metaphysical
Foundations was non-committal on the point, but what is surprising is that Kant thinks that
the ether can be established a priori (e.g., 21:222), which might seem to conflict with Kant's
project in the Critique of Pure Reason (or with his description of his position as “formal
idealism,” 4:337). (3) Kant also explores the idea that the subject must posit itself in positing
the various forces in matter, a doctrine that has come to be known as the Selbstsetzungslehre,
and attempts to incorporate it into his views on how man is situated in between the world of
experience and God, whose existence is a central requirement of morality.
The Opus postumum has long been a topic of interest especially insofar as it offers the hope
of clarification and development of central issues in Kant's Critical philosophy. While much
of the original literature focussing on it was in German (Adickes 1920, Hoppe 1969,
Tuschling 1971, and, more recently, Blasche, 1991 and Emundts, 2004), it has received
increased attention of late in English with discussions by Friedman (1992, chapter 5), Förster
(2000), Edwards (2000), and Guyer (2001).
4. Chemistry
In the Preface to the Metaphysical Foundations Kant claims that chemistry, at least as he
understood it in 1786, was not science “proper”, but such a claim leaves open the possibility
that chemistry could be fully scientific in some other sense or that, with time, it could develop
into science proper. Up through 1787, Kant accepted the fundamental tenets of Stahl's
chemical theory, according to which water and air are fundamental elements that function as
vehicles for change in both inflammable and “earthy” substances, and he commented on
particular issues in chemistry in his various physics lectures. As a result, it is clear that Kant
considers chemistry to be a science in some sense even during the Critical period. However,
beginning in the mid-1780s (and extending through the mid to late 1790s), Kant becomes
aware of significant new developments in chemistry (as evidenced in the Danziger Physik
and documented in Lichtenberg's revised, third edition of Erxleben's Anfangsgründe der
Naturlehre from which Kant lectured). In particular, Kant comes to reject Stahl's theory,
favoring Lavoisier's anti-phlogistic account of combustion and calcination, which relied on
his doctrines of latent heat and the caloric theory of the states of aggregation. While Kant
never explicitly claims that chemistry, so understood, can be considered science proper,
Kant's interest in these issues in the Opus postumum suggests that he was optimistic about
providing the kind of foundation that would be required for it to attain this status.
Kant's views on chemistry have not been widely discussed in the secondary literature.
However, outstanding discussions of Kant's views on the topic can be found in Carrier (1990,
2001) and Friedman (1992, chapter 5, III).
5. Psychology
Kant's views on psychology are intimately bound up with his more general position in the
philosophy of mind. (See the separate entry on this topic.) Still, one can take note of the fact
that Kant distinguishes between rational and empirical psychology and, in the Critique of
Pure Reason's Paralogisms, denies that rational psychology contains arguments that could
justify any substantive principles (especially concerning our immortality). Thus, only
empirical psychology, it seems, could be possible as a science. However, if Kant continues to
maintain that science requires a pure part, and denies that rational psychology contains any
substantive knowledge that might constitute the pure part of psychology, then it follows that
empirical psychology cannot qualify as science proper either.
At the same time, Kant's own project in the Critique of Pure Reason requires what one might
call transcendental psychology, that is, the study of those faculties that are required for us to
have cognition. Transcendental psychology thus differs from rational psychology insofar as
the former presupposes that we have experience (albeit of a very general sort), whereas the
latter is restricted to the mere concept “I think”. Thus, it would seem that many of Kant's
most important claims in the Critique of Pure Reason would fall under the domain of
transcendental psychology. As is well known, Kant's Critique of Pure Reason came under
attack immediately after its publication (most notably by Johann Georg Hamann) for not
presenting an explicit account of how we obtain knowledge of our transcendental faculties.
Just as Kant held that chemistry could either be scientific in less than the strictest sense or
perhaps become science proper (depending on how it develops), the same could be said for
psychology. For one, Kant may simply be objecting to psychology as it was practiced in the
18th century, e.g., as based on introspection (a method that both rules out the application of
mathematics and gives rise to other difficulties, 4:471). Further, in the Critique of Pure
Reason, Kant holds that although objects can never be given in intuition that would
correspond to ideas of reason, such ideas never the less function as regulative principles that
direct our understanding with regard to what it should inquire into next. Thus, our idea of the
world as a totality is supposed to drive us to look for smaller and smaller parts to bodies and
objects in further regions of space and at earlier moments of time. But if we have an idea of
our soul, then it too should guide our scientific inquiry into our own representations, which
seems to imply that psychology is a legitimate scientific practice (even if it does ultimately
fall short of being science proper).
Kitcher (1990) presents a detailed argument for the role of transcendental psychology within
Kant's Critique of Pure Reason. Sturm (2001) argues that Kant's critical comments about
psychology are primarily directed against introspection-based conceptions of psychology.
6. Other Sciences: History, Physical Geography, and Anthropology
Though Kant is sometimes quite strict about what qualifies as science “proper,” we have seen
that he clearly accepts that other disciplines, such as chemistry and psychology, can be
scientific in some other sense. (Kant is also extremely interested in providing an explanation
of the nature and origin of organisms, which is central to an account of biology. For a more
detailed description of Kant's stance on biology, see the entry on his views on aesthetics and
teleology.) However, his explicit reflections about science (taken broadly so as to encompass
not only “Wissenschaft” but any kind of “Lehre” or “Kunde”) extend even further so as to
include bodies of cognition such as history, physical geography, and anthropology. (Even in
the Metaphysical Foundations Kant leaves room for such “scientific cognition” by dividing
the doctrine of nature into natural science — e.g., physics — and the historical doctrine of
nature, which is separated further into natural description and natural history, 4:468). In the
case of historical sciences, Kant views its cognitions as related not by subordination (as is the
case in physics), but rather by coordination, since historical facts cannot be derived from each
other, but rather merely related to each other in space (geographically) and time
(chronologically). Kant views anthropology as having the same subject matter as empirical
psychology. What enables the coordination of facts in each of these domains is not a
theoretical, but rather a practical idea. In the case of history, it is the idea of reason (or
freedom) that provides a guiding principle of coordination (cf. Kant's “Idea for a Universal
History from a Cosmopolitan Point of View”), while anthropology is framed by the
cosmopolitan moral ideal of the world-best. Thus, it is clear that Kant's emphasis on physics
throughout his career did not blind him to the value of other sciences, nor did it keep him
from reflecting creatively on how best to account for them from the perspective of his Critical
philosophy.
Kant's views on history are discussed by Yovel (1980) and, more recently, Kleingeld (1995
and 1999). Kant's Physical Geography has been discussed by Adickes (1911). Brandt (1999),
Wood (1999), and Makkreel (2001) have recently published important discussions of Kant's
anthropology.
The Copernican Revolution
The Copernican revolution in science, occurring roughly at the beginning of the sixteenth
century, also had far-reaching consequences for philosophy. Until that time science and
philosophy of science had operated by rational thought far more than by empirical
observation. But the Copernican revolution, based as it was on observation of the world more
than on thought and thought processes, meant that science became much more empirically
and experimentally oriented than it had been in the Medieval-Aristotelian-Scholastic period,
and a philosophy of science based on an acknowledgement of the primary role—or at least
the large role—of observation and experiment in science came to the fore. Here a split
between England and the European continent arose, with British philosophy becoming
strongly empiricist, especially in the work of John Locke, George Berkeley, and David
Hume, and continental philosophy being more rationalist, especially in the philosophy of
Rene Descartes, Baruch Spinoza, and Gottfried Leibniz.
The Nineteenth Century
In the nineteenth century a great opposition arose between two camps of philosophers of
science, and that opposition continued up into the post-World War II era .One group, often
called inductivists, claimed that scientific hypothesis formation is an inductive process arising
from observation of particular items of evidence, and that causation is nothing but
observation of regularity. According to this view, a supposed inductive logic and procedure
give rise to and justify hypotheses. John Stuart Mill can be taken as the nineteenth-century
exemplar of that view; some (among many) others who took that position were Ernst Mach,
Pierre Duhem, Philipp Frank, Carl Hempel, Rudolf Carnap, R.B. Braithwaite, and most
members of the Vienna Circle. This group of philosophers is sometimes called reductionists,
and their attitude reductionism. Other forms of this view are sometimes called
instrumentalism, meaning that scientific theories are just instruments for making predictions,
and operationalism, which claims to reduce all empirical statements to experimental
operations. In addition, many exponents of this view, including Mill, held that the axioms of
mathematics are nothing more than generalizations from experience. Later developers of set
theory would try to prove – ultimately unsuccessfully – that mathematics could be derived
solely from set theory and was thus only a branch of formal logic (this view of mathematics
is sometimes called formalism), without any embodiment of any synthetic a priori
knowledge, as Kant had held.
A second group of philosophers held that scientific knowledge is not derived solely from the
senses but is a combination of sensation and ideas, and scientific hypotheses are not arrived at
solely through inductive logic or procedures; the knower contributes something. In this view,
“A scientist makes a discovery when he finds that he can, without strain, add an organizing
idea to a multitude of sensations.”[2] William Whewell can be taken as the nineteenth-century
exemplar of this view. This view is sometimes called realism because it holds that a scientific
statement or hypothesis expresses something true—some realists have gone so far as to hold
that it expresses a necessary truth and not just a contingent one—about a real external world,
and not just something about the observer’s empirical experience. Ludwig Boltzmann, N.R.
Campbell, and many of today’s post-positivist philosophers who hold to some version of
scientific realism can be understood as being in this camp.
Scientific realism is the view that the universe really is as explained by scientific statements.
Realists hold that things like electrons and magnetic fields actually exist. It is naïve in the
sense of taking scientific models at face value, and is the view that many scientists adopt.
In contrast to realism, instrumentalism holds that our perceptions, scientific ideas and theories
do not necessarily reflect the real world accurately, but are useful instruments to explain,
predict and control our experiences. To an instrumentalist, electrons and magnetic fields are
convenient ideas that may or may not actually exist. For instrumentalists, the empirical
method is used to do no more than show that theories are consistent with observations.
Instrumentalism is largely based on John Dewey's philosophy and, more generally,
pragmatism, which was influenced by philosophers such as Charles Peirce and William
James.
The empiricist/positivist attitude towards this nominalist-realist split was neatly summarized
by Rudolf Carnap:
Empiricists are in general rather suspicious with respect to any kind of abstract entities like
properties, classes, relations, numbers, propositions, etc. They usually feel much more in
sympathy with nominalists than with realists (in the medieval sense). As far as possible they
try to avoid any reference to abstract entities and to restrict themselves to what is sometimes
called a nominalist language, i.e. one not containing such references.[3]
From the Nineteenth and Into the Twentieth Century: Positivism
Explanations of observations and discoveries that were offered by the scientific revolution
were based on the characteristics, powers, and activities of nature itself and—supposedly,
anyway—not on any supernatural, transcendent, or occult powers or phenomena. If
something was observed in nature but no natural explanation for it was available at the time,
then it was assumed that a natural explanation existed and could be found later. In fact, the
demarcation between science and non-science, especially between science, on the one hand,
and religion and theology, on the other, was assumed to lie in this: It was supposed that
scientific knowledge or claims could be justified or proved without reference to anything
supernatural or transcendent, but religious and theological knowledge or claims necessarily
relied on supposed knowledge of what is extra-natural or transcendent. In fact, theory of
science, as it came to be developed, was primarily concerned to show both that and how
science has certainty in its knowledge without reliance on any transcendent or revealed or
metaphysical or occult knowledge.
One of the most widely asserted and held theories of science is known as positivism. This is,
William Reese has written, "A family of philosophies characterized by an extremely positive
evaluation of science and scientific method. In its earlier versions, the methods of science
were held to have the potential not only of reforming philosophy but society as well.”[4]
Following his teacher Saint-Simon, French philosopher Auguste Comte (1789-1857) founded
what he called the Positivistic Philosophy, which held, as Reese puts it, that "Every science,
and every society, must pass through theological and metaphysical states or stages, on the
way to the positive, scientific stage, which is their proper goal. In the theological stage
explanations are given in terms of the gods. In the metaphysical stage explanations are given
in terms of the most general abstractions. In the scientific stage explanations consist of
correlating the facts of observation with each other."
Austrian physicist-philosopher Ernst Mach (1836-1916) held that "Explanation consists in
calling attention to the sensations or 'neutral impressions' from which the concept was derived
and for which it stands. Any concept which does not relate to sensations in this manner is
metaphysical, and hence unacceptable." Moreover, "Scientific laws are abridged descriptions
of past experience designed to assist us in predicting future experience. Such abridged
descriptions are of the pattern of the past phenomena, and hence will be ... characteristic of
future instances of the same kind." In addition, the theoretical entities of science, "should not
be thought of as representing a world behind experience, but rather as instruments or tools
helping us formulate the mathematical relations leading to predictions." [5] Indeed, Mach tried
to get rid of unobservables altogether. He held that scientific investigations do not go past or
beyond observation to any “heart” of nature, and causality is merely the habit of mind in
seeing constant conjunction, as was held by David Hume. The goal of scientific investigation,
then, is to discover the relations between our sensations.
French philosopher Pierre Duhem (1861-1916) expressed views similar to those of Mach,
holding “a physical theory to be a system of mathematical propositions representing a set of
experimental laws…” and “The object of science is to discover the relations holding among
appearances.”[6] Science uses models, either actual ones or constructed ones, to represent
what underlies the observable phenomena.
Mathematics and Science
A major problem for empiricist and positivist accounts of science and knowledge that hoped
to eliminate metaphysics was posed by mathematics, for everyone recognized the central role
that mathematics plays in science. German philosopher Immanuel Kant—who, interestingly
enough, called his own work a "Copernican revolution"—had held that mathematics is both
synthetic (meaning that its statements tell us something about the world, as opposed to
analytic statements, which are true because of logic alone) and a priori (meaning that they are
known to be true prior to empirical experience). If Kant's claim is true, then things can be
known about the world before any observations are made. In addition, such notions as
absolute equality or perfect circle cannot be discovered by experiential observation because
there are no examples of perfect equality or perfect circle in the physical world. This shows
that mathematics and geometry must be metaphysical in nature in that they cannot be reduced
to or derived solely from the physical world.
This led to efforts to derive mathematics solely from logic, especially the logic of set theory,
and thereby disprove Kant's claim that mathematics is synthetic. The set theory of German
mathematician-logician Gottlob Frege (1848-1925) was the pioneering effort here, but British
logician-philosopher Bertrand Russell (1872-1970) found a fatal mistake in Frege's work.
This led Russell and his collaborator Alfred North Whitehead (1861-1947) to develop their
own consistent mathematical logic/set theory, published as Principia Mathematica (3 vols.),
and numerous people both before and since have worked on powerful forms of logic and
mathematical logic. But the effort to reduce even simple arithmetic to axiomatic logic was
shown decisively to be impossible by the work of Czech mathematician-logician Kurt Gödel
(1906-1978). Thus the effort to show that mathematics need not be metaphysical has failed.
Even today, some people still try to hold to empiricist, nominalist, or formalist accounts of
mathematics, but those efforts too have not been successful. Thus, those who hold to a realist
account of mathematics—to the view that there is a real world of numbers and mathematical
theorems and that mathematics cannot be reduced to generalizations from experience—have
a great deal of evidence on their side.
Verificationism and the Vienna Circle
The reductionistic views of Mach and Duhem were well formulated by the beginning of the
twentieth century and were highly influential in the developing philosophy of science. Indeed
the marriage of Mach’s positivism with the new powerful set-theoretic logic of Russell and
Alfred North Whitehead fed into the rise of logical positivism, also known as logical
atomism, expressed especially in the work of the members of the Vienna Circle. This was a
movement centering on the University of Vienna in the 1920s and 1930s, led by Moritz
Schlick (1882-1936). The goals of the Vienna Circle and the logical positivists were to
develop a positivistic account of science, to unify all the sciences into one unified system,
and to purify philosophy of all metaphysical elements, using logic as its means, and to
thereby also purify science of anything other than logic and empirical observation. In other
words, the logical positivists wished to draw a bright line or demarcation between science
and logic, on the valued and esteemed side, and anything else on the other side, especially
metaphysics and religion, and to banish everything that crossed that line because it was
deemed to be literally meaningless.
English philosopher A. J. Ayer (1910-1989) met with the Vienna Circle and returned to
England to write an enormously influential small book entitled Language, Truth, and Logic
(1936); this book amounted to being a gospel tract for logical positivism. In it Ayer
propounded what is known as the verification principle of meaning, according to which a
statement—and science is expressed in statements—is significant only if it is a statement of
logic (i.e. an analytic statement) or it could be verified by experience, meaning that it was
significant only if some empirical observation can be used to determine its truth or falsity.
Statements that did not meet those criteria of being either analytic or empirically verifiable
were judged to be nonsensical because they were not verifiable; such nonsensical and
meaningless statements included religious, metaphysical, ethical, and aesthetic statements.
Ayer's book provided a trenchant and arresting summary of logical positivism. It was read
and deeply absorbed by many thousands of people—not just philosophers, but many working
scientists and others—in the English-speaking world, and it had an enormous influence, so
that, for a time, many people were converted to positivism—to the view that only analytic or
empirical statements are meaningful, and that religious, metaphysical, ethical, and aesthetic
statements are meaningless. In fact, some people, including some working scientists, still
hold to that view even today.
Problems soon arose, however, about the verification principle itself: What is it and how shall
it be expressed? It is not analytic because it cannot be derived from logic. But it is not
empirical either, because it cannot be empirically discovered or verified. So what is it? If a
strict adherence to the verification principle is maintained, then the verification principle
itself is meaningless nonsense. That conclusion won't do, however, because the verification
principle is meaningful. So, again, what is it? The correct answer is that it is, in fact, a
metaphysical statement. But if it is both meaningful and metaphysical, this means that at least
some metaphysical statements are meaningful, and that refutes the verification principle. In
other words, investigation of the verification principle serves to refute that principle itself.
Proponents of logical positivism, when faced with this difficulty, resorted to saying that the
verification principle should be accepted as a proposal. But then others – those who believed
in religion, metaphysics, ethical, or aesthetic statements – could simply refuse to accept the
positivists' recommendation. The greatest or most thorough proponents of logical positivism,
such as Rudolf Carnap, Hans Reichenbach, Otto Neurath, and others, worked assiduously—
without ultimate success—to find some form or statement of the verification principle, or
confirmation principle as they tended to come to call it, that could be maintained.
Defenders of logical positivism also faced the problem that if the verification principle were
made strong it would destroy science itself, and the philosophical defense of science was
their major concern. One way of stating this problem is to note that scientific laws themselves
go beyond the evidence for them. Take, for example, the scientific law-like statement: “Pure
water freezes at 100 degrees Celsius.” That is a simple statement about a physical property of
water. But what is its evidential status? Have we frozen every sample of water in the
universe? Clearly not—we have frozen only a very small sample of all existent water. How
do we know that all samples will have the same freezing temperature as the ones we have
tested? We assume that all water is the same, but, in principle, we cannot empirically test that
assumption. We also assume that samples of water from elsewhere in the universe, besides
our planet, will melt at the same temperature. Why or what empirically verifiable principles
allow us to make these assumptions?
In short, our scientific law statements themselves go beyond the evidence that we have tested
in order to arrive at them because they necessarily make claims about untested things. Thus,
these statements must be regarded as being, at least to some extent, metaphysical. As John
Passmore has written, "Such [scientific] laws are, by the nature of the case, not conclusively
verifiable; there is no set of experiences such that having these experiences is equivalent to
the scientific law." So, in order to deal with this problem, the verification principle had to be
made weak enough to admit this dimension of scientific statements. But if it were weakened
in that way, then it would permit at least some other metaphysical statements, such as "Either
it is raining or the Absolute is not perfect." All attempts to solve this problem of having a
version of the verification principle (or confirmation principle) that admits all scientific
statements but excludes all metaphysical statements have met with failure.
A big problem for positivist and verificationist accounts of science arose when philosophers
pointed out the theory-dependence of observation. Observation involves perception, and so is
a cognitive process. That is, one does not make an observation passively, but is actively
involved in distinguishing the thing being observed from surrounding sensory data.
Therefore, observations depend on some underlying understanding of the way in which the
world functions, and that understanding may influence what is perceived, noticed, or deemed
worthy of consideration. (See the Sapir-Whorf hypothesis for an early version of this
understanding of the impact of cultural artifacts on our perceptions of the world.)
Empirical observation is supposedly used to determine the acceptability of some hypothesis
within a theory. When someone claims to have made an observation, it is reasonable to ask
them to justify their claim. Such a justification must make reference to the theory –
operational definitions and hypotheses – in which the observation is embedded. That is, the
observation is a component of the theory that also contains the hypothesis it either verifies or
falsifies. But this means that the observation cannot serve as a neutral arbiter between
competing hypotheses. Observation could only do this "neutrally" if it were independent of
the theory. Still another problem for verificationism arose from the Quine-Duhem thesis,
which pointed out that any theory could be made compatible with any empirical observation
by the addition of suitable ad hoc hypotheses. This is analogous to the way in which an
infinite number of curves can be drawn through any set of data points on a graph.
Confirmation holism, developed by Willard Van Orman Quine, states that empirical data is
not sufficient to make a judgment between theories. A theory can always be made to fit with
the available empirical data.
That empirical evidence does not serve to determine between alternate theories does not
imply that all theories are of equal value. Rather than pretending to use a universally
applicable methodological principle, the scientist is making a personal choice when choosing
one particular theory over another.
What all this means is that positivism, verificationism, or confirmationism cannot supply a
means whereby we can draw a bright or sharp line between genuine or true science on one
hand and metaphysics on the other. Science and metaphysics do merge into and encroach on
each other, at least so far as any positivist, verificationist or confirmationist account of
science can say.
Later in his life, Ayer himself admitted that the logical positivist program that he had so well
described and championed was all wrong.
Mounting evidence against it from all these considerations, and more, led eventually to the
demise of logical positivism and its program. As John Passmore wrote in the 1960s, “Logical
positivism, then, is dead, or as dead as a philosophical movement ever becomes.”[7] Among
other things, this means that the line between science and metaphysics—and possibly even
the line between science and religion—is not as clear or sharp as many people have wanted or
supposed it to be.
Popper and Falsificationism
Austrian-English philosopher Karl Popper (1902-1994) responded to the positivists by
proposing that both verificationism and inductivism should be replaced with falsificationism.
Popper was not a member of the Vienna Circle, but he met with some of its members and
discussed his work with them.
Early in his life Popper had become a Marxist and sometime later he became familiar with
the work of Sigmund Freud and of psychologist Alfred Adler. But then Albert Einstein came
and gave a lecture in Vienna, which Popper attended. He reports that he did not understand
what Einstein was saying, but that he did recognize in Einstein “an attitude utterly different
from the dogmatic attitude of Marx, Freud, Adler, and even more so than that of their
followers. Einstein was looking for crucial experiments whose agreement with his predictions
would by no means establish his theory, while a disagreement, as he was the first to stress,
would show his theory to be untenable.” This led Popper to the insight that would, he wrote,
guide his subsequent work in philosophy of science, "I arrived, by the end of 1919, at the
conclusion that the scientific attitude was the critical attitude, which did not look for
verifications but for crucial tests; tests which could refute the theory tested, although they
could never establish it."[8]
Positivist theory of science had been based on a theory of verification, on the view that
positive instances (observations or data), through a process of inductive logic or inductive
method, would serve to establish and verify a scientific theory, e.g. observation of white
swans would serve to positively verify the theory “All swans are white.”
Positivism and verificationism were, thus, based on acceptance of inductive logic and
inductive methods. But inductive methods and inductive logic had already been severely
criticized by Scottish philosopher David Hume (1711-1776), who, in his Treatise of Human
Nature, had pointed out that inferences drawn from experiences in the past and present
cannot be projected into the future unless one adopts a premise that the future will be like the
past.
That premise that the future will be like the past, however, cannot itself be justified
deductively, and any attempt to justify it inductively will lead to an infinite regress. Hume's
general point was that there is no logical sanction for an inference from observed cases to
unobserved ones.
A great deal of philosophers' thought and ink have been devoted to showing that Hume was
wrong and that inductive inferences can be logically good ones, but all that effort ultimately
has come to little avail. Some have gone so far as to declare that inductive logic is ipso facto
valid because it is the necessary pillar of scientific inquiry and discovery, which is surely to
get the cart before the horse! This "problem of induction" is sometimes called Hume's
problem.
Popper was concerned to do two things: to find a demarcation between (good or true) science
and pseudoscience—a distinction that had both epistemological and ethical import for him—
and to solve Hume's problem. He worked first on the problem of demarcation, and concluded
that what separates true or genuine science from pseudoscience—here his favorite example of
genuine science was the work of Einstein, and of pseudoscience Marxism (he had ceased
being a Marxist by the beginning of his 20s) and psychoanalysis—is that genuine science
makes predictions that can be tested; it is not dogmatic. According to this falsificationist
criterion of demarcation, a theory is genuinely scientific if we can specify what would cause
us to reject it.
Popper’s ultimate discovery, he claimed, was that adopting falsification, instead of
verification, would solve both of those problems. "Only after some time did I realize that
there was a close link, and that the problem of induction arose essentially from a mistaken
solution to the problem of demarcation—from the belief that what elevated science over
pseudoscience was the "scientific method" of finding true, secure, and justifiable knowledge,
and that this method was the method of induction: a belief that erred in more ways than one."
Popper noticed a logical asymmetry between verification and falsification. Attempts at
verification rely on the invalid and discredited inductive logic. But falsification relies on the
deductive and wholly valid inference of Modus Tollens.
Modus tollens has the form:
• Premise 1: If P then Q.
• Premise 2: Q is false.
• Therefore, Conclusion: P is false.
In using this logical step in the testing of scientific theories, we begin with some purported
scientific theory or law or law-like statement, such as, for example, "Silver melts at 961.93
degrees Celsius." From that we derive an observation or data statement; in our example it
would be something like: "This sample of silver [that I am about to test] will melt at 961.93
degrees C." Then we perform the test. If it turns out that the silver in our test does not melt at
961.93 degrees C, then the original law-like statement or theory is falsified. If the sample
does melt at the predicted temperature, then, tentatively, the law-like statement has passed the
test and is not falsified. Popper never said that a test confirms a theory; he did not believe in
theory confirmation.
Popper described falsifiability using the following observations, paraphrased from a 1963
essay on "Conjectures and Refutations":
• It is easy to confirm or verify nearly every theory — if we look for confirmations.
• Confirmations are significant only if they are the result of risky predictions; that is, if,
unenlightened by the theory, we should have expected an event which was
incompatible with the theory — an event which would have refuted the theory.
• "Good" scientific theories include prohibitions that forbid certain things to happen.
The more a theory forbids, the better it is.
• A theory that is not refutable by any conceivable event is non-scientific. Irrefutability
is not a virtue of a theory.
• Every genuine test of a theory is an attempt to falsify or refute it. Theories that take
greater "risks" are more testable, and more exposed to refutation.
• Confirming or corroborating evidence is only significant when it is the result of a
genuine test of the theory; "genuine" in this case means that it comes out of a serious
but unsuccessful attempt to falsify the theory.
• Some genuinely testable theories, when found to be false, are still upheld by their
advocates—for example by introducing ad hoc some auxiliary assumption, or by
reinterpreting the theory ad hoc in such a way that it escapes refutation. Such a
procedure is always possible, but it rescues the theory from refutation only at the price
of destroying, or at least lowering, its scientific status.
Popper was adamant that passing a test does not verify a scientific theory or law. "I hold," he
wrote, "that scientific theories are never fully justifiable or verifiable, but they are
nevertheless testable." He held that all our scientific knowledge is tentative and uncertain,
and that we test only those things we think need testing. Our knowledge, even the most
strongly held scientific "law", is like "piles driven into a swamp," in his apt metaphor,
because it is always subject to further testing should we think there is good reason to do so.
We do rely, psychologically and tentatively, on those statements or theories that have passed
many tests, but this never guarantees that such a theory will not fail in the future. When we
step onto an airplane, for example—a new example, not Popper's—we know that this airplane
has flown safely in the past, and in that sense it has passed many tests of its airworthiness, but
that is never a guarantee that in the present case it will take us to our destination safely.
The process of attempted falsification, Popper held, solves Hume's problem because there is
no induction, but, instead, science develops through the process of scientists making bold
conjectures of theories and then testing them. He also held that theory proposal and theory
testing are completely separate. There is no logic of finding and proposing theories; it is, he
claimed, a mysterious and extra-logical process. But, he held, the process of testing a theory,
once it is proposed, is a fully logical and objective process.
Almost all commentators have criticized naïve or simple falsificationism, the view that a
simple instance of disagreement or failed test will falsify a theory. But Popper himself had
written in his first book Logik der Forschung:[9]
We must distinguish between falsifiability and falsification.... We say that a theory is falsified
if we have accepted basic statements which contradict it. This condition is necessary, but not
sufficient; for we have seen that non-reproducible single occurrences are of no significance to
science. Thus a few basic statements contradicting a theory will hardly induce us to reject it
as falsified. We shall take it as falsified only if we discover a reproducible effect which
refutes the theory. In other words, we only accept the falsification if a low-level empirical
hypothesis which describes such an effect is proposed and corroborated.
Popper himself, and others such as Imre Lakatos, later discussed additional problems that
show that simple falsificationism cannot succeed. First, it relied on a clear distinction
between theoretical and observational terms or language. But various people, including
Popper, N.R. Hanson, and others, showed that observation is theory-laden; there is no simple
observation apart from a theory. Second, the objections against verificationism had shown
that no proposition, including observation statements, could be proved decisively to be true
by experiment. So in order to make simple falsification work we would have to rely on an
experiment that would disprove the theory under investigation. But experiments cannot prove
that either. Third, every test of a scientific theory relies on a ceteris paribus clause. (Latin for
"all other things being equal." Also understood as "within the realm of margin of error" or
"we assume for the time being.") So an attempted falsification could—and would, in the case
of a theory thought to be true or good—be dodged by saying that what had been falsified was
not the central theory but some part of the ceteris paribus clause.
These observations led to a more subtle theory of falsification, usually called methodological
falsificationism, and to a discussion of what were called "metaphysical research programs,"
or "scientific research programs." These programs contain the theory, plus a number of
ceteris paribus clauses; one Popperian called this a theory plus its "protective belt." In this
view, falsification is always tentative. If something seems to be falsified, we first look at the
"protective belt" and try to modify it. Only as a last resort do we consider the theory itself to
be falsified.
One telling objection to this is that methodological falsificationism cannot itself be falsified.
There is no decision procedure to tell us when we should hold onto and when to give up and
abandon a theory; we can always hold onto a seemingly falsified theory by declaring that it is
not the theory but some part of the "protective belt" that has to go. Lakatos expressed it as "a
game in which one has little hopes of winning [but]... it is still better to play than give up." In
addition, there is little evidence from the history of science—Sir John Eccles being a notable
counterexample—of any science procedure or research program being based on falsification,
simple or methodological.
Popper attempted to solve this problem in later writings by proposing a sophisticated version
of falsificationism in which a theory T1 is falsified only if three conditions are satisfied: (1)
there exists a theory T2 that has excess empirical content over T1, meaning that it predicts
novel facts not predicted by T1; (2) T2 explains everything that was explained by T1, and (3)
some of the new predictions of T2 have been confirmed by experiment.
Thus in this sophisticated version of falsification a theory itself is not rejected by a
falsification. Instead, we compare one theory with another, and we do not reject a theory until
a better one comes along. So, Popper could claim, scientific knowledge grows through a
process whereby a better theory replaces a previous one. The notion of a crucial falsifying
experiment was discarded in favor of tests to decide between two or more competing
theories. Moreover, in this view the notion of proliferating theories—numerous alternatives—
becomes important. So science does not deal with theories and falsifiers of those theories, but
with rival theories.
But this takes us far from the original notion that the demarcation problem could be solved
through falsificationism. The original idea had been that there was a simple methodological
or logical method—sometimes called a syntactic method, as opposed to a semantic one—of
making the demarcation. But that was true only for simple falsification, and we already have
seen that simple falsificationism fails and cannot be maintained. But demarcation cannot be
made using sophisticated falsificationism either because, in the sophisticated falsification
model, we must make decisions, relying on convention or choice, between what constitutes
"real" knowledge and what constitutes "background noise," and between what counts as a
genuine falsification of the theory under consideration and what counts as just a nick in the
"protective belt." In sophisticated falsificationism we can make the mistake of rejecting a true
theory or the mistake of holding on too long and tenaciously to a bad or false theory, and the
sophisticated falsificationist criteria themselves give us no sure logical or methodological
means of showing just which of those we are doing in any particular instance.
Thus, what is science and what is pseudoscience depends in a central way on convention or
decision. If that is so, then there can be unresolvable disagreements about what is genuine
science and what is pseudoscience, and there is no fully reliable or methodological way of
making this demarcation.
Imre Lakatos (1922-1974), a Hungarian refugee, brilliant mathematician, and philosopher of
science, arrived in 1960 at the London School of Economics, where Popper was a professor.
In the beginning Lakatos was a close associate of Popper and Popper praised him highly. But
they had a falling out—after which Popper bitterly denounced Lakatos and he became one of
Popper's most devastating critics.
Popper had noted, and Lakatos agreed, that the probability of any theory based on evidence
for it is zero, because there are an infinite number of theories that may account for any set of
data. This can be seen by thinking of dots on graph paper—those dots representing individual
bits of data—and then asking how many different lines can be drawn connecting all those
dots. The answer is an infinite number. Thus the probability that any one of them is the
correct one is zero, so inductivism cannot lead us to the right or correct theory, except
through a chance that has a zero probability.
But, Lakatos noted, sophisticated falsificationism relies on conventionalism, and any
convention can be saved through use of ad hoc hypotheses. No methodological rule exists for
distinguishing decisively between genuine or necessary auxiliary hypotheses and
degenerating or useless ad hoc ones, so conventionalism does not demarcate at all. As Paul
Newall has written, "The problem for conventionalism is what to do with established
theories. Since these are accepted by convention, we seem drawn into the conclusion that
experiment can refute a new theory but not an old one - that is, the power (and relevance) of
empirical investigation seems to lessen the more science develops."[10]
Finally, Lakatos turned to deeper investigation of the ceteris paribus clauses that always
accompany any important theory. As Newall writes, "Lakatos was able to show via
examples ...[from the history of science] that the insistence on falsification would render all
theories unscientific. Popper tried to avoid this—desperately, at times—by implicitly
claiming for himself the right to decide whether an anomaly is serious or not, even insisting
in a television interview that Mercury's perihelion was not. What we see, then, is that
Popper's falsificationism degenerates into a version of authoritarian conventionalism, since he
could not avoid relying on the judgement of scientists (or, more often, his own...) as to when
we should consider a falsifier strong enough to bring to bear all the weight of moral authority
in demanding the theory's rejection and when to 'wait and see' instead."
The ultimate conclusion is that falsificationism too fails in providing a demarcation criterion,
and also fails in giving a general account of science and scientific discovery.
Thomas Kuhn
No book received more attention in post-World War II twentieth-century philosophy and
history and sociology of science than Thomas S. Kuhn’s The Structure of Scientific
Revolutions.[11] This was the one book that all philosophers, historians, and sociologists of
science had to deal with, either positively or negatively, or simply because it was the most
common.
Kuhn had come not from philosophy or scientific methodology, but from the history of
science. Kuhn was also a post-positivist in that he did not accept the positivist or
verificationist program, and he did not hold that science is a cumulative process. Instead, he
argued that science goes through several kinds of episodes. In the normal course, science is
organized around what Kuhn called a paradigm, and this paradigm guides ongoing work in
the field at hand. This period was called “normal science” by Kuhn. But gradually over time
there accumulates enough anomalies in that paradigm that the paradigm is deemed to be
inadequate and a revolutionary episode occurs in which that (old) paradigm is given up and
supplanted by a new one; Kuhn calls these episodes “revolutionary science.” Moreover, Kuhn
uses language much like what is used for religious conversions to describe the movement
from one paradigm to a new one. Frequently it is younger people who can accept and use the
new paradigm, while older ones are still married to the older pre-revolutionary paradigm and
thus cannot undergo the conversion. Instead, they die off.
Kuhn summarized his thesis this way, “scientific revolutions are … those non-cumulative
developmental episodes in which an older paradigm is replaced in whole or in part by an
incompatible new one.”[12] Paradigms are “accepted examples of actual scientific practice –
examples which include law, theory, application, and instrumentation together – [which]
provide models from which spring particular coherent traditions of scientific research.”[13]
Kuhn contributed to the breakdown of the view that observation could adjudicate between
two competing theories by denying that it is ever possible to isolate the theory being tested
from the influence of the theory in which the observations are grounded. He argued that
observations always rely on a specific paradigm, and that it is not possible to evaluate
competing paradigms independently. More than one such paradigm, or logically consistent
construct can each paint a usable likeness of the world, but it is pointless to pit them against
each other, theory against theory. Neither is a standard by which the other can be judged.
Instead, the question is which "portrait" is judged by some set of people to promise the most
in terms of “puzzle solving."
Several things must be noted about Kuhn’s view. First, he does not say that the new paradigm
is “closer to the truth” than the old one; in fact, it is difficult to see how he can speak about
“the truth” in any ultimate sense, even as a beacon toward which science aims. Second, his
view is very much sociologically embedded and determined; it is the scientific community
that accepts one paradigm and conducts normal science during that period, but then
undergoes revolution and thereby adopts a new paradigm, and, following that, a new episode
of normal science. Third, although he may or may not have wanted it this way, Kuhn’s
account of science is deeply conventional: Science is whatever the reigning community holds
it to be at any given time. But we must also note that, whatever conclusions they may have
drawn about it—and conclusions differed a great deal—all philosophers of science did attend
to Kuhn’s work at great length and it has provoked an immense outpouring of literature
discussing it, both pro and con, from philosophers, sociologists of science, historians of
science, and many others.
A major development in recent decades has been the study of the formation, structure, and
evolution of scientific communities by sociologists and anthropologists including Michel
Callon, Elihu Gerson, Bruno Latour, John Law, Susan Leigh Star, Anslem Strauss, Lucy
Suchman, and others. Some of their work has been previously loosely gathered in actor
network theory. Here the approach to the philosophy of science is to study how scientific
communities actually operate.
The turn toward consideration of the scientific community has become much more
pronounced in recent decades. One area of interest among historians, philosophers, and
sociologists of science is the extent to which scientific theories are shaped by their social and
political context. This approach is usually known as social constructivism. Social
constructivism is in one sense an extension of instrumentalism that incorporates the social
aspects of science. In its strongest form, it sees science as merely a discourse between
scientists, with objective fact playing a small role if any. A weaker form of the constructivist
position might hold that social factors play a large role in the acceptance of new scientific
theories.
Demarcation Again
The failure of the demarcation attempts by both the positivists-verificationists and the
falsificationists, led by Popper, leaves us with no methodological way of making any sharp
demarcation between science and non-science. Both the positivist-verificationists and the
falsificationists had attempted to provide a logical or quasi-logical methodology for making
such a demarcation. But that has now collapsed, although some people still attempt to cling to
the view that such a methodology exists, either through some form of verificationism or some
form of falsificationsim. But Philosopher Michael Ruse, among many others, has noted that:
“It is simply not possible to give a neat definition—specifying necessary and sufficient
characteristics—which separates all and only those things that have ever been called
‘science’. …The concept ‘science’ is not as easily definable as, for example, the concept
‘triangle’.”
Amherst College philosopher Alexander George has written:
…the intelligibility of that [demarcation] task depends on the possibility of drawing a line
between science and non-science. The prospects for this are dim. Twentieth-century
philosophy of science is littered with the smoldering remains of attempts to do just that. . . .
Science employs the scientific method. No, there's no such method: Doing science is not like
baking a cake. Science can be proved on the basis of observable data. No, general theories
about the natural world can't be proved at all. Our theories make claims that go beyond the
finite amount of data that we've collected. There's no way such extrapolations from the
evidence can be proved to be correct. Science can be disproved, or falsified, on the basis of
observable data. No, for it's always possible to protect a theory from an apparently confuting
observation. Theories are never tested in isolation but only in conjunction with many other
extra-theoretical assumptions (about the equipment being used, about ambient conditions,
about experimenter error, etc.). It's always possible to lay the blame for the confutation at
the door of one of these assumptions, thereby leaving one's theory in the clear. And so forth. .
. . Let's abandon this struggle to demarcate and instead let's liberally apply the label
'science' to any collection of assertions about the workings of the natural world. …what has
a claim to being taught in the science classroom isn't all science, but rather the best science,
the claims about reality that we have strongest reason to believe are true....
He continues,
Science versus non-science seems like a much sharper dichotomy than better versus worse
science. The first holds out the prospect of an 'objective' test, while the second calls for
'subjective' judgment. But there is no such test, and our reliance on judgment is inescapable.
We should be less proprietorial about the unhelpful moniker 'science' but insist that only the
best science be taught in our schools.[14]
This does not mean that “anything goes,” or that just anything can call itself science, despite
Paul Feyerabend’s claim in his book Against Method. Ruse, for example, went on to claim,
“Science is a phenomenon that has developed through the ages — dragging itself apart from
religion, philosophy, superstition, and other bodies of human opinion and belief.” But that is
a sociological and communal demarcation instead of a methodological one, dependent on
what the community of scientists deems, at any given time, to be science and what it rules out
as pseudoscience or non-science. That is not a fully satisfactory answer to those who want a
strong line or demarcation between science and pseudoscience because what we are left with
is conventionalistic instead of methodological, and the community may be wrong or biased or
otherwise unreliable. But it seems to be the best answer now available.
Sociologists of science have spoken particularly to this issue. Paul Thagard, for example, has
proposed a soft or sociological demarcation, as opposed to methodological one, based on how
a particular group of theory proponents act as a community. Their proposal is
pseudoscientific if:
• It is less progressive than alternative theories over a long period of time and faces
many unsolved problems;
• The community of practitioners makes little attempt to develop the theory towards
solutions of these problems;
• The community shows no concern for attempts to evaluate the theory in relation to
others;
• The community is selective in considering confirmations and disconfirmations; and
• The community of practitioners does not apply appropriate safeguards against self-
deception and known pitfalls of human perception. Those safeguards include peer
review, blind or double blind studies, control groups, and other protections against
expectation and confirmation bias.
But that set of criteria is itself questionable because it may not exclude things that many
scientists want to rule out. Whether those criteria would exclude astrology, for example, is
doubtful because for each point, astrologers could be found who take pains to satisfy it.
Science, Subjectivity, and Objectivity
Although they may have had different ways of attempting to accomplish it, one of the goals
of positivism, verificationism, and falsificationism was to assert and preserve the ideal of the
objectivity of science. But, in the face of devastating critiques of each, the collapse of the
programs of positivism, verificationism, and falsificationism, along with the rise of
sociological and communal accounts of science, seems to close the door to the objectivity of
science and open the door to subjectivism. This became especially the case once it was shown
that observation is theory laden and is something in which the observer is an active
participant, not merely a passive receptor; the observing subject thus influences the outcome
of the observation. In fact, some theorists, such as Feyerabend, have directly embraced
subjectivism and declared science to be no different in principle from any other human
subjective pursuit. That tack, however, has been resisted by others who still want to preserve
the ideal of objectivity of science, even though there is no methodological way to guarantee
objectivity.
One theorist who took a novel approach to this problem was Michael Polanyi, in his books
Personal Knowledge and The Tacit Dimension. Polanyi declared that observation is a gestalt
accomplishment, and that we always know more than we can say, so that the tacit dimension
of knowing is greater than what we make explicit. Thus the subjective component is essential
and central to perception and knowing. But, Polanyi held, we reach objectivity through that
subjectivity because of a personal commitment that accompanies any of our declarations that
something is true. By that, we make a commitment that what we know and declare to be true
is something beyond our mere subjective knowing. We are, in effect, recommending to others
when we declare something to be true that they also subjectively accept it. Unfortunately,
Polanyi did not get much attention from theorists of science, but his work was influential in
some other fields.
Most serious philosophers of science have deplored the retreat into subjectivism that has
come about from the collapse mentioned above, and have attempted to find an objectivity to
science even though this cannot be guaranteed by any methodological means or rubric. Larry
Laudan has expressed his view on the problem in this way:
Feminists, religious apologists (including "creation scientists"), counterculturalists, neo-
conservatives, and a host of other curious fellow-travelers have claimed to find crucial grist
for their mills in, for instance, the avowed incommensurability and underdetermination of
scientific theories. The displacement of the idea that facts and evidence matter by the idea
that everything boils down to subjective interests and perspectives is...the most prominent
and pernicious manifestation of anti-intellectualism in our time.[15]
He has also declared,
“Insofar as our concern is to protect ourselves and our fellows from the cardinal sin of
believing what we wish were so rather than what there is substantial evidence for – and surely
that is what most forms of ‘quackery’ come down to – then our focus should be squarely on
the empirical and conceptual credentials for claims about the world. The ‘scientific’ status of
those claims is irrelevant.[16]
Reductionism in Science
One of the goals of some theorists has been reductionism. Reductionism in science can have
several different senses. One type of reductionism is the belief that all fields of study are
ultimately amenable to scientific explanation. Perhaps an historical event might be explained
in sociological and psychological terms, which in turn might be described in terms of human
physiology, which in turn might be described in terms of chemistry and physics. The
historical event will have been reduced to a physical event. This might be seen as implying
that the historical event was 'nothing but' the physical event, denying the existence of
emergent phenomena.
Daniel Dennett invented the term “greedy reductionism” to describe the assumption that such
reductionism was possible. He claims that it is just “bad science,” seeking to find
explanations that are appealing or eloquent, rather than those that are of use in predicting
natural phenomena.
Arguments made against greedy reductionism through reference to emergent phenomena rely
upon the fact that self-referential systems can be said to contain more information than can be
described through individual analysis of their component parts. Analysis of such systems is
necessarily information-destructive because the observer must select a sample of the system
that can be at best partially representative. Information theory can be used to calculate the
magnitude of information loss and is one of the techniques applied by chaos theory.
Continental Philosophy of Science
Although the members of the Vienna Circle were from Continental Europe, the logical
positivists and their program became mostly an Anglo-American phenomenon, especially
with the coming of the Nazis and World War II. In the Continental philosophical tradition,
science has been viewed more from a world-historical perspective than from a
methodological one. One of the first philosophers who supported this view was Georg
Wilhelm Friedrich Hegel. Philosophers such as Ernst Mach, Pierre Duhem and Gaston
Bachelard also wrote their works with this world-historical approach to science. Two other
approaches to science include Edmund Husserl's phenomenology and Martin Heidegger's
hermeneutics. All of these approaches involve a historical and sociological turn to science,
with a special emphasis on lived experience (Husserlian "life-world" or Heideggerian
"existential" approach), rather than a progress-based or anti-historical approach as was done
in the analytic tradition.

SCIENCE AND TECHNOLOGY IN AFRICA’S DEVELOPMENT:


FROM KNOWLEDGE TO WISDOM.

In the excitement over the unfolding of his scientific and technical powers, modern man has
built a system of production that ravages nature and a type of society that mutilates man.
Such is the philosophy of exaggerated materialism, which we here challenge as inadequate
for the purpose of human development. This philosophy promotes a narrow vision of reality
and emphasises the accumulation of goods and the manipulation of techniques, to the neglect
of the development of persons. In the past, scientists could disclaim direct responsibility for
the use to which mankind had put their disinterested discoveries. We cannot today take the
same attitude because as Uchii, Soshichi (2000 says, “the success which we have achieved
in the development of nuclear power is fraught with infinitely greater dangers than were all
the inventions of the past." The destructive forces of the modern world cannot be brought
under control simply by mobilizing more resources of wealth and knowledge – to fight
pollution and chemical hazards, preserve wild life, to discover new sources of energy, and to
arrive at more effective agreements on peaceful co-existence. Wealth and knowledge are
needed for any civilization though, a revision of the ends and aims which these means are
meant to serve must be in accord with the logic of right knowledge; wisdom. This implies,
above all else, the development of a life-style, which accords to material things their proper
and legitimate place. The paper views human development in Africa as the progressive
humanization of society, a movement away from the materialistic, and towards the
humanistic. The paper argues in particular that, the logic of development expressed through
the creative application of knowledge should be shifted in favour of a wisdom that is
anchored on the aims of science and technical knowledge that takes cognizance of the
knowledge systems of the receiving community. Such is what is argued here as a development
paradigm for Africa of today.
Introduction
Advocates of scientific neutrality argue that, science as a body of knowledge has no
moral or ethical quality substantially, value judgements, cultural biases and that, political
standpoints do not in any way influence or determine scientific knowledge. They argue
further that, there is nothing ‘good’ or ‘bad’ about scientific knowledge. Such position is
acknowledged by the great Galileo himself that “the conclusions of natural science are true
and necessary, and the judgement of man has nothing to do with them” (Joan Lipscombe &
Bill Williams 1979:6).
While acknowledging the quality and weight of such informed position, it suffices to
say that this position is a contradiction in terms for the simple reason that the pursuit of
knowledge in itself, which aim the scientist claim is the province of science, is in itself a
good thing. This inherent implication of the scientist’s claim is perhaps more reasonably
understood in the language of Black (1975) who draws a distinction between the pursuit of
knowledge as information and knowledge as understanding. He points out that the collection
of information in itself is a product of value judgements. Better still, human interaction has it
on record that science (or at least its application) could be a power for good or evil.
Indeed, the interaction of science and technology considered in the last chapter attest
to this fact; science has been seen as the means of relieving human burdens, and this, and not
the disinterested pursuit of knowledge, has often motivated scientists. It is perhaps this idea
of science and technology and its impact on man and society, and the consequences of such
impact that Bertrand Russell provocatively remarked that in, discussing the effects of science
upon human life we have therefore three more or less separate matters to examine (1) the
nature and scope of scientific knowledge (2) the increased power of manipulation derived
from scientific technique (3) the changes in social life and in traditional institutions which
must result from the scientific technique demands. This chapter concerns itself with the
second and third matter.
The understanding here is that, man, aided by science and technology, has the
capacity to make or mar a world of his choice. Two issues arise from this. The first is whether
the world of man’s choice may be the best possible world and, or, the most desired world for
the greatest number of people. The second issue is whether such a choice is a free one or can
be a free one that is blame worthy. Bertrand Russell argues that:
In so far as he is wise this new power is beneficent; in so far as he is foolish it is quite the
reverse. If therefore, a scientific civilization is to be a good civilization it is necessary that
increase in knowledge should be accompanied by increase in wisdom. I mean by wisdom, a
right conception of the ends of life. This is something science itself does not provide.
Increase in science itself, therefore, is not enough to guarantee any genuine progress though
it provides one of the ingredients which progress requires (Russell 1962:ix-x).
The implication of this thinking is that science and technology is a mixed blessing.
Such an explosive impact has far reaching consequences which, according to Jim Unah
(1998:344), “potend good and bad for man; consequences that spell good and evil for society;
consequences that snatched humankind out of the cruel forces of nature and yet threaten them
with collective suicide”. It means for us against this backdrop that science and technology
have both demonstrated that they constitute a double-edged sword, if man is wise in the use
of the instruments of his brains and hands, he would conquer nature and make it subserve his
essential interest. If, on the other hand, he becomes foolish, he would wipe out human
existence and the entire earth with its habitation.
Such is the nature of man that he can be described as a bundle of paradoxes, a being
empowered by God to create itself thus “you shall have the power to degenerate into the
lower forms of life, which are brutish. But you shall also have the power, out of your soul’s
judgement to be reborn into higher forms which are divine” (Ehusani, 1991:16).
This chapter is a critical exposition of the impact (positive and negative) of science
and technology on the human society. It argues that the phenomenal technological
advancement notwithstanding, our new world has seen “the emergence of the machine and
the disappearance of the person”. I.e. science and technology have both healed as well as
killed the society.
The Neutrality of Science and Technology?
The concept of neutrality associated to any human activity suggests an inherent
quality of perfection. In relation to science and technology, the neutrality theory argues for
itself the omniscience, which suggests and elevates the scientist (and to some extent the
technologist) to the role of a high priest expounding its truths. Whether such dogmatic
posturing is true or not, accepted or rejected, this theory grants to itself the self-contained
completeness of knowledge – truths especially when science is considered in the context of
what is normally defined as pure and applied science.
Sir Ernest Chain more clearly states the thesis of scientific neutrality thus: …science
as long as it limits itself to the descriptive study of the laws of nature, has no moral or ethical
quality, and this applies to the physical as well as the biological sciences (1970).
This position is traditionally inherent in scientific thinking more so that science seeks
to ascertain the truth about nature, which hypotheses which aim to move nearer and nearer to
an accurate description of natural laws, which are seen as universal truths. Such thinking is
also anchored on the fact that objective reasoning cannot deny scientific facts and all
scientists must inevitably reach the same conclusion.
It means, then, as Joan Lipscombe and Bill Williams (1979:6) posited, that “value
judgements, cultural biases or political standpoints do not in any way influence or determine
scientific knowledge. There is nothing ‘good’ or ‘bad’ about scientific knowledge”. Such
understanding of science has been carried forward and is strongly supported today which
perhaps may have informed the thought of Bronowski who attributes to science “an
unrelenting independence in the search for truth that pays no attention to received opinion or
expediency or political advantage (Bronowski, 1971:25).
The neutrality of technology unlike science does not very well find convenient
application. Indeed, there is no way in which we can talk about “the pursuit of knowledge for
its own sake” or the objectivity of observations, experiments and theory as applied to
technology, for it necessarily implies the application of science, invention and industry and or
commerce to matters which are of importance to our life style and must, therefore, have a
social effect. Notwithstanding such position, technology is, undoubtedly, commonly
regarded as being neutral in some senses of the word. Considered as a collection of
machines, techniques and tools, technology is here said to be neutral in the sense that in itself
it does not incorporate or imply any political or social values, and that it is neither ‘good’ nor
‘evil’.
Taken, therefore, as a blameless tool, any beneficial or harmful effect is said to arise
out of the motives of the people applying a particular piece of technology and the end to
which it is used. It means, then, that where a particular application, chosen for its beneficial
results, produces harmful side-effects, these are blamed either on inadequate social policies or
on lack of sophistication in the control of the effects of technology. Whichever is chosen as a
whipping boy, concludes Joan Lipscombe and Bill Williams (p. 19), technology itself is
‘neutral’.
But the most challenging question is, “to what extent is science and technology
neutral?” The question of the neutrality of science and technology is essentially the question
of the rationality of science and technology. This is perhaps where the essential link between
science and technology very clearly bears on man in his integral whole, in both his material
and spiritual life, but more so in the spiritual towards which the material must serve. Suffice
to say here that the argument in support of “an unrelenting independence [of science and
technology] in the search for truth that pays no attention to received opinion or expediency or
political advantage” is an exercise in the promotion of ignorance and scepticism. To quote
Andrew Efemini:
Anyone with scientific consciousness, understands the place of science in man’s struggle to
improve his living… (science is not) something that should be pursued for its own sake but
something that should be pursued for man’s benefit (Efemini, 1982:18).
It thus means that, traditionally, practical knowledge i.e. techne which is concerned
with making (recta ratio factibilium) directed to the perfection of the object of knowledge,
combines with theoretical knowledge i.e. scientia or episteme comprising also contemplation
of nature, which goal is the perfection of the subject (the knower) to bring about the ultimate
end in the perfection of the whole man. Such an endeavour is a conscious and goal oriented
one, which not only reflects the value systems of the society at that time, but are value laden
in themselves.
Granted that science is a move towards the unknown according to which “it is
impossible to foresee the practical results of any research in pure science”, it is neither a blind
move nor a goalless move. Matthew Nwoko aptly suggests here that:
At least a scientific research worthy of the name must be a planned venture. Even if the
scientist does not foresee the remote consequences of his venture, but the planned structure
of his work carries or must carry an ultimate intention of discovery for the good of man
(Nwoko, 1992:143).
It is, thus, the inherent vocation of the scientist to lay bare the richness of nature,
which practical use the technologist will bring to bear for the good of man. This is the
rationality of scientific inquiry, and such is the rationality of technological practice.
Understood as such, both the scientist and the technologist are said to be humanists who
“must not only reach out to the world’s wealth of knowledge and practice, but must also
pursue the solution of our problems (of industrial, manufacture, environmental pollution,
economic progress etc) with dedication, conviction and patriotism (Newswatch Feb. 12,
1990:14).
Furthermore, to argue that science is unaffected by extraneous factors, which
pontification justifies scientific neutrality, is an overstatement to say the least. The dialectics
of science is intertwined with theological, ethical, ideological and other non-scientific
arguments, which at some points become impossible to separate them, and stand-points on
reality were determined by considering all these aspects. R. M. Young (1971:31) thus, argues
that, “what people were prepared to accept as the ‘truth’ was not determined by science
alone” but also by subtle and often un-acknowledged influence of social factors.
The deliberate suppression of scientific knowledge or the active promotion of
particular theories, which conform to a specific political situation, similarly counts against the
neutrality theory of science. A ready example, here, is the Lysenko affair in Russia in which
a whole area of genetics was eliminated from Russian teaching and his theories imposed
because they were more supportive of the political system. Russian scientists worked within
the framework of these theories believing them to be ‘true’, at least as far as the existing
evidence was concerned (D. W. Caspari and R. E. Marshak 1965:275-278). The case of
Jeremiah Abalaka, a Nigerian, is another example in which the scientist’s search for the truth
and or scientific knowledge is substantially tempered with to bolster up the Nigerian/foreign
interest. Pursued to a logical conclusion, and in the extreme case, ‘scientific facts’ (if
Abalaka succumbs) may be the invention of a political regime instead of results from
disinterested pursuit of knowledge for its own sake.
Again, proponents of the neutrality theory say that science concerns itself purely with
a description of the world as it is, and so argue out the impossibility of scientific knowledge
giving rise to normative and evaluative statements. Arguably it cannot give rise to statement
about what should or should not be (normative), nor can it pass judgement on what is good or
bad (evaluative). Indeed orthodox philosophical argument has it that the only valid
conclusions of deductive arguments are the ones which contain only material which is
already in the premises, consequently scientific premises (factual) cannot lead to normative
or evaluative statements (Lipscombe and Williams 1979:8).
But, this argument collapses because of the difficulty in identifying which premises
are factual – normative statements, it is argued could be expressed in the same way as factual
ones, and there are considerable difficulties in clearly distinguishing one from the other.
Black thus asserts that:
Some normative evaluative propositions are objective (generally accepted and not subject to
individual values) and this removes the distinction which separates scientific propositions
from others (1975:40).
It is, thus, possible for science to provide factual statements that could lead to
normative or evaluative statements. An example of this could be:
Plant defoliants can cause food shortage (factual)
Food shortage lead to people starving (factual)
It is wrong that people should starve directly because of man’s action (Normative)
Therefore, plant defoliants should not be used (Normative)
The base of our argument, here, is that the scientist has a social responsibility for the
application of his work. This is informed by the logic of distinction between the abstract
concept of ‘science’ which argued position is the pursuit of knowledge for its own sake, from
the practical manifestation of that concept. This is science in the context of an overall
activity. Black mentions that:
Science as an overall activity can no longer be considered as the disinterested pursuit of
truth. Even where scientists are working on the purest science, which has no apparent
applications, scientists cannot escape the dilemma of responsibility because the speed of
development is such that discoveries are often harnessed very quickly to industrial, military
or other practical uses. (Black, 1975:40)
Besides, much of today’s pure research is consciously directed at serving specific
objectives and or solving some problems. Black argues further that:
It can no longer be considered neutral and is carried out with a definite purpose in mind: to
increase the profits of industry or strengthen the power of government. Scientists involved in
such projects know this and because the science is no longer neutral they have forfeited any
claim to moral neutrality; they cannot subsequently plead ‘not guilty’ when this purpose is
achieved and horror (or praise) is expressed at the results (ibid).
We may, perhaps, argue further that such state in which science has found itself, of
developing and applying its results to specific objectives, thus, removes it out of this old
argument of scientific neutrality (as for example the work in plasma physics carried out
specifically with the aim of generating electricity from nuclear fusion). In such
circumstances, there is no realistic way of separating basic research from its application and,
hence this harnessing of science to specific ends implies the end of scientific neutrality and
with it the end of any legitimate claim to moral neutrality.
W. H. Ferry convincingly that “technology has a career of its own, so far not subject
to the political guidance and restraints imposed on other enormously powerful institutions”
(Ferry, 1971:120). Obviously, such conclusion sounds like the distant drums of science which
pays no regard as to whether people accept the ‘truth’ it claims to have as part of its very
nature, or not. At best, such assumed neutrality of technology has arisen because certain
characteristics associated with science have, unjustifiably, been transferred to technology
hook, line and sinker. Richkover more correctly presents this scenario when he says:
A certain ruthlessness has been encouraged by the mistaken belief that to disregard human
consideration is as necessary in technology as it is in science (Richcover, 1965:154 ).
But such conclusion is founded on illicit premise. Technology understood as an art or
skill, entails in its essence the employment of means to accomplish some end: opposed to
nature, which in itself is a product of the rational faculty. Essentially, technology thus means
a set of principles, or rational method, in the production of something or in the achievement
of an end. By its nature, technology is not and cannot be neutral, because human needs and
values remain its essential ingredients. It is, perhaps, this conclusion that the German
language (especially in the philosophical usage) explains the term Technik as the utilization
of the knowledge of method or mode of production of material goods to serve human needs.
That is, technology by its nature is determined by the society. In the words of Dickson:

In general we can say that a society’s technology, when viewed as a social institution rather
than a heterogeneous collection of machines and tools, is structured in such a way that it
coincides with its dominant modes of action and interaction… Technology does not just
provide in its individual machines, the physical means by which a society supports and
promotes its power structure, it also reflects, as a social institution this social structure in its
design. A society’s technology can never be isolated from its power structure, and
technology can thus never be considered politically neutral (Dicson, 1974:25).

This explains the fact of our being dominated by technology and which our generation
has seen “the emergence of the machine, and the disappearance of the person”. The reason
for this, according to Dickson, is the political nature of technology. This dominating
technology, he says reflects the wishes of the ruling class to control their fellow men.
Looking back into history, we cannot but agree with this simple but thought provoking
truism. The very process of industrialization, for example, did not arise from an objective
assessment of production needs determined by economic factors. It arose from the desires of
the dominant social class, the providers of capital, to dominate and control both nature and
work force. Consequently, these set of values and desires were built into the design of the
machines and factories which benefits are confined to the members of a particular social
class.
To understand properly this argument of the neutrality of technology, it suffices to
clarify the distinction between science and technology. While the work which the scientists
do varies considerably along the spectrum from pure research to applied technology, that of
the technologists is concerned almost exclusively with developing and implementing specific
ideas with a definite end in mind. In putting this issue into proper perspective therefore, two
questions come to mind, namely, the question of the intended product of the work of the
technologist: how far, if at all, should the technologist make judgement about the desirability
or otherwise of the end product in considering whether or not to apply his technical skill to a
particular project? And the question which arises from the unpredicted harmful or
undesirable consequences which often arise from the application of a particular technology:
how far is the technologist ‘innocent’ or responsible for such consequences?
The answers to these questions are not far-fetched. It is argued here that, the
professional status of technologists makes them culpable for the work of their hands. If they
are truly professionals, they have a responsibility to relinquish their neutral role and to take
steps to limit the harmful consequences of their works. As professional automotive engineers
worth their salt, for example, they have the capacity to construct cars that may reduce road
deaths and injuries, high noise levels, congestion, pollution and despoliation of the
countryside among other harmful consequences. To argue the opposite view that
technologists are not in any way responsible for the intended consequences of the use of their
product is to create an artificial distinction between responsibility for the development of a
product and responsibility for the use of that product. Such distinction, is simply a question
of conscience which does not find relevance in this consideration. Einstein was here clear on
this issue when he forcefully writes: “we scientists whose tragic destination has been to help
in making the methods of annihilation more gruesome and more effective, must consider it
our solemn and transcendent duty to do all in our power in preventing these weapons from
being used” (Time, December 1999:59).
More unacceptable is even the question of unforeseen consequences of technological
invention and development. The question put in context is, is it still acceptable for the
technologist to plead innocence when his device intended for human benefit turns out to do
more harm than good? The answer in this regard is No! Hardin (1972 especially chapter 7) is
vehement on this score and, thus, introduces the concept of guilty until proven innocent,
suggesting further that this should be applied to all technical development. Professor C. S.
Momoh canvasses a similar idea in his “Philosophy and Moral Scientism”, according to
which all scientific and technical inventions are allowed to play out their effects in the
scientists’ laboratory, and those with harmful consequences disallowed to see the light of
application. He says:
For any scientific invention to be worth its salt, its consequences and purpose for mankind
and humanity must be seen to be moral… the simple test is: will the application of scientific
invention or discovery advance the moral worth of mankind? If ‘Yes’ such a discovery
should be developed and embraced. If the answer is ‘No’, such an invention should be left to
cool away in the laboratory. (Momoh, 2000:82)

The concern in both thinking is that, the burden of proof of both the effectiveness and
harmlessness has been placed on the proponent.
All this boils down to the fact that the technologist (and technology in general) is
clearly denied any shelter behind the neutrality shield. Technology is not and should no
longer be seen as, a neutral tool. It should be assumed to be harmful until proven otherwise.
While accepting that such action on the part of man is likely to delay benefits and so limit the
maximization of human creativity, it serves as a call for the critical examination, and re-
examination of the product of man’s mind and hands so as to confirm them as rational action
which they truly are, to serve man better and maximally.
Technology as part of the human culture, is for the good of man and his society,
which basis it is judged as valuable, as having proper meaning. M. I. Nwoko (1992:136)
confirms this lucidly as he says; So genuine technology draws its values from the good it
serves man. Actually, the goal of technology is its service to man to help him realize more
his being.
It is to be said, in conclusion that both science and technology have been mis-
presented over and over as neutral endeavours. Our postulations have proved this to be false
cultures seeking relevance in history. Indeed, science and technology in social context reveal
to us living and astonishing testimonies; that once generated by human culture, science or
technology in turn becomes a determinant factor in the social transformation, a
transformation which may be good or evil, beneficial or destructive, depending on the
manner of its appraisal in the society. Genuine science or technology is a system of rational
endeavour (action), which in itself presupposes organization of all the elements of the
endeavour (or work). Mudimbe and Appiah both acknowledge this intellectual attitude to
collapse the concept of scientific (and technological) neutrality. Hear them; scientists, like
the rest of us, hold on to theories longer than may be justified to: they suppress unconsciously
or half-consciously or consciously… evidence they do not know how to handle: lie a little
(Oluwole 1999:34).
As activities of our ingenious minds and hands, science and technology are not and
cannot be value-free. Hence, the burden of proof (of evidence of marked injury to man)
should lie on the man who wants to introduce any change (or scientific or technological
breakthrough) before the change or the breakthrough will proceed for public use. The
complicated planet, inhabited by more than a million and half species living together in a
more or less balanced equilibrium, in which they use and re-use the same “facilities” cannot,
and should not be improved by aimless and uninformed tinkering. Thus argued, “all changes
in a complex mechanism involve risk and should be undertaken only after careful study of all
the facts available. Changes should be made on a small scale first so as to provide a test
before they are widely applied. When information is incomplete, changes should stay close
to the natural processes which have in their favour the indisputable evidence of having
supported life for a very long time” (Schumacher 1979:130-131). Some two decades ago, the
emergence of nuclear energy was astonishingly progressive, and promised salvation and
solution to human production related problems. Today, the same product is at damnation best
described as a “hazard with a hitherto inexperienced ‘dimension’, endangering not only those
who might be directly affected by their radiation but their offspring as well”.
It is, thus, a disastrous abdication of moral responsibility for scientists or technologists
to ignore the consequences of their inventions or who resist questions of societal need and
engage in wilful blindness that cannot lead to the good of all, humans and the environment.
Whether it is science or technology, it concerns a body of organized rational modes of
theoretical foundation or making, involving also the analysis and judgement of the value
orientation of the patterning of the action whereby resources are mobilized for the attainment
of the common (greater) good.
Human Decline
Science and technology are two modes of human activity that are organised around
interaction with nature. Such interaction is neither random nor casual, but conscious and goal
oriented, which character derives from the need to understand nature in its diverse structures
and patterns of working. But, even within these two modes of human activity, there exists a
symbiotic relationship; science provides information to technology, and technology in turn
provides science with ingenious precision instruments, which extend the scope of human
sources of knowledge and also provides avenues for practical utilization of scientific theories.
It is, therefore, not an exaggeration to argue on this score that science and technology
have reached a pyramidal stage in our time, most probably because the scope of human needs
has attained weird and bizarre dimensions. This has led men and women into many
inventions and discoveries. Starting from the very humble beginning with the use of
railways, carriages and cars, then steamships and the invention of airplanes, man conquered
the land, water and the airspace. Today, man has penetrated outer space thus enabling
engineers, technicians and scientists to explore and exploit the outer space for limited periods
of time and to return to earth with the product of their effort. Today mankind has started to
make active use of outer space for its own purposes. Artificial satellites orbiting the earth are
employed to relay television programmes, transmit communication over long distances
forecast the weather, discover deposits of minerals and so on
These are clearly phenomena advancements in science and technology which qualify
the twentieth century to be described as the ‘fastest’ century in human history and in which
the human being of today could also rightly claim to be the most mobile homo-sapiens that
ever existed. Thus, the global scientifico-technological development has several
characteristic features though, one most important element in each of the great technological
breakthroughs is that each can be used to further the progress of man. No doubt then that
humanity is in common agreement that scientific technology has brought many good things
to man and society, which perhaps has elicited peacock’s conclusion that “in spite of
scepticism in certain circles about the long-term effects of science (and technology), our
contemporary world is a world of science (and technology) in the sense that science (and
technology) is/are generally thought to command the dominating heights of the cultural
landscape” (Peacocke, 1987:3), and shaping the outlook of mankind everywhere and
everything positively.
Put together therefore, the breakthroughs in science sourced from the advent of the
miletians culminating in the achievements of the 20th century homo faber in the fie4lds of
technological medicine, food and agricultural technology, communication and information
technology and even ammunition technology have altered the lives of humanity positively.
We now posses the tools to fulfil the creator’s charge to “be fruitful, multiply, fill the earth
and subdue it” (Gen. 1:28). In Africa, such western technological endeavours have changed
the once “dark continent” and opened it up to benefit from the most glamorous of human
civilizations in the areas of economics, education, medical care, communication and industry
among others. For,
without the possibilities offered by modern science and technology, life would be impossible
for many. The weak could become extremely vulnerable, since they would be unequipped to
deal with an otherwise hostile and unyielding nature (Ehusani, 1991:7).

The twentieth-century is arguably the most dehumanised, the ‘bloodiest’ in history as


evidenced in the internationalisation of military technology through the use of atomic bomb
that devastated the Japanese cities of Hiroshina and Nagasoki in 1945., which consideration
shall be our next concern.
Scientific and technological discoveries, sometimes assume international/
transnational properties, thus, impacting on peoples and cultures other than those of the
source of its origin. When this happens, such scientific or technological inventions are
internationalised and, thus, assume within it a quality of interdependence. As a globalized
endeavour, scientific technology entails mobility of the world technology. Generally
understood, globalization is mobility of the world economy, capital etc. It means simply a
shift from one land to another, and from one owner to another, everyday, of goods, people,
information and ideas (of a scientific or technological, political or economic nature).
Such an understanding portrays globalization as an ideology which postulates portend
innumerable advantages among which include “freedom, by which it implies, liberation of
the market relations from whatever fetters, including state regulation and even state
boundaries, with the usual seasoning of talks about democracy, civil society, human rights
and other attributes ascribed to the society of the western countries” (Alexei Vassiliev,
1999:2).

The question of grave concern is whether the African, European or Westerner,


concerned for his own problems and aspirations needs this information, and to what extent.
Professor Alexei Vassiliev adds his voice here, that, even if he needs a part of it, he is its
consumer, not producer. He asks in particular, “what about the common Africans, who have
to consume the spiritual food that is of little or no use to them? (Vassiliev, 1999:17) It means
here that such transposition of cultures via the use of global information systems help to
corrupt other values and since Africa is at the receiving end, it argues against her that the
African cultural heritage is affected most. A statement of the Catholic Bishops Conference of
England and Wales fully exposes the negative effect of the mass media that;
While Britain continues to enjoy standards of broadcasting which are rightly admired
elsewhere, those standards cannot be taken for granted. There is for instance, a constant
drift towards more screen violence greater use of obscene language and ever more explicit
depictions of intimate sexual activity (The Common Good, Feb 1997:25).

The argued position here is that, globalization opens new chances for developing
countries (but African countries in particular), but involves new challenges and dangers. Most
of what is given to the public as objective knowledge and or scientific technique portends
health risk or utter destruction due largely to public ignorance on the part of the consumers,
or greed on the part of the inventors. The American Journal of Medicine is reported to have
given a clean bill of health to cellullar phones (most obviously influenced by greed
economics), but recent studies in 1999 by Joseph Kallol has established a relationship
between the use of cellullar phones and brain damage. A Science News Programme
monitored on VOA (January 27, 2001) argues that there exists a genetic damage resulting
from the use of cellular phones, and that in children cells and bones tissues are affected most
due to radiation.
Understandably, it cannot be argued that broadcasters are merely responding to
changes in public taste, as they play a major part in shaping that taste. Information
technology, specially the hi-tech of the western world creates and maintains its market share
which accompanying foreign cultures have downward taste implication for Africans rather
than upwards, thus, resulting in the decay of public standards, dehumanisation of the mass
population and decline in regard for the common good.

Revolution in informatics and communication no doubt thinned the world and made
the world a truly global village, it has also fairly made knowledge a universal commodity,
but, it has also rubbished human dignity in this universal beneficence. This picture is more
eloquently stated by Professor Alexei Vassiliev thus:
The western media impose their own problems, their own world vision, their own system of
values, their own ethical and religious approaches to the Africans to whom they are totally
alien by and large… The inflow of show business and mass culture from the west breaks the
earlier ideas, distorts the population’s system of values and life orientation. It implants the
consumer ideology, sexual licentiousness, violence, worship of the golden calf, material
success at whatever price (Vassiliev, 1997:17).
Nothing could be added more than a statement of the fact that, combined with mass
(foreign) culture and advertisement, the mass media both western and African, dictate
people’s taste and behaviour, form their political, economic, religious and social likes and
dislikes, inculcate evaluations of events and facts. Our university campuses and urban towns
have become show rooms in this regard. It is not an exaggeration to say here that when the
television demonstrates an imaginary event, it becomes a fact that affects reality, though
haven happened in the visual world. Such is the state of mental slavery which modern
technical endeavour has afflicted on developing nations, but Africa in particular, which is yet
unable to overcome.
The paradox here is that, though surrounded by innumerable spiritual forces and
littered with abundant resources and fertile land-mass, and endowed with the capacity to turn
its many rivers into good and clean drinking water, Africans still look to foreign lands to
satisfy their religious ecstasy, she still depend on genetically engineered food in form of aids
to fed its teaming and ever growing population, and struggling to deal with the AIDS,
migration, orphan-hood and refugee problems resulting from terrorist and ethnic conflicts.
While all these are problems which should engage the intelligent mind of the African,
he still overlooks such problems in place of foreign television or internet programmes which
inspire his interest most, either to watch football, Michael Jackson or to source for more
information about the latest fashion style or video CD from the Hollywood, even when clean
drinking water is not available within his immediate environment.
The suffocating impact of this digital information technology on the developing world
(but Africa in particular) is best described by Ngugi Wa Thiongo as a Cultural Bomb which
effect is to annihilate a person’s belief in his cultural heritage and ultimately in himself. He
says:
The effect of this cultural bombs is to annihilate a person’s belief in their names, in their
language, in their heritage of struggle, in their unity, in their capacities and ultimately in
themselves, including their thought process (Wa Thiongo, 1986:4).

Similar attempt to promote balanced and professional use and development of


information technology in the areas of economics and information management has been far
reaching. While the development of information and communication technology has brought
into being products of global standard that have the capacity to solve human problems, as is
the case with computer application, it has on the other hand compounded many a human
problems.
Through the electronic banking, both buyers and suppliers are assisted to reduce risk,
and increase speed without much trouble of going through cashiers, and in the comfort of
their rooms or offices. Designed using Public Key Infrastructure (PKI) technology,
customers and buyers in the banking business are assisted via a device (electronic transaction
services) offered by COMPAQ, to validate their transaction via a major bank in a matter of
seconds. Said to be anchored on trust and security, this innovation; the e-commerce is said to
be supported by more than 10 of the world’s largest banks. The Post Express PC clinic
attractively presents its benefits thus:
Compaq indentures solution also includes key technology products and services to allow
financial institutions to seamlessly integrate existing applications with a new web enabled
front end handling business banking services, such as trade payments factoring and cash
flow management (The Post Express, April 6, 2000).
This is, no doubt, an achievement of no mean measure. It is, however, another thing
entirely to argue in its favour as to whether the human person for whom the service is meant
desires it, or his essential purpose is served. In all honesty, science or technical device
achieves its aim most when it serves man not when it helps to dehumanise him. The
technique of Compaq indentures – ready solution for verification infrastructures, which
functions where, and when man should, takes over man’s usefulness and relevance, and thus
renders him useless and irrelevant in the world, and without human value. Such
consideration, it is our argued position breads self-destruction.
Besides, such human tendency to “force things to appear”, kills rather than heals
humanity. Human nature is in harmony when it acts in the satisfaction of its essential needs.
Such it is that, it is bad culture to say that “we must live to eat”, instead of saying “we must
eat to live”. It must be said here that much of today’s goods and services which result from
our intelligent minds and works of our hands are least needed. They result from the
unhindered access to information technology, which creates in us a culture of consumerism.
Today therefore “people live to eat even when they are not hungry; they live to consume, to
produce more in order to consume more. All these, according Pope John Paul II (1988:33),
constitute the intrinsic contradiction of a development limited only to its economic element.
Such development, he says easily subjects the human person and his or her deepest needs to
the demands of economic planning and selfish profit.
It remains to be seen what has become of Africa and Africans, in, and out of this
global phenomena. Commentators have said that the present situation is much more serious.
They have argued unanimously that Africa’s civilization and cultural identity is in danger at
the turning point of the epoch. Globalisation, understood as internationalisation of interaction
of both human and material resource portend universal blessing with unaccountable wealth
though, still remains a distant social reality in Africa. In all areas, whether health care and
medicare, agriculture and industry, military technology or the information and
communication technology, the west remains a dominant overlord whose mass culture
destroys the achievement of other cultures. Thus, the dreamed global culture was and
remains an ideological fiction of neoliberalism rather than a social reality. It has advanced
the criminal and immoral interest of the west more than it has promoted material welfare,
spiritual upsurge, health and access to education, and human right and dignity.
According to Pope John Paul II, western countries “have betrayed their democratic
principles and are moving towards totalitarianism, and democracy has become a mere myth
and a cover for immorality” (John Paul II, March 1995). Perhaps the Pope was speaking
against the backdrop of injustices and dehumanising legacies of the west, which brought
more curses than blessings to Africans. In Africa, and for Africans, globalisation of the
world economy in relation to all technical endeavours, pushed ahead by the forces which
demand national boundaries be opened for trade and movement of capital, scientific
knowledge, technology and information and communication techniques, proved useful to
some people and marginalized much more other people, augmented inequality both within
the nations and among them. In what particular ways, and to what extent Africa and Africans
have been affected by the modern globalizing world is the question that shall be answered in
the next segment of our discussion.

The African Predicament


Western science and sophisticated techniques have no doubt made an explosive and
tremendous impact on human society. Such explosive impact has far reaching consequences;
consequences that spell good and evil for society; consequences that snatched humankind out
of the cruel forces of nature and yet threaten them with collective suicide. The expression
Paradox of Western Science and Technology means here that the many advances of science
and technical endeavours have improved the quality of human life, and are destroying life as
well. This increasingly paradoxical dimension of western science and technology most
obviously endears Einstein’s comment that, “concern for man himself and his fate must
always form the chief interest of all technical endeavours… in order that the creations of our
mind shall be blessing and not a curse to mankind” (Nwagwu, March 26, 1998:22).
But such right conception of the ends of life, which itself is wisdom seems not to part
of the defined project of western science, for while there is tremendous increase in
knowledge in scientific civilization, it is not accompanied by increase in wisdom. Today
therefore, our scientific civilization kills, destroys and dehumanises largely, hence science
and technology have demonstrated that they constitute a double-edged sword. In their
paradoxical dimensions, they have assumed more purposeful and purposeless, more
meaningful and bizarre, more useful and destructive, and while achievements in the enclave
of science and technology have served to prolong life, they have also served to provide
resources through which the foolish application of scientific techniques, man would be
exterminated from the surface of the earth.
These bundle of paradoxes of science and technology are said to be part of the
mysterious nature of man himself, the homo faber, who, within his same nature, has the
power to degenerate into lower forms of life which are brutish, but who also has the power to
be reborn into the higher forms which are noble. Thus, with the conquest of land, sea, and
outer space, science not only offers a one-dimensional image of the person, but also presents
the human person with the temptation to and self-deification, self destruction to the detriment
of the divine nature of man. One is thus inclined to assert the paradoxical nature of the
human person that, ours is an age that is marked by embarrassing contrast between the
spectacular scientific and technological achievements on the one hand, and a shameful
degradation of the human person on the other.
It is perhaps this unprecedented achievement of science in modern times that has
lured the African into the fool’s paradise, where he or she is unable to recognise his or her
nature as a paradox. Though richer materially, Africans have become morally and spiritually
poorer. It must be made clear here that, though efficiency and speed form the index of the hi-
technology which the African too is a beneficiary, the African continent is yet to experience
progress and development which are said to accompany such technology. In Africa, abject
poverty cohabit with stinking wealth, there is the phenomenon of starvation on the one hand,
and what is referred to as ‘influenza’. The new scientific and technological values have
destroyed the African humanistic value system. Today, Africans are no longer their brothers’
keepers. Though we live within the world of enormous wealth and unbridled luxury of a few,
a greater majority are experiencing utter poverty. The western scientific and technological
mindset has further destroyed African values which today allows humanity to destroy or to
squander goods that other people need for their lives. Ehusani very vividly captures this neo-
African spirit thus:
whereas the “Structural Adjustment Programme” embarked upon in the last few years by
African countries including Nigeria, has condemned thousands of the peasant population to
death by hunger and starvation, African millionaires have multiplied their ranks, as is
evident from the swelling account of individual Africans in European and American banks,
the number of ultra-modern mansions now springing up alongside shock and thatch huts in
African towns and villages, and the fleet of expensive foreign cars which now ply the scarcely
paved roads that run through these towns and villages. (Ehusani, 1991:11)

It may be said, here, that the atomistic view of western scientific technology and its
reductionist view of reality has encouraged and promoted collective selfishness of one class
of people against another, thus, reducing a vast segment of humanity to the culture of the
ghetto, making them more vulnerable to diseases and epidemics, drug addiction, crime and
countless social and psychiatric problems. At best, the legacy of western scientific and
technological civilization for the African could be summarised in what Thoreau says is an
“improved means to an unimproved end” (King Jr. 1968:172). The men and women of
Africa have been empowered with every technique of information and communication, yet
they remain unschooled ignoramus in the experience of communion. We have perfected and
erected bureaucratic structures where communication thrives and communion is nonexistent;
and so “every improvement in communication makes the bore more terrible” (De Marco,
1982:61). It is no wonder that today traditional African society experiences intense
loneliness and alienation, and the modern city dweller suffers while in the midst of many, and
city life in Africa is, millions of people being lonesome together.
These paradoxes of human nature, but in particular, those which prevails in Africa,
appears to inform the thinking of Vatican II when it says that; “there appears the dichotomy
of a world that is at once powerful and weak, capable of doing what is noble and what is
base, disposed to freedom and slavery” it continues: Man is growing conscious that the
forces, he has unleashed are in his own hand and that it is up to him to control them or be
enslaved by them. Here lies the dilemma (Vatican II, 1973:105).
The African predicament lies truly in this dilemma in the fact that the creations of
man (science and technology), have been more of a curse than a blessings, while Africans
never cease to speak of noble ideas, they watch the continent, as it were, helplessly
degenerating in humanity. The many wars, terrorist activities, ethnic and religious clashes in
which sophisticated weapons are freely used attest to this. Dr. King Jr. exposes what could
be described as the true African situation that, the African Heads of State continually issue
calls for world peace, yet, “they come to the peace table accompanied by bands of brigands
each bearing unsheathed swords”. On leaving the disarmament talk table, they go directly to
launch latest nuclear missiles (King Jr. p. 182).
Perhaps, the most devastating blow to the soul of Africa is located in the nineteenth
century when most of Africa was colonized by various European powers. The several years
of colonial experience sapped the African heritage, which involve both material exploitation,
cultural expropriation and anthropological impoverishment. Though highly certificated in the
disciplines of western thought and knowledgeable in the technique of the west, the African
suffers gross ego distortion. In general terms, the African continent has become the most
bastardised and misused continent, and they themselves have been milked of their self-
confidence. In one word, they have been dehumanised.
Perhaps the account of an American journalist reveals the African experience in more
greater details.
The colonialists left behind some schools and roads, some post offices and bureaucrats. But
their cruellest legacy on the African continent was a lingering inferiority complex, a
confused sense of identity. After all, when people are told for a century that they’re not as
clever or capable as their masters, they eventually believe it (Lamb 1986:140).

The implication here is that, the clashes of the two world views; western,
macrocosmic “superior” new world meant a displacement of the smaller “inferior” old order,
in place of which the new western “superior” order that succeeded it became a disaster. In
the language of Chinua Achebe, “the ‘Whiteman’ has indeed put a knife on the things that
held Africans together and they have fallen apart”. Western scientific and technological
civilization thus means for Africa, the collapse of a whole vision of life, of all beliefs, of
every authority, the loss for a people of their identity, i.e. the collapse of African humanistic
heritage.
This neo-technical culture has engendered wars and terrorist activities, tribal and
communal clashes in which lethal weapons are freely used. Today too, our towns and cities
are being brutally terrorised by armed robbers, hired assassins, thuggery and banditry.
African citizens have become prisoners in their homes, with high walls, iron bars and metal
gates. Africa has become a battle ground in which everybody is fighting everybody. Ehusani
captures this ugly scenario most vividly:
Thirty-years after independence of most African nations, not one of them is yet to boast of
political stability. As one country launches a return to democracy another reverts to military
dictatorship; as one country begins a national reconstruction after a bitter civil war, another
declares the onset of a religious war, and as the workers of one country return to work after
a period of total strike, the students of another country go on the rampage. Africa now
records the highest number of refugees, most of whom are not being displaced by natural
disasters but are rather on the run from totalitarian regimes, military dictators or rural
ethnic militia. It is a continent in turmoil (Ehusani, 1991:20).

The question to be answered here is whether the loss of humanity by Africans has got
something to show for it. Africans want scientific knowledge and technical know-how.
Though they have traded out their humanism, they have not been able to gain what they want;
scientific technology. They have lost their humanity, and so have become children without
heirs and so slaves of the creations of their minds.
This scenario is best described by K. C. Anyanwu as the crisis of science which
entails the crumbling of man’s beliefs, assumptions, and ideas about reality, a situation that
portends grave consequences on human conduct.
It means that reality no longer fits into our presuppositions about it, and this crisis has
profound consequences on our conduct… It means that we are no longer able to determine
the direction of change, to control events and to know how we are related to the world
(Anyanwu, 1983:70).

Perhaps, this state of affairs of science means also a crisis of perception i.e a condition
which prevents humanity from having a holistic view of reality that would enable it to
organise its actions positively; to determine the line between the permissible and the
forbidden, order and disorder, so as to deduce the principles of human association and
determine the standard of our values. But in particular, the crisis of western science is
founded on the mistaken assumption that there are absolute authorities in cultural modes of
thought, and that the Europeans and, or the west are dictators in this regard, who must lord it
over the rest of the world. In human situations, it must be said, all our cravings for truth, all
our disputes about knowledge and quarrels about conscience are cultural activities or cultural
quests, and they have all arisen from our desires as human beings to fulfil ourselves. This,
Macneile Dickson argues, is why “all reasoning is in a manner biased, and the bias is due to
the nature, surroundings and education of the thinker”. He posits further that,
There are in the realm of thought no absolute authorities, no dictators. No man, living or
dead, can claim oracular powers… All philosophies are in the end personal… systems of
thought are the shadows cast by different races, epochs, and civilizations (Dickson,
1958:13).

It is, perhaps this attempt by to superimpose western civilization and knowledge


systems and modes of thought that brings about the crisis in science which frontal offensive
has produced today the destruction of the African states, cultural, economic and social
institutions alongside its local elites, who are either destroyed or integrated into the western
(or international system), but who have lost their responsibility for their nation, all in the
name of progress.
Thus far, the negative impact of science and technology does not any longer portray
them as unified explanatory pattern of the world. The UNESCO science report says this much
when it states that, “faith in science that it makes clarity on all and makes all the universe
intelligible, according to a coherent order of the causes and effects has lost its strength”.
(UNESCO, Science Report, 1996:214).
Hans Zehrer is more forceful and vivid in his argument that not only has science
brought about the collapse of the European worldview, his world is undergoing
reconstruction. He says convincingly that:
Not only are we in the midst of a crisis of science today, but we have come to the end of that
scientific attitude which dominated the epochs of modern Europe…. We can lay it down that
scientific attitude which began to establish itself upon the Greek model at the period and
which determined the achievements and successes of that historical period was faced with a
reality in face of which it gives up, and in face of which its methods prove ineffectual; and we
begin to grasp that this attitude of mind has played out its role and can attain no more
success. If science be understood to mean what occurred within this epoch, then science is
played out; we are at the end of it today. (Zehrer, 1952:257)

Surely, the situation is worse in Africa, a continent which is outside the scientific
culture of the west. While there are still more discoveries and breakthroughs, the crises in
science still persist and human consciousness would not grasp their realities. Science has
power and knowledge, but lacks wisdom to use the power and knowledge properly. The
issue here is that, the basic assumptions about reality, the principles of its understanding, its
worldview, its methods and standards have collapsed. So, science and technology which are
said to be architectonics of progress are, themselves no longer regarded as sources of benefits
to humanity. In reality, they are the causes of new forms of evil variously expressed in
degradation of the environment, effects on human health, the dehumanising and the
robotizing of society, the deepening of social and political inequalities. Put paradoxically,
modern science, having endowed man with unrivalled powers of transformation of the world
has, at the same time, conferred on him an unrivalled potential for the destruction of the
planet. The human being has the capacity for good as well as for evil, for hate and conflict, as
well as for love and co-operation. In the present chaotic world of technology and mass
culture, these mixed qualities of humanities have been too freely exercised that the individual
too often feels lost and meaningless.

From Knowledge to Wisdom


The primary official intellectual aim of academic inquiry is to help humanity learn how to
solve its global problems and make progress towards a better world. The acquisition of
knowledge and technological know-how is thus said to be the primary way in which
academia can help humanity make progress towards a better world and applied to help solve
social problems. In order to be of benefit to humanity, academia must not just ensure that
authentic, objective, reliable knowledge is acquired, it must be the primary goal of science
and technology to advance the common good of all, in the first instance at least, to solving
problems of knowledge, so that knowledge that is acquired can, subsequently, be used to help
solve social problems of living. There can be no doubt whatsoever that the scientific pursuit
of knowledge has, over the centuries, helped transform the human condition, and has brought
immense benefits to our whole way of life. The modern world is quite simply inconceivable
without modern science. Nevertheless, the pursuit of knowledge dissociated from a more
fundamental concern with problems of living – with our global problems – as demanded by
knowledge-inquiry, despite the benefits that have resulted, has also had profoundly damaging
consequences. It has resulted in all our current global problems, including the lethal character
of modern terrorism.
In short, not only does the current devotion of academia to the pursuit of knowledge and
technological know-how prevent universities from taking their primary task to be to educate
the public about what our global problems are, and what we need to do about them. Even
worse, this immensely successful pursuit of knowledge dissociated from a more fundamental
concern with global problems of living is actually implicated in the creation of our current
global problems. In a perfectly reasonable sense of "cause", it could be convincingly said
that, our global problems have been caused by modern science and technology. In short, if by
the cause of an event we mean that prior change which led to that event occurring, then it is
the advent of modern science and technology that has caused all our current global crises. It
is not that people became greedier or more wicked in the 19th and 20th centuries; nor is it
that the new economic system of capitalism is responsible, as some historians and economists
would have us believe. The crucial factor is the creation and immense success of modern
science and technology. This has led to modern medicine and hygiene, to population growth,
to modern agriculture and industry, to world wide travel (which spreads diseases such as
AIDS), to global warming, and to the destructive might of the technology of modern war and
terrorism, conventional, chemical, biological, nuclear.
9/11 is a striking case in point. There is nothing exclusively modern about terrorism itself,
any more than there is about war: terrorism goes back at least to Biblical times. But what is
distinctively modern is the scale of the threat, and its impact. Those responsible for 9/11 used
nothing more high-tech than knives, but they were able to exploit modern technology so as
vastly to increase the enormity of their action, and the scale of its impact. They exploited
aeroplanes with which to do the deed, and relied on television and modern communications
to spread news and images of what they had done round the world instantly, as the horror
unfolded. It was modern technology which made the immediate global impact of 9/11
possible.
Before the advent of modern science and technology, lack of wisdom – lack of the capacity to
resolve our problems of living intelligently and humanely – did not matter too much. We
lacked the power to do too much damage to ourselves, or to the planet (although some
damage we did do). But now that we (or some of us) possess unprecedented powers, thanks
to modern science, lack of wisdom has become a menace. Humanity urgently needs to learn
how to solve its problems more intelligently and humanely than it has done up to the present,
and for that, as I have said, we urgently need to develop public institutions of learning
rationally designed and devoted to achieving this goal.
But how is this to be done? Who could get academics to agree to transform the whole
academic enterprise in the way that is, it seems, required? What guidelines could there be for
creating a kind of inquiry rationally devoted to promoting wisdom? Might not the whole
endeavour be a disaster, in that the only outcome would be the undermining of the
objectivity, the intellectual integrity, of science, and thus its human value? Is it not an absurd
over-reaction to cry for the transformation of academia so that the public may be better
educated about the problems of the world? Is it not hopelessly utopian to think, in any case,
that it is possible for humanity to learn wisdom?
Correcting the Blunders.
A perfectly acceptable answer to these questions stares us in the face. And yet it is one that
almost everyone overlooks. Modern science has met with astonishing success in improving
our knowledge of the natural world. It is this very success, as we have seen, that is the cause
of our current problems. But instead of merely blaming science for our troubles, as some are
inclined to do, we need, rather, to try to learn from the success of science. We need to learn
from the manner in which science makes progress towards greater knowledge how we can
make social progress towards a better, wiser world.
This is not a new idea. It goes back to the Enlightenment of the 18 th century, especially the
French Enlightenment. Voltaire, Diderot, Condorcet and the other philosophers of the
Enlightenment had the profoundly important idea that it might be possible to learn from
scientific progress how to achieve social progress towards an enlightened world. They did not
just have the idea: they did everything they could to put the idea into practice in their lives.
They fought dictatorial power, superstition, and injustice with weapons no more lethal than
those of argument and wit. They gave their support to the virtues of tolerance, openness to
doubt, readiness to learn from criticism and from experience. Courageously and energetically
they laboured to promote reason and enlightenment in personal and social life. And in doing
so they created, in a sense, the modern world, with all its glories and disasters.
The philosophers of the Enlightenment had their hearts in the right place. But in developing
the basic Enlightenment idea intellectually the philosophers, unfortunately, blundered. They
botched the job. And it is this that we are suffering from today. The philosophers thought that
the proper way to implement the Enlightenment Programme of learning from scientific
progress how to achieve social progress towards an enlightened world is to develop the social
sciences alongside the natural sciences. If it is important to acquire knowledge of natural
phenomena to better the lot of mankind, as Francis Bacon had insisted, then (so, in effect, the
philosophes thought) it must be even more important to acquire knowledge of social
phenomena. First, knowledge must be acquired; then it can be applied to help solve social
problems. They thus set about creating and developing the social sciences: economics,
psychology, anthropology, history, sociology, political science.
This traditional version of the Enlightenment Programme, despite being damagingly
defective, was immensely influential. It was developed throughout the 19th century, by men
such as Saint-Simon, Comte, Marx, Mill and many others, and was built into the intellectual-
institutional structure of academic inquiry in the first part of the 20 th century with the
creation of departments of the social sciences in universities all over the world.
Academic inquiry today, devoted primarily to the pursuit of knowledge and technological
know-how, is the outcome of two past revolutions: the scientific revolution of the 16th and
17th centuries which led to the development of modern natural science, and the later
profoundly important but very seriously defective Enlightenment revolution. It is this
situation which calls for the urgent need to bring about a third revolution to put right the
structural defects we have inherited from the Enlightenment.
But what, it may be asked, is wrong with the traditional Enlightenment Programme?
Almost everything. In order to implement properly the basic Enlightenment idea of learning
from scientific progress how to achieve social progress towards a civilized world, it is
essential to get the following three things right.
1. The progress-achieving methods of science need to be correctly identified.
2. These methods need to be correctly generalized so that they become fruitfully applicable to
any worthwhile, problematic human endeavour, whatever the aims may be, and not just
applicable to the one endeavour of acquiring knowledge.
3. The correctly generalized progress-achieving methods then need to be exploited correctly
in the great human endeavour of trying to make social progress towards an enlightened, wise
world.
Unfortunately, the philosophers of the Enlightenment got all three points wrong. And as a
result these blunders, undetected and uncorrected, are built into the intellectual-institutional
structure of academia as it exists today. Academia today is, in other words, the outcome of a
botched attempt to learn from scientific progress how to make social progress towards a
better world.
First, the philosophes failed to capture correctly the progress-achieving methods of natural
science. From D’Alembert in the 18 th century to Popper in the 20 th, the widely held view,
amongst both scientists and philosophers, has been (and continues to be) that science
proceeds by assessing theories impartially in the light of evidence, no permanent assumption
being accepted by science about the universe independently of evidence. But this standard
empiricist view is untenable. If taken literally, it would instantly bring science to a standstill.
For, given any accepted scientific theory, T, Newtonian theory say, or quantum theory,
endlessly many rivals can be concocted which agree with T about observed phenomena but
disagree arbitrarily about some unobserved phenomena. Science would be drowned in an
ocean of such empirically successful rival theories if empirical considerations alone
determined which theories are accepted, which rejected.
In practice, these rivals are excluded because they are disastrously disunified. Two
considerations govern acceptance of theories in science: empirical success and unity. But in
persistently accepting unified theories, to the extent of rejecting disunified rivals that are just
as, or even more, empirically successful, science makes a big persistent assumption about the
universe. Science assumes that the universe is such that all disunified theories are false. The
universe has some kind of unified dynamic structure. It is physically comprehensible in the
sense that explanations for phenomena exist to be discovered.
But this untestable (and thus metaphysical) assumption that the universe is comprehensible is
profoundly problematic. How can we possibly know that the universe is comprehensible?
Science is obliged to assume, but does not know, that the universe is comprehensible. Much
less does it know that the universe is comprehensible in this or that way. A glance at the
history of physics reveals that ideas about how the universe may be comprehensible have
changed dramatically over time. In the 17 th century there was the idea that the universe
consists of corpuscles, minute billiard balls, which interact only by contact. This gave way to
the idea that the universe consists of point-particles surrounded by rigid, spherically
symmetrical fields of force, which in turn gave way to the idea that there is one unified self-
interacting field, varying smoothly throughout space and time. Nowadays we have the idea
that everything is made up of minute quantum strings embedded in ten or eleven dimensions
of space-time. Some kind of assumption along these lines must be made but, given the
historical record, and given that any such assumption concerns the ultimate nature of the
universe, that of which we are most ignorant, it is only reasonable to conclude that it is
almost bound to be false.
The way to overcome this fundamental dilemma, inherent in the scientific enterprise, is to
construe science as making a hierarchy of metaphysical assumptions concerning the
comprehensibility and knowability of the universe, these assumptions asserting less and less
as one goes up the hierarchy, and thus becoming more and more likely to be true. In this way
a framework of relatively insubstantial, unproblematic, fixed assumptions and associated
methods is created within which much more substantial and problematic assumptions and
associated methods can be changed, and indeed improved, as scientific knowledge improves.
Put another way, a framework of relatively unspecific, unproblematic, fixed aims and
methods is created within which much more specific and problematic aims and methods
evolve as scientific knowledge evolves. (A basic aim of science is to discover in what precise
way the universe is comprehensible, this aim evolving as assumptions about
comprehensibility evolve.) There is positive feedback between improving knowledge, and
improving aims-and-methods, improving knowledge-about-how-to-improve-knowledge. This
is the nub of scientific rationality, the methodological key to the unprecedented success of
science. Science adapts its nature to what it discovers about the nature of the universe. ( see
Maxwell, 1998; 2001, chapter 3 and appendix 3; and 2004, chapter 1 and 2 and appendix;
2007, chapter 14).
Second, having failed to identify the methods of science correctly, the philosophers naturally
failed to generalize these methods properly. They failed to appreciate that the idea of
representing the problematic aims (and associated methods) of science in the form of a
hierarchy can be generalized and applied fruitfully to other worthwhile enterprises besides
science. Many other enterprises have problematic aims; these would benefit from employing
a hierarchical methodology, generalized from that of science, thus making it possible to
improve aims and methods as the enterprise proceeds. There is the hope that, in this way,
some of the astonishing success of science might be exported into other worthwhile human
endeavours, with aims quite different from those of science.
Third, and most disastrously of all, the philosophers failed completely to try to apply such
generalized progress-achieving methods to the immense, and profoundly problematic
enterprise of making social progress towards an enlightened, wise world. The aim of such an
enterprise is notoriously problematic. For all sorts of reasons, what constitutes a good world,
an enlightened, wise or civilized world, attainable and genuinely desirable, must be
inherently and permanently problematic. Here, above all, it is essential to employ the
generalized version of the hierarchical, progress-achieving methods of science, designed
specifically to facilitate progress when basic aims are problematic.
Properly implemented, in short, the Enlightenment idea of learning from scientific progress
how to achieve social progress towards an enlightened world would involve developing
social inquiry as social methodology, or social philosophy, not primarily as social science. A
basic task would be to get into personal and social life, and into other institutions besides that
of science – into government, industry, agriculture, commerce, the media, law, education,
international relations – hierarchical, progress-achieving methods (designed to improve
problematic aims) arrived at by generalizing the methods of science. A basic task for
academic inquiry as a whole would be to help humanity learn how to resolve its conflicts and
problems of living in more just, cooperatively rational ways than at present. This task would
be intellectually more fundamental than the scientific task of acquiring knowledge. Social
inquiry would be intellectually more fundamental than physics. Academia would be a kind of
people’s civil service, doing openly for the public what actual civil services are supposed to
do in secret for governments. Academia would have just sufficient power (but no more) to
retain its independence from government, industry, the press, public opinion, and other
centres of power and influence in the social world. It would seek to learn from, educate, and
argue with the great social world beyond, but would not dictate. Academic thought would be
pursued as a specialized, subordinate part of what is really important and fundamental: the
thinking that goes on, individually, socially and institutionally, in the social world, guiding
individual, social and institutional actions and life. The fundamental intellectual and
humanitarian aim of inquiry would be to help humanity acquire wisdom – wisdom being the
capacity to realize (apprehend and create) what is of value in life, for oneself and others,
wisdom thus including knowledge and technological know-how but much else besides.
One important consequence flows from the point that the basic aim of inquiry would be to
help us discover what is of value, namely that our feelings and desires would have a vital
rational role to play within the intellectual domain of inquiry. If we are to discover for
ourselves what is of value, then we must attend to our feelings and desires. But not
everything that feels good is good, and not everything that we desire is desirable. Rationality
requires that feelings and desires take fact, knowledge and logic into account, just as it
requires that priorities for scientific research take feelings and desires into account. In
insisting on this kind of interplay between feelings and desires on the one hand, knowledge
and understanding on the other, the conception of inquiry that we are considering resolves the
conflict between Rationalism and Romanticism, and helps us to acquire what we need if we
are to contribute to building civilization: mindful hearts and heartfelt minds.
Another outcome of getting into social and institutional life the kind of aim-evolving,
hierarchical methodology indicated above, generalized from science, is that it becomes
possible for us to develop and assess rival philosophies of life as a part of social life,
somewhat as theories are developed and assessed within science. Such a hierarchical
methodology “provides a framework within which diverse philosophies of value – diverse
religions, political and moral views – may be cooperatively assessed and tested against the
experience of personal and social life. There is the possibility of cooperatively and
progressively improving such philosophies of life (views about what is of value in life and
how it is to be achieved) much as theories are cooperatively and progressively improved in
science. In science diverse universal theories are critically assessed with respect to each
other, and with respect to experience (observational and experimental results). In a somewhat
analogous way, diverse philosophies of life may be critically assessed with respect to each
other, and with respect to experience – what we do, achieve, fail to achieve, enjoy and suffer
– the aim being so to improve philosophies of life (and more specific philosophies of more
specific enterprises within life such as government, education or art) that they offer greater
help with the realization of value in life” (Maxwell, 1984, p. 254).
All in all, if the Enlightenment revolution had been carried through properly, the three steps
indicated above being correctly implemented, the outcome would have been a kind of
academic inquiry very different from what we have at present. We would possess what we so
urgently need, and at present so dangerously and destructively lack, institutions of learning
well-designed from the standpoint of helping us create a better, a wiser world.
We have travelled far from our initial topic, the disastrous “war on terrorism”. And yet, the
transformation in our instruments of public learning that I have (briefly) argued for, are
highly relevant to our capacity to deal effectively and humanely with terrorism. What our
initial discussion of the eight principles that need to be observed in combating terrorism
revealed is that, again and again, the current “war on terrorism” is achieving the very
opposite of what was intended. Terrorism is being actively promoted, even implemented, not
contained and curtailed. The aim of combating terrorism, like so many other aims in life, is
inherently problematic. If we do not proceed intelligently, learning from past mistakes, it is
all too likely that we will achieve the very opposite of what we seek. Hence the fundamental
importance of a kind academic inquiry, a kind of learning, which emphasizes the need to
subject problematic aims to sustained criticism and improvement.
It would be absurd, of course, to argue that we need to transform academia so that we can
learn how to combat terrorism intelligently. That is not what I have argued. Rather, my claim
is that international terrorism is one of a number of global problems that confront us and that,
if we are to tackle these problems intelligently, humanely and democratically (as we must
do), people quite generally must have a much better understanding of what these problems
are, and what needs to be done about them, than they do at present, this in turn requiring a
kind of inquiry rationally designed to promote such public education about our problems, this
in turn requiring a revolution in our schools and universities. Learning how to tackle
terrorism more intelligently would be a beneficiary along with learning how to tackle more
intelligently our other global problems.
The reasons for the revolution in the aims and methods of inquiry according to Nicholas
Maxwell are not only humanitarian, they are also absolutely decisive intellectual reasons.
The kind of inquiry that would emerge – wisdom-inquiry, Both in his view are more rigorous
intellectually, and of greater human value, than what we have at present. The revolution is
needed in the interests both of the intellectual and the practical aspects of inquiry.
But will it happen? I first spelled out the argument over thirty years ago (Maxwell, 1976). It
was spelled out again, in very much greater detail, in my second book (Maxwell, 1984). This
received many excellent reviews, in particular a glowing review from Christopher Longuet-
Higgins in Nature, who remarked, during the course of his review, “ Maxwell is advocating
nothing less than a revolution (based on reason, not on religious or Marxist doctrine) in our
intellectual goals and methods of inquiry ... There are altogether too many symptoms of
malaise in our science-based society for Nicholas Maxwell's diagnosis to be ignored"
(Longuet-Higgins, 1984). Unfortunately it has been ignored. With agonizing slowness, in a
wholly piecemeal and confused fashion, some changes have taken place in science, and in
academia more generally, that are somewhat in the direction that I have argued for, but in
complete ignorance of my argument (and often masked by other changes that take things in
the opposite direction): see Maxwell (2007, chapters 6 and 12); see, also, Iredale (2007).
Academia is supposed to be about innovation but, when it comes to the rules of the game,
dogmatic conservatism tends to take over. It is difficult, too, to arouse public interest in the
current damaging irrationality of academia. In the popular mind, “academic” is almost
synonymous with “irrelevant” or “pointless”. That judgement is part of the problem.
On the other hand, the revolution that we need might be compared in significance to the
Renaissance, to the scientific revolution, or to the 18th century Enlightenment. Intellectual
revolutions as profound and far-reaching as these do not happen overnight. Thirty years of
inaction, when the matter is viewed in that light, is perhaps not such a long time interval.
But the question that haunts us today is: Given the state of the world today, given the
enormity of the problems that face us, can humanity afford to put off any longer creating
institutions of learning rationally designed to help us discover how to tackle our problems in
wiser, more cooperatively rational ways? To such and many other similar questions, Claude
Levi-Strauss, (1997:45) answers,
The more knowledge makes progress, the more it understands why it cannot come to
anything. Whenever we have the feeling to make progress in knowledge we see that it raises
other problems… knowledge becomes convinced of its disability (Claude Levi-Strauss,
1997:45)
The point in question here is that, Science and technology are beneficial in many ways
though, they have conferred on man the power to destroy himself and the environment in
which he finds himself. The question still remains as to whether the loss of humanity by
Africans could be so regenerated through more knowledge. It is argued here that, all is not
lost. Nicholas Maxwell’s From Knowledge to Wisdom which argues for a movement away
from knowledge to wisdom provides a genuine insight into the motives and character of
human knowledge and comes up with sensible proposals as to how the problems generated by
science should be tackled. It is an exemplary contribution to wisdom-inquiry; to create a kind
of inquiry rationally designed to help humanity learn how to create and or remake a better
world.
1. There needs to be a change in the basic intellectual aim of inquiry, from the growth of
knowledge to the growth of wisdom — wisdom being taken to be the capacity to realize what
is of value in life, for oneself and others, and thus including knowledge, understanding and
technological know-how.
2. There needs to be a change in the nature of academic problems, so that problems of living
are included, as well as problems of knowledge. Furthermore, problems of living need to be
treated as intellectually more fundamental than problems of knowledge.
3. There needs to be a change in the nature of academic ideas, so that proposals for action are
included as well as claims to knowledge. Furthermore, proposals for action need to be treated
as intellectually more fundamental than claims to knowledge.
4. There needs to be a change in what constitutes intellectual progress, so that progress-in-
ideas-relevant-to-achieving-a-more-civilized-world is included as well as progress in
knowledge, the former being indeed intellectually fundamental.
5. There needs to be a change in the idea as to where inquiry, at its most fundamental, is
located. It is not esoteric theoretical physics, but rather the thinking we engage in as we seek
to achieve what is of value in life.
6. There needs to be a dramatic change in the nature of social inquiry (reflecting points 1 to
5). Economics, politics, sociology, and so on, are not, fundamentally, sciences, and do not,
fundamentally, have the task of improving knowledge about social phenomena. Instead, their
task is threefold. First, it is to articulate problems of living, and propose and critically assess
possible solutions, possible actions or policies, from the standpoint of their capacity, if
implemented, to promote wiser ways of living. Second, it is to promote such cooperatively
rational tackling of problems of living throughout the social world. And third, at a more basic
and long-term level, it is to help build the hierarchical structure of aims and methods of aim-
oriented rationality into personal, institutional and global life, thus creating frameworks
within which progressive improvement of personal and social life aims-and-methods
becomes possible. These three tasks are undertaken in order to promote cooperative tackling
of problems of living — but also in order to enhance empathic or “personalistic”
understanding between people as something of value in its own right. Acquiring knowledge
of social phenomena is a subordinate activity, engaged in to facilitate the above three
fundamental pursuits.
7. Natural science needs to change, so that it includes at least three levels of discussion:
evidence, theory, and research aims. Discussion of aims needs to bring together scientific,
metaphysical and evaluative consideration in an attempt to discover the most desirable and
realizable research aims.
8. There needs to be a dramatic change in the relationship between social inquiry and natural
science, so that social inquiry becomes intellectually more fundamental from the standpoint
of tackling problems of living, promoting wisdom.
9. The way in which academic inquiry as a whole is related to the rest of the human world
needs to change dramatically. Instead of being intellectually dissociated from the rest of
society, academic inquiry needs to be communicating with, learning from, teaching and
arguing with the rest of society — in such a way as to promote cooperative rationality and
social wisdom. Academia needs to have just sufficient power to retain its independence from
the pressures of government, industry, the military, and public opinion, but no more.
Academia becomes a kind of civil service for the public, doing openly and independently
what actual civil services are supposed to do in secret for governments.
10. There needs to be a change in the role that political and religious ideas, works of art,
expressions of feelings, desires and values have within rational inquiry. Instead of being
excluded, they need to be explicitly included and critically assessed, as possible indications
and revelations of what is of value, and as unmasking of fraudulent values in satire and
parody, vital ingredients of wisdom.
11. There need to be changes in education so that, for example, seminars devoted to the
cooperative, imaginative and critical discussion of problems of living are at the heart of all
education from five-year-olds onwards. Politics, which cannot be taught by knowledge-
inquiry, becomes central to wisdom-inquiry, political creeds and actions being subjected to
imaginative and critical scrutiny.
12. There need to be changes in the aims, priorities and character of pure science and
scholarship, so that it is the curiosity, the seeing and searching, the knowing and
understanding of individual persons that ultimately matters, the more impersonal, esoteric,
purely intellectual aspects of science and scholarship being means to this end. Social inquiry
needs to give intellectual priority to helping empathic understanding between people to
flourish (as indicated in 6 above).
13. There need to be changes in the way mathematics is understood, pursued and taught.
Mathematics is not a branch of knowledge at all. Rather, it is concerned to explore
problematic possibilities, and to develop, systematize and unify problem-solving methods.
14. Literature needs to be put close to the heart of rational inquiry, in that it explores
imaginatively our most profound problems of living and aids personalistic understanding in
life by enhancing our ability to enter imaginatively into the problems and lives of others.
15. Philosophy needs to change so that it ceases to be just another specialized discipline and
becomes instead that aspect of inquiry as a whole that is concerned with our most general and
fundamental problems — those problems that cut across all disciplinary boundaries.
Philosophy needs to become again what it was for Socrates: the attempt to devote reason to
the growth of wisdom in life.

Conclusion
This is the revolution we need to bring about in our traditions and institutions of learning, if
they are to be properly and rationally designed to help us learn how to make progress towards
a wiser world.

SOCRATES, HEIDEGGER AND HUSSERL ON THE ETHICS OF SCIENTIFIC


TECHNOLOGY AND SUSTAINABLE ENVIRONMENT.

Introduction
“Technology and its instruments are appreciated not as extensions of man’s physical
faculties but as participating in his intellectual insight with its spiritual values” (Mclean,
1984:11).
Technology like any rational work of man has as its effect the achievement of
the destiny of man, which destiny includes the good and happiness of man. So, it is the fruit
of both the spiritual and material life of man. Today however, the interplay of science and
technology stands in great confusion and increasingly assuming paradoxical dimensions,
more purposeful and purposeless, more meaningful and bizarre, and more useful and
destructive. While the achievements in science and technology have served to prolong life,
they have also served to provide resources for its brutal extermination. Science and
technology provide the material ingredients which human development requires though,
happiness, ethical values, spiritual well being and wholesomeness of the human person are no
less needed as important elements of a humane society.
This paper argues here that, scientific technology (i.e. human creativity), interacting
with nature (i.e. natural environment) is not and should not be “a journey outward away from
home but a homecoming”; a discovery of the essence of ourselves on earth, and within our
environment in the world. Such an endeavour is uniquely the function of man whose active
life involves a rational principle; an activity of the soul. Man’s moral action, it is contended,
entails the conscious, rational control and guidance of the irrational part of the soul in its
conception of ideas, and or active creation and use of technique for sustainable humaniniy.
Four philosophers, namely, Socrates, Edmund Husserl, Martin Heidegger and Herbert
Marcuse shall be our focus as we attempt the evolution of an ethical approach to sustainable
human environment.
Socrates and Environmental Ethics
That Socrates was at once a moral and intellectual reformer is not, and cannot be an
issue in dispute. History books state in non-contradictory tones his dogged preoccupation in
transforming and restoring moral conduct through knowledge. Human conduct, it was his
belief, is central to every other human activity. For Socrates therefore virtue i.e. knowledge
of the good, is science as much as science is virtue. This thesis is further and better
established in a celebrated dialogue between Socrates and Aristippus (Xenophon, 1925. Bk
III, Ch. VIII).
Thus understood, the basis and content of Socratic ethics is fundamentally relational.
That is, the idea of utility is essentially the idea of a relationship between a means and an end;
nothing, he says, is useful intrinsically, it is useful for something or someone. He thus echoes
in the dialogue above that “nothing is good in itself, that all good is relative”. Without
regressing into the intellectual dogmas of the different strands of ethical theories of
objectivism, subjectivism, relativism and individualism, and their psudo adequacy debates,
one would want to argue that Socrates ethics as a science of the sciences transcends such
limited analysis of the contemporary ethicists. The informed thinking of Socrates is founded
on the nature and function of MAN, who according to him is a soul and not a body.
Accordingly, our attempt at determining the good of man must itself involve considering the
good of the soul and not the good of the body.
It is to be argued though that, there cannot be a soul without the body, and essentially
too that, the good of the soul makes no meaning without corporeality. Truly, such argument
is sound only to the extent that man is not studied dualistically which fallacy Albert
Schweitzer (1961:20) laments that western civilization is a disaster because it is far
developed materially than spiritually, but that it balance is disturbed. The central argument of
Socrates is not that which is canvassed in ethical relativism. Far from that, it rest in the
reasoning that, at the individual level, every body must take care of himself, hence the maxim
know yourself. Thus, the good of man will consist in developing his reason by controlling as
much as possible the desires of his body which are disastrous for the health of the soul.
Such is the overwhelming position of the Socratic ethics that instruments in the hands
of man are said to be neither good in themselves, but worthy relatively to the use we make of
them. Apparently, Socrates is a candidate of the neutrality theory of science and technology.
But this is not the true interpretation of Socratic philosophy. Interpreted to mean wisdom or
reflection, virtue signifies excellence which upon further investigation has nothing in
common with modern day endeavours of science and technology which have no self-limiting
measures or restraints. Such adumbrations by the revered philosophers argues cogently for a
grund norm which universal application will engender biospheric harmony. The philosopher
himself had argued that “to be virtuous is to be fully developed; being good at something,
realizing one’s power. Professor E. K. Ogundowole more clearly understands this state of
affairs as liberation, which according to him is self-liberation, hence, self-reliance supported
by a mental disposition. This mental disposition he argues “must be such that eschew
exploitation of the abilities, enterprise, intelligence and hard work of others, deplore
acquisitiveness for the purpose of gaining and or consolidating power, and reject personal
wealth accumulated or concentrated as to be tantamount to, or effect a vote of, “no
confidence in the social system” (Ogundowole, 1992:255).
Obviously, the development type inspired by such unethical paradigms contradicts the
essential nature of man whose unique and true good is to grow more and more reasonable.
Fundamentally, such moral basis and content as promoted and propagated by Socrates
is definitive of the human environment which essential features of civilization he consistently
points out does not lie in material achievement but in the moral and spiritual development of
the individual i.e. the good of the soul not the good of the body. Placid Tempels (1959:172)
also echoes similarly that “material possessions; housing, increase in professional skills are
no doubt useful and even necessary values. But do they constitute civilization? Is not
civilization above all else progress in human personality?” It is understood here that
‘progress in human personality’ entails a liberated individual with a creative approach to the
human environment, who is constantly guided by the good, and able to consistently live up to
its demands.
Argued as such, ethics (Socratic ethics) is the greatest science (knowledge), and it
identifies virtue with knowledge (science) which true science is architectonic to the essence
of man; “to become a good man.” In what seems to be a global challenge, Socrates queried:
What is it good for to know all the rest, if you do not know the only thing which is essential?
What use will you make of a science if you do not know how to use it for the good? It will be
in your possession like a tool in the hands of a man without experience he manipulates it a
random and injures himself more than he makes progress at work (Diogenes, 1925:179).
By interpretative analysis, Socrates enunciates a true science as encased in the domain
of ethics, the science of excellence, which knowledge can promote human interaction; within
human beings on the one hand, and between human beings and other beings in the biosphere.
That humanity has the capacity to do everything and to be everything. Most rightly enthused,
it is in ourselves that we find the science of good and evil. It is through the examination of
our inner state that we learn and we must seek for whatever we must avoid. The inner
reflection provides us all the solutions sought (Ahoyo, 1997:58).
Truly, science and technology have powerfully helped man to free himself from the
immediate material constraints imposed by the search for security though, they have similarly
caused new evils like degradation of the environment, effects on man’s health, the
dehumanising robotizing of society and the deepening of social inequalities among others.
Prevalence of such noticeable evils of science according to Socrates is a product of ignorance
“Know yourself and you will know what convenes you” is what Socrates commands.
What then counts as an ethical approach for sustainable human development is
founded on the Socratic assumption that all men have the same nature and whatever is good
for one is also good for the other. Methodically, humanity engages in self-search to unravel
objective values, that self-introspection engenders a higher practical value which according to
Hegel is self-discovery. It is a Socratic principle which aim is that,
…man must discover in himself, his destination, his end, the ultimate end of the world, the
truth that is what is in itself for itself, he must attain by himself the truth. It is the return of
self-conscious which is on the contrary determined as getting out the particular subjectivity.
It is thereby that it is eliminated the accidental character of consciousness, the particular
whim, the particularity, by having deep down oneself, this exit, having what is in itself and
for itself. Objectivity has in this context the sense of universality, that is in itself and for itself
and not an external objectivity (Ahoyo 1997:62-63).
Self-knowledge which here means a rigorously rational introspection obviously
avoids contradictions but promotes harmony between convictions and actions. Such
condition is what life is said to be a moral one. Thus, as a basis for human activity in a
biosphere, ethics acts as a guide in the promotion of a true moral life. Human endeavours,
which results from self-consciousness, does not (and cannot) disrupt the link between
conviction (belief) and action.
In truth, such ethical approach more properly defines authentic human beings and
hence sustainable human development. Understandably, ethical knowledge (self knowledge)
amount to good ethical conduct which knowledge unites conviction with will, thought with
action, under the guidance of an inner lucidity, of reason, or of reflective wisdom (Ahoyo, p.
64). This knowledge guides (or should guide) the products of our brains and the works of our
hands to avoid contradictions, and so to be in tune with human existence. But human
existence, it must be unequivocally stated demands meaning in the universe. The
meaningfulness or meaningless of the universe itself starts from the meaningfulness or
meaninglessness of human existence. Every human endeavour, using this ethical approach as
a guide must be subordinated to the human person long acknowledged by Socrates as the
focal point of philosophy. It is here argued that, absolute devaluation of the human person as
is common in todays techno-polis is most unethical. Sustainable human development process
with its purview an invitation to the understanding of the nature and value of the human
person to which Professor J. I. Omoregbe (1990:196) readily provides; that,
man is the key to the understanding of the whole reality. The human person transcends the
infra-human world. The human person posses an inviolable dignity an inalienable liberty
and an inseparable moral responsibility.
This high premium on the centrality of the human person as the absolute value and the
Supreme Being in the universe isolates him out never to be used simply as a means to an end.
In the thinking of Socrates, virtue, which quality is self-knowledge can set us free from the
illusion of reliance on individual ability, and so liberate us from the servitude of the
selfishness, calculation and anti-social ego to fit into the universality of moral laws where in
contradictions are non existent, with man always thinking and acting rightly in the promotion
of the common good. Arguably, such a civilization is wholistic, which human (sustainable)
development, individuals are able to express their inner talents fully in the creation of a happy
and peaceful community, just as they bring about an ecologically prosperous natural
environment, which nurtures them. Such is what is argued as an ‘ethical approach” towards
the evolution of a sustainable human development, wherein, the interests of the individual
and society and humans and nature become congruent. The question is, how does the
SCIENCE of Socrates regulate the modern sciences (and technologies) in the achievement of
this noble goal of sustainable human development?
To answer this all-important question suggests to us a little knowledge of the person
of Socrates. Socrates, we are told was not a metaphysician, but a practitioner, a physician of
souls. It is business was not to construct a system, but to make men think and act morally.
He calls this endeavour the only true science, which engenders the good of man. Captured in
fragments as handed down to us by Plato and Xenopho, Socrates dictates such a true science
as is flavoured by narrowly utilitarian motives thus:
What I ought to do is, what is good for me, and what is good for me is what is useful to me –
really useful (Jacques Maritain, 1979:51).

It is to be understood here that, Socratic ethics seems at first sight to have been
dictated by narrowly utilitarian motives though, he went beyond utilitarianism of every
description. “What is good for me is what is useful to me – really useful” means only that,
the good is not just the material, physical or transient things, but what is really useful to man;
and at this point Socrates compelled his hearers to acknowledge that man’s true utility can
only be determined by reference to a good, absolute and incorruptible i.e. man’s sovereign
good which is his last end. Regulated as such, Socrates seems to be arguing that, humanity is
saved from the catastrophe which trails the trend of development of human knowledge
(science) and skills (technology) that are constantly in the direction of seeking more
comforts, conveniences and control on the natural environment.
More than ever, humanity is today confronted with a new reality, the increasing
knowledge of nature and the ready capability to manipulate it which capability and
understanding have conferred on him a power able to destroy the delicate network which he,
is himself, as a creature of the nature, involved for better for worse Ahoyo (1997:76) argues
in support here that, “to that effect, he (man) has stored in his armouries forces of nature
which, if they escape his control, could annihilate the whole mankind.” When and where this
happens, humanity is said to be acting in the fashion of cancer cells, which when they run
amok and burst out of the prostrate and take over the liver and lymph glands, it kills
everything in the body including the cancer cells themselves.
Obviously, modern science and technology has given today’s humanity more than he
bargained for; serious and burning problems ranging from ecology, exhaustion of the natural,
non-renewable raw materials and the problems of scarcity, starvation and misery of the great
majority of people in the third-world. But as it is said, “where the danger is, grows also the
saving power”, which saving power is the ethical approach of Socrates. This approach
emphasises inwardness, subjectivity and self-knowledge. It is perhaps the absence of this self
knowledge, this self-consciousness that blinds our knowledge of human essence s graphically
presented by Eric Fromm. He says:
He (man) works and strives, but has an obscure consciousness of the usefulness of his action.
Whereas his power on the matter increases, he witnesses his powerlessness on the twofold
level of personal and social life…. Becoming master of the nature, he has become slave of
the machine he has made with his hands. His knowledge about matter is great, but his
knowledge about himself is nil (Ahoyo 1997:138).
Rightly self-consciousness or introspection which quality is self-examination and
hence the capacity to realise what is more authentic in man, is for us the saving power. This
endeavour in human knowledge remains undirected towards the inward dimensions of man
offers the only gateway to the true essence of man on the true human condition. Working
within the framework of this true science (ethics), human aspirations are made to rule self-
interests and short-range perspective, and profitability subordinates sustainability. For,
“nature has to be considered as the whole of which human beings form one component. As a
very important component, they are meant to serve nature rather than make it subservient to
their own needs and wants, for each generation must pass on what it has received in good
order to the next.
The argued conclusion here is that science is truly useful to human kind only and only
as it is ethically sensitive. Correctly rephrased, science without conscience is but ruin in the
soul. This subordination of science to the human spirit is lucidly interpreted by Pope John
Paul II (The Common Good, 1997:31) to signify the kingship and dominion of man over the
visible world, which task consists in “the priority of ethics over technology, in the primacy of
the person over things, and the superiority of spirit over matter.” Humanity totals, and society
tumbles in the event that there is the growing priority of technology over ethics, in the
growing primacy of things over persons and in the growing superiority of matter over spirit.
This is a contradiction of the human will resulting from absence of self-knowledge. In order
to act well, which thought links with action, the stake, according to Socrates is to acquire the
science of the good, and virtue is that science. This according to Socrates the good of the
whole man; the truncated man who is caught between two poles; a material pole, which, in
reality, does not concern the true person but rather the shadow of personality of what in the
strict sense, is called individuality, and a spiritual pole, which concern true personality.
Sustainable human development is derivable from this spiritual pole, the source of
liberty, meaning and bountifulness of man, the form or soul of the whole man. Material
entities have their meaning or rationality because of the impress of the form or soul
(metaphysical energy) the spirit is ordained to inform matter. This is the primary duty of
philosophy which Socrates has recasted in his principle of self-examination which functions
to control the excesses of the sciences by critique and controversy in the attainment of the
ultimate good of man. As the sciences are ever developing and progressing, and responding
to the diverse needs and expectations of Homo technos, ethics (philosophy), the supreme
science must ever trail them, judging and governing them to accord with the pursuit of the
common good, even against strong economic forces that would deny it so as the feared evil of
turning science into an endeavour that devotes itself to organised murder and mass
dehumanisation. The perfect thought of St. Thomas Aquinas may here suffice, that, “any
culture or society or age that does not submit the sciences to the critical leadership of
philosophy (ethics) heads to confusion and low rationality” (Nwoko, 1992:12). Meaning then
that, public life needs rescuing from utilitarian expediency and the pursuit of self-interest. In
human affairs, the twin principles of solidarity and subsidiary need to be applied
systematically to the reform of the institutions of public life.
Husserl and Heidegger on the Ethical Approach
Phenomenology as adopted and used by both Husserl and his student Heidegger
suggest a method of investigation where from the essences of Beings are made known as they
are in themselves as they are. While Husserl insists that phenomenology as a method is
characterised by ‘what’ of the object of philosophical investigation as to its subject matter,
Heidegger argues otherwise that it is the ‘How’ of that investigation. Notwithstanding their
special emphasis, phenomenology etymologically formulated means “to let that which shows
itself be seen from itself in the very way in which it shows itself from itself” (Heidegger,
1978:58). It means by this descriptive statement that, true knowledge (science) of being is
possible only through phenomenology; that ‘phenomenon’ is the being of entities, its
meaning, its modifications and derivatives. Thus, the argued conviction of both Husserl and
Heidegger is that, behind the phenomenon there is nothing else, but since the phenomenon
itself can be hidden – proximally and for the most part, there is, need for phenomenology
which maxim is “to the things themselves”.
But as human knowledge has made progress, it has not and cannot come to anything;
it has rather raised more problems than solutions. Such paradoxical situation, to which
knowledge (science and technology) has led us to, convinces us that knowledge itself is a
disability. This is what we call the crisis of science and technology to which Husserl and
Heidegger offers a phenomenological rescue mission. Science and technology are products
of the essentially metaphysical character of the western intellectual tradition which
technocratic reduction of everything to planning, calculation and predictable laws, wrest
objectivity from what is, the quest for certainty in our ways of knowing and the passion for
totality or the total dominance of everything.
Such is the real source of the problem of modern science (and technology). As Dr Jim
Unah rightly alludes, “by forcing things to appear which he (man) does not need, man turns
himself into the conqueror of nature, into an overlord who wills to thoroughly exploit and
dominate the earth”. But he concludes rightly too that, “he who exploits and dominates the
earth ends up thoroughly debasing the earth” and destroying the entire biosphere, including
himself (1998:362).
It remains to be seen how Husserl and Heidegger have adopted the phenomenological
method as an approach to true humanism and hence removing “Abstacles to the Building of a
Beautiful World” (Read Easlea, B, 1973). They both argued that, what leads to a distortion
of reality is not any inherence of a distorting element in things themselves, but the way we
position ourselves to view them. They argued further that we can position ourselves to view
things and relate with objects and see the objects the way they are, without bias, prejudice,
preconceptions and predispositions of particular circumstances. Thus inquisitional
methodology for Husserl is epoch and phenomenological reduction, while goes for the
explication of Dasien. Phenomenology, they argued in conviction, promises to be a vehicle
for authenticity as it purges the metaphysical attitude of viewing what is presented to ones
consciousness from the cognitive imposition of another.

Edmund Husserl (1859-1938)


Phenomenal technological advancement is said to be the most important noticeable
index of the twentieth century. As lavishly described in chapter two, the twentieth, but
beginning with the nineteenth century witnessed spectacular achievement in virtually every
area of human endeavour thus earning our generation the appellation the best of times. But
the necessary natural law of Nyohon yuan which is exemplified in the principle of opposites;
good and bad, positive and negative have on the other hand brought to bear on humanity
deterioration in our ecological system, widespread abortion and the ever looming threat of
nuclear holocaust through war or accidental detonation. On this negative note, our century
could also be described as the worst of times.
It is to be said that the forces of the techno-scientific economy are threatening the very
existence of human life, even while they create unheard-of material bounties for a minority of
humanity. These same forces are giving rise to ever more complex social, political and moral
questions. Rightly described, our century has become in the words of Eric Hobsbawn
(1996:6), an era of “decomposition uncertainty and crisis which for Edmund Husserl means
that science has lost its importance for life. The question of science, but the domineering
spirit of the positive sciences in particular, with its spirit of exaggerated and blinded
materialism meant that man was diverting himself with indifference from the questions which
are decisive for an authentic humanity. Perhaps the problem which could best described as
global in nature is most comprehensively listed to include among others the following:
Uncontrolled human proliferation, chaos and division in society, social injustice, hunger and
malnutrition, widespread poverty, the mania for growth inflation, energy crisis, international
trade and monetary disruptions, protectionism, illiteracy and anachronistic education, youth
rebellion, alienation, uncontrolled urban spread and decay, crime and drugs, violence and
brutality, torture and terrorism, disregard for law and order, nuclear folly, sclerosis and
inadequacy of institutions, corruption, bureaucratisation, degradation of environment,
decline of moral values, loss of faith, sense of instability, lack of understanding of the above.
Problems and their interrelationship. (Aurelio Peccei, 1979)
This prevalence of the problems have reduced human thinking to the concluding that
science no longer has anything to say to humanity in the distress of their live. It is within the
bounds of this intellectual tradition that Edmund Husserl argues that science has ignored its
most crucial traditional function. As he alludes, the questions it excludes in principle from its
field of concern are precisely questions which are the most burning for our unfortunate times,
for a humanity abandoned to the upheavals of destiny. They are questions related to the
sense or the absence of sense of all human existence. These questions, he argues further
require in their generality and necessity that we carefully and adequately consider them so as
to find answers to them which come from a rational view; he says
the evil in the positivist approach of science consists in excluding subjectivity from its
domain of research; but then all that concerns man himself is precisely to be found in this
subjectivity, this spirituality (Ahoyo, 1997:79)
Husserl implies by this assertion that the person is the basis of judgement of techno-
science, and that it is the absence of the concept of person that science has found itself in the
present mess, losing sight the original foundation on which it has been built. Personalism is
perhaps the best watch-doctrine and the most wholesome in the presentation of man against
the truncated conception of man. This philosophical knowledge of man which involves the
true meaning, dignity and destiny of man may have most obviously informed. Boethiu
definition of the person as naturae rationalis individual substantia (the individual substance
of rational nature) (Boethius ). This definition implies that man is a natural unity, a unity of
the individual man, a unique entity of self. The person is the totality which the self achieves
in the individual entity in the unity of his spiritual and physical aspects.
The threatening symptoms of the crisis in European culture is viewed by Husserl as
eroding this unique understanding of man. He thus undertakes through a critical and deep
analysis of the philosophic thought which has lost its human dimension. He thus jettisoned
the traditional reduction of human knowledge to objectives scientific knowledge leaving
aside the vast domain of sensitive and immediately subjective knowledge. Using this
approach, Husserl sought to merge sensitivity and understanding, subjective emotion and
concept into a single whole in an attempt to strike a relationship between the activity of
consciousness and human essence. In his “The Idea of Phenomenology (1970) Husserl
convincingly argues out this possibility. That when the mind or ego is purified, and so
effectively carried out, a zero – attachment is achieved, hence consciousness is poised to
“see’ the thing as it truly is. This is properly speaking what Husserl calls reduction, which is
a cognitive process of arriving at the essence of a thing through the extraction of intellectual,
doctrinal and particular colorations, which procedure meaning is intuited, leading the human
person to transcendental subjectivism. It is to be acknowledged that such a sense of human
existence in its wholeness is founded on Husserl’s strong belief that the human person is the
basis of our practical judgement of the good. Indeed, it is actually the principle of goodness
alive in the world. Such is which strong faith Husserl had in the human project that he
brilliantly captures in his paper entitled “The Crisis of European Humanity” when he says
that “misled humanity could be called back to reason and that phenomenology could reveal
its authentic image.”
Interestingly, Husserl sought a reverse in methodology. Reality he says is found
through the eye-glasses of phenomenology which search for immutable foundations of
philosophy directs knowledge towards pure consciousness”, towards total subjectivity
(Ahoyo 1997:110) Husserl argues then that the new task of philosophy is to restore the sense
of human wholeness which positivism has destroyed; to rediscover the sense of wholeness.
Such is why he sees philosophers as the civil servants of humanity, to restore that which is
most noble and most perfect in all of nature, to integrate all the existing values and
potentialities in the world towards a transcendent goal which find expression in his concept of
intersubjectivity, that i, the coalescing together of subjectivities, which “agreement that a
certain thing is the case becomes objective” (Unah, 1998).
For Husserl therefore, the human person is central to what counts as development and
that which is good is that which is subsumed in the concept of the human person who is the
centre of complementation and communion of all created worldly values; the natural social,
the universal values, all values: material and spiritual have their ultimate meaning only in
reference to the person. Sustainable development only is to the extent that it makes the whole
of man; his material and spiritual values, a focal point. For St. Thomas Aquinas most rightly
posits that, the human person signifies what is most perfect in all nature, hence Agenda 21
Principles (UN Briefing Paper, 1997:27) correctly adumbrates the point further that, “human
beings are at the centre of concerns for sustainable development. They are entitled to a
healthy and productive life in harmony with nature.”

Martin Heidegger
The thoughts of Heidegger on science (and technology) are not too expressly distinct
from those of Edmund Husserl. But more than Husserl, Heidegger made a successful attempt
at distinguishing between modern science (or technology) and ancient science (or
technology). He thus argued like Husserl that, modern science had developed losing sight of
the original foundation on which it has been erected and that neglect was responsible for the
crisis it is getting across in spite of its success. According to Heidegger, (1997:3-37)”
technology (in its everyday sense) is not equivalent to the essence of technology” to be free
of misunderstandings, to relate technology intelligently, we must fund its central meaning and
that can be done only by discovering its essence; we must think of its relationships with all
else. To view technology as a complex of contrivances and technical skills, put forth by
human activity and developed as a means to our ends is an error of judgement. “We are
delivered over to it in the worst possible way when we regard it as something neutral”, he
says. On the contrary, the essence of technology reveals it as something far from neutral or
merely an instrument of human control; it is an autonomous organizing activity within which
humans themselves are organised.
The argued position of Heidegger is that the true essence of technology is to be
located in the modes of occasioning, the four causes; Causa materialis, causa formalis causa
finalis and causa efficiens. As he put it “every occasion for whatever passes over and goes
forward into presencing from that which is not precencing is poiesis, is bringing-forth”
(Heidegger 1977:10). This bringing-forth is, in its most generally understood sense what the
Greeks called aletheia, which Heidegger expressed in the German word Entbergen and his
English translators have expressed in the word ‘revealing’, that is truth revealing which
objective significance is the expression of the actual coming into presence of something.
Put in proper perspective, Heidegger here locates technology within its Greek
etymology as essentially that which belonging to the general notion of bringing-forth,
Poiesis. he thus adumbrates this position further and better thus;
techne… reveals whatever does not bring itself forth and does not yet lie here before us,
whatever can look and turn out now one way and now another… Thus what is decisive in
techne does not lie at all in making or manipulating nor using of means, It is a revealing, and
not as manufacturing, that, techne is a bringing-forth (1977:13).
By this assertion Heidegger means to insist that the basis essence of technology has
remained the same unchanged and that this essence is most readily observed in the Greek
origins of our thinking about these things. The problem with modern technology and the
dangers of modern science and technology is that they have evolved outside this essential
nature as a mode of revealing. What we understand as modern technology can hardly be
recognised as having a common origin with the fine arts or craft. Instead modern technology
is distinguished in having made its alliance with modern physical science rather than with the
arts and crafts.
It is thus not farfetched to conclude that modern technology destroys, and
dehumanizes. Indeed, humanity journeys and involves with nature to the point of intrusion
upon it. Thus, instead of diverting the natural course co-operatively (wherein lies the essence
of technology) modern technology emphrames and achieves the unnatural by force. Not only
is it achieved by force but it is achieved by placing nature in our subjective context, setting
aside natural processes entirely, and conceiving of all revealing as being relevant only in
human subjective needs.
The essence of technology originally was a revealing of life and nature in which
human intervention deflected the natural course while still regarding nature as the teacher
and, for that matter, the keeper. The essence of modern technology is a revealing of
phenomena, often far removed from anything that resembles ‘life and nature’ in which human
intrusion not only diverts nature but fundamentally changes it. As a mode of revealing,
technology today is challenging – forth of nature so that the technologically altered nature of
things is always a situation in which nature and objects wait, standing in reserve for our use.
We pump crude oil from the ground and we ship it to refineries where it is fractionally
distilled into volatile substance and we ship these to gas stations around the world where they
reside in huge underground tanks, standing ready to power our automobiles or airplanes.
Technology has intruded upon nature in a far more active mode that represents a consistent
direction of domination. Everything is viewed as “standing-reserve” and, in that, loses its
natural objective identity. The river for instance, is not seen as a river, it is seen as a source
of hydro-electric power, as a water supply, or as an avenue of navigation through which to
contact inland markets. In the era of techne humans were relationally involved with other
objects in coming to presence; in the era of modern technology, humans challenge forth the
subjectively valued elements of the universe so that, within this new form of revealing,
objects lose their significance to anything but their subjective status of standing-ready for
human design. Thus everything in the universe, including humans, have been transformed in
significance leading to a loss of humanity. It may be said to that extent that, humanity has
been conducted out of its own essence.
Obviously, our attempt at converting ‘science and technology as tools of human
development but which have become standing reserves has effected the greatest threat to
humanity by carrying humanity away from its essential nature. On the one hand we consider
ourselves, rightfully, the most advanced humans that have peopled the earth but, on the other
hand, we can see, when we care to that our way of life has also become the most profound
threat to life that the earth has yet witnessed. Medical science and technology, it is argued,
have even begun to suggest that we may learn enough about disease and processes of aging in
the human body tat we might extend individual human lives indefinitely. In this respect we
have not only usurped the god’s rights of creation and destruction of species, but we may
even usurp the most sacred and terrifying of the god’s rights, the determination of mortality
or immortality (Tad Beckman, 2000:13) Thus maternally and spiritually, human life and its
environment have been profoundly transformed, and humanity no longer has a correct
relationship with the environment.
For Heidegger therefore human development is not and cannot be a product of
modern technology, for it has lost its essence. As human beings become progressively more
involved as the orders of reality conceived as standing reserve, they too become standing
reserve at a higher level of organization. That is, as human beings come to see other beings
in the world only for their potential applications to human dispositions, humans themselves
come to mirror this shallowness of “being” and to see themselves merely in terms of potential
resources to the dispositions of others. Understood within this human disposition, our
essence as human beings falls into concealment which activity Heidegger calls enframing.
As Tad Beckman is to argue in explication,
Emframing challenges us forth in the decisive role as organizer and challenger of all that is
in such a way that human life withdraws from its essential nature. Within this role the
essence of our humanity fall into concealment; we can no longer grasp the real nature of life.
We withdraw into a conception of reality that is subjective and isolated (Beckman 2000:15)

But, Heidegger asserts that the human essence is not a being in isolation. Human
beings unlike most beings that are simply in existence with no relationship to one another, no
consciousness, are unique, they are beings among beings, beings who witness other beings.
Such essence of human life is founded in the facticity, or objectivity of Dasein; not only do
we humans come into relationship with other beings through our characteristic consciousness
but they come into their own beings as objects through us. They are witnessed by us. This is
why Heidegger insisted that from the position of our own essence, “we can never encounter
only [ourselves’” (Adams; 1946:27). So argued, any conception of our environment that
perceives only ourselves and our dispositions is necessarily flawed from the point of view of
essential human nature.
But is there a way out of this human predicament? The answer to this complex and
difficult question may simply be YES. We agree with Holderlin that “where the danger is,
grows also the saving power”. We must stare into the depths of all that is and was and can be
and recognise, above all, that what humans essentially are is, in some mysterious way, a
“grant.” So Heidegger says, “only what is granted endures. That which endures primally out
of the earliest beginnings is what grants” (Heidegger 1977:31) If technology is seen as an
imminent threat to humans, it comes to focus attention upon that which is granted to human
life, since what is granted is precisely what is most threatened. Thus Heidegger suggests that
the saving power begins to grow precisely within the greatest danger. The saving power is
found in the arts, which saving power is “more than hauling something back to its original
form; instead, it should be construed as bringing something back into its essence. Thus the
saving power that arises through art and within the danger of modern technology must be a
power to bring humanity back into their essence. Heidegger’s attempt could be summarised
within this thinking, that “we must proceed into the future, as we interact in the techno-polis,
from where we stand but, while we proceed, we should use these things and our talents to
come back into our own essential nature.
Technology carries humanity outward from ourselves and to that extent humanity fails
in the essential task of human fulfilment as beings whose very essence is to be – there, to
witness the whole of what is. Through the art, which nature is homecoming, that is,
discovering the essence of ourselves on earth and within our environment in the world, we
are healed by coming back into our on essence. It is not an exaggeration to say that the art
does bring us the power that can save us from the people that we have become. Art might be
able in some way to drawn us back into a more original form of bringing – things forth. It is
perhaps to be understood that Heidegger is well informed on this consistent and well
developed picture of art, especially the art of poetry. In this picture,
art is a mode in which life is experienced in which life is experienced in which truth happens
for us. …art is a mode of revealing, a setting forth, in which humans and other objects-
beings come to presence in an organization that is far closer to the essential nature of human
life on this earth. (Heidegger, 1971:25).
As a saving power that returns humanity to its essential nature, art carries us into the
essential tension between earth and world and to the essential need of humans to fund a
joyous home within for just as technology in the epoch of enframing has effected the greatest
threat to us by carrying us away from our essential nature, art possesses the capacity to
become the mastering theme of a new epoch in which we are healed by coming back into our
own essence.
Such is what Heidegger calls a bringing-forth which means the liberation of man from
the hold of technology and a modification or a redefinition of our relationship with them.
Instead of being fascinated and dominated by them, we can in using them normally keep a
certain distance to them, that is allowing them to reveal themselves the way they are in
themselves as they are in themselves. This condition of science and technology is sine qua
non for human fulfilment nay sustainable human development. For Heidegger therefore,
We can say “yes” to the inevitable use of technology but at the same time say “no”, which
means that we should impede them to monopolise us and thus to miss stifle and finally empty
our Being (Heidegger, 1977:49)
Such temperament is what Heidegger calls the “serenity of the soul” which condition
entails a communion between the body and the mind. Since a human person is possessed of
both mind and body requiring both spiritual and material fulfilment pursuit of wealth and the
satisfaction of the physical needs of man must be tempered by the cultivation of the mind.
Outer satisfactions of a material kind should be enhanced by the inner satisfaction of the
mind and spirit. This is the goal of wholistic human development which the physical needs
of man are achieved through science and technology (from nature) though, they are not used
in a manner that they will dominate us and finally empty our Being. They are used in a way
that we are at peace and a piece of nature, at peace with our emotional needs by maintaining
peace between the individual and society, from which we also derive intellectual and spiritual
peace.

A Phenomenological Rescue Mission


In an essay entitled “Remembrance of the Poet”, first published in 1943, Heidegger
(1979:233-269) explicates an analysis of Holderlin’s elegy “Homecoming” in which he tells
the story of a man who returns from his youthful travels to the town of his birth, his home.
He sails across Lake Constance and out of the shade of the ALPS to the little town, where he
finds familiar places and congenial faces. As Heidegger saw it “Homecoming” tells a deeper
story of a poet who is finding the significance of his homeland and, hence, of home itself.
One most significant aspect of this story is Heidegger’s conception of the poet’s
journey in life as wholly a matter of “homecoming”. Life, Heidegger argues, “really consists
solely in the people of the country becoming at home in the still – withheld essence of home”
(1979:245). Homecoming is the return into the proximity of the source, it is the essence of
our being on the earth and that towards which we should work in our lives.
Such contemplation is truly which philosophy informs the phenomenological
principle to which Edmund Husserl and Martin Heidegger both subscribe. That total
domination of the earth by anybody or any group would not save and preserve the earth. “If
our true project is to save the earth from cosmic disaster” opines Dr Jim Unah, “all we require
is to have or learn to have a phenomenological access to the world to things and people”
(1998:363). This attitude is founded in the facticity or objectivity of man who has
relationship with other beings, the only being who observes both himself and other beings.
Thus, phenomenological access to what is, entails a bringing-forth, it requires an attitude of
the mind, which allow things to manifest themselves without forcing them into our
straitjackets. It is letting things be, and letting what is reveal itself without coercion; a
revealing that heals humanity and conducts it back into its real essence on earth. This
inherent disposition of phenomenology is what Jim Unah (ibid) tagged, phenomenological
rescue mission, but which Holderlin calls “Homecoming”; a uniquely vital journey into the
basic human issue of finding the essence of home (man’s original state of existence) within
life on this earth.

Conclusion
The task of philosophy, it is often said is the critical examination of the ideas we live
by. This supposition further argues that, philosophy has always announced and justified the
task of a rational reorganization of the world, which implies the recognition of the specific or
at least the potential rationality of the universe. Candidates of this school of thought are
quick to conclude here that, one might rationalise the existing, though it is not immediately
rational. That, what the rationalization of the world has led us to through science and
technology is the worsening of human condition, perhaps without a saving power. The
present alienation of human condition, they argued is a by-product of this condition which
remains caught in the trap of positivism whose evils it has so brilliantly unmasked.
In our preceding analysis, we have shown that philosophy is not only the critical and
rational examination of the ideas we live by, but that it is also the saving power of the ideas
that govern human existence. Basking in the era of the crisis of science and technology,
which consequent effect is the disappearance of “the person, but the emergence of the
machine”, we have argued in a reverse order that, the ‘person’ is the measure of all things.
The person, it is argued in this work, is the totality which the self achieves in the individual
entity in the unity of his spiritual and physical aspects understood as such, the human person
is the basis of our practical judgement of the good.
It is based on this thinking that we conclude that “science without conscience is ruin
of the soul.” The implication here is that, science and technology necessarily needs to be
inward directed so as to avoid the feared danger of conducting humanity out of its real
essence on this earth. Sustainable human development is more than mere growth and
progress in material terms, it means growth and progress in reference to the human person
who is the reason for all values in the world; material and spiritual. This, to us is what counts
as an ethical approach to sustainable development.
Informed by the thinking that humanity always poses problems that it can solve, we
proposed three options as a way ahead (out of) the present crisis in science and technology.
Option one argues that the way out of the radical upheaval caused by science is a
return to morality. Humanity, Socrates says, is at the crossroad and can only be returned
back into natural human essence through his moral philosophy which characteristic features
are inwardness, subjectivity and self-knowledge. Ethics he says, is the queen of all sciences
and without ethics science cannot stand.
Option two argues out the rehumanization and healing of the positivist contagion
through phenomenology. Pioneered by Edmund Husserl, this, line of thought argues that the
restoration of human wholeness which positivism has destroyed is possible through
transcendental consciousness. Phenomenology, in its search for immutable foundations of
philosophy, he says, directs knowledge towards. “Pure consciousness”, towards total
subjectivity. Philosophy, hitherto, he says, has remained naively objectivist; the ontological
constants brought out by that philosophy are without any relationship with human existence.
Hence, it is the task of philosophy to rediscover this sense of wholeness. So, he says
philosophers are the civil servants of humanity (Husserl, 1965:23).
Option four closely associates itself with option two, and extends its garb to
existentialism using it as a war against the project of a scientific philosophy propounded by
Martin Heidegger, a student of Husserl, existentialism like phenomenology argues the role of
philosophy in the restoration of moral and spiritual confidence of men, against the
dehumanisation and scepticism sowed by science and technology.
Whether it is the moral philosophy of Socrates, or the phenomenology of Husserl, or
the existentialism of Heidegger, the synergy is that, ethics define relationships between
beings; beings who witness other beings as beings among beings, and that, the nature of a
relationship defines the essential human nature or otherwise. Such conviction may have
informed the observations of the French philosopher and mathematician, Michel Serres, that,
…we (i.e. the scientists) have henceforth the responsibility to manage the infinite cone of the
possible that the ethics of our fathers named reality, inventory, speculative activity seems to
pose non-ethical problems; choice on the contrary, a serious one. Once you have for a long
time combined, you have to choose what may pass from the possible to the actual (Ahoyo,
1997:122-123).
What this observation reduces to is that, the age long paradgm of scientific neutrality
is replaced with the phrase “to know amounts to choose”, for as Heidegger rightly suggests,
“from the position of our own essence, we can never encounter only ourselves”. We have
thus argued that, human action must always be informed by an attitude of the mind, in the
promotion of the human person. Kant’s formulation of the categorical imperative may
suffice here, that human action should be always directed “as to treat humanity, whether in
thy own person or in that of any other, in every case as an end in all, never as means only”
(Ozumba, 2001:89). While we support Kant’s position that the motive of the will is good if
and only if its motive is solely one that emanates from duty, we hasten to say that such
motive should not emanate from duty for duty sake. Such sweeping conclusion has the
capacity of promoting the culture of scientism.
It is thus argued that, human dignity, of everyman should be topmost in human
presencing. Arguably, human presence is crucial to other beings coming-to-presence, to truth
happening. That is to say that, the human essence is fundamentally involved in all revealing,
in all objects coming into unconcealment. Technology as a mode of revealing, is one part
within many possible parts that open up within the essential nature of that human role; each
of these parts develops a specific aspect of our relations to beings. That relationship is
always reciprocated in the sense that, in so far as being-there is our essential mature, the way
that we are there, the way that we relate, is the way that we ourselves come into being during
that period. Heidegger, an authority in this insight allays with this conclusion and adds in
particular that, “the way we treat other things is determinant of the way we ourselves will be
treated.
True, science and technology have made tremendous progress and growth, we have
mastered gravity and space, we have driven back the limits of life or death, we can now
choose the sex of our children and may tomorrow reproduce our own kind asexually and treat
any type of complicated disease, thanks to the breakthrough in the study of genes. But
herein, that power, lies all our problems. It is thus no longer what could I know , which is the
question of science, but what should I know and do which is the question of sense (ethics).
What is being argued for is a responsible human environment in which humanity is called
upon to integrate in its present actions the care to preserve the life of its descendants, nay its
environment.
In what appears as a summary of our position, Hans Jonas has formulated in a Kantian
formula the following ethical imperative:
Act in such a way that the effect of your action be compatible with the permanence of an
authentically human life on earth. (Ahoyo 1997:136).
This imperative itself is a call for a meaningful relationship of openness and dialogue
within human being on the one hand and, with nature; the environment on the other. Yersu
Kim (1999:42) in his “A Common Framework for the Ethics of the 21st Century” provides in
addition, a four point agenda in this regard:
(i) The view of nature as accessible through causal mechanistic law has enabled
humanity to control nature and provide for itself the good life on earth. The same
view has contributed to the destruction of the natural environment and alienation
of human beings. We must therefore seek a balance such that we may maintain a
sustainable harmonious relationship between the human species and nature.
(ii) As nature is a finite quality, we must learn to manage the economy to sustain the
complexity and stability of nature while at the same time to manage nature so as
to sustain our economy. As our desires are insatiable, we must learn to
accommodate our desires to the limits nature sets, not to push the limits of nature
beyond its capacity for generation.
(iii) Humanity needs to develop economically and technologically in order to deal with
the problem of poverty in which a great majority of human beings still live.
Continuation of economic development at the present rate endangers the rights of
the future generations to life and a healthy environment. We must therefore, learn
to balance short-term thinking and immediate gratification with long term thinking
for future generations by shifting the balance towards quality rather than quantity.
(iv) Consumption contributes to human well-being when it enlarges the capabilities
and enriches the lives of the people. Consumption, when excessive, undermines
the resource base and exacerbates inequalities. Consumption therefore must be
such as to ensure basic needs for all, without compromising the well-being of
others and without mortgaging the choices of future generations.

This is the agenda for sustainable development which corpus entails that nature has to
be considered as the whole, of which human beings form one component, which important
component, they are meant to serve nature rather make it subservient to their own needs and
wants. The human species, with all its attributes of intelligence, inventiveness and capacity
of intervention is called upon to use these qualities in a positive manner to serve the whole of
which they are a part. Instead of exploiting nature in a manner of forcing things to appear
which man does not need, instead of dominating nature which action backfires and ends up
thoroughly debasing the earth with man inclusive, humanity should act as sentinels of nature
and help maintain the multifarious delicate webs of the eco-systems that make it function in a
sustainable manner. “We could learn from the bees” recommends Dr Devendra Kumar, “the
manner to serve nature and get its sustenance simultaneously. The more the honey it collects
from the flowers, the more it serves, in the propagation of the plants by helping in their
fertilization. We could emulate the bees by fulfilling our needs through a similar symbiotic
relationship with nature.” (Kumar 2001:2). Perhaps too, the Delphic Method of Rushworth
Kidder, the founder of the “Institute for Global Ethics” (USA) could help reinvent a new
world order for sustainable human development. In his Shared Values for a Troubled World,
Kidder (1994), identifies a number of cross-cultural core values: love, truthfulness, fairness,
freedom, unity, tolerance, responsibility and respect for life as architectonics of sustainable
human development; a wholistic development which entails a combination of the physical,
emotional, intellectual and spiritual dimensions. This should be in a way that humanity is at
peace with nature; at peace with our emotional, intellectual, and spiritual needs. This, John
XXIII (1963) argues can be established only if the order between men and nations, laid down
by God, and rooted in the nature and dignity of the human person is observed.
This is a call for the regulation of human activity which activity proceeds from man,
and which human activity is also ordered to him. The development of his life through his
mind and his works should not only transform matter and society, but it should also fulfil
him, his spiritual realm, for it is what a person is rather than what he has that counts. Thus,
technical progress is an important compliment of human development though, it is of less
value than advances towards greater justice, wider brotherhood and a more humane social
environment. It is here argued that, the norm for human activity is to harmonise with the
authentic interest of the human race, in accordance with God’s will and design, and to enable
men as individuals and members of society to pursue and fulfil their total vocation – the
better ordering of human society.

References
Achebe, C. (1959) Things Fall Apart, Greenwich Ct., Fawcett Pub.
Adams (1946) The Educatioin of Henry Adams, Intro; Adams, James Truslaw, New York, Random House.
Anyanwu, K.C. (1983) The African Experience in the American Market Place, New York, Exposition
Press.
Anyidoho, K (2000) “Culture: the Human Factor in African Development” in, Ghana: Changing Value,
Changing Technologies (Ghanaian Philosophical Studies II) H. Lauer (ed) Washington D.C.
The council for Research in value and Philosophy.
Easlea, B. (1973) Liberation and the Aims of Science: An Essay on Obstacles to the Building of a
Beautiful world. London, Chatto and Windows for Sussex University Press.
Ekei, J.C. (2001) Justice in Communalism: A foundation in African Philosophy, Lagos. Realm
Communication Ltd.
Fromm, E. (1968) The Revolution of Hope Towards a Humanized Technology, New York, Harper and
Row.
Heidegger, M. (1977) “The Question Concerning Technology” in The Question Concerning Technology
and Other Essays, New York, Harper and Row, pp. 3-35.
Heidegger, M. (1979) “Holderlin and the Essence of Poetry” in Existence and Being, Int. Brock, Werner.
South Bend, Ind: Regnery/Gateway, pp. 233-269.
Husserl, E (1965) Philosophy as a Rigorous Science and Philosophy and the Crisis of European Man,
New York, Harper and Row.
Husserl, E (1976) The Crisis of European Science and The Transcendental Phenomenology, Gallimard.
Kim, Y. (1999) A Common Framework for Ethics of the 21st Century; Paris: Division of Philosophy and
Ethics, UNESCO.
Marcuse, H. (1964) One Dimensional Man: Studies in the Ideology of Advance Industrial Society, (3rd
Impression) London, Routledge & Kegan Paul Ltd.
Mclean, G. (1964) “The Contemporary Philosopher and His Technological Culture in G.E.
McLean (ed.) Philosophy in a Technological Culture, Washington D.C., C.U.A. Press.
Momoh, C.S. (ed) (2000) Philosophy For All Disciplines, Vol. II, Lagos, Department of Philosophy
Publications.
Nwoko, M.I. (1992) Philosophy of Technology and Nigeria, Nekede, Maryland, Claretian Institute of
Philosophy.
Ogundowole, F.K. (ed) (2002) Man, History and Philosophy of Science: A Compendium of
Reradings, Lagos Department of Philosophy, University of Lagos.
Omoregbe, J.I. (1990) Knowing Philosophy, Lagos, Joja Educational Research and Publishers Ltd.
Onuobia, O.N. (ed) (1991) History and Philosophy of Science, Aba Maiden Educational Publishers Ltd.
Ozumba, G.O. (2001) A Course Text on Ethics Lagos, Obaroh and Ogbinaka Publishers ltd.
Russell, B. (1962) “The Taming of Power” In the Basic Writings of Bertrand Russell (1903-1959), R.E.
Egner, & Dennon (eds), New York: Simon & Schuster .
Schweitzer, A. (1961) Civilization and Ethics, London: Unwin Books.
Temples, P. (1959) Bantu Philosophy, Paris Presence Africaine.
Unah, J.I. (ed) (1998): Philosophical Science For General Studies, Lagos, Foresight Press
Kumar, D. (2001) Excerpt from his Award Lecture, the Indian National Science Academy (INSA) Annual
B.D. Tilak Award for Rural Development. Available at http://www.insa.org/
John Paul II (Pope) (1979) Redemptor Hominis 16. Boston, Daughters of St. paul.
John Paul II (Pope) (1988) Solliciludo Rei Socialis, Boston, Daughters of St. Paul.
1. Dennett, Daniel. 1995. Darwin's Dangerous Idea. New York: Simon & Schuster. Reprint
edition, 1996. ISBN 068482471X
2. Harré, R. 1967. “History of Philosophy of Science,” in Encyclopedia of Philosophy, edited
by Paul Edwards. Vol. 6, p. 289.
3. Carnap, Rudolf. "Empiricism, Semantics, and Ontology." Revue Internationale de
Philosophy IV (1950): p. 20.
4. Reese, William L. Dictionary of Philosophy and Religion: Eastern and Western Thought.
New and enlarged edition, 1996. p. 596. Atlantic Highlands, NJ: Humanities Press
International. ISBN 0391038648
1. ↑ Reese.
2. ↑ Reese.
3. ↑ “Logical Positivism” in Encyclopedia of Philosophy. Edited by Paul Edwards. Vol.
5, p. 56.
4. ↑ Popper, Karl R. 1976. Unended Quest: An Intellectual Autobiography. Revised
edition, 1982. La Salle, IL: Open Court Publishing Company. ISBN 0875483437
5. ↑ Vienna, 1934. It was not translated and published in English until the mid-1950s,
when it appeared as The Logic of Scientific Discovery. New paperback edition, 2002.
London: Routledge. ISBN 0415278449
6. ↑ Newall, Paul. “Lakatos and the Demarcation Problem.”
7. ↑ Kuhn, Thomas S. 1962. The Structure of Scientific Revolutions. Third edition, 1996.
Chicago: University of Chicago Press. ISBN 0226458083
8. ↑ Kuhn 1962, 91.
9. ↑ Kuhn 1962, 10.
10. ↑ George, Alexander. “What's wrong with intelligent design and with its critics.”
11. ↑ Laudan, Larry. 1990. Science and Relativism: Some Key Controversies in the
Philosophy of Science. Chicago: University of Chicago Press. ISBN 0226469492
12. ↑ Laudan, Larry. “The Demise of the Demarcation Problem,” in Michael Ruse (ed.).
1996. But is it Science?: The Philosophical Question in the Creation/Evolution
Controversy. Amherst, NY: Prometheus Books. ISBN 1573920878
Black, M. (1975) “Is Scientific Neutrality a Myth?” In Lipscombe, J. and Williams Bill (eds) Are
Science and Technology Neutral? London, Btterworths.
Bronowski, J. (1971) “The Disestablishmentof Science”, in Watson Fuller (ed) The social Impact
of Modern Biology, London, Routledge and Kegan Paul.
Caspari, D.W. and Marshak, R.E. (1965) “The Rise and Fall of Lysenko”, In Science, 16 July.
Dampier, C.W. (1966) A History of Science and Its relation with Philosophy and Religion. Great
Britain, Cambridge University Press.
Dickson, D. (1974) Alternative Technology and the Politics of Technical Change. London,
Fontana, Chapters 1,2,3,7.
Dixon, W.M. (1958) The Human situation, Harmondsworth, Middlesex, England; Penguin
Efemini, A. (1996). Scientific and Technological Development in Nigeria: A Dialectical
Approach”, in Uduigwomen, A.F. (ed.) A Textbook of History and Philosophy of Science,
Aba, A.A.U. Industries: Publishing Division.
Ernest C. (Sir), (1970) “Social Responsibility and the Scientist” in New Scientist, 22 October, pp.
166-170.
Feyerabend, P. (1925) Against Method, London, Lowe and Brydone Printers Ltd.
Kuhn, T. (1970) The Structure of Scientific Revolutions, Chicago, University Press.
Lewis, J. (1977) History of Philosophy, London, The English University Press Ltd.
Lipscombe, J. and William, B. (1979) Are Science and Technology Neutral? London,
Butterworth & Co. (Publishers) Ltd.
Momoh, C.S. (ed) (2000) The Substance of African Philosophy, Auchi, African Philosophy
Projects Publications.
Moore, G.E. (1903) Principia Ethica, Cambridge, Cambridge University Press.
Newswatch Feb. 12,
Nwoko, M.I. (1992) Philosophy of Technology and Nigeria, Nekede, Maryland, Claretian
Institute of Philosophy.
Oluwole, S.B. (1999) Philosophy and Oral Tradition, Lagos, African Research Konsultancy
(ARK)
Popper, K.R. (1945) Open Societies and Its Enemies, London, Routledge and Kegan Paul.
Popper, K.R. (1963) Conjectures and Refutation: The Growth of Scientific Knowledge, New
York, Harper and Row Publishers.
Richter, M. (1972) Science as a Cultural Process, London, Schenkman Publishing Co. Inc.
Russell, B. (1927) Selected Papers of Bertrand Russell, New York.
Russell, B. (1962) “The Taming of Power” In the Basic Writings of Bertrand Russell (1903-
1959), R.E. Egner, & Dennon (eds), New York: Simon & Schuster .
Schumacher, E.F. (1979) Small is Beautiful; New York, Harper and Row Publishers.
Time, December 1999:59).
Young, R.M. (1971) “Evolutionary Biology and Ideology”, in Watson fuller (ed) The Social
Impact of the Modern Biology, London, Routledge and Kegan paul.

Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate talks about such non-empirical factors on non-empirical factors on pp.124-127.
Tyson, Neil D. “The structure of science.”
“science, philosophy of.”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 121.
Schick, Theodore “The end of science?”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 105.; Ratzsch, Del The Philosophy of Science pp.14-20—see this source for
further discussion.
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 104.
Goodstein “Conduct, Misconduct, and the Structure of Science.”

Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 111
Ratzsch, Del The Philosophy of Science chapter 2
“hypothetico-deductive method.”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate p. 110
Ibid. p. 108; “hypothetico-deductive method.”
I briefly talk about why we don't have any hard proof that memory, testimony, and sense
experience of being reliable at
http://www.angelfire.com/mn2/tisthammerw/rlgnphil/skepticism.html.
Schick, Theodore “The end of science?”
Ibid.; Rosen, Kenneth H. Discrete Mathematics and its applications p. 172
Ibid.; Glasner, David “Karl Popper, critical rationalist. (modern philosopher).”
Wolpert, Lewis “Science: The art of the insoluble?”
“Karl Popper.”
Glasner, David “Karl Popper, critical rationalist.(modern philosopher)”
Popper, Karl Conjectures and Refutations: The Growth of Scientific Knowledge p. 46
Schick, Theodore “The end of science?”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Ibid.; Glasner, David “Karl Popper, critical rationalist.(modern philosopher)”
Schick, Theodore “The end of science?”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Wolpert, Lewis “Science, the art of the insoluble?”; Wolpert, Lewis “Hypotheses”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Schick, Theodore “The end of science?”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Ibid.
Carroll, Robert Todd “Skeptic’s Dictionary: ad hoc hypothesis.”
Trinklein, Frederick E. “Modern Physics.” p. 56
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. pp. 131,132
Carroll, Robert Todd “Skeptic’s Dictionary: Occam’s razor.”
— Regarding an apparently irrational theory conforming to simplicity, he merely dismisses
the theory as “simpleminded” without giving further reasoning to why it should be
rejected.
Miller, David “Skepticism and Relativism.”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 191
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. p. 110
National Academy of Science 1996, p. 201; Ratzsch, Del The Battle of Beginnings: Why
Neither Side is Winning the Creation- Evolution Debate. p. 126
Harre, Rom Obituary: Professor Thomas S. Kuhn
“Science” The Columbia Encyclopedia
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. pp. 127-136
Ibid. pp. 130
Goodstein “Conduct, Misconduct, and the Structure of Science.”; Goodstein, David “What do
we mean when we use the term ‘science fraud’?”
— The author calls the mistaken belief “the Myth of the Noble Scientist.”
Goodstein, David “What do we mean when we use the term ‘science fraud’?”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation- Evolution
Debate. pp. 116, 128-129, 174
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
I know I wouldn’t.
Schafersman, Steven D. “An Introduction to Science.”
— Here the author is an example.
Dembski, William “Disbelieving Darwin—And Feeling No Shame!”; Milton, Richard
“Alternative Science.”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
— The authors suggest that tenacity may be necessary for scientific developments to be
successful.
Dembski, William “Disbelieving Darwin—And Feeling No Shame!”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Shapin, Steven “History of Science and Its Sociological Reconstructions.” p. 160; Gould,
Stephen Jay “Bathybius Meets Eozoon.” p. 18
Klotz, Irving “The N-Ray Affair.” p. 168ff
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation-Evolution
Debate p. 124
Goodstein, David “What Do We Mean When We Use the Term ‘Science Fraud?’”
Ibid.
Gould, Stephen Jay “Smith Woodward’s Folly.”
Kathloff, Mark “God and Creation: A Historical Look at Encounters Between Christianity
and Science.” In Bauman pp. 5-30
Davidson, Aaron “Science as a Belief System.”
Wertheim, Margaret “Science & Religion: Blurring the Boundaries.”
Gingerich, Owen “How Galileo Changed the Rules of Science.”
Witham, Larry “Creation-evolution debate takes on a less-shrill tone.”
— Niles Eldredge, a curator at the American Museum of Natural History holds this view.
Wertheim, Margaret “Science & Religion: Blurring the Boundaries.”
Wolpert, Lewis “Hypotheses.”
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation-Evolution
Debate. p. 166
Helweg, Otto J. “SCIENTIFIC FACTS: Compatible with Christian Faith?”
Ibid.
Ibid.
Bede “Christianity and the rise of modern science”
Helweg, Otto J. “SCIENTIFIC FACTS: Compatible with Christian Faith?”
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”; Bridgman, Percy W. “On Scientific Method”
Ibid.
American Association for the Advancement of Science 1993, p. 7; Ibid.
Woodward, James and David Goodstein “Conduct, Misconduct, and the Structure of
Science.”
Ibid.
Ibid.
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation-Evolution
Debate. p. 127
Wipond, Rob “The World is Round (and other mythologies of modern science). (Exploring
the Foundations of Humanism)”
Ratzsch, Del “Recapitulations”; Ratzsch, Del The Battle of Beginnings: Why Neither Side is
Winning the Creation-Evolution Debate. p. 165
Ratzsch, Del The Battle of Beginnings: Why Neither Side is Winning the Creation-Evolution
Debate. p. 132
Reines, Frederick “Who Needs Science?”
Ibid.
• Adickes, E., 1911, Untersuchungen zu Kants Physischer Geographie, Tübingen:
Mohr.
• –––, 1920, Kants Opus postumum, dargestellt und beurteilt, Berlin: Reuther &
Reichard.
• –––, 1924-1925, Kant als Naturforscher, (2 vols.) Berlin: De Gruyter.
• Blasche, S., 1991, Übergang. Untersuchungen zum Spätwerk Immanuel Kants,
Frankfurt: Klostermann.
• Brandt, R., 1999, Kritischer Kommentar zu Kants Anthropologie in pragmatischer
Hinsicht, Hamburg: Felix Meiner Verlag.
• Brittan, G., 1978, Kant's Theory of Science, Princeton: Princeton University Press.
• –––, 1995, “The Continuity of Matter,” in Proceedings of the Eighth International
Kant Congress, H. Robinson (ed.), Milwaukee: Marquette University Press, vol. I/2 ,
pp. 611-618.
• Buchdahl, G., 1968, “Gravity and Intelligibility: Newton to Kant,” in The
Methodological Heritage of Newton, R. Butts & J. Davis (eds.), Toronto: University
of Toronto Press, pp. 74-102; reprinted in Buchdahl 1992.
• –––, 1969, Metaphysics and the Philosophy of Science, Cambridge: Belknap Press.
• –––, 1986, “Kant's 'Special Metaphysics' and The Metaphysical Foundations of
Natural Science,” in Kant's Philosophy of Physical Science, R. Butts (ed.), Dordrecht:
D. Reidel Publishing Company, pp. 127-161.
• –––, 1992, Kant and the Dynamics of Reason, Oxford and Malden: Blackwell.
• Butts, R., 1986, “The Methodological Structure of Kant's Metaphysics of Science,” in
Kant's Philosophy of Physical Science, R. Butts (ed.), Dordrecht: D. Reidel
Publishing Company, pp. 163-199.
• –––, 1986, (ed.), Kant's Philosophy of Physical Science, Dordrecht: D. Reidel
Publishing Company.
• Carrier, M., 1990, “Kants Theorie der Materie und ihre Wirkung auf die
zeitgenössische Chemie,” Kant-Studien, 81: 170-210.
• –––, 1992, “Kant's Relational Theory of Absolute Space,” Kant-Studien, 83: 399-416.
• –––, 2001, “Kant's Theory of Matter and His Views on Chemistry,” in Kant and the
Sciences, E. Watkins (ed.), New York: Oxford University Press, pp. 205-230.
• Cramer, K., 1985, Nicht-reine synthetische Urteile a priori. Ein Problem der
Transzendentalphilosophie Immanuel Kants, Heidelberg: Winter.
• Duncan, H., 1984, “Inertia, the Communication of Motion, and Kant's Third Law of
Mechanics,” Philosophy of Science, 51: 93-119.
• Edwards, J., 2000, Substance, Force, and the Possibility of Knowledge: On Kant's
Philosophy of Material Nature, Berkeley: University of California Press.
• Emundts, D., 2004, Kants Uebergangzkonzeption im Opus postumum, Berlin: De
Gruyter.
• Engelhard, K., 2005, Das Einfache und die Materie, Berlin: De Gruyter.
• Förster, E., 2000, Kant's Final Synthesis: An Essay on the Opus postumum,
Cambridge: Harvard University Press.
• Friedman, M., 1989, “Kant on Laws of Nature and the Foundations of Newtonian
Science,” in Proceedings of the Sixth International Kant Congress, G. Funke and T.
Seebohm (eds.), Washington: University Press of America, Vol. II/2, pp. 97-107.
• –––, 1990, “Kant and Newton: Why Gravity is Essential to Matter,” in Philosophical
Perspectives on Newtonian Science, P. Bricker and R.I.G. Hughes (eds.), Cambridge:
pp. 185-202.
• –––, 1992, Kant and the Exact Sciences, Cambridge: Harvard University Press.
• –––, 1995, “Matter and Material Substance in Kant's Philosophy of Nature: The
Problem of Infinite Divisibility,” in Proceedings of the Eighth International Kant
Congress, Vol. I/2, Milwaukee: Marquette University Press, pp. 595-610.
• –––, 2001, “Matter and Motion in the Metaphysical Foundations and the First
Critique: The Empirical Concept of Matter and the Categories,” in Kant and the
Sciences, E. Watkins (ed.), New York: Oxford University Press, pp. 53-69.
• –––, 2002, “Kant on Science and Experience,” in Kant und die Berliner Aufklärung,
V. Gerhardt, R.-P. Horstmann, and R. Schumacher (eds.), Berlin: De Gruyter, Vol. I,
pp. 233-245.
• Gloy, K., 1976, Die Kantische Theorie der Naturwissenschaft. Eine Strukturanalyse
ihrer Möglichkeit, ihres Umfangs und ihrer Grenzen, Berlin: de Gruyter.
• Guyer, P., 2001, “Organisms and the Unity of Science,” in Kant and the Sciences, E.
Watkins (ed.), New York: Oxford University Press, pp. 259-281.
• Harman, P., 1982, Metaphysics and Natural Philosophy, Brighton/Sussex: Harvester
Press.
• Hoppe, H., 1969, Kants Theorie der Physik. Eine Untersuchung über das Opus
postumum von Kant, Frankfurt: Klostermann.
• Kitcher, P., 1990, Kant's Transcendental Psychology, New York: Oxford University
Press.
• Kitcher, Ph., 1983, “Kant's Philosophy of Science,” in Midwest Studies in Philosophy
V. VIII Contemporary Perspectives on the History of Philosophy, P. French, T.
Uehling, and H. Wettstein (eds.), Minneapolis: University of Minnesota Press, pp.
387-408.
• Kleingeld, P., 1995, Fortschritt und Vernunft: Zur Geschichtsphilosophie Kants,
Würzburg: Königshausen and Neumann.
• –––, 1999, “Kant, History, and the Idea of Moral Development,” History of
Philosophy Quarterly, 16: 59-80.
• Kuehn, M., 2001, “Kant's Teachers in the Exact Sciences,” in Kant and the Sciences,
E. Watkins (ed.), New York: Oxford University Press, pp.11-30.
• Laywine, A., 1993, Kant's Early Metaphysics and the Origins of the Critical
Philosophy, Vol. 3 in North American Kant Society Studies in Philosophy,
Atascadero: Ridgeview Publishing Company.
• Lefevre, W., and Wunderlich, F., 2000, Kants naturtheoretische Begriffe (1747-
1780), Berlin: De Gruyter.
• –––, 2001, (ed.) Between Leibniz, Newton and Kant: Philosophy and Science in the
18th Century, Dordrecht: Kluwer Academic Publishers.
• Makkreel, R., 2001, “Kant on the Scientific Status of Psychology, Anthropology, and
History,” in Kant and the Sciences, E. Watkins (ed.), New York: Oxford University
Press, pp. 185-201.
• Malzkorn, W., 1998, “Kant über die Teilbarkeit der Materie,” Kant-Studien, 89: 385-
409.
• Palter, R., 1972, “Kant's Formulation of the Laws of Motion,” Synthese, 24: 96-111.
• Plaass, P., 1965, Kants Theorie der Naturwissenschaft. Eine Untersuchung zur
Vorrede von Kants ‘Metaphysischen Anfangsgründen der Naturwissenschaft’,
Göttingen: Vandenhoek & Ruprecht; available in English translation, Kant's Theory
of Natural Science, A. Miller and M. Miller (translators), Dordrecht: Kluwer, 1994.
• Pollok, K., 2001, Kants 'Metaphysische Anfangsgründe der Naturwissenschaft'. Ein
kritischer Kommentar, Hamburg: Felix Meiner Verlag.
• –––, 2002, “Fabricating a World in Accordance with Mere Fantasy...? The Origins of
Kant's Critical Theory of Matter,” Review of Metaphysics, 56: 61-97.
• Schäfer, L., 1966, Kants Metaphysik der Natur, Berlin: De Gruyter.
• Schönfeld, M., 2000, The Philosophy of the Young Kant: The Precritical Project,
New York: Oxford University Press.
• Sturm, T., 2001, “Kant on Empirical Psychology: How Not to Investigate the Human
Mind,” in Kant and the Sciences, E. Watkins (ed.), New York: Oxford University
Press, pp. 163-184.
• Tuschling, B., 1971, Metaphysische und Transzendentale Dynamik in Kants Opus
postumum, Berlin: De Gruyter.
• Walker, R.C.S., 1974, “The Status of Kant's Theory of Matter”, in Kant's Theory of
Knowledge, L.W. Beck (ed.), Dordrecht: Reidel, pp. 151-156.
• Daniel Warren, 2001, Reality and Impenetrability in Kant's Philosophy of Nature,
London: Routledge.
• Watkins, E., 1997, “The Laws of Motion from Newton to Kant,” Perspectives on
Science, 5: 311-348.
• –––, 1998a, “The Argumentative Structure of Kant's Metaphysical Foundations of
Natural Science,” Journal of the History of Philosophy, 36: 567-593.
• –––, 1998b, “Kant's Justification of the Laws of Mechanics,” in Studies in History
and Philosophy of Science, 29: 539-60.
• –––, 2001, (ed.), Kant and the Sciences, New York: Oxford University Press.
• –––, 2001, “Kant on Force and Extension: Critical Appropriations of Leibniz and
Newton” in Between Leibniz, Newton and Kant: Philosophy and Science in the 18th
Century, W. Lefevre (ed.), Dordrecht: Kluwer Academic Publishers, pp. 111-127.
• –––, 2003, “Forces and Causes in Kant's Early Pre-Critical Writings,” Studies in
History and Philosophy of Science, 33: 5-27.
• Westphal, K., 1995, “Does Kant's Metaphysical Foundations of Natural Science fill a
gap in the Critique of Pure Reason?” Synthese, 103:43-86.
• Wood, A., 1999, Kant's Ethical Thought, New York: Cambridge University Press.
• Yovel, Y., 1980, Kant and the Philosophy of History, Princeton: Princeton University
Press.

Das könnte Ihnen auch gefallen