atomism

atomism
atomist, n.atomistic, atomistical, adj.atomistically, adv.
/at"euh miz'euhm/, n.
1. Also called atomic theory. Philos. the theory that minute, discrete, finite, and indivisible elements are the ultimate constituents of all matter.
2. Psychol. a method or theory that reduces all psychological phenomena to simple elements.
[1670-80]

* * *

Philosophical doctrine that material objects are aggregates of simpler parts known as atoms.

Atomism in the strict sense is characterized by three points: the atoms are absolutely indivisible, qualitatively identical apart from shape, size, and motion, and combinable with each other only by juxtaposition. Atomism is usually associated with realism and mechanism; it is mechanistic because it maintains that all observable changes can be reduced to changes in the configuration of the atoms that constitute matter. It is opposed to holism because it holds that the properties of any whole can be explained in terms of those of its parts.

* * *

Introduction

      any doctrine that explains complex phenomena in terms of aggregates of fixed particles or units. This philosophy has found its most successful application in natural science: according to the atomistic view, the material universe is composed of minute particles, which are considered to be relatively simple and immutable and too small to be visible. The multiplicity of visible forms in nature, then, is based upon differences in these particles and in their configurations; hence any observable changes must be reduced to changes in these configurations.

The basic nature of atomism
      Atomism is in essence an analytical doctrine. It regards observable forms in nature not as intrinsic wholes but as aggregates. In contrast to holistic theories, which explain the parts in terms of qualities displayed by the whole, atomism explains the observable properties of the whole by those of its components and of their configurations.

      In order to understand the historical development of atomism and, especially, its relation to modern atomic theory, it is necessary to distinguish between atomism in the strict sense and other forms of atomism. Atomism in the strict sense is characterized by three points: the atoms are absolutely indivisible, qualitatively identical (i.e., distinct only in shape, size, and motion), and combinable with each other only by juxtaposition. Other forms of atomism are less strict on these points.

      Atomism is usually associated with a “realistic (realism)” and mechanistic (mechanism) view of the world. It is realistic in that atoms are not considered as subjective constructs of the mind employed for the sake of getting a better grip upon the phenomena to be explained; instead, atoms exist in actual reality. By the same token, the mechanistic view of things, which holds that all observable changes can be reduced to changes of configuration, is not merely a matter of employing a useful explanatory model; the mechanistic thesis holds, instead, that all observable changes are caused by motions of the atoms. Finally, as an analytic doctrine atomism is opposed to organismic doctrines, which teach that the nature of a whole cannot be discovered by dividing it into its component parts and studying each part by itself.

Various senses of atomism
      The term atomism is derived from the Greek word atoma—“things that cannot be cut or divided.”

Two basic types of atomism
      The history of atomism can be divided into two more or less distinct periods, one philosophical and the other scientific, with a transition period between them (from the 17th to the 19th century). This historical fact justifies the distinction between philosophical and scientific atomism.

Philosophical atomism
      In philosophical atomism, which is as old as Greek philosophy, attention was focussed not on the detailed explanation of all kinds of concrete phenomena but on some basic general aspects of these phenomena and on the general lines according to which a rational explanation of these aspects was possible. These basic aspects were the existence in nature of a manifold of different forms and of continuous change. In what way could these features be explained? Philosophical atomism offered a general answer to that question. It did not, however, strictly confine itself to the general problem of explaining the possibility of change and multiplicity—not even in ancient Greek atomism, for in Greek thought philosophy and science still formed a unity. Consequently, atomists also tried to give more detailed explanations of concrete phenomena, such as evaporation, though these explanations were meant more to endorse the general doctrine of atomism than to establish a physical theory in the modern sense of the word. Such a theory was not yet possible, because a physical theory must be based upon indirect or direct information about the concrete properties of the atoms involved, and such information was not then available.

      With the development of a scientific atomic theory, the general philosophical problems gradually disappear into the background. All attention is focussed on the explanation of concrete phenomena. The properties of the atoms are determined in direct relationship with the phenomena to be explained. For this reason the chemical atomic theory of the 19th century supposes that each identified chemical element has its own specific atoms and that each chemical compound has its own molecules (molecule) (fixed combinations of atoms). What particles act as unchanged and undivided units depends upon what kind of process is involved. Some phenomena, such as evaporation, are explained by a process in which the molecules remain unchanged and identical. In chemical reactions, however, the molecules lose their identity. Their structures are broken up, and the composing atoms, while retaining their own identity, are rearranged into new molecules. With nuclear reactions a new level is reached, on which the atoms themselves are no longer considered as indivisible: more elementary particles than the atoms appear in the explanations of nuclear reactions.

Extensions to other fields
      Whereas classical atomism spoke mainly of material atoms (i.e., of particles of matter), the success of the atomic doctrine encouraged the extension of the general principles of atomism to other phenomena, more or less removed from the original field of application. Rather plausible, for example, was the extension of atomism to the phenomena of electricity. There were reasons to suppose the existence of an elementary charge of electricity associated with an elementary material particle, the electron (19th century). A second fruitful extension concerned energetic processes (20th century). Some experimental data suggested the hypothesis that energy can exist only in amounts that are whole multiples of an elementary quantity of energy. Extensions of the idea of an atomic structure to amounts of gravitation and even to time have been attempted but have not been sufficiently confirmed.

      More removed from the original field of application of atomism is a theory known as Logical Atomism (developed by the eminent philosopher and logician Bertrand Russell (Russell, Bertrand) and by the philosopher of language Ludwig Wittgenstein (Wittgenstein, Ludwig)), which supposes that a perfect isomorphism exists between an “atom” of language (i.e., an atomic proposition) and an atomic fact; i.e., that for each atomic fact there is a corresponding atomic proposition. An atomic proposition is one that asserts that a certain thing has a certain quality; e.g., “this is red.” An atomic fact is the simplest kind of fact and consists in the possession of a quality by some individual thing.

      Another application of atomism (albeit in a moot sense) lies in the monadology (monad) of the philosopher-scientist G.W. Leibniz (Leibniz, Gottfried Wilhelm). According to Leibniz the atoms of Democritus, who provides the paradigm case of ancient Greek atomism, were not true unities; possessing size and shape, they still were divisible in principle. The ultimate constituents of things must, therefore, be points, said Leibniz—not mathematical but metaphysical points; i.e., points of real existence. They are indeed a kind of soul, which he came to call “monads.”

      In psychology, atomism is a doctrine about perception. It holds that what man perceives is a mosaic of atomic sensations, each independent and unconnected with any other sensation. According to the early modern Empiricist David Hume (Hume, David) and the pre-World War I father of experimental psychology Wilhelm Wundt (Wundt, Wilhelm), the fact that man nevertheless experiences an ordered whole formed from the unordered “atoms” of perception is caused by the mind's capacity to combine them by “ association.”

Diverse philosophical characterizations of atomism

The intrinsic nature of the atoms
      In 1927 the Belgian astronomer Georges Lemaître formulated the hypothesis that the present high degree of differentiation of matter in space and the complexity of forms displayed by the various astronomical objects must have resulted from a violent explosion (big-bang model) and subsequent dispersal of an originally highly compressed homogeneous material, a kind of “primitive atom,” containing all of the matter that exists. From the philosophical viewpoint this hypothesis is interesting. By its attempt to reduce the manifold to unity it recalls the beginning of Greek philosophy, which was also inspired by a thesis of the unity of being, propounded by the Eleatic Parmenides. Even apart from their respective contexts, there is, of course, a great difference between Lemaître's and Parmenides' conceptions of the unity of being, for the latter combined the thesis of the unity of being with that of the immutability of being.

      Although it would be wrong to classify Parmenides among the atomists, it is nonetheless appropriate that in an introduction to the diverse forms of atomism, his conception of reality as just one being should be mentioned. Parmenides' thesis is not only historically but also logically the cornerstone of atomistic thought. Any atomic theory can be interpreted as an attempt to reconcile the thesis of the unity and immutability of being with the fact that the senses observe multiplicity and change. The different ways in which the unity and immutability are understood characterize the different forms of atomism.

Atoms as lumpish corpuscles
      As corpuscles (minute particles), atoms can either be endowed with intrinsic qualities or be inherently qualityless.

      The most striking basic differences in the material world, which lead to a first classification of substances in nature (nature, philosophy of), are those between solids, liquids, gases, and fire. These differences are an observed datum that must be accounted for by every scientific theory of nature. It is, therefore, only natural that one of the first attempts to explain the phenomena of nature was based upon these differences and proclaimed that there are four qualitatively different primitive constituents of everything, namely, the four elements: earth, water, air, and fire ( Empedocles, 5th century BC)—a theory that dominated physics and chemistry until the 17th century.

      Although the theory of the four elements is not necessarily an atomistic theory, it obviously lends itself to interpretation in atomistic terms, namely, when the elements are conceived as smallest parts that are immutable. In this case, all observable changes are reduced to the separation and commingling of the primitive elementary substances. Thus Parmenides' thesis that being is immutable is maintained, whereas the absolute unity of being is abandoned. Yet, the fact that the infinite variety of forms and changes in nature is reduced to just one type of process between only four elementary kinds of atoms shows its affinity with the thesis of the unity of all being.

      Notwithstanding the great disparity between the theory of the four elements and modern chemistry, it is clear that modern chemistry with its approximately 100 qualitatively different atoms falls into the same class of atomic theories as that of Empedocles. There are differences, of course, but these will be deferred for later discussion.

      More removed from the original thesis of Parmenides was the theory of his contemporary Anaxagoras of Clazomenae, which assumed as many qualitatively different “atoms” as there are different qualitied substances in nature. Inasmuch as these atoms, which Anaxagoras called “seeds,” were eternal and incorruptible, this theory still contains an idea borrowed from Parmenides. A special feature of Anaxagoras' theory was that every substance contains all possible kinds of seeds and is named after the kind of seed that predominates in it. Since the substance contains also other kinds of seed, it can change into something else by the separation of its seeds.

      Another interesting form of atomism with inherently qualitied atoms, also based on the doctrine of the four elements, was proposed by Plato. On mathematical grounds he determined the exact forms that the smallest parts of the elements must have. Fire has the form of a tetrahedron, air of an octahedron, water of an icosahedron, and earth of a cube. Inasmuch as he characterized the atoms of the four elements by different mathematical forms, Plato's conception can be considered as a transition between the qualitative and quantitative types of atomism.

      The most significant system of atomism in ancient philosophy was that of Democritus (5th century BC). Democritus agreed with Parmenides on the impossibility of qualitative change but did not agree with him on that of quantitative change. This type of change, he maintained, is subject to mathematical reasoning and therefore possible. By the same token, Democritus also denied the qualitative multiplicity of visible forms but accepted a multiplicity based on purely quantitative differences. In order to reduce the observable qualitative differences to quantitative differences, Democritus postulated the existence of invisible atoms, characterized only by quantitative properties: size, shape, and motion. Observed qualitative changes are based upon changes in the combination of the atoms, which themselves remain intrinsically unchanged. Thus Democritus arrived at a position that was defined above as atomism in the strict sense. In order to make the motion of atoms possible, this atomism had to accept the existence of the void (empty space) as a real entity in which the atoms could move and rearrange themselves. By accepting the void and by admitting a plurality of beings, even an infinite number of them, Democritus seemed to abandon—even more than Empedocles did—the unity of being. Nevertheless, there are sound reasons to maintain that, in spite of this doctrine of the void, Democritus' theory remained close to Parmenides' thesis of the unity of being. For Democritus' atoms were conceived in such a way that almost no differences can be assigned to them. First of all, there are no qualitative differences; the atoms differ only in shape and size. Secondly, the latter difference is characterized by continuity; there are no privileged shapes and no privileged sizes. All shapes and sizes exist, but they could be placed in a row in such a manner that there would be no observable difference between successive shapes and sizes. Thus not even the differences in shape and size seem to offer any ground explaining why atoms should be different. By accepting an infinite number of atoms, Democritus retained as much as possible the principle that being is one. With respect to the acceptance of the void, it must be stressed that the void in the eyes of Democritus is more nonbeing than being. Thus even this acceptance does not seriously contradict the unity of being.

Atoms as sheer extension
      Democritus had declared quantitative differences to be intelligible, because they were subject to mathematical reasoning. Precisely this relationship between quantitative differences and mathematics made it impossible for Descartes (Descartes, René) (17th century) to think along the atomistic lines of Democritus. If the only thing that is clearly understandable in matter is mathematical proportions, then matter and spatial extension are the same—a conclusion that Descartes did not hesitate to draw. Consequently, he rejected not only the idea of indivisible atoms but also that of the void. In his eyes the concept “void” is a contradiction in terms. Where there is space, there is by definition extension and, therefore, matter.

      Yet, however strange it may seem in view of his identification of matter with extension, Descartes offered nonetheless a fully developed theory of smallest particles. To the questions that arise immediately as to how these particles are separated and distinct from each other, Descartes answered that a body or a piece of matter is all of that which moves together. In the beginning of the world all matter was divided into particles of equal size. These particles were in constant motion and filled all of space. As, however, there was no empty space for moving particles to move into, they could only move by taking the places vacated by other particles that, however, were themselves in motion. Thus the motion of a single particle involved the motion of an entire closed chain of particles, called a vortex. As a result of the original motion, some particles were gradually ground into a spherical form, and the resulting intermediary space became filled with the surplus splinters or “grindings.” Ultimately, three main types of particles were formed: (1) the splinter materials, which form the finest matter and possess the greatest velocity; (2) the spherical particles, which are less fine and have a smaller velocity; and (3) the biggest particles, which originated from those original particles that were not subject to grinding and became united into larger parts.

      Thus Descartes could construct an atomic theory without atoms in the classical sense. Although this theory as such has not been of great value for the scientific atomic theory of modern times, its general tendency was not without importance. However arbitrarily and speculatively Descartes may have proceeded in the derivation of the different kind of corpuscles, he finally arrived at corpuscles characterized by differences in mass, velocity, amount of motion, etc.—properties that could be treated mathematically.

Atoms as centres of force: dynamic particles
      Most systems of atomism depict the action between atoms in terms of collision—i.e., as actual contact. In Newton's theory of gravitation (Newton's law of gravitation), however, action between bodies is supposed to be action at a distance—which means that the body in question acts everywhere in space. As its action is the expression of its existence, it is difficult to confine its existence to the limited space that it is supposed to occupy according to its precise shape and size. There is, therefore, no reason for a sharp distinction between occupied and empty space. Consequently, the mind finds it natural to consider the atoms not as extended particles but as point-centres of force. This conception was worked out by the Dalmatian scientist R.G. Boscovich (Boscovich, Ruggero Giuseppe) (1711–87), who attempted to account for all known physical effects in terms of action at a distance between point-particles, dynamic centres of force.

Atoms as psychophysical monads
      The idea of applying the atomistic conceptions not only to material but also to psychical phenomena is as old as atomism itself. Democritus had spoken of the atoms of the soul. According to the principles of his doctrine these atoms could differ only quantitatively from those of the body: they were smoother, rounder, and finer. This made it easy for them to move into all parts of the body. Basically, however, the atoms of the soul were no less material than other atoms.

      In Leibniz's (Leibniz, Gottfried Wilhelm) monadology (monad) the situation was quite different. Leibniz did not first conceive of material atoms and then only later interpret the soul in terms of these atoms; from the beginning he conceived his “atoms,” the monads, in terms of an analogy with the soul. A monad is much more a spiritual than a material substance. Monads have no extension; they are centres of action but not, first of all, in the physical sense. Each of the monads is gifted with some degree of perception; each mirrors the universe in its own way. Monads differ from each other, however, in the degree of perception of which they are capable.

The immutability of atoms
      By their nature all atomic theories accept a certain degree of immutability of the atoms. For without any fixed units no rational analysis of complex phenomena is possible. At least with respect to the stable factors in the analysis involved, the atoms have to be considered as immutable. According to atomism in the strict sense, this immutability had to be interpreted in an absolute way.

      The same absolute interpretation appeared in classical chemistry, although its atomic theory deviated from atomism in the strict sense by assuming qualitatively different atoms and by assuming molecules (rather stable aggregates of atoms). The decisive point, however, is that molecules were formed by mere juxtaposition of atoms without any intrinsic change of the qualities of the atoms. Modern atomic theory, in contrast, gives a less rigid interpretation of the immutability of elementary particles: the particles that build up an atom do not retain their identity in an absolute way.

      In some philosophical atomistic theories, the immutability of the atoms has been understood in a highly relative sense. This interpretation arose mainly in the circles of those Aristotelian (Aristotelianism) philosophers who tried to combine atomistic principles with the principle of Aristotle that elements changed their nature when entering a chemical compound. The combination of both principles led to the doctrine known as the minima naturalia theory, which holds that each kind of substance has its specific minima naturalia, or smallest entities in nature. Minima naturalia are not absolutely indivisible: they can be divided but then become minima naturalia of another substance; they change their nature. In a chemical reaction the minima of the reagents change into the minima of the substances that result from the reaction.

Other differences
      Atomisms also differ regarding the number of atoms, whether they occupy a void, and how they relate to one another.

Number of atoms
      As has already been mentioned, Democritus introduced the hypothesis that the atoms are infinite in number. Although one may question whether the term infinite has to be taken in its strict sense, there is no doubt that by using this term Democritus wanted not merely to express the triviality that, on account of their smallness, there had to be an enormous quantity of atoms. Democritus also had a strong rational argument for postulating a strictly infinite quantity of atoms: only thus could he exclude the existence of atoms that specifically differed from each other.

      When in modern science the problem of the number of atoms arises, the situation is quite different from that of the Greek atomists. There is now much more detailed information about the properties of the atoms and of the elementary particles, and there is also in astrophysical cosmology some information about the universe as a whole. Consequently, the attempt to calculate the total number of atoms that exist is not entirely impossible, although it remains a highly speculative matter. In a time (around 1930) when all chemical atoms were supposed to be composed of electrons and protons, the pioneering joint-relativity-quantum astrophysicist A.S. Eddington (Eddington, Sir Arthur Stanley) calculated the number of these elementary particles to be 2 × 136 × 2256, or approximately 1079, arguing that, since matter curves space, this is just the number of particles required barely to close the universe up into a hypersphere and to fill up all possible existence states.

The existence of the void
      To Democritus the existence of the void was a necessary element in atomistic theory. Without the void the atoms could not be separated from each other and they could not move. In the 17th century Descartes rejected the existence of the void, whereas Newton's (Newton, Sir Isaac) conception of action at a distance was in perfect harmony with the acceptance of the void and the drawing of a sharp distinction between occupied and nonoccupied space.

      The success of the Newtonian law of gravitation was one of the reasons that atomic theories came to prevail in the 18th century. Even with respect to the phenomena of light, the corpuscular and hence atomic theory of Newton, which held that light is made of tiny particles, was adopted almost universally, in spite of Huygens' brilliant development of the wave hypothesis.

      When in the beginning of the 19th century the corpuscular theory of light in its turn was abandoned in favour of the wave theory (electromagnetic radiation), the case for the existence of the void had to be reopened, for the proponents of the wave theory did not think in terms of action at a distance; the propagation of waves seemed to presuppose, instead, a medium with not only geometrical properties but with physical ones as well. At first the physical properties of the medium, the ether, were described in the language of mechanics; later they were described in that of the electromagnetic field theory of J.C. Maxwell. Yet, to a certain extent the old dichotomy between occupied and nonoccupied space continued to exist. For, according to the ether theory, the atoms moved without difficulty in the ether, whereas the ether pervaded all physical bodies.

      In contemporary science this dichotomy has lost its sharpness, owing to the fact that the distinction between material phenomena, which were supposed to be discontinuous, and the phenomena of light, which were supposed to be continuous, appears to be only a relative one. In conclusion it can be claimed that although modern theories still speak of space and even of “empty” space, this “emptiness” is not absolute: space has come to be regarded as the seat of the electromagnetic field, and it certainly is not the void in the sense in which the term was used by Democritus.

Atoms in external aggregation versus in internal relationship
      In most forms of atomism it is a matter of principle that any combination of atoms into a greater unity can only be an aggregate of these atoms. The atoms remain intrinsically unchanged and retain their identity. The classical atomic theory of chemistry was based upon the same principle: the union of the atoms into the molecules of a compound (chemical compound) was conceived as a simple juxtaposition. Each chemical formula (e.g., H2O, H2SO4, NaCl, etc.) reflects this principle through the tacit implication that each atom is still an H, O, or S, etc., even when in combination to form a molecule.

      Chemistry had a twofold reason for adopting this principle. One reason was an observational, the other a philosophical one. The fact that some of the properties of a chemical compound could, by simple juxtaposition, be derived from those of the elements (the molecular weight, for example, equals the simple sum of the respective atomic weights) was a strong factual argument in favour of the principle. Many properties of the components, however, could not be determined in this way. In fact, most chemical properties of compounds differed considerably from those of the composing elements. Consequently, the principle of juxtaposition could not be based on factual data alone. It was in need of a more general support. This support was offered by the philosophical idea that inspired all atomism, viz., that if complex phenomena cannot be explained in terms of aggregates of more elementary factors, they cannot be explained at all.

      For the evaluation of this idea, the development of the scientific atomic theory is highly interesting, especially with respect to the interpretation of the concept of an aggregate. Is the only interpretation of this concept that of an assemblage in which the components preserve their individuality—like, for instance, a heap of stones?

      Modern atomic theory offers an answer to this question. This theory still adheres to the basic principle that a complex structure has to be explained in terms of aggregates of more elementary factors, but it interprets the term “aggregate” in such a way that it is not limited to a mere juxtaposition of the components. In modern theories atomic and molecular structures are characterized as associations of many interacting entities that lose their own identity. The resulting aggregate originates from the converging contributions of all of its components. Yet, it forms a new entity, which in its turn controls the behaviour of its components. Instead of mere juxtaposition of components, there is an internal relationship between them. Or, expressed in another way: in order to know the properties of the components, one has to study not only the isolated components but also the structures into which they enter. To a certain extent modern atomic theory has bridged the gap between atomistic and holistic thought.

History and major representatives of the various atomisms

Philosophical atomism
      From the ancient Greeks through the 16th century, atomism remained mainly philosophical.

Ancient Greek atomism
      It is characteristic of the importance of Greek philosophy that, already in the foregoing exposition of the different aspects of atomism, several Greek philosophers had to be introduced. Not only the general idea of atomism but also the whole spectrum of its different forms originated in ancient Greece. As early as the 5th century BC atomism in the strict sense (Leucippus and Democritus) is found, along with various qualitative forms of atomism: that of Empedocles, based on the doctrine of the four elements, and that of Anaxagoras, with as many qualitatively different atoms as there are different substances.

      Yet, in spite of its successful start, atomism did not gain preeminence in Greek thought. This is mainly because Plato and Aristotle were not satisfied with the atomistic solution of the problems of change as a general solution. They refused to reduce the whole of reality, including man, to a system that knew nothing but moving atoms. Even with respect to the problems of the material world, atomism seemed to offer no sufficient explanation. It did not explain the observable fact that, notwithstanding continual changes, a total order of specific forms continued to exist. For this reason Aristotle, with Plato, was more interested in the principle of order than in that of the material elements. In his own analysis of change, which resulted in the matter-form doctrine, Aristotle explicitly rejected the thesis of Democritus that in a chemical reaction the component parts retain their identity. According to Aristotle, the elements that entered into a composite with each other did not remain what they were but became a compound. Although there is some indication that in Aristotle's chemical theory smallest particles played a role, it was certainly not a very important one.

      Meanwhile, atomistic ideas remained known in Greek thought. Their opponents paid much attention to them, and there were also a few adherents of Democritean atomism in later times, such as the Greek hedonist Epicurus (c. 341–279 BC) and the Roman poet Lucretius Carus (Lucretius) (c. 95–55 BC) who, through his famous didactic poem De rerum natura (“On the Nature of Things”), introduced atomism into the Latin world.

The elachista of the early Aristotelian commentators
      Empedocles had suggested an atomism with qualitatively different atoms, based upon the doctrine of the four elements. Aristotle adopted the latter doctrine but without its atomistic suggestion. Certain Greek commentators on the works of Aristotle, however, viz., Alexander of Aphrodisias (2nd century AD), Themistius (4th century AD), and Philoponus (6th century AD), combined the Aristotelian theory of chemical reactions with atomistic conceptions. In their systems the atoms were called elachista (“very small” or “smallest”). The choice of this term is connected with the Aristotelian rejection of the infinite divisibility of matter. Each substance had its own minimum of magnitude below which it could not exist. If such a minimum particle were to be divided, then it would become a minimum of another substance.

The minima naturalia of the Averroists
      The Latin commentators on Aristotle translated the term elachista into its Latin equivalent minima or also into minima naturalia; i.e., minima determined by the nature of each substance. In fact, for most medieval Aristotelians the minima acquired little more reality than the theoretical limit of divisibility of a substance; and in their descriptions of physical and chemical processes, they paid no attention to the minima. With the Averroists (Latin Averroism)—followers of the Arab Aristotelian Averroës (1126–98)—an interesting development occurred. Agostino Nifo (Nifo, Agostino) (1473–1538), for example, explicitly stated that in a substance the minima naturalia are present as parts; they are physical entities that actually play a role in certain physical and chemical processes. Because the minima had acquired more physical reality, it then became necessary to know how the properties of the minima could be connected with the sensible properties of a substance. Speculations in this direction were developed by the Italian physician, philosopher, and litterateur Julius Caesar Scaliger (Scaliger, Julius Caesar) (1484–1558).

Modern scientific atomism: early pioneering work
      Modern atomism arose with the flowering of science in the present sense of the word.

The 17th century
      In the history of atomism the 17th century occupies a special place for two reasons: it saw the revival of Democritean atomism, and it saw the beginning of a scientific atomic theory.

      The revival of Democritean atomism was the work of the ambiguous Epicureo-Christian thinker Pierre Gassendi (Gassendi, Pierre) (1592–1655), who made his contemporaries not only better acquainted with atomism but also succeeded in divesting it of the materialistic interpretation with which it was hereditarily infected. This reintroduction of Democritus was well timed. Because of its quantitative character Democritus' atomism invited for its elucidation the application of mathematics and mechanics, which in the 17th century were sufficiently developed to answer this invitation. In point of fact, the 17th century was more interested in the possibilities that atomism offered for a physical theory than it was in the philosophical differences between the different atomistic systems. For this reason it saw, for example, hardly any difference between the systems of Gassendi and Descartes, although the latter explicitly rejected some of the fundamentals of Democritus, such as the existence of the void and the indivisibility of the atoms, as noted above (see Atoms as sheer extension (atomism)).

      In the case of scientists mainly interested in the chemical aspects, the same shift of emphasis from philosophical to scientific considerations can be discerned. According to the physician and philosopher of nature Daniel Sennert (1572–1637), Democritus' atomism and the minima theory really amounted to the same thing. As far as philosophy was concerned, Sennert was only interested in the general idea of atomism; the precise content of an atomic doctrine in his view ought to be a matter of chemical experimentation. His own experience as a chemist taught him the specific differences existing between the atoms. In this respect Sennert continued the minima tradition. His own contribution to the chemical atomic theory lay in the clear distinction that he made between elementary atoms and the prima mista, or atoms of chemical compounds.

      The early modern experimentalist Robert Boyle (Boyle, Robert) (1627–91) followed the same line of thought as Sennert, but he was much more aware of the discrepancy between Democritus' atomism and an atomic theory suitable for chemical purposes. Boyle's solution to this problem was the thesis that the atoms of Democritus were normally associated into primary concretions, which did not easily dissociate and which acted as elementary atoms in the chemical sense. These primary concretions can combine to form compounds of a higher order, which may be compared to Sennert's prima mista and to the molecules of modern chemistry.

Founding of modern atomism
      The 17th century had laid the theoretical foundations for a scientific atomic theory. For its further development it was in need of better chemical insights, especially concerning the problem of what substances should be considered as chemical elements. Boyle had shown conclusively that the traditional four “elements” were certainly not elementary substances, but at the same time he confessed that he did not yet see any satisfactory method to determine which substances were true elements. This method was provided by another of the principal founders of modern chemistry A.-L. Lavoisier (Lavoisier, Antoine-Laurent) (1743–94): a chemical element is a substance that cannot be further analyzed by known chemical methods.

      John Dalton (Dalton, John) (1766–1844), usually regarded as the father of modern atomic theory, applied the results of Lavoisier's chemical work to atomistic conceptions. When Dalton spoke of elementary atoms, he did not have a merely theoretical idea in mind but the chemical elements as determined by Lavoisier. Dalton held that there are as many different kinds of elementary atoms as there are chemical elements. The atoms of a certain element are perfectly alike in weight, figure, etc.; and the same point applies to the atoms of a certain compound. As weight was the decisive characteristic in Lavoisier's theory, Dalton stressed the importance of ascertaining the relative weights of atoms and the number of elementary atoms that constituted one compound “atom.” As to the question of the way in which the atoms are combined in a compound, Dalton conceived this combination as a simple juxtaposition with each atom under the influence of Newtonian forces of attraction. The atoms retain their identity through a chemical reaction. In this one point the founder of the chemical atomic theory did not differ from Democritus.

Recent and contemporary scientific atomism
      Until its development in the third decade of the 20th century, the scientific atomic theory did not differ philosophically very much from that of Dalton, although at first sight the difference may appear large. Dalton's atoms were no longer considered to be immutable and indivisible; new elementary particles sometimes appeared on the scene; and molecules were no longer seen as a mere juxtaposition of atoms—when entering into a compound atoms became ions. Yet, these differences were only accidental; the atoms revealed themselves as composed of more elementary particles—protons, neutrons, and electrons—but these particles themselves were considered then as immutable. Thus the general picture remained the same. The material world was still thought to be composed of smallest particles, which differed in nature and which in certain definite ways could form relatively stable structures (atoms). These structures were able to form new combinations (molecules (molecule)) by exchanging certain component parts (electrons). The whole process was ruled by well-known mechanical and electrodynamic laws.

      In contemporary atomic theory the differences from Dalton are much more fundamental. The hypothesis of the existence of immutable elementary particles (subatomic particle) has been abandoned: elementary particles can be transformed into radiation and vice versa. And when they combine into greater units, the particles do not necessarily preserve their identity; they can be absorbed into a greater whole.

Atomism in the thought of India
      It is interesting to note that atomistic conceptions are not restricted to Western philosophy and science. Examples of qualitative atomism, based upon the doctrine of the four elements, are also found in Indian philosophy. In some Indian systems the atoms are not absolutely indivisible but only relatively so. In certain aspects Indian atomism is, therefore, more related to the minima doctrine than to the atomism of Democritus. Indian atomism has, however, not developed into a scientific theory.

Foundational issues posed by atomism

Atomism as a metaphysical (metaphysics) system
      In discussing atomism, one particular system, that of Democritus, has been here distinguished as atomism in the strict sense because of the fact that in no other system have the foundational issues of atomism been so clearly expressed. Atomism in the strict sense is not merely one of the historical forms of atomism, one of the many possible scientific attempts at explaining certain physical phenomena; it is, first of all, a metaphysical system: it has been presented as the only possible explanation of change and multiplicity. And as a metaphysic it is rationalistic (Rationalism), mechanistic, and realistic.

      It is rationalistic because it has so much confidence in reason that, in order to explain observed phenomena, it does not hesitate to postulate the existence of unobservable atoms: i.e., of things that are in principle unobservable by the human senses and can be known only by a process of reasoning. atomists go even further, for they not only are convinced of the existence of atoms but they also think it possible to deduce in a rational way their fundamental properties. Moreover, the description of these properties in mechanistic terms is not just a matter of convenience; it is supposed to be the adequate expression of reality.

      This rationalistic and mechanistic metaphysics is not only characteristic of Democritus' atomism but also of the early forms of scientific atomism. The clearest expression of this metaphysics is found in Descartes. For Democritus mechanistic concepts are clear and distinct ideas, so that any further experimental investigation is superfluous. It should be stressed that the atomistic assumption that the human mind, by mere reasoning, can know the properties of the atoms is a necessary consequence of the idea that atoms are not subject to internal change; for the changeless can never be a subject of experimentation. The great weakness of the mechanistic concept of immutable atoms was that it forced the analyzing experiments to stop at the atoms; but this weakness could reveal itself only after, in the course of the further development of science, the fundamentally experimental character of human knowledge had become evident.

      This weakness, in fact, was precisely one of the reasons why Aristotle rejected the atomism of Democritus, viz., that the latter had postulated atoms that were not subject to change. For Aristotle the very essence of matter was its being subject to change; hence to him the concept of immutable atoms was a contradiction in terms.

      Aristotle's criticism of atomism was clearly directed against its mechanistic metaphysics, not against its realism. This latter characteristic was the target, however, of an attack launched by the incomparable 18th-century epistemologist Immanuel Kant (Kant, Immanuel). In a famous argument, known as the antinomy of the continuum, Kant tried to prove that the acceptance of the reality of spatial extension, the cornerstone of atomism, led to contradictions. His argument can be summarized as follows: It is possible to prove that any compound must be composed of simple things (for if not, there would be nothing but composition). On the other hand, it also is possible to prove that no material thing can be simple, for the very reason that a part of an extended being is always extended and is thus open to division. Hence, every allegedly simple part is at once simple and nonsimple. Consequently, spatial extension cannot be real. Extension is therefore not a property of the material world itself; it is a form imposed upon reality by man's perception.

      By his argument Kant did not intend to reject atomic theories as such; he rejected only their realistic pretensions. Kant was deeply convinced that man had to think of nature by way of analogy with a mechanism, but he denied that knowledge construed in such a fashion could reach reality itself.

      In the 19th century, scientists were, as a rule, hardly impressed by Kant's attack on the realistic pretensions of human knowledge. Scientists had already learned to go their own way and no longer worried about philosophical considerations. Only when an internal crisis in science itself arose were they prepared to reflect upon their presuppositions. Such a crisis occurred in the 20th century when science was forced to accept the relativity of both classical models: the wave and the particle models.

      To a certain extent, the problem of whether a scientific model is nothing but a subjective construct in which the scientist unites his experience is the same as the problem that Kant had in mind. One of the differences, however, is that in Kant's time science was still rather exclusively theory. Its close connection with praxis (practice, doing) had not yet been discovered. For this reason the Kantian epistemological (or human knowledge) problem could centre on such a question as: what guarantee does the knowing subject have that his “models” of reality reflect reality itself? Inasmuch as, in an exclusively theoretical science, the only contact that one has with reality is afforded by means of one's knowledge, the problem seems to be insoluble.

      The development of science from a theoretical to an experimental discipline forces philosophy to view the epistemological problem in a new way. For in an experimental science man is in a twofold contact with reality, viz., by his knowledge and by his experimental praxis. Modern atomic theory is one of the best examples to illustrate this point. It was this theory that was most directly confronted with the problem of the realistic value of its models. It could take up this challenge because of the theory's effectiveness for experimental praxis, which is the final judge of the realistic value of the theoretical models. It has confirmed the audacious rational speculations of ancient atomism; but at the same time it has revealed that, in order to be really effective, reason is in need of experimental assistance.

Ancient Greek atomism versus contemporary scientific atomism
      In comparing Greek atomism and modern atomic theories, it should be recalled that in Greek thought philosophy and science still formed a unity. Greek atomism was inspired as much by the desire to find a solution for the problems of mutability and plurality in nature as by the desire to provide scientific explanations for specific phenomena. While it is true that some of the Greek atomists' ideas can rightly be considered as precursors of later physics, the main importance of the old atomistic doctrines for modern science does not lie in these primitive anticipations. Much more important is the attempt to take seriously the variety and mutability discerned by sense experience and yet to reconcile it with the thesis of Parmenides about the unity and the immutability of matter. In its search for universal and unchangeable laws, modern science is to a great extent inspired by the same idea as Parmenides, since universal laws presuppose a certain unity in the material world and unchangeable laws cannot be established without the presupposition that something unchangeable must be hidden behind all changes. By the same token, without this latter presupposition experiments would not make any sense at all. For if the diversity of reactions occurring under different conditions is to reveal anything, these reactions must be the expression of an immutable nature. The differences have to indicate something about that which remains the same. The great achievement of the Greek philosophers was, therefore, that they took a general view of nature as a whole that made a scientific attitude possible. To this, both the quantitative and the qualitative forms of atomism contributed, the former drawing attention to the mathematical aspects of the problem, the latter to the observational.

      A comparison of ancient Greek atomism with scientific atomism merely on the basis of their respective scientific contents would therefore do a great injustice to Greek atomism; it would in fact misjudge its main value. Such a comparison would, however, also take too narrow a view of modern scientific atomism. It would imply the philosophical irrelevance of the latter. It has here been shown, however, that the later development of the scientific atomic theory has clarified many philosophical problems that, as basic issues, divided atomism in the strict sense from other forms of atomism. To mention only a few examples: the development of the scientific atomic theory has deepened man's insights into the relationship between a whole and its components, into the relative character of ultimate particles, and into certain fundamental epistemological problems.

Evaluation of atomism
      The success of the atomic theory shows the value of the idea of atomism: the explanation of the complex in terms of aggregates of fixed particles or units. Its history also shows, however, the inherent danger of this idea, namely, that of absolutism. History has corrected this absolutism: the unitary factors have to be conceived as ultimate only with respect to the complex under consideration, and their union into aggregates need not occur only by way of juxtaposition.

Andrew G.M. van Melsen

Additional Reading
E. Cantore, Atomic Order: An Introduction to the Philosophy of Microphysics (1969); F. Copleston, A History of Philosophy, 8 vol. (1950–66); E.J. Dijksterhuis, Die Mechanisierung des Weltbildes (1956; Eng. trans., The Mechanization of the World Picture, 1961), a history of science from antiquity to the 17th century; A.S. Eddington, The Philosophy of Physical Science (1939); K. Lasswitz, Geschichte der Atomistik vom Mittelalter bis Newton, 2 vol. (1890, reprinted 1963), a 19th-century classic; A.G.M. van Melsen, Van atomos naar atoom (1949; Eng. trans., From Atomos to Atom: The History of the Concept Atom, 2nd ed. 1960), including references for the primary sources; L.K. Nash, The Atomic-Molecular Theory (1950), a discussion of the first phase of the chemical atomic theory; E.T. Whittaker, History of the Theories of Aether and Electricity, rev. ed., 2 vol. (1951–54), only for readers with a solid background in science; L.L. Whyte, Essay on Atomism: From Democritus to 1960 (1961), a brief introduction to the idea of atomism and its history. See also Mrinalkauti Gangopadhyaya, Indian Atomism: History and Sources (1980).

* * *


Universalium. 2010.

Игры ⚽ Поможем написать реферат

Look at other dictionaries:

  • Atomism — (from ancient Greek atomos, meaning uncuttable ) is a natural philosophy that developed in several ancient traditions. The atomists theorized that the natural world consists of two fundamental parts: indivisible atoms and empty void. According to …   Wikipedia

  • Atomism — • The system of those who hold that all bodies are composed of minute, indivisible particles of matter called atoms Catholic Encyclopedia. Kevin Knight. 2006. Atomism     Atomism      …   Catholic encyclopedia

  • ATOMISM — ATOMISM, theory that physical bodies consist ultimately of minute, irreducible, and homogeneous particles called atoms (Greek atomos/atomon = indivisible). In medieval Arabic and Hebrew works atomism derives from Greek (Democritus, Epicurus) and… …   Encyclopedia of Judaism

  • atomism — ATOMÍSM s.n. 1. Concepţie materialist mecanicistă care considera că materia este formată din atomi indivizibili. 2. Teoria ştiinţifică modernă a structurii şi proprietăţilor atomilor. 3. Cercetare ştiinţifică sau concepţie care reduce un ansamblu …   Dicționar Român

  • Atomism — At om*ism, n. [Cf. F. atomisme.] The doctrine of atoms. See {Atomic philosophy}, under {Atomic}. [1913 Webster] …   The Collaborative International Dictionary of English

  • atomism — [at′əm iz΄əm] n. [ ATOM + ISM] Philos. a theory that the universe is made up of tiny, simple, indivisible particles that cannot be destroyed atomist n., adj …   English World dictionary

  • Atomism — the theory that all the objects in the universe are composed of very small, indestructible elements. (This is the case for the Western [i.e., Greek] theories of atomism. Buddhists also have well developed theories of atomism, and which involve… …   Mini philosophy glossary

  • atomism — A philosophical doctrine at least as old as Democritus, and plausibly viewed as an attempt to combine an a priori conviction of the unchangeable and immutable nature of the world with the variety and change of things as we know them. This is the… …   Philosophy dictionary

  • atomism — A philosophical position which views the world as composed of discrete atomistic elements, and reduces knowledge to observation of the smallest indivisible elements, such as human beings but not social structures and social institutions. In… …   Dictionary of sociology

  • atomism —   n. theory of atoms.    ♦ atomize, v.t. reduce to atoms or a fine spray; treat as individual units rather than as a whole.    ♦ atomizer, n. pump like instrument producing fine spray.    ♦ atomistic, a. of atoms or atomism; divided into units …   Dictionary of difficult words

Share the article and excerpts

Direct link
Do a right-click on the link above
and select “Copy Link”