science, history of

science, history of

Introduction

      the history of science from its beginnings in prehistoric times to the 20th century.

      On the simplest level, science is knowledge of the world of nature. There are many regularities in nature that mankind has had to recognize for survival since the emergence of Homo sapiens as a species. The Sun and the Moon periodically repeat their movements. Some motions, like the daily “motion” of the Sun, are simple to observe; others, like the annual “motion” of the Sun, are far more difficult. Both motions correlate with important terrestrial events. Day and night provide the basic rhythm of human existence; the seasons determine the migration of animals upon which humans depended for millennia for survival. With the invention of agriculture, the seasons became even more crucial, for failure to recognize the proper time for planting could lead to starvation. Science defined simply as knowledge of natural processes is universal among mankind, and it has existed since the dawn of human existence.

      The mere recognition of regularities does not exhaust the full meaning of science, however. In the first place, regularities may be simply constructs of the human mind. Humans leap to conclusions; the mind cannot tolerate chaos, so it constructs regularities even when none objectively exists. Thus, for example, one of the astronomical “laws” of the Middle Ages was that the appearance of comets presaged a great upheaval, as the Norman Conquest of Britain followed the comet of 1066. True regularities must be established by detached examination of data. Science, therefore, must employ a certain degree of skepticism to prevent premature generalization.

      Regularities, even when expressed mathematically as laws of nature, are not fully satisfactory to everyone. Some insist that genuine understanding demands explanations of the causes of the laws, but it is in the realm of causation that there is the greatest disagreement. Modern quantum mechanics, for example, has given up the quest for causation and today rests only on mathematical description. Modern biology, on the other hand, thrives on causal chains that permit the understanding of physiological and evolutionary processes in terms of the physical activities of entities such as molecules, cells, and organisms. But even if causation and explanation are admitted as necessary, there is little agreement on the kinds of causes that are permissible, or possible, in science. If the history of science is to make any sense whatsoever, it is necessary to deal with the past on its own terms, and the fact is that for most of the history of science natural philosophers appealed to causes that would be summarily rejected by modern scientists. Spiritual and divine forces were accepted as both real and necessary until the end of the 18th century and, in areas such as biology, deep into the 19th century as well.

      Certain conventions governed the appeal to God or the gods or to spirits. Gods and spirits, it was held, could not be completely arbitrary in their actions; otherwise the proper response would be propitiation, not rational investigation. But since the deity or deities were themselves rational, or bound by rational principles, it was possible for humans to uncover the rational order of the world. Faith in the ultimate rationality of the creator or governor of the world could actually stimulate original scientific work. Kepler's laws, Newton's absolute space, and Einstein's rejection of the probabilistic nature of quantum mechanics were all based on theological, not scientific, assumptions. For sensitive interpreters of phenomena, the ultimate intelligibility of nature has seemed to demand some rational guiding spirit. A notable expression of this idea is Einstein's statement that the wonder is not that mankind comprehends the world, but that the world is comprehensible.

      Science, then, is to be considered in this article as knowledge of natural regularities that is subjected to some degree of skeptical rigour and explained by rational causes. One final caution is necessary. Nature is known only through the senses, of which sight, touch, and hearing are the dominant ones, and the human notion of reality is skewed toward the objects of these senses. The invention of such instruments as the telescope, the microscope, and the Geiger counter has brought an ever-increasing range of phenomena within the scope of the senses. Thus, scientific knowledge of the world is only partial, and the progress of science follows the ability of humans to make phenomena perceivable.

      This article provides a broad survey of the development of science as a way of studying and understanding the world, from the primitive stage of noting important regularities in nature to the epochal revolution in our notion of what constitutes reality that has occurred in 20th-century physics. More detailed treatments of the histories of specific sciences, including developments of the later 20th century, may be found in the articles biology; Earth science (Earth sciences); and physical science.

Science as natural philosophy

Precritical science
      Science, as it has been defined above, made its appearance before writing. It is necessary, therefore, to infer from archaeological remains what was the content of that science. From cave paintings and from apparently regular scratches on bone and reindeer horn, it is known that prehistoric humans were close observers of nature who carefully tracked the seasons and times of the year. About 2500 BC there was a sudden burst of activity that seems to have had clear scientific importance. Great Britain and northwestern Europe contain large stone structures from that era, the most famous of which is Stonehenge on the Salisbury Plain in England, that are remarkable from a scientific point of view. Not only do they reveal technical and social skills of a high order—it was no mean feat to move such enormous blocks of stone considerable distances and place them in position—but the basic conception of Stonehenge and the other megalithic (megalith) structures also seems to combine religious and astronomical purposes. Their layouts suggest a degree of mathematical sophistication that was first suspected only in the mid-20th century. Stonehenge is a circle, but some of the other megalithic structures are egg-shaped and, apparently, constructed on mathematical principles that require at least practical knowledge of the Pythagorean theorem that the square of the hypotenuse of a right triangle is equal to the sum of the squares of the other two sides. This theorem, or at least the Pythagorean numbers that can be generated by it, seems to have been known throughout Asia, the Middle East, and Neolithic Europe two millennia before the birth of Pythagoras.

      This combination of religion and astronomy was fundamental to the early history of science. It is found in Mesopotamia, Egypt, China (although to a much lesser extent than elsewhere), Central America, and India. The spectacle of the heavens, with the clearly discernible order and regularity of most heavenly bodies highlighted by extraordinary events such as comets and novae and the peculiar motions of the planets, obviously was an irresistible intellectual puzzle to early mankind. In its search for order and regularity, the human mind could do no better than to seize upon the heavens as the paradigm of certain knowledge. Astronomy was to remain the queen of the sciences (welded solidly to theology) for the next 4,000 years.

      Science, in its mature form, developed only in the West. But it is instructive to survey the protoscience that appeared in other areas, especially in light of the fact that until quite recently this knowledge was often, as in China, far superior to Western science.

      As has already been noted, astronomy seems everywhere to have been the first science to emerge. Its intimate relation to religion gave it a ritual dimension that then stimulated the growth of mathematics. Chinese savants, for example, early devised a calendar and methods of plotting the positions of stellar constellations. Since changes in the heavens presaged important changes on the Earth (for the Chinese considered the universe to be a vast organism in which all elements were connected), astronomy and astrology were incorporated into the system of government from the very dawn of the Chinese state in the 2nd millennium BC. As the Chinese bureaucracy developed, an accurate calendar became absolutely necessary to the maintenance of legitimacy and order. The result was a system of astronomical observations and records unparalleled elsewhere, thanks to which there are, today, star catalogs and observations of eclipses and novae that go back for millennia.

      In other sciences, too, the overriding emphasis was on practicality, for the Chinese, almost alone among ancient peoples, did not fill the cosmos with gods and demons whose arbitrary wills determined events. Order was inherent and, therefore, expected. It was for man to detect and describe this order and to profit from it. Chemistry (or, rather, alchemy), medicine, geology, geography, and technology were all encouraged by the state and flourished. Practical knowledge of a high order permitted the Chinese to deal with practical problems for centuries on a level not attained in the West until the Renaissance.

      Astronomy was studied in India for calendrical purposes to set the times for both practical and religious tasks. Primary emphasis was placed on solar and lunar motions, the fixed stars serving as a background against which these luminaries moved. Indian mathematics seems to have been quite advanced, with particular sophistication in geometrical and algebraic techniques. This latter branch was undoubtedly stimulated by the flexibility of the Indian system of numeration that later was to come into the West as the Hindu-Arabic numerals.

      Quite independently of China, India, and the other civilizations of Europe and Asia, the Maya of Central America, building upon older cultures, created a complex society in which astronomy and astrology played important roles. Determination of the calendar, again, had both practical and religious significance. Solar and lunar eclipses were important, as was the position of the bright planet Venus. No sophisticated mathematics are known to have been associated with this astronomy, but the Mayan calendar was both ingenious and the result of careful observation.

The Middle East
      In the cradles of Western civilization in Egypt (Egypt, ancient) and Mesopotamia (Mesopotamia, history of), there were two rather different situations. In Egypt there was an assumption of cosmic order guaranteed by a host of benevolent gods. But unlike China, whose rugged geography often produced disastrous floods, earthquakes, and violent storms that destroyed crops, Egypt was surpassingly placid and delightful. Egyptians found it difficult to believe that all ended with death; enormous intellectual and physical labour, therefore, was devoted to preserving life after death. Both Egyptian theology and the pyramids are testaments to this preoccupation. All of the important questions were answered by religion, so the Egyptians did not concern themselves overmuch with speculations about the universe. The stars and the planets had astrological significance in that the major heavenly bodies were assumed to “rule” the land when they were in the ascendant (from the succession of these “rules” came the seven-day week, after the five planets and the Sun and the Moon), but astronomy was largely limited to the calendrical calculations necessary to predict the annual life-giving flood of the Nile. None of this required much mathematics, and there was, consequently, little of any importance.

      Mesopotamia was more like China. The life of the land depended upon the two great rivers, the Tigris and the Euphrates, as that of China depended upon the Huang Ho (Yellow River) and the Yangtze. The land was harsh and made habitable only by extensive damming and irrigation works. Storms, insects, floods, and invaders made life insecure. To create a stable society required both great technological skill, for the creation of hydraulic works, and the ability to hold off the forces of disruption. These latter were early identified with powerful and arbitrary gods who dominated Mesopotamian theology. The cities of the plain were centred on temples run by a priestly caste whose functions included the planning of major public works, like canals, dams, and irrigation systems, the allocation of the resources of the city to its members, and the averting of a divine wrath that could wipe everything out.

       mathematics and astronomy thrived under these conditions. The number system, probably drawn from the system of weights and coinage, was based on 60 (it was in ancient Mesopotamia that the system of degrees, minutes, and seconds developed) and was adapted to a practical arithmetic. The heavens were the abode of the gods, and because heavenly phenomena were thought to presage terrestrial disasters, they were carefully observed and recorded. Out of these practices grew, first, a highly developed mathematics that went far beyond the requirements of daily business, and then, some centuries later, a descriptive astronomy that was the most sophisticated of the ancient world until the Greeks took it over and perfected it.

      Nothing is known of the motives of these early mathematicians for carrying their studies beyond the calculations of volumes of dirt to be removed from canals and the provisions necessary for work parties. It may have been simply intellectual play—the role of playfulness in the history of science should not be underestimated—that led them onward to abstract algebra. There are texts from about 1700 BC that are remarkable for their mathematical suppleness. Babylonian mathematicians knew the Pythagorean relationship well and used it constantly. They could solve simple quadratic equations and could even solve problems in compound interest involving exponents. From about a millennium later there are texts that utilize these skills to provide a very elaborate mathematical description of astronomical phenomena.

      Although China and Mesopotamia provide examples of exact observation and precise description of nature, what is missing is explanation in the scientific mode. The Chinese assumed a cosmic order that was vaguely founded on the balance of opposite forces (yin–yang) and the harmony of the five elements (water, wood, metal, fire, and earth). Why this harmony obtained was not discussed. Similarly, the Egyptians found the world harmonious because the gods willed it so. For Babylonians and other Mesopotamian cultures, order existed only so long as all-powerful and capricious gods supported it. In all these societies, humans could describe nature and use it, but to understand it was the function of religion and magic, not reason. It was the Greeks who first sought to go beyond description and to arrive at reasonable explanations of natural phenomena that did not involve the arbitrary will of the gods. Gods might still play a role, as indeed they did for centuries to come, but even the gods were subject to rational laws.

The birth of natural philosophy
      There seems to be no good reason why the Hellenes, clustered in isolated city-states in a relatively poor and backward land, should have struck out into intellectual regions that were only dimly perceived, if at all, by the splendid civilizations of the Yangtze, the Tigris and Euphrates, and the Nile valleys. There were many differences between ancient Greece and the other civilizations, but perhaps the most significant was religion. What is striking about Greek religion, in contrast to the religions of Mesopotamia and Egypt, is its puerility. Both of the great river civilizations evolved complex theologies that served to answer most, if not all, of the large questions about mankind's place and destiny. Greek religion did not. It was, in fact, little more than a collection of folk tales, more appropriate to the campfire than to the temple. Perhaps this was the result of the collapse of an earlier Greek civilization, now called Mycenaean, toward the end of the 2nd millennium BC, when a dark age descended upon Greece that lasted for three centuries. All that was preserved were stories of gods and men, passed along by poets, that dimly reflected Mycenaean values and events. Such were the great poems of Homer, the Iliad and the Odyssey, in which heroes and gods mingled freely with one another. Indeed, they mingled too freely, for the gods appear in these tales as little more than immortal adolescents whose tricks and feats, when compared with the concerns of a Marduk or Jehovah, are infantile. There really was no Greek theology in the sense that theology provides a coherent and profound explanation of the workings of both the cosmos and the human heart. Hence, there were no easy answers to inquiring Greek minds. The result was that ample room was left for a more penetrating and ultimately more satisfying mode of inquiry. Thus were philosophy and its oldest offspring, science, born.

      The first natural philosopher, according to Hellenic tradition, was Thales of Miletus, who flourished in the 6th century BC. We know of him only through later accounts, for nothing he wrote has survived. He is supposed to have predicted a solar eclipse in 585 BC and to have invented the formal study of geometry in his demonstration of the bisecting of a circle by its diameter. Most importantly, he tried to explain all observed natural phenomena in terms of the changes of a single substance, water, which can be seen to exist in solid, liquid, and gaseous states. What for Thales guaranteed the regularity and rationality of the world was the innate divinity in all things that directed them to their divinely appointed ends. From these ideas there emerged two characteristics of classical Greek science. The first was the view of the universe as an ordered structure (the Greek kósmos means “order”). The second was the conviction that this order was not that of a mechanical contrivance but that of an organism; all parts of the universe had purposes in the overall scheme of things, and objects moved naturally toward the ends they were fated to serve. This motion toward ends is called teleology and, with but few exceptions, it permeated Greek as well as much later science.

      Thales inadvertently made one other fundamental contribution to the development of natural science. By naming a specific substance as the basic element of all matter, Thales opened himself to criticism, which was not long in coming. His own disciple, Anaximander, was quick to argue that water could not be the basic substance. His argument was simple: water, if it is anything, is essentially wet; nothing can be its own contradiction. Hence, if Thales were correct, the opposite of wet could not exist in a substance, and that would preclude all of the dry things that are observed in the world. Therefore, Thales was wrong. Here was the birth of the critical tradition that is fundamental to the advance of science.

      Thales' conjectures set off an intellectual explosion, most of which was devoted to increasingly refined criticisms of his doctrine of fundamental matter. Various single substances were proposed and then rejected, ultimately in favour of a multiplicity of elements that could account for such opposite qualities as wet and dry, hot and cold. Two centuries after Thales, most natural philosophers accepted a doctrine of four elements: earth (cold and dry), fire (hot and dry), water (cold and wet), and air (hot and wet). All bodies were made from these four.

      The presence of the elements only guaranteed the presence of their qualities in various proportions. What was not accounted for was the form these elements took, which served to differentiate natural objects from one another. The problem of form was first attacked systematically by the philosopher and cult leader Pythagoras in the 6th century BC. Legend has it that Pythagoras became convinced of the primacy of number when he realized that the musical notes produced by a monochord were in simple ratio to the length of the string. Qualities (tones) were reduced to quantities (numbers in integral ratios). Thus was born mathematical physics, for this discovery provided the essential bridge between the world of physical experience and that of numerical relationships. Number provided the answer to the question of the origin of forms and qualities.

      Hellenic science was built upon the foundations laid by Thales and Pythagoras. It reached its zenith in the works of Aristotle and Archimedes. Aristotle represents the first tradition, that of qualitative forms and teleology. He was, himself, a biologist (biology) whose observations of marine organisms were unsurpassed until the 19th century. Biology is essentially teleological—the parts of a living organism are understood in terms of what they do in and for the organism—and Aristotle's biological works provided the framework for the science until the time of Charles Darwin. In physics, teleology is not so obvious, and Aristotle had to impose it on the cosmos. From Plato, his teacher, he inherited the theological proposition that the heavenly bodies (stars and planets) are literally divine and, as such, perfect. They could, therefore, move only in perfect, eternal, unchanging motion, which, by Plato's definition, meant perfect circles. The Earth, being obviously not divine, and inert, was at the centre. From the Earth to the sphere of the Moon, all things constantly changed, generating new forms and then decaying back into formlessness. Above the Moon the cosmos consisted of contiguous and concentric crystalline spheres moving on axes set at angles to one another (this accounted for the peculiar motions of the planets) and deriving their motion either from a fifth element that moved naturally in circles or from heavenly souls resident in the celestial bodies (celestial mechanics). The ultimate cause of all motion was a prime, or unmoved, mover (God) that stood outside the cosmos.

      Aristotle was able to make a great deal of sense of observed nature by asking of any object or process: what is the material involved, what is its form and how did it get that form, and, most important of all, what is its purpose? What should be noted is that, for Aristotle, all activity that occurred spontaneously was natural. Hence, the proper means of investigation was observation. Experiment, that is, altering natural conditions in order to throw light on the hidden properties and activities of objects, was unnatural and could not, therefore, be expected to reveal the essence of things. Experiment was thus not essential to Greek science.

      The problem of purpose did not arise in the areas in which Archimedes made his most important contributions. He was, first of all, a brilliant mathematician whose work on conic sections and on the area of the circle prepared the way for the later invention of the calculus. It was in mathematical physics, however, that he made his greatest contributions to science. His mathematical demonstration of the law of the lever was as exact as a Euclidean proof in geometry. Similarly, his work on hydrostatics introduced and developed the method whereby physical characteristics, in this case specific gravity, which Archimedes discovered, are given mathematical shape and then manipulated by mathematical methods to yield mathematical conclusions that can be translated back into physical terms.

 In one major area the Aristotelian and the Archimedean approaches were forced into a rather inconvenient marriage. astronomy was the dominant physical science throughout antiquity, but it had never been successfully reduced to a coherent system. The Platonic-Aristotelian astral religion required that planetary orbits be circles. But, particularly after the conquests of Alexander the Great had made the observations and mathematical methods of the Babylonians available to the Greeks, astronomers found it impossible to reconcile theory and observation. Astronomy then split into two parts: one was physical and accepted Aristotelian theory in accounting for heavenly motion; the other ignored causation and concentrated solely on the creation of a mathematical model that could be used for computing planetary positions. Ptolemy, in the 2nd century AD, carried the latter tradition to its highest point in antiquity in his Hē mathēmatikē syntaxis (“The Mathematical Collection,” better known under its Greek-Arabic title, Almagest). (See Theories of the Solar System—>.)

      The Greeks not only made substantial progress in understanding the cosmos but also went far beyond their predecessors in their knowledge of the human body. Pre-Greek medicine had been almost entirely confined to religion and ritual. Disease (human disease) was considered the result of divine disfavour and human sin, to be dealt with by spells, prayers, and other propitiatory measures. In the 5th century BC a revolutionary change came about that is associated with the name of Hippocrates. It was Hippocrates and his school who, influenced by the rise of natural philosophy, first insisted that disease was a natural, not a supernatural, phenomenon. Even maladies as striking as epilepsy, whose seizures appeared to be divinely caused, were held to originate in natural causes within the body.

      The height of medical science in antiquity was reached late in the Hellenistic period. Much work was done at the museum of Alexandria, a research institute set up under Greek influence in Egypt in the 3rd century BC to sponsor learning in general. The heart and the vascular system were investigated, as were the nerves and the brain. The organs of the thoracic cavity were described, and attempts were made to discover their functions. It was on these researches, and on his own dissections of apes and pigs, that the last great physician of antiquity, Galen Of Pergamum, based his physiology. It was, essentially, a tripartite system in which so-called spirits—natural, vital, and animal—passed respectively through the veins, the arteries, and the nerves to vitalize the body as a whole. Galen's attempts to correlate therapeutics with his physiology were not successful, and so medical practice remained eclectic and a matter of the physician's choice. Usually the optimal choice was that propounded by the Hippocratics, who relied primarily on simple, clean living and the ability of the body to heal itself.

Science in Rome (ancient Rome) and Christianity
      The apogee of Greek science in the works of Archimedes and Euclid coincided with the rise of Roman power in the Mediterranean. The Romans were deeply impressed by Greek art, literature, philosophy, and science, and after their conquest of Greece many Greek intellectuals served as household slaves tutoring noble Roman children. The Romans were a practical people, however, and, while they contemplated the Greek intellectual achievement with awe, they also could not help but ask what good it had done the Greeks. Roman common sense was what kept Rome great; science and philosophy were either ignored or relegated to rather low status. Even such a Hellenophile as the statesman and orator Cicero used Greek thought more to buttress the old Roman ways than as a source of new ideas and viewpoints.

      The spirit of independent research was quite foreign to the Roman mind, so scientific innovation ground to a halt. The scientific legacy of Greece was condensed and corrupted into Roman encyclopaedias whose major function was entertainment rather than enlightenment. Typical of this spirit was the 1st-century-AD aristocrat Pliny the Elder, whose Natural History was a multivolume collection of myths, odd tales of wondrous creatures, magic, and some science, all mixed together uncritically for the titillation of other aristocrats. Aristotle would have been embarrassed by it.

      At its height Rome incorporated a host of peoples with different customs, languages, and religions within its empire. One religious sect that proved more significant than the rest was Christianity. Jesus (Jesus Christ) and his kingdom were not of this world, but his disciples and their followers were. This world could not be ignored, even though concern with worldly things could be dangerous to the soul. So the early Christians approached the worldly wisdom of their time with ambivalence: on the one hand, the rhetoric and the arguments of ancient philosophy were snares and delusions that might mislead the simple and the unwary; on the other hand, the sophisticated and the educated of the empire could not be converted unless the Christian message was presented in the terms and rhetoric of the philosophical schools. Before they knew it, the early Christians were enmeshed in metaphysical arguments, some of which involved physics. What, for example, was the nature of Jesus, in purely physical terms? How was it possible that anybody could have two different essential natures, as was claimed for Jesus? Such questions revealed how important knowledge of the arguments of Greek thinkers on the nature of substance could be to those engaged in founding a new theology.

      Ancient learning, then, did not die with the fall of Rome and the occupation of the Western Empire by tribes of Germanic barbarians. To be sure, the lamp of learning burned very feebly, but it did not go out. Monks in monasteries faithfully copied out classics of ancient thought and early Christianity and preserved them for posterity. Monasteries continued to teach the elements of ancient learning, for little beyond the elementary survived in the Latin West. In the East, the Byzantine Empire remained strong, and there the ancient traditions continued. There was little original work done in the millennium following the fall of Rome, but the ancient texts were preserved along with knowledge of the ancient Greek language. This was to be a precious reservoir of learning for the Latin West in later centuries.

Science in Islām (Islāmic world)
      The torch of ancient learning passed first to one of the invading groups that helped bring down the Eastern Empire. In the 7th century the Arabs, inspired by their new religion, burst out of the Arabian peninsula and laid the foundations of an Islāmic empire that eventually rivalled that of ancient Rome. To the Arabs, ancient science was a precious treasure. The Qurʿān (Qurʾān), the sacred book of Islām, particularly praised medicine as an art close to God. Astronomy and astrology were believed to be one way of glimpsing what God willed for mankind. Contact with Hindu mathematics and the requirements of astronomy stimulated the study of numbers and of geometry. The writings of the Hellenes were, therefore, eagerly sought and translated, and thus much of the science of antiquity passed into Islāmic culture. Greek medicine, Greek astronomy and astrology, and Greek mathematics, together with the great philosophical works of Plato and, particularly, Aristotle, were assimilated in Islām by the end of the 9th century. Nor did the Arabs stop with assimilation. They criticized and they innovated. Islāmic astronomy and astrology were aided by the construction of great astronomical observatories that provided accurate observations against which the Ptolemaic predictions could be checked. Numbers fascinated Islāmic thinkers, and this fascination served as the motivation for the creation of algebra (from Arabic al-jabr) and the study of algebraic functions.

Medieval (Middle Ages) European science
      Medieval Christendom confronted Islām chiefly in military crusades, in Spain and the Holy Land, and in theology. From this confrontation came the restoration of ancient learning to the West. The Reconquista in Spain gradually pushed the Moors south from the Pyrenees, and among the treasures left behind were Arabic translations of Greek works of science and philosophy. In 1085 the city of Toledo, with one of the finest libraries in Islām, fell to the Christians. Among the occupiers were Christian monks who quickly began the process of translating ancient works into Latin. By the end of the 12th century much of the ancient heritage was again available to the Latin West.

      The medieval world was caricatured by thinkers of the 18th-century Enlightenment as a period of darkness, superstition, and hostility to science and learning. On the contrary, it was one of great technological vitality. The advances that were made may appear today as trifling, but that is because they were so fundamental. They included the horseshoe and the horse collar, without which horsepower cannot be efficiently exploited. The invention of the crank, the brace and bit, the wheelbarrow, and the flying buttress made possible the great Gothic cathedrals. Improvements in the gear trains of waterwheels and the development of windmills harnessed these sources of power with great efficiency. Mechanical ingenuity, building on experience with mills and power wheels, culminated in the 14th century in the mechanical clock, which not only set a new standard of chronometrical accuracy but also provided philosophers with a new metaphor for nature itself.

      An equal amount of energy was devoted to achieving a scientific understanding of nature, but it is essential to understand to what use medieval thinkers put this kind of knowledge. As the fertility of the technology shows, medieval Europeans had no deep prejudices against utilitarian knowledge. But the areas in which scientific knowledge could find useful expression were few. Instead, science was viewed chiefly as a means of understanding God's creation and, thereby, the Godhead itself. The best example of this attitude is found in the medieval study of optics. light, as Genesis makes clear, was among the first creations of God. The 12th–13th-century cleric-scholar Robert Grosseteste (Grosseteste, Robert) saw in light the first creative impulse. As light spread it created both space and matter, and, in its reflection from the outermost circle of the cosmos, it gradually solidified into the heavenly spheres. To understand the laws of the propagation of light was to understand, in some slight way, the nature of the creation.

      In the course of studying light, particular problems were isolated and attacked. What, for example, is the rainbow? It is impossible to get close enough to a rainbow to see clearly what is going on, for as the observer moves, so too does the rainbow. It does seem to depend upon the presence of raindrops, so medieval investigators sought to bring the rainbow down from the skies into their studies. Insight into the nature of the rainbow could be achieved by simulating the conditions under which rainbows occur. For raindrops the investigators substituted hollow glass balls filled with water, so that the rainbow could be studied at leisure. Valid conclusions about rainbows could then be drawn by assuming the validity of the analogy between raindrops and water-filled globes. This involved the implicit assumptions that nature was simple (i.e., governed by a few general laws) and that similar effects had similar causes. Such a nature was what could be expected of a rational, benevolent deity; hence, the assumption could be persuasively adopted.

      Medieval philosophers were not content, as the above example shows, to repeat what the ancients had said. They subjected ancient texts to close critical scrutiny. Usually the intensity of the criticism was directly proportional to the theological significance of the problem involved. Such was the case with motion. Medieval philosophers examined all aspects of motion with great care, for the nature of motion had important theological implications. Thomas Aquinas (Aquinas, Thomas, Saint) used Aristotle's dictum, that everything that moves is moved by something else, to show that God must exist, for otherwise the existence of any motion would imply an infinite regression of prior causal motions.

      It should be clear that there was no conscious conflict between science and religion in the Middle Ages. As Aquinas pointed out, God was the author of both the book of Scripture and the book of nature. The guide to nature was reason, the faculty that was the image of God in which mankind was made. Scripture was direct revelation, although it needed interpretation, for there were passages that were obscure or difficult. The two books, having the same author, could not contradict each other. For the short term, science and revelation marched hand in hand. Aquinas carefully wove knowledge of nature into his theology, as in his proof from motion of the existence of God. But if his scientific concepts of motion should ever be challenged, there would necessarily be a theological challenge as well. By working science into the very fabric of his theology, he virtually guaranteed that someday there would be conflict. Theologians would side with theology and scientists with science, to create a breach that neither particularly desired.

      The glory of medieval science was its integration of science, philosophy, and theology into a magnificent and comprehensible whole. It can be best contemplated in the greatest of all medieval poems, Dante's Divine Comedy (Divine Comedy, The). Here was an essentially Aristotelian cosmos, finite and easily understood, over which God, his Son, and his saints reigned. Humanity and the Earth occupied the centre, as befitted their centrality in God's plan. The nine circles of hell were populated by humans whose exercise of their free will had led to their damnation. Purgatory contained lesser sinners still capable of salvation. The heavenly spheres were populated by the saved and the saintly. The natural hierarchy gave way to the spiritual hierarchy as one ascended toward the throne of God. Such a hierarchy was reflected in the social and political institutions of medieval Europe, and God, the supreme monarch, ruled his creation with justice and love. All fit together in a grand cosmic scheme, one not to be abandoned lightly.

The rise of modern science

The authority of phenomena
      Even as Dante was writing his great work, deep forces were threatening the unitary cosmos he celebrated. The pace of technological innovation began to quicken. Particularly in Italy, the political demands of the time gave new importance to technology, and a new profession emerged, that of civil and military engineer (engineering). These people faced practical problems that demanded practical solutions. Leonardo da Vinci is certainly the most famous of them, though he was much more as well. A painter of genius, he closely studied human anatomy in order to give verisimilitude to his paintings. As a sculptor he mastered the difficult techniques of casting metal. As a producer-director of the form of Renaissance dramatic production called the masque, he devised complicated machinery to create special effects. But it was as a military engineer that he observed the path of a mortar bomb being lobbed over a city wall and insisted that the projectile did not follow two straight lines—a slanted ascent followed by a vertical drop—as Aristotle had said it must. Leonardo and his colleagues needed to know nature truly; no amount of book learning could substitute for actual experience, nor could books impose their authority upon phenomena. What Aristotle and his commentators asserted as philosophical necessity often did not gibe with what could be seen with one's own eyes. The hold of ancient philosophy was too strong to be broken lightly, but a healthy skepticism began to emerge.

      The first really serious blow to the traditional acceptance of ancient authorities was the discovery of the New World at the end of the 15th century. Ptolemy, the great astronomer and geographer, had insisted that only the three continents of Europe, Africa, and Asia could exist, and Christian scholars from St. Augustine on had accepted it, for otherwise men would have to walk upside down at the antipodes. But Ptolemy, St. Augustine, and a host of other authorities were wrong. The dramatic expansion of the known world also served to stimulate the study of mathematics, for wealth and fame awaited those who could turn navigation into a real and trustworthy science.

      In large part the Renaissance was a time of feverish intellectual activity devoted to the complete recovery of the ancient heritage. To the Aristotelian texts that had been the foundation of medieval thought were added translations of Plato, with his vision of mathematical harmonies, of Galen, with his experiments in physiology and anatomy, and, perhaps most important of all, of Archimedes, who showed how theoretical physics could be done outside the traditional philosophical framework. The results were subversive.

      The search for antiquity turned up a peculiar bundle of manuscripts that added a decisive impulse to the direction in which Renaissance science was moving. These manuscripts were taken to have been written by or to report almost at first hand the activities of the legendary priest, prophet, and sage Hermes Trismegistos. Hermes was supposedly a contemporary of Moses, and the Hermetic writings contained an alternative story of creation that gave man a far more prominent role than the traditional account. God had made man fully in his image: a creator, not just a rational animal. Man could imitate God by creating. To do so, he must learn nature's secrets, and this could be done only by forcing nature to yield them through the tortures of fire, distillation, and other alchemical manipulations. The reward for success would be eternal life and youth, as well as freedom from want and disease. It was a heady vision, and it gave rise to the notion that, through science and technology, man could bend nature to his wishes. This is essentially the modern view of science, and it should be emphasized that it occurs only in Western civilization. It is probably this attitude that permitted the West to surpass the East, after centuries of inferiority, in the exploitation of the physical world.

      The Hermetic tradition also had more specific effects. Inspired, as is now known, by late Platonist mysticism, the Hermetic writers had rhapsodized on enlightenment and on the source of light, the Sun (sun worship). Marsilio Ficino (Ficino, Marsilio), the 15th-century Florentine translator of both Plato and the Hermetic writings, composed a treatise on the Sun that came close to idolatry. A young Polish student visiting Italy at the turn of the 16th century was touched by this current. Back in Poland, he began to work on the problems posed by the Ptolemaic astronomical system. With the blessing of the church, which he served formally as a canon, Nicolaus Copernicus (Copernicus, Nicolaus) set out to modernize the astronomical apparatus by which the church made such important calculations as the proper dates for Easter and other festivals.

The scientific revolution
 In 1543, as he lay on his deathbed, Copernicus finished reading the proofs of his great work; he died just as it was published. His De revolutionibus orbium coelestium libri VI (“Six Books Concerning the Revolutions of the Heavenly Orbs”) was the opening shot in a revolution whose consequences were greater than those of any other intellectual event in the history of mankind. The scientific revolution radically altered the conditions of thought and of material existence in which the human race lives, and its effects are not yet exhausted.

 All this was caused by Copernicus' daring in placing the Sun, not the Earth, at the centre of the cosmos. Copernicus actually cited Hermes Trismegistos to justify this idea, and his language was thoroughly Platonic. But he meant his work as a serious work in astronomy, not philosophy, so he set out to justify it observationally and mathematically. The results were impressive. At one stroke, Copernicus reduced a complexity verging on chaos to elegant simplicity. The apparent back-and-forth movements of the planets, which required prodigious ingenuity to accommodate within the Ptolemaic system, could be accounted for just in terms of the Earth's own orbital motion added to or subtracted from the motions of the planets. Variation in planetary brightness was also explained by this combination of motions. The fact that Mercury and Venus were never found opposite the Sun in the sky Copernicus explained by placing their orbits closer to the Sun than that of the Earth. Indeed, Copernicus was able to place the planets in order of their distances from the Sun by considering their speeds and thus to construct a system of the planets (Copernican system), something that had eluded Ptolemy. This system had a simplicity, coherence, and aesthetic charm that made it irresistible to those who felt that God was the supreme artist. His was not a rigorous argument, but aesthetic considerations are not to be ignored in the history of science.

      Copernicus did not solve all of the difficulties of the Ptolemaic system. He had to keep some of the cumbrous apparatus of epicycles and other geometrical adjustments, as well as a few Aristotelian crystalline spheres. The result was neater, but not so striking that it commanded immediate universal assent. Moreover, there were some implications that caused considerable concern: Why should the crystalline orb containing the Earth circle the Sun? And how was it possible for the Earth itself to revolve on its axis once in 24 hours without hurling all objects, including humans, off its surface? No known physics could answer these questions, and the provision of such answers was to be the central concern of the scientific revolution.

      More was at stake than physics and astronomy, for one of the implications of the Copernican system struck at the very foundations of contemporary society. If the Earth revolved around the Sun, then the apparent positions of the fixed stars should shift as the Earth moves in its orbit. Copernicus and his contemporaries could detect no such shift (called stellar parallax), and there were only two interpretations possible to explain this failure. Either the Earth was at the centre, in which case no parallax was to be expected, or the stars were so far away that the parallax was too small to be detected. Copernicus chose the latter and thereby had to accept an enormous cosmos consisting mostly of empty space. God, it had been assumed, did nothing in vain, so for what purposes might he have created a universe in which the Earth and mankind were lost in immense space? To accept Copernicus was to give up the Dantean cosmos. The Aristotelian hierarchy of social place, political position, and theological gradation would vanish, to be replaced by the flatness and plainness of Euclidean space. It was a grim prospect and not one that recommended itself to most 16th-century intellectuals, and so Copernicus' grand idea remained on the periphery of astronomical thought. All astronomers were aware of it, some measured their own views against it, but only a small handful eagerly accepted it.

      In the century and a half following Copernicus, two easily discernible scientific movements developed. The first was critical, the second, innovative and synthetic. They worked together to bring the old cosmos into disrepute and, ultimately, to replace it with a new one. Although they existed side by side, their effects can more easily be seen if they are treated separately.

 The critical tradition began with Copernicus. It led directly to the work of Tycho Brahe (Brahe, Tycho), who measured stellar and planetary positions more accurately than had anyone before him. But measurement alone could not decide between Copernicus and Ptolemy, and Tycho insisted that the Earth was motionless. Copernicus did persuade Tycho to move the centre of revolution of all other planets to the Sun. To do so, he had to abandon the Aristotelian crystalline spheres that otherwise would collide with one another. Tycho also cast doubt upon the Aristotelian doctrine of heavenly perfection, for when, in the 1570s, a comet and a new star appeared, Tycho showed that they were both above the sphere of the Moon. Perhaps the most serious critical blows struck were those delivered by Galileo after the invention of the telescope. In quick succession, he announced that there were mountains on the Moon, satellites circling Jupiter, and spots upon the Sun. Moreover, the Milky Way was composed of countless stars whose existence no one had suspected until Galileo saw them. Here was criticism that struck at the very roots of Aristotle's system of the world.

 At the same time Galileo was searching the heavens with his telescope, in Germany Johannes Kepler (Kepler, Johannes) was searching them with his mind. Tycho's precise observations permitted Kepler to discover that Mars (and, by analogy, all the other planets) did not revolve in a circle at all, but in an ellipse, with the Sun at one focus. Ellipses tied all the planets together in grand Copernican harmony. The Keplerian cosmos was most un-Aristotelian, but Kepler hid his discoveries by burying them in almost impenetrable Latin prose in a series of works that did not circulate widely.

      What Galileo and Kepler could not provide, although they tried, was an alternative to Aristotle that made equal sense. If the Earth revolves on its axis, then why do objects not fly off it? And why do objects dropped from towers not fall to the west as the Earth rotates to the east beneath them? And how is it possible for the Earth, suspended in empty space, to go around the Sun—whether in circles or ellipses—without anything pushing it? The answers were long in coming.

 Galileo attacked the problems of the Earth's rotation and its revolution by logical analysis. Bodies do not fly off the Earth because they are not really revolving rapidly, even though their speed is high. In revolutions per minute, any body on the Earth is going very slowly and, therefore, has little tendency to fly off. Bodies fall to the base of towers from which they are dropped because they share with the tower the rotation of the Earth. Hence, bodies already in motion preserve that motion when another motion is added. So, Galileo deduced, a ball dropped from the top of a mast of a moving ship would fall at the base of the mast. If the ball were allowed to move on a frictionless horizontal plane, it would continue to move forever. Hence, Galileo concluded, the planets, once set in circular motion, continue to move in circles forever. Therefore, Copernican orbits exist. Galileo never acknowledged Kepler's ellipses; to do so would have meant abandoning his solution to the Copernican problem.

      Kepler realized that there was a real problem with planetary motion. He sought to solve it by appealing to the one force that appeared to be cosmic in nature, namely magnetism. The Earth had been shown to be a giant magnet by William Gilbert in 1600, and Kepler seized upon this fact. A magnetic force, Kepler argued, emanated from the Sun and pushed the planets around in their orbits, but he was never able to quantify this rather vague and unsatisfactory idea.

      By the end of the first quarter of the 17th century Aristotelianism was rapidly dying, but there was no satisfactory system to take its place. The result was a mood of skepticism and unease, for, as one observer put it, “The new philosophy calls all in doubt.” It was this void that accounted largely for the success of a rather crude system proposed by René Descartes (Descartes, René). Matter and motion were taken by Descartes to explain everything by means of mechanical models of natural processes, even though he warned that such models were not the way nature probably worked. They provided merely “likely stories,” which seemed better than no explanation at all.

      Armed with matter and motion, Descartes attacked the basic Copernican problems. Bodies once in motion, Descartes argued, remain in motion in a straight line unless and until they are deflected from this line by the impact of another body. All changes of motion are the result of such impacts. Hence, the ball falls at the foot of the mast because, unless struck by another body, it continues to move with the ship. Planets move around the Sun because they are swept around by whirlpools of a subtle matter filling all space. Similar models could be constructed to account for all phenomena; the Aristotelian system could be replaced by the Cartesian. There was one major problem, however, and it sufficed to bring down Cartesianism. Cartesian matter and motion had no purpose, nor did Descartes's philosophy seem to need the active participation of a deity. The Cartesian cosmos, as Voltaire later put it, was like a watch that had been wound up at the creation and continues ticking to eternity.

 The 17th century was a time of intense religious feeling, and nowhere was that feeling more intense than in Great Britain. There a devout young man, Isaac Newton, was finally to discover the way to a new synthesis in which truth was revealed and God was preserved.

 Newton was both an experimental and a mathematical genius, a combination that enabled him to establish both the Copernican system and a new mechanics. His method was simplicity itself: “from the phenomena of motions to investigate the forces of nature, and then from these forces to demonstrate the other phenomena.” Newton's genius guided him in the selection of phenomena to be investigated, and his creation of a fundamental mathematical tool—the calculus (simultaneously invented by Gottfried Leibniz)—permitted him to submit the forces he inferred to calculation. The result was Philosophiae Naturalis Principia Mathematica (Mathematical Principles of Natural Philosophy, usually called simply the Principia), which appeared in 1687. Here was a new physics that applied equally well to terrestrial and celestial bodies. Copernicus, Kepler, and Galileo were all justified by Newton's analysis of forces. Descartes was utterly routed.

      Newton's three laws of motion and his principle of universal gravitation sufficed to regulate the new cosmos, but only, Newton believed, with the help of God. Gravity, he more than once hinted, was direct divine action, as were all forces for order and vitality. Absolute space, for Newton, was essential, because space was the “sensorium of God,” and the divine abode must necessarily be the ultimate coordinate system. Finally, Newton's analysis of the mutual perturbations of the planets caused by their individual gravitational fields predicted the natural collapse of the solar system unless God acted to set things right again.

The diffusion of scientific method
      The publication of the Principia marks the culmination of the movement begun by Copernicus and, as such, has always stood as the symbol of the scientific revolution. There were, however, similar attempts to criticize, systematize, and organize natural knowledge that did not lead to such dramatic results. In the same year as Copernicus' great volume, there appeared an equally important book on anatomy: Andreas Vesalius' (Vesalius, Andreas) De humani corporis fabrica (“On the Fabric of the Human Body,” called the De fabrica), a critical examination of Galen's anatomy in which Vesalius drew on his own studies to correct many of Galen's errors. Vesalius, like Newton a century later, emphasized the phenomena, i.e., the accurate description of natural facts. Vesalius' work touched off a flurry of anatomical work in Italy and elsewhere that culminated in the discovery of the circulation of the blood by William Harvey (Harvey, William), whose Exercitatio Anatomica De Motu Cordis et Sanguinis in Animalibus (An Anatomical Exercise Concerning the Motion of the Heart and Blood in Animals) was published in 1628. This was the Principia of physiology that established anatomy and physiology as sciences in their own right. Harvey showed that organic phenomena could be studied experimentally and that some organic processes could be reduced to mechanical systems. The heart and the vascular system could be considered as a pump and a system of pipes and could be understood without recourse to spirits or other forces immune to analysis.

      In other sciences the attempt to systematize and criticize was not so successful. In chemistry, for example, the work of the medieval and early modern alchemists had yielded important new substances and processes, such as the mineral acids and distillation, but had obscured theory in almost impenetrable mystical argot. Robert Boyle (Boyle, Robert) in England tried to clear away some of the intellectual underbrush by insisting upon clear descriptions, reproducibility of experiments, and mechanical conceptions of chemical processes. Chemistry, however, was not yet ripe for revolution.

      In many areas there was little hope of reducing phenomena to comprehensibility, simply because of the sheer number of facts (information science) to be accounted for. New instruments like the microscope and the telescope vastly multiplied the worlds with which man had to reckon. The voyages of discovery brought back a flood of new botanical and zoological specimens that overwhelmed ancient classificatory schemes. The best that could be done was to describe new things accurately and hope that someday they could all be fitted together in a coherent way.

      The growing flood of information put heavy strains upon old institutions and practices. It was no longer sufficient to publish scientific results in an expensive book that few could buy; information had to be spread widely and rapidly. Nor could the isolated genius, like Newton, comprehend a world in which new information was being produced faster than any single person could assimilate it. Natural philosophers had to be sure of their data, and to that end they required independent and critical confirmation of their discoveries. New means were created to accomplish these ends. Scientific societies sprang up, beginning in Italy in the early years of the 17th century and culminating in the two great national scientific societies that mark the zenith of the scientific revolution: the Royal Society of London for the Promotion of Natural Knowledge, created by royal charter in 1662, and the Académie des Sciences (Sciences, Academy of) of Paris, formed in 1666. In these societies and others like them all over the world, natural philosophers could gather to examine, discuss, and criticize new discoveries and old theories. To provide a firm basis for these discussions, societies began to publish scientific papers. The Royal Society's Philosophical Transactions, which began as a private venture of its secretary, was the first such professional scientific journal. It was soon copied by the French academy's Mémoires, which won equal importance and prestige. The old practice of hiding new discoveries in private jargon, obscure language, or even anagrams gradually gave way to the ideal of universal comprehensibility. New canons of reporting were devised so that experiments and discoveries could be reproduced by others. This required new precision in language and a willingness to share experimental or observational methods. The failure of others to reproduce results cast serious doubts upon the original reports. Thus were created the tools for a massive assault on nature's secrets.

      Even with the scientific revolution accomplished, much remained to be done. Again, it was Newton who showed the way. For the macroscopic world, the Principia sufficed. Newton's three laws of motion and the principle of universal gravitation were all that was necessary to analyze the mechanical relations of ordinary bodies, and the calculus provided the essential mathematical tools. For the microscopic world, Newton provided two methods. Where simple laws of action had already been determined from observation, as the relation of volume and pressure of a gas (Boyle's law, pv = k), Newton assumed forces (force) between particles that permitted him to derive the law. He then used these forces to predict other phenomena, in this case the speed of sound in air, that could be measured against the prediction. Conformity of observation to prediction was taken as evidence for the essential truth of the theory. Second, Newton's method made possible the discovery of laws of macroscopic action that could be accounted for by microscopic forces. Here the seminal work was not the Principia but Newton's masterpiece of experimental physics, the Opticks, published in 1704, in which he showed how to examine a subject experimentally and discover the laws concealed therein. Newton showed how judicious use of hypotheses could open the way to further experimental investigation until a coherent theory was achieved. The Opticks was to serve as the model in the 18th and early 19th centuries for the investigation of heat, light, electricity, magnetism, and chemical atoms.

The Classic Age of Science
      Just as the Principia preceded the Opticks, so, too, did mechanics maintain its priority among the sciences in the 18th century, in the process becoming transformed from a branch of physics into a branch of mathematics. Many physical problems were reduced to mathematical ones that proved amenable to solution by increasingly sophisticated analytical methods. The Swiss Leonhard Euler (Euler, Leonhard) was one of the most fertile and prolific workers in mathematics and mathematical physics. His development of the calculus of variations provided a powerful tool for dealing with highly complex problems. In France, Jean Le Rond d'Alembert (Alembert, Jean Le Rond d') and Joseph-Louis Lagrange (Lagrange, Joseph-Louis, comte de l'Empire) succeeded in completely mathematizing mechanics, reducing it to an axiomatic system requiring only mathematical manipulation.

      The test of Newtonian mechanics was its congruence with physical reality. At the beginning of the 18th century it was put to a rigorous test. Cartesians insisted that the Earth, because it was squeezed at the Equator by the etherial vortex causing gravity, should be somewhat pointed at the poles, a shape rather like that of an American football; Newtonians, arguing that centrifugal force was greatest at the Equator, calculated an oblate sphere that was flattened at the poles and bulged at the Equator. The Newtonians were proved correct after careful measurements of a degree of the meridian were made on expeditions to Lapland and to Peru. The final touch to the Newtonian edifice was provided by Pierre-Simon, marquis de Laplace (Laplace, Pierre-Simon, marquis de), whose masterly Traité de mécanique céleste (1798–1827; Celestial Mechanics) systematized everything that had been done in celestial mechanics under Newton's inspiration. Laplace went beyond Newton by showing that the perturbations of the planetary orbits caused by the interactions of planetary gravitation are in fact periodic and that the solar system is, therefore, stable, requiring no divine intervention.

      Although Newton was unable to bring to chemistry the kind of clarification he brought to physics, the Opticks did provide a method for the study of chemical phenomena. One of the major advances in chemistry in the 18th century was the discovery of the role of air, and of gases generally, in chemical reactions. This role had been dimly glimpsed in the 17th century, but it was not fully seen until the classic experiments of Joseph Black (Black, Joseph) on magnesia alba (basic magnesium carbonate) in the 1750s. By extensive and careful use of the chemical balance, Black showed that an air with specific properties could combine with solid substances like quicklime and could be recovered from them. This discovery served to focus attention on the properties of “air,” which was soon found to be a generic, not a specific, name. Chemists discovered a host of specific gases and investigated their various properties: some were flammable, others put out flames; some killed animals, others made them lively. Clearly, gases had a great deal to do with chemistry.

      The Newton of chemistry was Antoine-Laurent Lavoisier (Lavoisier, Antoine-Laurent). In a series of careful balance experiments Lavoisier untangled combustion reactions to show that, in contradiction to established theory, which held that a body gave off the principle of inflammation (called phlogiston) when it burned, combustion actually involves the combination of bodies with a gas that Lavoisier named oxygen. The chemical revolution was as much a revolution in method as in conception. Gravimetric methods made possible precise analysis, and this, Lavoisier insisted, was the central concern of the new chemistry. Only when bodies were analyzed as to their constituent substances was it possible to classify them and their attributes logically and consistently.

The imponderable fluids
      The Newtonian method of inferring laws from close observation of phenomena and then deducing forces from these laws was applied with great success to phenomena in which no ponderable matter figured. light, heat, electricity, and magnetism were all entities that were not capable of being weighed, i.e., imponderable. In the Opticks, Newton had assumed that particles of different sizes could account for the different refrangibility of the various colours of light. Clearly, forces of some sort must be associated with these particles if such phenomena as diffraction and refraction are to be accounted for. During the 18th century heat, electricity, and magnetism were similarly conceived as consisting of particles with which were associated forces of attraction or repulsion. In the 1780s, Charles-Augustin de Coulomb (Coulomb, Charles-Augustin de) was able to measure electrical and magnetic forces, using a delicate torsion balance of his own invention, and to show that these forces follow the general form of Newtonian universal attraction. Only light and heat failed to disclose such general force laws, thereby resisting reduction to Newtonian mechanics.

Science and the Industrial Revolution
      It has long been a commonsensical notion that the rise of modern science and the Industrial Revolution were closely connected. It is difficult to show any direct effect of scientific discoveries upon the rise of the textile or even the metallurgical industry in Great Britain, the home of the Industrial Revolution, but there certainly was a similarity in attitude to be found in science and nascent industry. Close observation and careful generalization leading to practical utilization were characteristic of both industrialists and experimentalists alike in the 18th century. One point of direct contact is known, namely James Watt's (Watt, James) interest in the efficiency of the Newcomen steam engine, an interest that grew from his work as a scientific-instrument maker and that led to his development of the separate condenser that made the steam engine an effective industrial power source. But in general the Industrial Revolution proceeded without much direct scientific help. Yet the potential influence of science was to prove of fundamental importance.

      What science offered in the 18th century was the hope that careful observation and experimentation might improve industrial production significantly. In some areas, it did. The potter Josiah Wedgwood (Wedgwood, Josiah) built his successful business on the basis of careful study of clays and glazes and by the invention of instruments like the pyrometer with which to gauge and control the processes he employed. It was not, however, until the second half of the 19th century that science was able to provide truly significant help to industry. It was then that the science of metallurgy permitted the tailoring of alloy steels to industrial specifications, that the science of chemistry permitted the creation of new substances, like the aniline dyes, of fundamental industrial importance, and that electricity and magnetism were harnessed in the electric dynamo and motor. Until that period science probably profited more from industry than the other way around. It was the steam engine that posed the problems that led, by way of a search for a theory of steam power, to the creation of thermodynamics. Most importantly, as industry required ever more complicated and intricate machinery, the machine tool industry developed to provide it and, in the process, made possible the construction of ever more delicate and refined instruments for science. As science turned from the everyday world to the worlds of atoms and molecules, electric currents and magnetic fields, microbes and viruses, and nebulae and galaxies, instruments increasingly provided the sole contact with phenomena. A large refracting telescope driven by intricate clockwork to observe nebulae was as much a product of 19th-century heavy industry as were the steam locomotive and the steamship.

      The Industrial Revolution had one further important effect on the development of modern science. The prospect of applying science to the problems of industry served to stimulate public support for science. The first great scientific school of the modern world, the École Polytechnique in Paris, was founded in 1794 to put the results of science in the service of France. The founding of scores more technical (technology, history of) schools in the 19th and 20th centuries encouraged the widespread diffusion of scientific knowledge and provided further opportunity for scientific advance. Governments, in varying degrees and at different rates, began supporting science even more directly, by making financial grants to scientists, by founding research institutes, and by bestowing honours and official posts on great scientists. By the end of the 19th century the natural philosopher following his private interests had given way to the professional scientist with a public role.

The Romantic (Romanticism) revolt
      Perhaps inevitably, the triumph of Newtonian mechanics elicited a reaction, one that had important implications for the further development of science. Its origins are many and complex, and it is possible here to focus on only one, that associated with the German philosopher Immanuel Kant (Kant, Immanuel). Kant challenged the Newtonian confidence that the scientist can deal directly with subsensible entities such as atoms, the corpuscles of light, or electricity. Instead, Kant insisted, all that the human mind can know is forces (force). This epistemological axiom freed Kantians from having to conceive of forces as embodied in specific and immutable particles. It also placed new emphasis on the space between particles; indeed, if one eliminated the particles entirely, there remained only space containing forces. From these two considerations were to come powerful arguments, first, for the transformations and conservation of forces and, second, for field theory as a representation of reality. What makes this point of view Romantic is that the idea of a network of forces in space tied the cosmos into a unity in which all forces were related to all others, so that the universe took on the aspect of a cosmic organism. The whole was greater than the sum of all its parts, and the way to truth was contemplation of the whole, not analysis.

      What Romantics, or nature (nature, philosophy of) philosophers, as they called themselves, could see that was hidden from their Newtonian colleagues was demonstrated by Hans Christian Ørsted (Ørsted, Hans Christian). He found it impossible to believe that there was no connection between the forces of nature. Chemical affinity, electricity, heat, magnetism, and light must, he argued, simply be different manifestations of the basic forces of attraction and repulsion. In 1820 he showed that electricity and magnetism were related, for the passage of an electrical current through a wire affected a nearby magnetic needle. This fundamental discovery was explored and exploited by Michael Faraday (Faraday, Michael), who spent his whole scientific life converting one force into another. By concentrating on the patterns of forces produced by electric currents and magnets, Faraday laid the foundations for field theory, in which the energy of a system was held to be spread throughout the system and not localized in real or hypothetical particles.

      The transformations of force necessarily raised the question of the conservation of force. Is anything lost when electrical energy is turned into magnetic energy, or into heat or light or chemical affinity or mechanical power? Faraday, again, provided one of the early answers in his two laws of electrolysis, based on experimental observations that quite specific amounts of electrical “force” decomposed quite specific amounts of chemical substances. This work was followed by that of James Prescott Joule (Joule, James Prescott), Robert Mayer, and Hermann von Helmholtz (Helmholtz, Hermann von), each of whom arrived at a generalization of basic importance to all science, the principle of the conservation of energy (energy, conservation of).

      The nature philosophers were primarily experimentalists who produced their transformations of forces by clever experimental manipulation. The exploration of the nature of elemental forces benefitted as well from the rapid development of mathematics. In the 19th century the study of heat was transformed into the science of thermodynamics, based firmly on mathematical analysis; the Newtonian corpuscular theory of light was replaced by Augustin-Jean Fresnel's (Fresnel, Augustin-Jean) mathematically sophisticated undulatory theory; and the phenomena of electricity and magnetism were distilled into succinct mathematical form by William Thomson (Lord Kelvin (Kelvin, William Thomson, Baron)) and James Clerk Maxwell (Maxwell, James Clerk). By the end of the century, thanks to the principle of the conservation of energy and the second law of thermodynamics, the physical world appeared to be completely comprehensible in terms of complex but precise mathematical forms describing various mechanical transformations in some underlying ether.

      The submicroscopic world of material atoms (atom) became similarly comprehensible in the 19th century. Beginning with John Dalton's (Dalton, John) fundamental assumption that atomic species differ from one another solely in their weights, chemists were able to identify an increasing number of elements and to establish the laws describing their interactions. Order was established by arranging elements according to their atomic weights and their reactions. The result was the periodic table, devised by Dmitry Mendeleyev (Mendeleyev, Dmitry Ivanovich), which implied that some kind of subatomic structure underlay elemental qualities. That structure could give rise to qualities, thus fulfilling the prophecy of the 17th-century mechanical philosophers, was shown in the 1870s by Joseph-Achille Le Bel (Le Bel, Joseph-Achille) and Jacobus van't Hoff (Hoff, Jacobus Henricus van 't), whose studies of organic chemicals showed the correlation between the arrangement of atoms or groups of atoms in space and specific chemical and physical properties.

The founding of modern biology
      The study of living matter lagged far behind physics and chemistry, largely because organisms are so much more complex than inanimate bodies or forces. Harvey had shown that living matter could be studied experimentally, but his achievement stood alone for two centuries. For the time being, most students of living nature had to be content to classify living forms as best they could and to attempt to isolate and study aspects of living systems.

      As has been seen, an avalanche of new specimens in both botany and zoology put severe pressure on taxonomy. A giant step forward was taken in the 18th century by the Swedish naturalist Carl von Linné (Linnaeus, Carolus)—known by his Latinized name, Linnaeus—who introduced a rational, if somewhat artificial, system of binomial nomenclature. The very artificiality of Linnaeus' system, focussing as it did on only a few key structures, encouraged criticism and attempts at more natural systems. The attention thus called to the organism as a whole reinforced a growing intuition that species are linked in some kind of genetic relationship, an idea first made scientifically explicit by Jean-Baptiste, chevalier de Lamarck (Lamarck, Jean-Baptiste).

      Problems encountered in cataloging the vast collection of invertebrates at the Museum of Natural History in Paris led Lamarck to suggest that species change (evolution) through time. This idea was not so revolutionary as it is usually painted, for, although it did upset some Christians who read the book of Genesis literally, naturalists who noted the shading of natural forms one into another had been toying with the notion for some time. Lamarck's system failed to gain general assent largely because it relied upon an antiquated chemistry for its causal agents and appeared to imply a conscious drive to perfection on the part of organisms. It was also opposed by one of the most powerful paleontologists and comparative anatomists of the day, Georges Cuvier (Cuvier, Georges, Baron), who happened to take Genesis quite literally. In spite of Cuvier's opposition, however, the idea remained alive and was finally elevated to scientific status by the labours of Charles Darwin (Darwin, Charles). Darwin not only amassed a wealth of data supporting the notion of transformation of species, but he also was able to suggest a mechanism by which such evolution could occur without recourse to other than purely natural causes. The mechanism was natural selection, according to which minute variations in offspring were either favoured or eliminated in the competition for survival, and it permitted the idea of evolution to be perceived with great clarity. Nature shuffled and sorted its own productions, through processes governed purely by chance, so that those organisms that survived were better adapted to a constantly changing environment.

      Darwin's On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life, published in 1859, brought order to the world of organisms. A similar unification at the microscopic level had been brought about by the cell theory announced by Theodor Schwann (Schwann, Theodor) and Matthias Schleiden (Schleiden, Mathias Jacob) in 1838, whereby cells were held to be the basic units of all living tissues. Improvements in the microscope during the 19th century made it possible gradually to lay bare the basic structures of cells, and rapid progress in biochemistry permitted the intimate probing of cellular physiology. By the end of the century the general feeling was that physics and chemistry sufficed to describe all vital functions and that living matter, subject to the same laws as inanimate matter, would soon yield up its secrets. This reductionist view was triumphantly illustrated in the work of Jacques Loeb (Loeb, Jacques), who showed that so-called instincts in lower animals are nothing more than physicochemical reactions, which he labelled tropisms (tropism).

      The most dramatic revolution in 19th-century biology was the one created by the germ theory of disease (human disease), championed by Louis Pasteur (Pasteur, Louis) in France and Robert Koch (Koch, Robert) in Germany. Through their investigations, bacteria were shown to be the specific causes of many diseases. By means of immunological methods first devised by Pasteur, some of mankind's chief maladies were brought under control.

The 20th-century revolution
      By the end of the 19th century, the dream of the mastery of nature for the benefit of mankind, first expressed in all its richness by Sir Francis Bacon, seemed on the verge of realization. Science was moving ahead on all fronts, reducing ignorance and producing new tools for the amelioration of the human condition. A comprehensible, rational view of the world was gradually emerging from laboratories and universities. One savant went so far as to express pity for those who would follow him and his colleagues, for they, he thought, would have nothing more to do than to measure things to the next decimal place.

      But this sunny confidence did not last long. One annoying problem was that the radiation emitted by atoms proved increasingly difficult to reduce to known mechanical principles. More importantly, physics found itself relying more and more upon the hypothetical properties of a substance, the ether, that stubbornly eluded detection. Within a span of 10 short years, roughly 1895–1905, these and related problems came to a head and wrecked the mechanistic system the 19th century had so laboriously built. The discovery of X rays and radioactivity revealed an unexpected new complexity in the structure of atoms. Max Planck's solution to the problem of thermal radiation introduced a discontinuity into the concept of energy that was inexplicable in terms of classical thermodynamics. Most disturbing of all, the enunciation of the special theory of relativity by Albert Einstein (Einstein, Albert) in 1905 not only destroyed the ether and all the physics that depended on it but also redefined physics as the study of relations between observers and events, rather than of the events themselves. What was observed, and therefore what happened, was now said to be a function of the observer's location and motion relative to other events. Absolute space was a fiction. The very foundations of physics threatened to crumble.

      This modern revolution in physics has not yet been fully assimilated by historians of science. Suffice it to say that scientists managed to come to terms with all of the upsetting results of early 20th-century physics but in ways that made the new physics utterly different from the old. Mechanical models were no longer acceptable, because there were processes (like light) for which no consistent model could be constructed. No longer could physicists speak with confidence of physical reality, but only of the probability of making certain measurements.

      All this being said, there is still no doubt that science in the 20th century has worked wonders. The new physics—relativity, quantum mechanics, particle physics—may outrage common sense, but it enables physicists to probe to the very limits of physical reality. Their instruments and mathematics permit modern scientists to manipulate subatomic particles with relative ease, to reconstruct the first moment of creation, and to glimpse dimly the grand structure and ultimate fate of the universe.

      The revolution in physics has spilled over into chemistry and biology and led to hitherto undreamed-of capabilities for the manipulation of atoms and molecules and of cells and their genetic structures. Chemists perform molecular tailoring today as a matter of course, cutting and shaping molecules at will. genetic engineering makes possible active human intervention in the evolutionary process and holds out the possibility of tailoring living organisms, including the human organism, to specific tasks. This second scientific revolution may prove to be, for good or ill, the most important event in the history of mankind.

L. Pearce Williams

Additional Reading

General works
George Sarton, A History of Science, 2 vol. (1952–59, reissued 1993), and Introduction to the History of Science, 3 vol. in 5 (1927–48, reprinted 1975), embody the legacy of this founder of the discipline of the history of science. Though more philosophical in tone, the immensely influential work by Thomas Kuhn, The Structure of Scientific Revolutions, 3rd ed. (1996), is essential reading for the student of the history of science. Of several surveys of the entire field, J.D. Bernal, Science in History, new ed., 4 vol. (1969, reissued 1979), although Marxist-inflected, is the most easily accessible and useful—vol. 1 and 2 are the strongest. David C. Lindberg, The Beginnings of Western Science (1992), is the best recent overview of Western science to 1400, with a thorough bibliography. Joseph Needham et al., Science and Civilisation in China (1954– ), is the indispensable history of Chinese science and technology; Colin Ronan, The Shorter Science and Civilisation in China (1980– ), is an abridgment.General studies of specific fields include Dirk J. Struik, A Concise History of Mathematics, 4th rev. ed. (1987); Dirk J. Struik (ed.), A Source Book in Mathematics, 1200–1800 (1969, reprinted 1986); on astronomy, Timothy Ferris, Coming of Age in the Milky Way (1988); and John North, The Fontana History of Astronomy and Cosmology (also published as The Norton History of Astronomy and Cosmology, 1994); Peter J. Bowler, The Fontana History of the Environmental Sciences (1992; also published as The Norton History of the Environmental Sciences, 1993), a history of biological and ecological thought; and William H. Brock, The Fontana History of Chemistry (1992; also published as The Norton History of Chemistry, 1993). The relationship between science and religion is surveyed in John Hedley Brooke, Science and Religion: Some Historical Perspectives (1991); and David C. Lindberg and Ronald L. Numbers (eds.), God and Nature (1986), a valuable collection of historical essays.Charles Coulton Gillispie (ed.), Dictionary of Scientific Biography, 16 vol. (1970–80), is the definitive biographical reference source for the field. Roy Porter (ed.), The Biographical Dictionary of Scientists, 2nd ed. (1994), offers briefer biographies. A series of annotated bibliographies of literature on the history of science and technology includes, for example, Gavin Bridson, The History of Natural History (1994); Bernard S. Finn, The History of Electrical Technology (1991); and David F. Channell, The History of Engineering Science (1989). Isis Current Bibliography of the History of Science and Its Cultural Influences (annual) surveys the most recent literature.Ongoing research is reported in a number of journals. Isis (quarterly) is the leading U.S. journal; and Osiris (annual) is also published in the United States; while History of Science (quarterly) and The British Journal for the History of Science (quarterly) are good British journals with wide coverage, bibliographies, and essay reviews. More specialized journals include Historical Studies in the Physical and Biological Sciences (semiannual); Journal of the History of Biology (3/yr.); Journal for the History of Astronomy (quarterly); Ambix (3/yr.), on the history of chemistry and alchemy; Studies in History and Philosophy of Science (quarterly); and Social Studies of Science (quarterly).

Ancient and medieval science
Mott T. Greene, Natural Knowledge in Preclassical Antiquity (1992), is a general survey of preclassical sciences; while O. Neugebauer, The Exact Sciences in Antiquity, 2nd ed. (1957, reissued 1993), focuses on ancient mathematics and astronomy. The best general history of Greek science is G.E.R. Lloyd, Early Greek Science: Thales to Aristotle (1970), and Greek Science After Aristotle (1973); Marshall Clagett, Greek Science in Antiquity (1955, reissued 1994), is also of value. G.E.R. Lloyd, Adversaries and Authorities (1996), is a comparative study of Greek and Chinese science. Roman science is treated in William H. Stahl, Roman Science: Origins, Development, and Influence to the Later Middle Ages (1962, reprinted 1978); and Roger French and Frank Greenaway (eds.), Science in the Early Roman Empire: Pliny the Elder, His Sources and Influences (1986).There is no book-length survey of Islāmic science, but A.I. Sabra, “Science, Islamic,” in Joseph R. Strayer (ed.), Dictionary of the Middle Ages (1988), vol. 11, pp. 81–89, is a good short overview. Astronomy and physics are discussed in A.I. Sabra, Optics, Astronomy, and Logic: Studies in Arabic Science and Philosophy (1994); David A. King, Astronomy in the Service of Islam (1993); and Edward Grant, Planets, Stars, and Orbs: The Medieval Cosmos, 1200–1687 (1994). Toby E. Huff, The Rise of Early Modern Science (1993), compares science in the medieval Islāmic world, China, and the West. A.C. Crombie, Augustine to Galileo, 2nd rev. ed., 2 vol. (1959, reissued as The History of Science from Augustine to Galileo, 2 vol. in 1, 1995), is still a useful introduction to medieval science. Edward Grant, Physical Science in the Middle Ages (1971); and Marshall Clagett, The Science of Mechanics in the Middle Ages (1959, reissued 1979), discuss medieval physics. Lynn White, “The Historical Roots of Our Ecological Crisis,” Science, 155:1203–07 (March 10, 1967), presents an influential argument about medieval attitudes toward nature. Joyce E. Salisbury, The Beast Within (1994), discusses medieval attitudes toward animals.

The scientific revolution
The vast literature on the scientific revolution is surveyed in H. Floris Cohen, The Scientific Revolution: A Historiographical Inquiry (1994). Allen G. Debus, Man and Nature in the Renaissance (1978); and Steven Shapin, The Scientific Revolution (1996), are the best short overviews. Additional surveys include Margaret C. Jacob, The Cultural Meaning of the Scientific Revolution (1988, reissued 1993), strong on comparative and social issues; David C. Lindberg and Robert S. Westman (eds.), Reappraisals of the Scientific Revolution (1990), a cross section of current scholarship; and A. Rupert Hall, The Scientific Revolution, 1500–1800: The Formation of the Modern Scientific Attitude, 2nd ed. (1962, reissued 1972), a classic intellectual history. The relationship between science, politics, and society in early modern England has been the subject of numerous important studies, including Robert K. Merton, Science, Technology & Society in Seventeenth Century England (1938, reissued 1993), which argues for a “Puritan spur to science”; while Charles Webster, The Great Instauration: Science, Medicine, and Reform, 1626–1660 (1975), presents a case for more radical interest in science in England. Steven Shapin and Simon Schaffer, Leviathan and the Air-Pump: Hobbes, Boyle, and the Experimental Life (1985); and Steven Shapin, A Social History of Truth: Civility and Science in Seventeenth-Century England (1994), trace the development of the experimental method. Samuel Y. Edgerton, Jr., The Heritage of Giotto's Geometry (1991), explores the relationship between art and science in the Renaissance. Physical science and astronomy are discussed in Alexandre Koyré, From the Closed World to the Infinite Universe (1957, reissued 1994); Thomas S. Kuhn, The Copernican Revolution: Planetary Astronomy in the Development of Western Thought (1957, reissued 1985); and Richard S. Westfall, The Construction of Modern Science: Mechanisms and Mechanics (1971). Natural history and biology are treated in Martin J.S. Rudwick, The Meaning of Fossils: Episodes in the History of Palaeontology, 2nd rev. ed. (1976, reprinted 1985). The links between Hermeticism, magic, and science are explored in Frances A. Yates, Giordano Bruno and the Hermetic Tradition (1964, reissued 1991); Keith Thomas, Religion and the Decline of Magic (1971); and Brian Vickers (ed.), Occult and Scientific Mentalities in the Renaissance (1984). Alchemy is examined in Betty Jo Teeter Dobbs, The Janus Faces of Genius: The Role of Alchemy in Newton's Thought (1991); William R. Newman, Gehennical Fire: The Lives of George Starkey, an American Alchemist in the Scientific Revolution (1994); and Pamela H. Smith, The Business of Alchemy: Science and Culture in the Holy Roman Empire (1994). Studies of gender and Renaissance science include Evelyn Fox Keller, Reflections on Gender and Science (1985, reissued 1995), an essential starting point; Carolyn Merchant, The Death of Nature: Women, Ecology, and the Scientific Revolution (1980, reissued 1990); and the controversial work by David F. Noble, A World Without Women: The Christian Clerical Culture of Western Science (1992).

Modern science
Thomas L. Hankins, Science and the Enlightenment (1985), is an excellent survey of 18th-century science. Joseph Ben-David, The Scientist's Role in Society (1971, reprinted with a new introduction, 1984), surveys the development of scientific institutions and communities in Europe and America. Margaret C. Jacob (ed.), The Politics of Western Science, 1640–1990 (1994); and Peter Galison and Bruce Hevly (eds.), Big Science: The Growth of Large-Scale Research (1992), are useful introductions to the politics of science. Works on gender and science include Margaret Alic, Hypatia's Heritage: A History of Women in Science from Antiquity Through the Nineteenth Century (1986), which is strongest on the modern period; Londa L. Schiebinger, Nature's Body: Gender in the Making of Modern Science (1993); Ludmilla Jordanova, Sexual Visions: Images of Gender in Science and Medicine Between the Eighteenth and Twentieth Centuries (1989, reissued 1993); and Margaret W. Rossiter, Women Scientists in America: Struggles and Strategies to 1940 (1982), and Women Scientists in America: Before Affirmative Action, 1940–1972 (1995). Surveys of science in the national and imperial context include Nancy Stepan, Beginnings of Brazilian Science (1976, reissued 1981); Lewis Pyenson, Cultural Imperialism and the Exact Sciences: German Expansion Overseas, 1900–1930 (1985); Daniel R. Headrick, The Tentacles of Progress: Technology Transfer in the Age of Imperialism, 1850–1940 (1988); James R. Bartholomew, The Formation of Science in Japan (1989); and Loren R. Graham, Science in Russia and the Soviet Union (1993).The literature on specific sciences since the 18th century is voluminous. The physical sciences are discussed in J.L. Heilbron, Electricity in the 17th and 18th Centuries (1979); Christa Jungnickel and Russell McCormmach, Intellectual Mastery of Nature: Theoretical Physics from Ohm to Einstein, 2 vol. (1986), strong on the late 19th century; Daniel J. Kevles, The Physicists (1978, reprinted 1995), a history of American physics; P.M. Harman, Energy, Force, and Matter (1982), on 19th-century developments; and three important collections of essays: Peter Galison, How Experiments End (1987), on high-energy physics; David Gooding, Trevor Pinch, and Simon Schaffer (eds.), The Uses of Experiment: Studies in the Natural Sciences (1989); and Jed Z. Buchwald (ed.), Scientific Practice: Theories and Stories of Doing Physics (1995). Henry C. King, The History of the Telescope (1955, reprinted 1979), is still useful as a history of astronomy; while Owen Gingerich (ed.), Astrophysics and Twentieth-Century Astronomy to 1950, vol. 1 (1984), is a more recent survey. Karl Hufbauer, Exploring the Sun: Solar Science Since Galileo (1991), focusses on solar physics. David H. DeVorkin, The History of Modern Astronomy and Astrophysics (1982), is an annotated bibliographic guide. Quantification and mathematics are discussed in Tore Frängsmyr, J.L. Heilbron, and Robin E. Rider (eds.), The Quantifying Spirit in the 18th Century (1990); Lorraine Daston, Classical Probability in the Enlightenment (1988); Theodore M. Porter, The Rise of Statistical Thinking, 1820–1900 (1986), and Trust in Numbers: The Pursuit of Objectivity in Science and Public Life (1995); and Joan L. Richards, Mathematical Visions: The Pursuit of Geometry in Victorian England (1988). Voyages of discovery are studied in Harry Woolf, The Transits of Venus: A Study of Eighteenth-Century Science (1959); Bernard Smith, European Vision and the South Pacific, 2nd ed. (1985), and Imagining the Pacific (1992); and William H. Goetzmann, New Lands, New Men: America and the Second Great Age of Discovery (1986, reissued 1995). Life sciences are discussed in David Elliston Allen, The Naturalist in Britain (1976, reissued 1994), a social history of natural history and collecting; W.F. Bynum, Science and the Practice of Medicine in the Nineteenth Century (1994); Garland E. Allen, Life Science in the Twentieth Century (1975), a general survey; and Ronald Rainger, Keith R. Benson, and Jane Maienschein (eds.), The American Development of Biology (1988, reissued 1991). Evolution is the subject of Peter J. Bowler, Evolution: The History of an Idea, rev. ed. (1989); and Adrian Desmond, The Politics of Evolution: Morphology, Medicine, and Reform in Radical London (1989, reissued 1992). Genetics and molecular biology are discussed in Horace Freeland Judson, The Eighth Day of Creation: Makers of the Revolution in Biology (1979); James D. Watson, The Double Helix (1968), available also in a critical edition ed. by Gunther S. Stent (1980), an interesting and controversial autobiographical account of the discovery of DNA structure; Adele E. Clarke and Joan H. Fujimura (eds.), The Right Tools for the Job (1992), on instruments and animals in 20th-century biology; and Robert E. Kohler, From Medical Chemistry to Biochemistry (1982), and Lords of the Fly (1994), on Drosophila melanogaster and genetics research. The social sciences are discussed in Christopher Fox, Roy Porter, and Robert Wokler (eds.), Inventing Human Science: Eighteenth-Century Domains (1995); Dorothy Ross, The Origins of American Social Science (1991); Daniel J. Kevles, In the Name of Eugenics: Genetics and the Uses of Human Heredity (1985, reissued 1995); George W. Stocking, Jr., Victorian Anthropology (1987); and Donna Haraway, Primate Visions: Gender, Race, and Nature in the World of Modern Science (1989), a history of primatology.L. Pearce Williams Ed.

* * *


Universalium. 2010.

Игры ⚽ Поможем написать курсовую

Look at other dictionaries:

  • National Museum of Nuclear Science & History — Coordinates: 35°03′58″N 106°32′02″W / 35.0660°N 106.5339°W / 35.0660; 106.5339 …   Wikipedia

  • Infoage Science/History Learning Center — The Infoage Science/History Learning Center is a science center founded by Fred Carl that will be opening on the grounds of the former Camp Evans military base in Wall Township, New Jersey, approximately 30 miles from New York City. Camp Evans is …   Wikipedia

  • Museum of Science & History — The Museum of Science History (MOSH) in Jacksonville, Florida, USA, is a private, non profit located on the Southbank Riverwalk and is the most visited museum in town. [http://www.coj.net/About+Jacksonville/Museums.htm] Primarily intended as a… …   Wikipedia

  • History of science — History of science …   Wikipedia

  • Science and the Church — • Dicsusses the relationship between the two subjects Catholic Encyclopedia. Kevin Knight. 2006. Science and the Church     Science and the Church      …   Catholic encyclopedia

  • History of astronomy — History of science …   Wikipedia

  • Science chretienne — Science chrétienne Église de la Science chrétienne à Boston La Science chrétienne (ou Christian Science) est un mouvement religieux fondé par Mary Baker Eddy en 1879 et rassemblant de nos jours 1 800 églises dans 82 pays …   Wikipédia en Français

  • Science in the medieval Islamic world — This article is about the history of science in the Islamic civilization between the 8th and 16th centuries. For information on science in the context of Islam, see Islam and science …   Wikipedia

  • Science (Philosophies of) — Philosophies of science Mach, Duhem, Bachelard Babette E.Babich THE TRADITION OF CONTINENTAL PHILOSOPHY OF SCIENCE If the philosophy of science is not typically represented as a ‘continental’ discipline it is nevertheless historically rooted in… …   History of philosophy

  • History of biology — For the video game, see History of Biology (video game). The frontispiece to Erasmus Darwin s evolution themed poem The Temple of Nature shows a goddess pulling back the veil from nature (in the person of Artemis). Allegory and metaphor have… …   Wikipedia

Share the article and excerpts

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