physical science

physical science
physical scientist.
1. any of the natural sciences dealing with inanimate matter or with energy, as physics, chemistry, and astronomy.
2. these sciences collectively.
[1835-45]

* * *

Introduction

      the systematic study of the inorganic world, as distinct from the study of the organic world, which is the province of biological science. Physical science is ordinarily thought of as consisting of four broad areas: astronomy, physics, chemistry, and the Earth sciences. Each of these is in turn divided into fields and subfields. This article discusses the historical development—with due attention to the scope, principal concerns, and methods—of the first three of these areas. The Earth sciences are discussed in a separate article.

      Physics, in its modern sense, was founded in the mid-19th century as a synthesis of several older sciences—namely, those of mechanics, optics, acoustics, electricity, magnetism, heat, and the physical properties of matter. The synthesis was based in large part on the recognition that the different forces of nature are related and are, in fact, interconvertible because they are forms of energy.

      The boundary between physics and chemistry is somewhat arbitrary. As it has developed in the 20th century, physics is concerned with the structure and behaviour of individual atoms and their components, while chemistry deals with the properties and reactions of molecules. These latter depend on energy, especially heat, as well as on atoms; hence, there is a strong link between physics and chemistry. Chemists tend to be more interested in the specific properties of different elements and compounds, whereas physicists are concerned with general properties shared by all matter.

      Astronomy is the science of the entire universe beyond the Earth; it includes the Earth's gross physical properties, such as its mass and rotation, insofar as they interact with other bodies in the solar system. Until the 18th century, astronomers were concerned primarily with the Sun, Moon, planets, and comets. During the last two centuries, however, the study of stars, galaxies, nebulas, and the interstellar medium has become increasingly important. Celestial mechanics, the science of the motion of planets and other solid objects within the solar system, was the first testing ground for Newton's laws of motion and thereby helped to establish the fundamental principles of classical (that is, pre-20th-century) physics. Astrophysics, the study of the physical properties of celestial bodies, arose during the 19th century and is closely connected with the determination of the chemical composition of those bodies. In the 20th century physics and astronomy have become more intimately linked through cosmological theories, especially those based on the theory of relativity.

Heritage of antiquity and the Middle Ages
      The physical sciences ultimately derive from the rationalistic Materialism that emerged in classical Greece (ancient Greek civilization), itself an outgrowth of magical and mythical views of the world. The Greek philosophers of the 6th and 5th centuries BC abandoned the animism of the poets and explained the world in terms of ordinarily observable natural processes. These early philosophers posed the broad questions that still underlie science: How did the world order emerge from chaos? What is the origin of multitude and variety in the world? How can motion and change be accounted for? What is the underlying relation between form and matter? Greek philosophy answered these questions in terms that provided the framework for science for approximately 2,000 years.

Ancient Middle Eastern and Greek astronomy
      Western astronomy had its origins in Egypt (Egypt, ancient) and Mesopotamia (Mesopotamia, history of). Egyptian astronomy, which was neither a very well-developed nor an influential study, was largely concerned with time reckoning. Its main lasting contribution was the civil calendar (Egyptian calendar) of 365 days, consisting of 12 months of 30 days each and five additional festival days at the end of each year. This calendar played an important role in the history of astronomy, allowing astronomers to calculate the number of days between any two sets of observations.

      Babylonian (Babylonia) astronomy, dating back to about 1800 BC, constitutes one of the earliest systematic, scientific treatments of the physical world. In contrast to the Egyptians, the Babylonians were interested in the accurate prediction of astronomical phenomena, especially the first appearance of the new Moon. Using the zodiac as a reference, by the 4th century BC, they developed a complex system of arithmetic progressions and methods of approximation by which they were able to predict first appearances. At no point in the Babylonian astronomical literature is there the least evidence of the use of geometric models. The mass of observations they collected and their mathematical methods were important contributions to the later flowering of astronomy among the Greeks.

      The Pythagoreans (Pythagoreanism) (5th century BC) were responsible for one of the first Greek astronomical theories. Believing that the order of the cosmos is fundamentally mathematical, they held that it is possible to discover the harmonies of the universe by contemplating the regular motions of the heavens (celestial mechanics). Postulating a central fire about which all the heavenly bodies including the Earth and Sun revolve, they constructed the first physical model of the solar system. Subsequent Greek astronomy derived its character from a comment ascribed to Plato, in the 4th century BC, who is reported to have instructed the astronomers to “save the phenomena” in terms of uniform circular (orbit) motion. That is to say, he urged them to develop predictively accurate theories using only combinations of uniform circular motion. As a result, Greek astronomers never regarded their geometric models as true or as being physical descriptions of the machinery of the heavens. They regarded them simply as tools for predicting planetary positions.

       Eudoxus of Cnidus (4th century BC) was the first of the Greek astronomers to rise to Plato's challenge. He developed a theory of homocentric spheres, a model that represented the universe by sets of nesting concentric spheres the motions of which combined to produce the planetary and other celestial motions. Using only uniform circular motions, Eudoxus was able to “save” the rather complex planetary motions with some success. His theory required four homocentric spheres for each planet and three each for the Sun and Moon. The system (geocentric system) was modified by Callippus, a student of Eudoxus, who added spheres to improve the theory, especially for Mercury and Venus. Aristotle, in formulating his cosmology, adopted Eudoxus' homocentric spheres as the actual machinery of the heavens. The Aristotelian cosmos was like an onion consisting of a series of some 55 spheres nested about the Earth, which was fixed at the centre. In order to unify the system, Aristotle added spheres in order to “unroll” the motions of a given planet so that they would not be transmitted to the next inner planet.

      The theory of homocentric spheres failed to account for two sets of observations: (1) brightness changes suggesting that planets are not always the same distance from the Earth, and (2) bounded elongations (i.e., Venus is never observed to be more than about 48° and Mercury never more than about 24° from the Sun). Heracleides of Pontus (Heracleides Ponticus) (4th century BC) attempted to solve these problems by having Venus and Mercury revolve about the Sun, rather than the Earth, and having the Sun and other planets revolve in turn about the Earth, which he placed at the centre. In addition, to account for the daily motions of the heavens, he held that the Earth rotates on its axis. Heracleides' theory had little impact in antiquity except perhaps on Aristarchus Of Samos (3rd century BC), who apparently put forth a heliocentric (heliocentric system) hypothesis similar to the one Copernicus was to propound in the 16th century.

       Hipparchus (fl. 130 BC) made extensive contributions to both theoretical and observational astronomy. Basing his theories on an impressive mass of observations, he was able to work out theories of the Sun and Moon that were more successful than those of any of his predecessors. His primary conceptual tool was the eccentric circle, a circle in which the Earth is at some point eccentric to the geometric centre. He used this device to account for various irregularities and inequalities observed in the motions of the Sun and Moon. He also proved that the eccentric circle is mathematically equivalent to a geometric figure called an epicycle-deferent system, a proof probably first made by Apollonius of Perga a century earlier.

      Among Hipparchus' observations, one of the most significant was that of the precession of the equinoxes—i.e., a gradual apparent increase in longitude between any fixed star and the equinoctial point (either of two points on the celestial sphere where the celestial equator crosses the ecliptic). Thus the north celestial pole, the point on the celestial sphere defined as the apparent centre of rotation of the stars, moves relative to the stars in its vicinity. In the heliocentric theory, this effect is ascribed to a change in the Earth's rotational axis, which traces out a conical path around the axis of the orbital plane.

      Claudius Ptolemy (fl. AD 140) applied the theory of epicycles to compile a systematic (Ptolemaic system) account of Greek astronomy. He elaborated theories for each of the planets, as well as for the Sun and Moon. His theory generally fitted the data available to him with a good degree of accuracy, and his book, the Almagest, became the vehicle by which Greek astronomy was transmitted to astronomers of the Middle Ages and Renaissance. It essentially molded astronomy for the next millennium and a half.

Greek physics
      Several kinds of physical theories emerged in ancient Greece, including both generalized hypotheses about the ultimate structure of nature and more specific theories that considered the problem of motion from both metaphysical and mathematical points of view. Attempting to reconcile the antithesis between the underlying unity and apparent multitude and diversity of nature, the Greek atomists Leucippus (mid-5th century BC), Democritus (late 5th century BC), and Epicurus (late 4th and early 3rd century BC) asserted that nature consists of immutable atoms (atomism) moving in empty space. According to this theory, the various motions and configurations of atoms and clusters of atoms are the causes of all the phenomena of nature.

      In contrast to the particulate universe of the atomists, the Stoics (Stoicism) (principally Zeno, of Citium, bridging 4th and 3rd centuries BC, Chrysippus [3rd century BC], and Poseidonius of Apamea [fl. c. 100 BC]) insisted on the continuity of nature, conceiving of both space and matter as continuous and as infused with an active, airlike spirit—pneuma—which serves to unify the frame of nature. The inspiration for the Stoic emphasis on pneumatic processes probably arose from earlier experiences with the “spring” (i.e., compressibility and pressure) of the air. Neither the atomic theory nor Stoic physics survived the criticism of Aristotle and his theory.

      In his physics, Aristotle was primarily concerned with the philosophical question of the nature of motion as one variety of change. He assumed that a constant motion requires a constant cause; that is to say, as long as a body remains in motion, a force must be acting on that body. He considered the motion of a body through a resisting medium as proportional to the force producing the motion and inversely proportional to the resistance of the medium. Aristotle used this relationship to argue against the possibility of the existence of a void, for in a void resistance is zero, and the relationship loses meaning. He considered the cosmos to be divided into two qualitatively different realms, governed by two different kinds of laws. In the terrestrial realm, within the sphere of the Moon, rectilinear up-and-down motion is characteristic. Heavy bodies, by their nature, seek the centre and tend to move downward in a natural motion. It is unnatural for a heavy body to move up, and such unnatural or violent motion requires an external cause. Light bodies, in direct contrast, move naturally upward. In the celestial realm, uniform circular motion is natural, thus producing the motions of the heavenly bodies.

       Archimedes (3rd century BC) fundamentally applied mathematics to the solution of physical problems and brilliantly employed physical assumptions and insights leading to mathematical demonstrations, particularly in problems of statics and hydrostatics. He was thus able to derive the law of the lever rigorously and to deal with problems of the equilibrium of floating bodies.

Islāmic and medieval science
      Greek science reached a zenith with the work of Ptolemy in the 2nd century AD. The lack of interest in theoretical questions in the Roman (ancient Rome) world reduced science in the Latin West to the level of predigested handbooks and encyclopaedias that had been distilled many times. Social pressures, political persecution, and the anti-intellectual bias of some of the early Church Fathers drove the few remaining Greek scientists and philosophers to the East. There they ultimately found a welcome when the rise of Islām (Islāmic world) in the 7th century stimulated interest in scientific and philosophical subjects. Most of the important Greek scientific texts were preserved in Arabic translations. Although the Muslims did not alter the foundations of Greek science, they made several important contributions within its general framework. When interest in Greek learning revived in western Europe during the 12th and 13th centuries, scholars turned to Islāmic Spain for the scientific texts. A spate of translations resulted in the revival of Greek science in the West and coincided with the rise of the universities. Working within a predominantly Greek framework, scientists of the late Middle Ages reached high levels of sophistication and prepared the ground for the scientific revolution of the 16th and 17th centuries.

      Mechanics was one of the most highly developed sciences pursued in the Middle Ages. Operating within a fundamentally Aristotelian framework, medieval physicists criticized and attempted to improve many aspects of Aristotle's physics.

      The problem of projectile motion was a crucial one for Aristotelian (Aristotelianism) mechanics, and the analysis of this problem represents one of the most impressive medieval contributions to physics. Because of the assumption that continuation of motion requires the continued action of a motive force, the continued motion of a projectile after losing contact with the projector required explanation. Aristotle himself had proposed explanations of the continuation of projectile motion in terms of the action of the medium. The ad hoc character of these explanations rendered them unsatisfactory to most of the medieval commentators, who nevertheless retained the fundamental assumption that continued motion requires a continuing cause.

      The most fruitful alternative to Aristotle's attempts to explain projectile motion resulted from the concept of impressed force. According to this view, there is an incorporeal motive force that is imparted to the projectile, causing it to continue moving. Such views were espoused by John Philoponus of Alexandria (fl. 6th century), Avicenna, the Persian philosopher (d. 1037), and the Arab Abū al Barakāt al-Baghdādi (d. 1164). In the 14th century the French philosopher Jean Buridan (Buridan, Jean) developed a new version of the impressed-force theory, calling the quality impressed on the projectile “impetus.” Impetus, a permanent quality for Buridan, is measurable by the initial velocity of the projectile and by the quantity of matter contained in it. Buridan employed this concept to suggest an explanation of the everlasting motions of the heavens.

      During the 1300s certain Oxford scholars pondered the philosophical problem of how to describe the change that occurs when qualities increase or decrease in intensity and came to consider the kinematic (kinematics) aspects of motion. Dealing with these problems in a purely hypothetical manner without any attempt to describe actual motions in nature or to test their formulas experimentally, they were able to derive the result that in a uniformly accelerated (acceleration) motion, distance increases as the square of the time.

      Although medieval science was deeply influenced by Aristotle's philosophy, adherence to his point of view was by no means dogmatic. During the 13th century, theologians at the University of Paris were disturbed by certain statements in Aristotle that seemed to imply limitations of God's powers as well as other statements, such as the eternity of the world, which stood in apparent contradiction to scripture. In 1277 Pope John XXI condemned 219 propositions, many from Aristotle and St. Thomas Aquinas, which had clearly theological consequences. Many of these condemned propositions had scientific implications as well. For example, one of these propositions states, “That the first cause (i.e., God) could not make several worlds.” Although it is unlikely that anyone in the Middle Ages actually asserted the existence of many worlds, the condemnation led to the discussion of that possibility, as well as other important problems such as the possibility that the Earth moved.

The scientific revolution
      During the 15th, 16th, and 17th centuries, scientific thought underwent a revolution. A new view of nature emerged, replacing the Greek view that had dominated science for almost 2,000 years. Science became an autonomous discipline, distinct from both philosophy and technology, and it came to be regarded as having utilitarian goals. By the end of this period, it may not be too much to say that science had replaced Christianity as the focal point of European civilization. Out of the ferment of the Renaissance and Reformation there arose a new view of science, bringing about the following transformations: the reeducation of common sense in favour of abstract reasoning; the substitution of a quantitative for a qualitative view of nature; the view of nature as a machine (mechanism) rather than as an organism; the development of an experimental method that sought definite answers to certain limited questions couched in the framework of specific theories; the acceptance of new criteria for explanation, stressing the “how” rather than the “why” that had characterized the Aristotelian search for final causes.

 The scientific revolution began in astronomy. Although there had been earlier discussions of the possibility of the Earth's motion, the Polish astronomer Nicolaus Copernicus (Copernicus, Nicolaus) was the first to propound a comprehensive heliocentric theory, equal in scope and predictive capability to Ptolemy's geocentric system. Motivated by the desire to satisfy Plato's dictum, Copernicus (Copernican system) was led to overthrow traditional astronomy because of its alleged violation of the principle of uniform circular motion and its lack of unity and harmony as a system of the world. Relying on virtually the same data as Ptolemy had possessed, Copernicus turned the world inside out, putting the Sun (heliocentric system) at the centre and setting the Earth into motion around it. Copernicus' theory, published in 1543, possessed a qualitative simplicity that Ptolemaic astronomy appeared to lack. To achieve comparable levels of quantitative precision, however, the new system became just as complex as the old. Perhaps the most revolutionary aspect of Copernican astronomy lay in Copernicus' attitude toward the reality of his theory. In contrast to Platonic instrumentalism, Copernicus asserted that to be satisfactory astronomy must describe the real, physical system of the world.

      The reception of Copernican astronomy amounted to victory by infiltration. By the time large-scale opposition to the theory had developed in the church and elsewhere, most of the best professional astronomers had found some aspect or other of the new system indispensable. Copernicus' book De revolutionibus orbium coelestium libri VI (“Six Books Concerning the Revolutions of the Heavenly Orbs”), published in 1543, became a standard reference for advanced problems in astronomical research, particularly for its mathematical techniques. Thus, it was widely read by mathematical astronomers, in spite of its central cosmological hypothesis, which was widely ignored. In 1551 the German astronomer Erasmus Reinhold published the Tabulae prutenicae (“Prutenic Tables”), computed by Copernican methods. The tables were more accurate and more up-to-date than their 13th-century predecessor and became indispensable to both astronomers and astrologers.

  During the 16th century the Danish astronomer Tycho Brahe (Brahe, Tycho), rejecting both the Ptolemaic and Copernican systems, was responsible for major changes in observation, unwittingly providing the data that ultimately decided the argument in favour of the new astronomy. Using larger, stabler, and better calibrated instruments, he observed regularly over extended periods, thereby obtaining a continuity of observations that were accurate for planets to within about one minute of arc—several times better than any previous observation. Several of Tycho's observations contradicted Aristotle's system: a nova that appeared in 1572 exhibited no parallax (meaning that it lay at a very great distance) and was thus not of the sublunary sphere and therefore contrary to the Aristotelian assertion of the immutability of the heavens; similarly, a succession of comets appeared to be moving freely through a region that was supposed to be filled with solid, crystalline spheres. Tycho devised his own world system—a modification of Heracleides'—to avoid various undesirable implications of the Ptolemaic and Copernican systems.

 At the beginning of the 17th century, the German astronomer Johannes Kepler (Kepler, Johannes) placed the Copernican hypothesis on firm astronomical footing. Converted to the new astronomy as a student and deeply motivated by a neo-Pythagorean desire for finding the mathematical principles of order and harmony according to which God had constructed the world, Kepler spent his life looking for simple mathematical relationships that described planetary motions. His painstaking search for the real order of the universe forced him finally to abandon the Platonic ideal of uniform circular motion in his search for a physical basis for the motions of the heavens.

      In 1609 Kepler (Kepler's laws of planetary motion) announced two new planetary laws derived from Tycho's data: (1) the planets travel around the Sun in elliptical orbits, one focus of the ellipse being occupied by the Sun; and (2) a planet moves in its orbit in such a manner that a line drawn from the planet to the Sun always sweeps out equal areas in equal times. With these two laws, Kepler abandoned uniform circular motion of the planets on their spheres, thus raising the fundamental physical question of what holds the planets in their orbits. He attempted to provide a physical basis for the planetary motions by means of a force analogous to the magnetic force, the qualitative properties of which had been recently described in England by William Gilbert (Gilbert, William) in his influential treatise, De Magnete, Magneticisque Corporibus et de Magno Magnete Tellure (1600; “On the Magnet, Magnetic Bodies, and the Great Magnet of the Earth”). The impending marriage of astronomy and physics had been announced. In 1618 Kepler stated his third law, which was one of many laws concerned with the harmonies of the planetary motions: (3) the square of the period in which a planet orbits the Sun is proportional to the cube of its mean distance from the Sun.

  A powerful blow was dealt to traditional cosmology by Galileo Galilei, who early in the 17th century used the telescope, a recent invention of Dutch lens grinders, to look toward the heavens. In 1610 Galileo announced observations that contradicted many traditional cosmological assumptions. He observed that the Moon is not a smooth, polished surface, as Aristotle had claimed, but that it is jagged and mountainous. Earthshine on the Moon revealed that the Earth, like the other planets, shines by reflected light. Like the Earth, Jupiter was observed to have satellites; hence, the Earth had been demoted from its unique position. The phases of Venus proved that that planet orbits the Sun, not the Earth.

      The battle for Copernicanism was fought in the realm of mechanics as well as astronomy. The Ptolemaic–Aristotelian system stood or fell as a monolith, and it rested on the idea of Earth's fixity at the centre of the cosmos. Removing the Earth from the centre destroyed the doctrine of natural motion and place, and circular motion of the Earth was incompatible with Aristotelian physics.

      Galileo's contributions to the science of mechanics were related directly to his defense of Copernicanism. Although in his youth he adhered to the traditional impetus physics, his desire to mathematize in the manner of Archimedes led him to abandon the traditional approach and develop the foundations for a new physics that was both highly mathematizable and directly related to the problems facing the new cosmology. Interested in finding the natural acceleration of falling bodies, he was able to derive the law of free fall (free-fall) (the distance, s, varies as the square of the time, t2). Combining this result with his rudimentary form of the principle of inertia, he was able to derive the parabolic path of projectile motion. Furthermore, his principle of inertia enabled him to meet the traditional physical objections to the Earth's motion: since a body in motion tends to remain in motion, projectiles and other objects on the terrestrial surface will tend to share the motions of the Earth, which will thus be imperceptible to someone standing on the Earth.

      The 17th-century contributions to mechanics of the French philosopher René Descartes (Descartes, René), like his contributions to the scientific endeavour as a whole, were more concerned with problems in the foundations of science than with the solution of specific technical problems. He was principally concerned with the conceptions of matter and motion as part of his general program for science—namely, to explain all the phenomena of nature in terms of matter and motion. This program, known as the mechanical philosophy, came to be the dominant theme of 17th-century science.

      Descartes rejected the idea that one piece of matter could act on another through empty space; instead, forces must be propagated by a material substance, the “ether,” that fills all space. Although matter tends to move in a straight line in accordance with the principle of inertia, it cannot occupy space already filled by other matter, so the only kind of motion that can actually occur is a vortex in which each particle in a ring moves simultaneously.

      According to Descartes, all natural phenomena depend on the collisions of small particles, and so it is of great importance to discover the quantitative laws of impact. This was done by Descartes's disciple, the Dutch physicist Christiaan Huygens, who formulated the laws of conservation of momentum and of kinetic energy (the latter being valid only for elastic collisions).

 The work of Sir Isaac Newton (Newton, Sir Isaac) represents the culmination of the scientific revolution at the end of the 17th century. His monumental Philosophiae Naturalis Principia Mathematica (1687; Mathematical Principles of Natural Philosophy) solved the major problems posed by the scientific revolution in mechanics and in cosmology. It provided a physical basis for Kepler's laws, unified celestial and terrestrial physics under one set of laws, and established the problems and methods that dominated much of astronomy and physics for well over a century. By means of the concept of force, Newton was able to synthesize two important components of the scientific revolution, the mechanical philosophy and the mathematization of nature.

      Newton was able to derive all these striking results from his three laws of motion (Newton's laws of motion):

1. Every body continues in its state of rest or of motion in a straight line unless it is compelled to change that state by force impressed on it;
2. The change of motion is proportional to the motive force impressed and is made in the direction of the straight line in which that force is impressed;
3. To every action there is always opposed an equal reaction: or, the mutual actions of two bodies upon each other are always equal.

      The second law was put into its modern form F = ma (where a is acceleration) by the Swiss mathematician Leonhard Euler in 1750. In this form, it is clear that the rate of change of velocity is directly proportional to the force acting on a body and inversely proportional to its mass.

      In order to apply his laws to astronomy, Newton had to extend the mechanical philosophy beyond the limits set by Descartes. He postulated a gravitational force acting between any two objects in the universe, even though he was unable to explain how this force could be propagated.

      By means of his laws of motion and a gravitational force proportional to the inverse square of the distance between the centres of two bodies, Newton could deduce Kepler's laws of planetary motion. Galileo's law of free fall is also consistent with Newton's laws. The same force that causes objects to fall near the surface of the Earth also holds the Moon and planets in their orbits.

      Newton's physics led to the conclusion that the shape of the Earth is not precisely spherical but should bulge at the Equator. The confirmation of this prediction by French expeditions in the mid-18th century helped persuade most European scientists to change from Cartesian to Newtonian physics. Newton also used the nonspherical shape of the Earth to explain the precession of the equinoxes, using the differential action of the Moon and Sun on the equatorial bulge to show how the axis of rotation would change its direction.

      The science of optics in the 17th century expressed the fundamental outlook of the scientific revolution by combining an experimental approach with a quantitative analysis of phenomena. Optics had its origins in Greece (ancient Greek civilization), especially in the works of Euclid (c. 300 BC), who stated many of the results in geometric optics that the Greeks had discovered, including the law of reflection: the angle of incidence is equal to the angle of reflection. In the 13th century, such men as Roger Bacon, Robert Grosseteste, and John Pecham, relying on the work of the Arab Alhazen (d. 1039), considered numerous optical problems, including the optics of the rainbow. It was Kepler, taking his lead from the writings of these 13th-century opticians, who set the tone for the science in the 17th century. Kepler introduced the point by point analysis of optical problems, tracing rays from each point on the object to a point on the image. Just as the mechanical philosophy (Materialism) was breaking the world into atomic parts, so Kepler approached optics by breaking organic reality into what he considered to be ultimately real units. He developed a geometric theory of lenses, providing the first mathematical account of Galileo's telescope.

      Descartes sought to incorporate the phenomena of light into mechanical philosophy by demonstrating that they can be explained entirely in terms of matter and motion. Using mechanical analogies, he was able to derive mathematically many of the known properties of light, including the law of reflection and the newly discovered law of refraction.

      Many of the most important contributions to optics in the 17th century were the work of Newton, especially the theory of colours (colour). Traditional theory considered colours to be the result of the modification of white light. Descartes, for example, thought that colours were the result of the spin of the particles that constitute light. Newton upset the traditional theory of colours by demonstrating in an impressive set of experiments that white light is a mixture out of which separate beams of coloured light can be separated. He associated different degrees of refrangibility with rays of different colours, and in this manner he was able to explain the way prisms produce spectra of colours from white light.

      His experimental method was characterized by a quantitative approach, since he always sought measurable variables and a clear distinction between experimental findings and mechanical explanations of those findings. His second important contribution to optics dealt with the interference phenomena that came to be called “ Newton's rings.” Although the colours of thin films (e.g., oil on water) had been previously observed, no one had attempted to quantify the phenomena in any way. Newton observed quantitative relations between the thickness of the film and the diameters of the rings of colour, a regularity he attempted to explain by his theory of fits of easy transmission and fits of easy reflection. Notwithstanding the fact that he generally conceived of light as being particulate, Newton's theory of fits involves periodicity and vibrations of ether, the hypothetical fluid substance permeating all space (see above).

      Huygens (Huygens, Christiaan) was the second great optical thinker of the 17th century. Although he was critical of many of the details of Descartes's system, he wrote in the Cartesian tradition, seeking purely mechanical explanations of phenomena. Huygens regarded light as something of a pulse phenomenon, but he explicitly denied the periodicity of light pulses. He developed the concept of wave front, by means of which he was able to derive the laws of reflection and refraction from his pulse theory and to explain the recently discovered phenomenon of double refraction.

      Chemistry had manifold origins, coming from such diverse sources as philosophy, alchemy, metallurgy, and medicine. It emerged as a separate science only with the rise of mechanical philosophy in the 17th century. Aristotle had regarded the four elements earth, water, air, and fire as the ultimate constituents of all things. Transmutable each into the other, all four elements were believed to exist in every substance. Originating in Egypt and the Middle East, alchemy had a double aspect: on the one hand it was a practical endeavour aimed to make gold from baser substances, while on the other it was a cosmological theory based on the correspondence between man and the universe at large. Alchemy contributed to chemistry a long tradition of experience with a wide variety of substances. Paracelsus, a 16th-century Swiss natural philosopher, was a seminal figure in the history of chemistry, putting together in an almost impenetrable combination the Aristotelian (Aristotelianism) theory of matter, alchemical correspondences, mystical forms of knowledge, and chemical therapy in medicine. His influence was widely felt in succeeding generations.

      During the first half of the 17th century, there were few established doctrines that chemists generally accepted as a framework. As a result, there was little cumulative growth of chemical knowledge. Chemists tended to build detailed systems, “chemical philosophies,” attempting to explain the entire universe in chemical terms. Most chemists accepted the traditional four elements (air, earth, water, fire), or the Paracelsian principles (salt, sulfur, mercury), or both, as the bearers of real qualities in substances; they also exhibited a marked tendency toward the occult.

      The interaction between chemistry and mechanical philosophy altered this situation by providing chemists with a shared language. The mechanical philosophy had been successfully employed in other areas; it seemed consistent with an experimental empiricism and seemed to provide a way to render chemistry respectable by translating it into the terms of the new science. Perhaps the best example of the influence of the mechanical philosophy is the work of Robert Boyle (Boyle, Robert). The thrust of his work was to understand the chemical properties of matter, to provide experimental evidence for the mechanical philosophy, and to demonstrate that all chemical properties can be explained in mechanical terms. He was an excellent laboratory chemist and developed a number of important techniques, especially colour-identification tests.

Science from the Enlightenment to the 20th century
      Seminal contributions to science are those that change the tenor of the questions asked by succeeding generations. The works of Newton formed just such a contribution. The mathematical rigour of the Principia and the experimental approach of the Opticks became models for scientists of the 18th and 19th centuries. Celestial mechanics developed in the wake of his Principia, extending its scope and refining its mathematical methods. The more qualitative, experimental, and hypothetical approach of Newton's Opticks influenced the sciences of optics, electricity and magnetism, and chemistry.

Celestial mechanics and astronomy
Impact of Newtonian theory
      Eighteenth-century theoretical astronomy in large measure derived both its point of view and its problems from the Principia. In this work Newton had provided a physics for the Copernican worldview by, among other things, demonstrating the implications of his gravitational theory for a two-body system consisting of the Sun and a planet. While Newton himself had grave reservations as to the wider scope of his theory, the 18th century witnessed various attempts to extend it to the solution of problems involving three gravitating bodies.

      Early in the 18th century the English astronomer Edmond Halley (Halley, Edmond), having noted striking similarities in the comets (Halley's Comet) that had been observed in 1531, 1607, and 1682, argued that they were the periodic appearances every 75 years or so of but a single comet that he predicted would return in 1758. Months before its expected return, the French mathematician Alexis Clairaut employed rather tedious and brute-force mathematics to calculate the effects of the gravitational attraction of Jupiter and Saturn on the otherwise elliptical orbit of Comet Halley. Clairaut was finally able to predict in the fall of 1758 that Comet Halley would reach perihelion in April of 1759, with a leeway of one month. Its actual return, in March, was an early confirmation of the scope and power of the Newtonian theory.

      It was, however, the three-body problem of either two planets and the Sun or the Sun–Earth–Moon system that provided the most persisting and profound test of Newton's theory. This problem, involving more regular members of the solar system (i.e., those describing nearly circular orbits having the same sense of revolution and in nearly the same plane), permitted certain simplifying assumptions and thereby invited more general and elegant mathematical approaches than the comet problem. An illustrious group of 18th-century continental mathematicians (including Clairaut; the Bernoulli family and Leonhard Euler (Euler, Leonhard) of Switzerland; and Jean Le Rond d'Alembert, Joseph-Louis, comte de Lagrange, and Pierre-Simon, marquis de Laplace, of France) attacked these astronomical problems, as well as related ones in Newtonian mechanics, by developing and applying the calculus of variations as it had been formulated by Leibniz (Leibniz, Gottfried Wilhelm). It is a lovely irony that this continental exploitation of Leibniz' mathematics—which was itself closely akin to Newton's version of calculus, which he called fluxions—was fundamental for the deepening establishment of the Newtonian theory to which Leibniz had objected because it reintroduced, according to Leibniz, occult forces into physics.

      In order to attack the lunar theory, which also commanded attention as the most likely astronomical approach to the navigational problem of determining longitude at sea, Clairaut was forced to adopt methods of approximation, having derived general equations that neither he nor anyone else could integrate. Even so, Clairaut was unable to calculate from gravitational theory a value for the progression of the lunar apogee greater than 50 percent of the observed value; therefore, he supposed in 1747 (with Euler) that Newton's inverse-square law was but the first term of a series and, hence, an approximation not valid for distances as small as that between the Earth and Moon. This attempted refinement of Newtonian theory proved to be fruitless, however, and two years later Clairaut was able to obtain, by more detailed and elaborate calculations, the observed value from the simple inverse-square relation.

      Certain of the three-body problems, most notably that of the secular acceleration of the Moon, defied early attempts at solution but finally yielded to the increasing power of the calculus of variations in the service of Newtonian theory. Thus it was that Laplace (Laplace, Pierre-Simon, marquis de)—in his five-volume Traité de mécanique céleste (1798–1827; Celestial Mechanics)—was able to comprehend the whole solar system (celestial mechanics) as a dynamically stable, Newtonian gravitational system. The secular acceleration of the Moon reappeared as a theoretical problem in the middle of the 19th century, persisting into the 20th century and ultimately requiring that the effects of the tides be recognized in its solution.

      Newtonian theory was also employed in much more dramatic discoveries that captivated the imagination of a broad and varied audience. Within 40 years of the discovery of Uranus in 1781 by the German-born British astronomer William Herschel (Herschel, Sir William), it was recognized that the planet's motion was somewhat anomalous. In the next 20 years the gravitational attraction of an unobserved planet was suspected to be the cause of Uranus' persisting deviations. In 1845 Urbain-Jean-Joseph Le Verrier (Le Verrier, Urbain-Jean-Joseph) of France and John Couch Adams (Adams, John Couch) of England independently calculated the position of this unseen body; the visual discovery (at the Berlin Observatory in 1846) of Neptune in just the position predicted constituted an immediately engaging and widely understood confirmation of Newtonian theory. In 1915 the American astronomer Percival Lowell (Lowell, Percival) published his prediction of yet another outer planet to account for further perturbations of Uranus not caused by Neptune. Pluto was finally discovered by sophisticated photographic techniques in 1930 and observed visually in 1950.

      In the second half of the 19th century, the innermost region of the solar system also received attention. In 1859 Le Verrier calculated the specifications of an intra-mercurial planet to account for a residual advance in the perihelion of Mercury's (Mercury) orbit (38 seconds of arc per century), an effect that was not gravitationally explicable in terms of known bodies. While a number of sightings of this predicted planet were reported between 1859 and 1878—the first of these resulting in Le Verrier's naming the new planet Vulcan—they were not confirmed by observations made either during subsequent solar eclipses or at the times of predicted transits of Vulcan across the Sun.

      The theoretical comprehension of Mercury's residual motion involved the first successful departure from Newtonian gravitational theory. This came in the form of Einstein's formulation of the principle of the general of relativity, one consequence of which was a correction term for a body in motion. The correction involves the square of the ratio of the body's velocity to the enormous velocity of light and hence, among the planets, is appreciable only for the fastest: Mercury. Einstein's theory accounted for the residual effect, which by 1915 was calculated to be 43 seconds. This achievement, combined with the 1919 observation of the bending of a ray of light passing near a massive body (another consequence of general relativity theory), constitutes the main experimental verification of that theory.

New discoveries
      Astronomy of the 18th, 19th, and early 20th centuries was not quite so completely Newtonian, however. Herschel's discovery of Uranus, for example, was not directly motivated by gravitational considerations. Nine years earlier, a German astronomer, Johann D. Titius (Titius, Johann Daniel), had announced a purely numerical sequence, subsequently refined by another German astronomer, Johann E. Bode (Bode, Johann Elert), that related the mean radii of the planetary orbits—a relation entirely outside gravitational theory. The sequence, called Bode's law (or the Bode–Titius law), is given by 0 + 4 = 4, 3 + 4 = 7, 3 × 2 + 4 = 10, 3 × 4 + 4 = 16, and so on, yielding additional values of 28, 52, and 100. If the measured radius of the Earth's orbit is defined as being 10, then to a very good approximation that of Mercury is 4, Venus is 7, Mars is 15 plus, Jupiter is 52, and Saturn is 95 plus. The fit where it can be made is good and continues since the next number in the sequence is 196 and the measured radius for Uranus' orbit is 191, but no planet had been observed to correspond to the Bode–Titius law value of 28. Powerful computational methods finally verified the existence of the asteroid Ceres and confirmed the law. The Bode–Titius law subsequently provided simplifying assumptions for the calculations of the predicted positions, which led to the observations of Neptune and Pluto, while the novel properties of the asteroids (asteroid) (nearly 500 of which had been discovered by the end of the century) stimulated star charts of the zodiacal regions, provided the means for improved measurements of solar-system distances, and forced astronomers, by their very number and variety, to face the question of the allocation of resources.

      Regularities in the structure of the solar system, such as the Bode–Titius law, and the fact that all planets move in the same direction around the Sun suggested that the system might originally have been formed by a simple mechanistic process. Laplace proposed that this process was driven by the cooling of the hot, extended, rotating atmosphere of the primitive Sun. As the atmosphere contracted, it would have to rotate faster (to conserve angular momentum), and when centrifugal force exceeded gravity at the outside, a ring of material would be detached, later to condense into a planet. The process would be repeated several times and might also produce satellites. After Herschel suggested that the nebulas he observed in the sky were condensing to stars, the Laplace theory became known as the “nebular hypothesis.” It was the favoured theory of the origin of the solar system throughout the 19th century. During this period the associated idea that the Earth was originally a hot fluid ball that slowly cooled down while forming a solid outer crust dominated geologic speculation.

      Attempts to detect the motion of the Earth caused investigators of the 18th and 19th centuries observational problems that were directly motivated by the Copernican theory. In 1728 the English astronomer James Bradley (Bradley, James) attributed annual changes that he observed in stellar positions to a slight tilting of the telescope with respect to the true direction of the star's light, a tilting that compensated for the Earth's motion. This effect, which depends also on the ratio of the Earth's velocity to the velocity of light, is the so-called aberration of light.

      In 1838 the long-sought “stellar parallax” effect—the apparent motion of nearby stars due to the Earth's annual motion around the Sun—was discovered by the German astronomer Friedrich Wilhelm Bessel. While anticlimactic as a verification of the Copernican hypothesis, the measurement of parallax provided for the first time a direct quantitative estimate of the distances of a few stars.

      While attention has been focused on the more positional aspects of astronomy, mention should be made of two other broad areas of investigation that in their 19th-century form derived largely from the work of William Herschel. These areas, dealing with more structural features of the heavens and with the physical character of the stars, developed in large measure with advancements in physics.

      Since they provided the principal basis for subsequent investigations, Newton's optical views were subject to close consideration until well into the 19th century. From his researches into the phenomena of colour, Newton became convinced that dispersion necessarily accompanies refraction and that chromatic aberration (colour distortion) could therefore be eliminated by employing reflectors, rather than refractors, as telescopes. By the mid-18th century Euler and others had theoretical arguments against Newton, and Euler offered the human eye as an example of an achromatic lens system. Although he was virtually alone in this, Euler also rejected Newton's essentially corpuscular theory of the nature of light by explaining optical phenomena in terms of vibrations in a fluid ether. The dominance of Newton's theory throughout the 18th century was due partly to its successful direct application by Newton and his followers and partly to the comprehensiveness of Newton's thought. For example, Bradley's observations found an immediate and natural explanation in terms of the corpuscular theory that also was supported by the accelerating success of Newton's gravitational theory involving discrete particles of matter.

      At the turn of the century, Thomas Young (Young, Thomas), an English physician studying the power of accommodation of the eye (i.e., its focusing power), was led gradually to extensive investigations and discoveries in optics, including the effect of interference. By means of a wave theory of light, Young was able to explain both this effect, which in its most dramatic manifestation results in two rays of light canceling each other to produce darkness, and also the various colour phenomena observed by Newton. The wave theory of light was developed from 1815 onward in a series of brilliant mathematical and experimental memoirs of the physicist Augustin-Jean Fresnel (Fresnel, Augustin-Jean) but was countered by adherents of the corpuscular theory, most notably by a group of other French scientists, Pierre-Simon Laplace, Siméon-Denis Poisson, Étienne Malus (Malus, Étienne-Louis), and Jean-Baptiste Biot, and most strikingly in connection with Malus' discovery (1808) of the polarization of light by reflection. Following Young's suggestion in 1817, Fresnel was able to render polarization effects comprehensible by means of a wave theory that considered light to be a transverse rather than a longitudinal wave, as the analogy with sound had suggested.

      The propagation of a transverse wave, the velocity of which through various media and under a variety of conditions was measured terrestrially with increasing accuracy from mid-century onward, seemed to require an ether having the properties of a highly elastic solid (e.g., such as steel), which, however, offered no resistance to the planetary motions. These bizarre properties stimulated a number of mechanical models of the ether, most notably those of the English physicist William Thomson, Lord Kelvin. In order to encompass the aberration of light by means of his wave theory, Fresnel had assumed that the motionless ether freely permeated the opaque Earth and thus remained unaffected by its motions. Furthermore, he derived as a theoretical consequence (verified experimentally in mid-century by Armand-Hippolyte-Louis Fizeau) that the ether was partially, and only partially, dragged along by a moving transparent substance depending on the index of refraction of the substance. However, all subsequent investigators (most notably the American scientists A.A. Michelson and Edward W. Morley, in 1887) failed in their attempts to measure the required ether drift. It was just to escape this difficulty of a necessary but undetected ether drift that George Francis FitzGerald (FitzGerald, George Francis) of England and the Dutch theorist Hendrik Antoon Lorentz (Lorentz, Hendrik Antoon) independently, at the close of the century, postulated the contraction of moving bodies in the direction of their motion through the ether. The Lorentz–FitzGerald contraction involves the square of the ratio of the velocity of the body to the velocity of light and ensures theoretically the experimental undetectability of the ether drift. It was the seeming necessity of arbitrary postulations of this kind that was eliminated by Einstein's formulation of relativity theory.

Electricity and magnetism
      Until the end of the 18th century, investigations in electricity and magnetism exhibited more of the hypothetical and spontaneous character of Newton's Opticks than the axiomatic and somewhat forbidding tone of his Principia. Early in the century, in England Stephen Gray and in France Charles François de Cisternay DuFay studied the direct and induced electrification of various substances by the two kinds of electricity (then called vitreous and resinous and now known as positive and negative), as well as the capability of these substances to conduct the “effluvium” of electricity. By about mid-century, the use of Leyden jars (to collect charges) and the development of large static electricity machines brought the experimental science into the drawing room, while the theoretical aspects were being cast in various forms of the single-fluid theory (by the American Benjamin Franklin and the German-born physicist F.M.U.T.H. Aepinus, among others) and the two-fluid theory.

      By the end of the 18th century, in England, Joseph Priestley (Priestley, Joseph) had noted that no electric effect was exhibited inside an electrified hollow metal container and had brilliantly inferred from this similarity that the inverse-square law (of gravity) must hold for electricity as well. In a series of painstaking memoirs, the French physicist Charles-Augustin de Coulomb (Coulomb, Charles-Augustin de), using a torsion balance that Henry Cavendish had used in England to measure the gravitational force (Coulomb's law), demonstrated the inverse-square relation for electrical and magnetic attractions and repulsions. Coulomb went on to apply this law to calculate the surface distribution of the electrical fluid in such a fundamental manner as to provide the basis for the 19th-century extensions by Poisson and Lord Kelvin.

      The discovery of galvanic electricity and the development of voltaic electricity opened whole new areas of investigation for the 19th century by providing convenient sources of sustained electrical current. The Danish physicist Hans Christian Ørsted's (Ørsted, Hans Christian) discovery, in 1820, of the magnetic effect accompanying an electric current led almost immediately to quantitative laws of electromagnetism (electromagnetic field) and electrodynamics. By 1827, André-Marie Ampère (Ampère, André-Marie) had published a series of mathematical and experimental memoirs on his electrodynamic theory that not only rendered electromagnetism comprehensible but also ordinary magnetism, identifying both as the result of electrical currents. Ampère solidly established his electrodynamics by basing it on inverse-square forces (which, however, are directed at right angles to, rather than in, the line connecting the two interacting elements) and by demonstrating that the effects do not violate Newton's third law of motion, notwithstanding their transverse direction.

      Michael Faraday's (Faraday, Michael) discovery in 1831 of electromagnetic induction (Faraday's law of induction) (the inverse of the effect discovered by Ørsted), his experimental determination of the identity of the various forms of electricity (1833), his discovery of the rotation of the plane of polarization of light by magnetism (1845), in addition to certain findings of other investigators—e.g., the discovery by James Prescott Joule in 1843 (and others) of the mechanical equivalent of heat (the conservation of energy)—all served to emphasize the essential unity of the forces of nature. Within electricity and magnetism attempts at theoretical unification were conceived in terms of either gravitational-type forces acting at a distance, as with Ampère, or, with Faraday, in terms of lines of force and the ambient medium in which they were thought to travel. The German physicists Wilhelm Eduard Weber (Weber, Wilhelm Eduard) and Rudolph Kohlrausch, in order to determine the coefficients in his theory of the former kind, measured the ratio of the electromagnetic and electrostatic units of electrical charge to be equal to the velocity of light.

      The Scottish physicist James Clerk Maxwell (Maxwell, James Clerk) developed his profound mathematical electromagnetic theory from 1855 onward. He drew his conceptions from Faraday and thus relied fundamentally on the ether required by optical theory, while using ingenious mechanical models. One consequence of Maxwell's mature theory was that an electromagnetic wave must be propagated through the ether with a velocity equal to the ratio of the electromagnetic to electrostatic units. Combined with the earlier results of Weber and Kohlrausch, this result implied that light is an electromagnetic phenomenon. Moreover, it suggested that electromagnetic waves of wavelengths other than the narrow band corresponding to visible light should exist in nature or could be artificially generated.

      Maxwell's theory received direct verification in 1886, when Heinrich Hertz (Hertz, Heinrich) of Germany detected the predicted electromagnetic waves. Their use in long-distance communication—“radio”—followed within two decades, and gradually physicists became acquainted with the entire electromagnetic spectrum.

      Eighteenth-century chemistry was derived from and remained involved with questions of mechanics, light, and heat as well as with iatrochemistry and notions of medical therapy and the interaction between substances and the formation of new substances. Chemistry took many of its problems and much of its viewpoint from the Opticks and especially the “Queries” with which that work ends. Newton's suggestion of a hierarchy of clusters of unalterable particles formed by virtue of the specific attractions of its component particles led directly to comparative studies of interactions and thus to the tables of affinities of the physician Herman Boerhaave (Boerhaave, Hermann) and others early in the century. This work culminated at the end of the century in the Swede Torbern Bergman's (Bergman, Torbern Olof) table that gave quantitative values of the affinity of substances both for reactions when “dry” and when in solution and that considered double as well as simple affinities.

      Seventeenth-century investigations of “airs” or gases (gas), combustion and calcination, and the nature and role of fire were incorporated by the chemists Johann Joachim Becher (Becher, Johann Joachim) and Georg Ernst Stahl (Stahl, Georg Ernst) of Sweden into a theory of phlogiston. According to this theory, which was most influential after the middle of the 18th century, the fiery principle, phlogiston, was released into the air in the processes of combustion, calcination, and respiration. The theory explained that air was simply the receptacle for phlogiston, and any combustible or calcinable substance contained phlogiston as a principle or element and thus could not itself be elemental. Iron, in rusting, was considered to lose its compound nature and to assume its elemental state as the calx of iron by yielding its phlogiston into the ambient air.

      Investigations that isolated and identified various gases in the second half of the 18th century, most notably the English chemist Joseph Black's quantitative manipulations of “fixed air” (carbon dioxide) and Joseph Priestley's discovery of “dephlogisticated air” (oxygen), were instrumental for the French chemist Antoine Lavoisier's (Lavoisier, Antoine-Laurent) formulation of his own oxygen theory of combustion and rejection of the phlogiston theory (i.e., he explained combustion not as the result of the liberation of phlogiston, but rather as the result of the combination of the burning substance with oxygen). This transformation coupled with the reform in nomenclature at the end of the century (due to Lavoisier and others)—a reform that reflected the new conceptions of chemical elements, compounds, and processes—constituted the revolution in chemistry.

      Very early in the 19th century, another study of gases, this time in the form of a persisting Newtonian approach to certain meteorological problems by the British chemist John Dalton (Dalton, John), led to the enunciation of a chemical atomic (atom) theory. From this theory, which was demonstrated to agree with the law of definite proportions and from which the law of multiple proportions was derived, Dalton was able to calculate definite atomic weights by assuming the simplest possible ratio for the numbers of combining atoms. For example, knowing from experiment that the ratio of the combining weights of hydrogen to oxygen in the formation of water is 1 to 8 and by assuming that one atom of hydrogen combined with one atom of oxygen, Dalton affirmed that the atomic weight of oxygen was eight, based on hydrogen as one. At the same time, however, in France, Joseph-Louis Gay-Lussac (Gay-Lussac, Joseph-Louis), from his volumetric investigations of combining gases, determined that two volumes of hydrogen combined with one of oxygen to produce water. While this suggested H2O rather than Dalton's HO as the formula for water, with the result that the atomic weight of oxygen becomes 16, it did involve certain inconsistencies with Dalton's theory.

      As early as 1811 the Italian physicist Amedeo Avogadro (Avogadro, Amedeo) was able to reconcile Dalton's atomic theory with Gay-Lussac's volumetric law by postulating that Dalton's atoms were indeed compound atoms, or polyatomic. For a number of reasons, one of which involved the recent successes of electrochemistry, Avogadro's hypothesis was not accepted until it was reintroduced by the Italian chemist Stanislao Cannizzaro half a century later. From the turn of the century, the English scientist Humphry Davy and many others had employed the strong electric currents of voltaic piles for the analysis of compound substances and the discovery of new elements. From these results, it appeared obvious that chemical forces were essentially electrical in nature and that two hydrogen atoms, for example, having the same electrical charge, would repel each other and could not join to form the polyatomic molecule required by Avogadro's hypothesis. Until the development of a quantum-mechanical theory of the chemical bond, beginning in the 1920s, bonding was described by empirical “valence” rules but could not be satisfactorily explained in terms of purely electrical forces.

      Between the presentation of Avogadro's hypothesis in 1811 and its general acceptance soon after 1860, several experimental techniques and theoretical laws were used by various investigators to yield different but self-consistent schemes of chemical formulas and atomic weights. After its acceptance, these schemes became unified. Within a few years of the development of another powerful technique, spectrum analysis, by the German physicists Gustav Kirchhoff (Kirchhoff, Gustav Robert) and Robert Bunsen (Bunsen, Robert Wilhelm) in 1859, the number of chemical elements whose atomic weights and other properties were known had approximately doubled since the time of Avogadro's announcement. By relying fundamentally but not slavishly upon the determined atomic weight values and by using his chemical insight and intuition, the Russian chemist Dmitry Ivanovich Mendeleyev provided a classification scheme that ordered much of this burgeoning information and was a culmination of earlier attempts to represent the periodic repetition of certain chemical and physical properties of the elements.

      The significance of the atomic weights themselves remained unclear. In 1815 William Prout, an English chemist, had proposed that they might all be integer multiples of the weight of the hydrogen atom, implying that the other elements are simply compounds of hydrogen. More accurate determinations, however, showed that the atomic weights are significantly different from integers. They are not, of course, the actual weights of individual atoms, but by 1870 it was possible to estimate those weights (or rather masses) in grams by the kinetic theory of gases and other methods. Thus one could at least say that the atomic weight of an element is proportional to the mass of an atom of that element.

Margaret J. Osler J. Brookes Spencer Stephen G. Brush

Developments and trends of the 20th century

      Some of the most spectacular advances in modern astronomy have come from research on the large-scale structure and development of the universe. This research goes back to William Herschel's observations of nebulas at the end of the 18th century. Some astronomers considered them to be “island universes”—huge stellar systems outside of and comparable to the Milky Way Galaxy, to which the solar system belongs. Others, following Herschel's own speculations, thought of them simply as gaseous clouds—relatively small patches of diffuse matter within the Milky Way Galaxy, which might be in the process of developing into stars and planetary systems, as described in Laplace's nebular (nebula) hypothesis.

      In 1912 Vesto Melvin Slipher (Slipher, Vesto Melvin) began at the Lowell Observatory in Arizona an extensive program to measure the velocities of nebulas, using the Doppler shift of their spectral lines. (Doppler shift is the observed change in wavelength of the radiation from a source that results from the relative motion of the latter along the line of sight.) By 1925 he had studied about 40 nebulas, most of which were found to be moving away from the Earth according to the red shift (displacement toward longer wavelengths) of their spectra.

      Although the nebulas were apparently so far away that their distances could not be measured directly by the stellar parallax method, an indirect approach was developed on the basis of a discovery made in 1908 by Henrietta Swan Leavitt (Leavitt, Henrietta Swan) at the Harvard College Observatory. Leavitt studied the magnitudes (magnitude) (apparent brightnesses) of a large number of variable stars, including the type known as Cepheid variables. Some of them were close enough to have measurable parallaxes so that their distances and thus their intrinsic brightnesses could be determined. She found a correlation between brightness and period of variation. Assuming that the same correlation holds for all stars of this kind, their observed magnitudes and periods could be used to estimate their distances.

      In 1923 the American astronomer Edwin P. Hubble (Hubble, Edwin Powell) identified a Cepheid variable in the so-called Andromeda Nebula. Using Leavitt's period–brightness correlation, Hubble estimated its distance to be approximately 900,000 light-years. Since this was much greater than the size of the Milky Way system, it appeared that the Andromeda Nebula must be another galaxy (island universe) outside of our own.

      In 1929 Hubble combined Slipher's measurements of the velocities of nebulas with further estimates of their distances and found that on the average such objects are moving away from the Earth with a velocity proportional to their distance. Hubble's velocity–distance relation suggested that the universe of galactic nebulas is expanding (expanding universe), starting from an initial state about 2,000,000,000 years ago in which all matter was contained in a fairly small volume. Revisions of the distance scale in the 1950s and later increased the “Hubble age” of the universe to more than 10,000,000,000 years.

      Calculations by Aleksandr A. Friedmann in the Soviet Union, Willem de Sitter in The Netherlands, and Georges Lemaître in Belgium, based on Einstein's general theory of relativity, showed that the expanding universe could be explained in terms of the evolution of space itself. According to Einstein's theory, space is described by the non-Euclidean geometry proposed in 1854 by the German mathematician G.F. Bernhard Riemann. Its departure from Euclidean space is measured by a “ curvature” that depends on the density of matter. The universe may be finite, though unbounded, like the surface of a sphere. Thus the expansion of the universe refers not merely to the motion of extragalactic stellar systems within space but also to the expansion of the space itself.

      The beginning of the expanding universe was linked to the formation of the chemical elements in a theory developed in the 1940s by the physicist George Gamow (Gamow, George), a former student of Friedmann who had emigrated to the United States. Gamow proposed that the universe began in a state of extremely high temperature and density and exploded outward—the so-called big bang (big-bang model). Matter was originally in the form of neutrons, which quickly decayed into protons and electrons; these then combined to form hydrogen and heavier elements.

      Gamow's students Ralph Alpher and Robert Herman estimated in 1948 that the radiation left over from the big bang should by now have cooled down to a temperature just a few degrees above absolute zero (0 K, or -459° F). In 1965 the predicted cosmic background radiation was discovered by Arno A. Penzias and Robert W. Wilson of the Bell Telephone Laboratories as part of an effort to build sensitive microwave-receiving stations for satellite communication. Their finding provided unexpected evidence for the idea that the universe was in a state of very high temperature and density sometime between 10,000,000,000 and 20,000,000,000 years ago.

Evolution of stars and formation of chemical elements
      Just as the development of cosmology relied heavily on ideas from physics, especially Einstein's general theory of relativity, so did theories of stellar structure and evolution depend on discoveries in atomic physics. These theories also offered a fundamental basis for chemistry by showing how the elements could have been synthesized in stars (star).

      The idea that stars are formed by the condensation of gaseous clouds was part of the 19th-century nebular hypothesis (see above). The gravitational energy released by this condensation could be transformed into heat, but calculations by Hermann von Helmholtz and Lord Kelvin indicated that this process would provide energy to keep the Sun shining for only about 20,000,000 years. Evidence from radiometric dating, starting with the work of the British physicist Ernest Rutherford in 1905, showed that the Earth is probably several billion years old. Astrophysicists were perplexed: what source of energy has kept the Sun shining for such a long time?

      In 1925 Cecilia Payne, a graduate student from Britain at Harvard College Observatory, analyzed the spectra of stars using statistical atomic theories that related them to temperature, density, and composition. She found that hydrogen and helium are the most abundant elements in stars, though this conclusion was not generally accepted until it was confirmed four years later by the noted American astronomer Henry Norris Russell. By this time Prout's hypothesis that all the elements are compounds of hydrogen had been revived by physicists in a somewhat more elaborate form. The deviation of atomic weights from exact integer values (expressed as multiples of hydrogen) could be explained partly by the fact that some elements are mixtures of isotopes with different atomic weights and partly by Einstein's relation between mass and energy (taking account of the binding energy of the forces that hold together the atomic nucleus). The German physicist Werner Heisenberg (Heisenberg, Werner) proposed in 1932 that, whereas the hydrogen nucleus consists of just one proton, all heavier nuclei contain protons and neutrons. Since a proton can be changed into a neutron by fusing it with an electron, this meant that all the elements could be built up from protons and electrons—i.e., from hydrogen atoms.

      In 1938 the German-born physicist Hans Bethe (Bethe, Hans) proposed the first satisfactory theory of stellar energy generation based on the fusion of protons to form helium and heavier elements. He showed that once elements as heavy as carbon had been formed, a cycle of nuclear reactions could produce even heavier elements. Fusion of hydrogen into heavier elements would also provide enough energy to account for the Sun's energy generation over a period of billions of years. Although Bethe's theory, as extended by Fred Hoyle, Edwin E. Salpeter, and William A. Fowler, is the best one available, there is still some doubt about its accuracy because the neutrinos supposedly produced by the fusion reactions have not been observed in the amounts predicted.

      According to the theory of stellar evolution developed by the Indian-born American astrophysicist Subrahmanyan Chandrasekhar and others, a star will become unstable after it has converted most of its hydrogen to helium and may go through stages of rapid expansion and contraction. If the star is much more massive than the Sun, it will explode violently, giving rise to a supernova. The explosion will synthesize heavier elements and spread them throughout the surrounding interstellar medium, where they provide the raw material for the formation of new stars and eventually of planets and living organisms.

      After a supernova explosion, the remaining core of the star may collapse further under its own gravitational attraction to form a dense star composed mainly of neutrons. This so-called neutron star, predicted theoretically in the 1930s by the astronomers Walter Baade and Fritz Zwicky, is apparently the same as the pulsar (a source of rapid, very regular pulses of radio waves), discovered in 1967 by Jocelyn Bell of the British radio astronomy group under Antony Hewish at Cambridge University.

      More massive stars may undergo a further stage of evolution beyond the neutron star: they may collapse to a black hole, in which the gravitational force is so strong that even light cannot escape. The black hole as a singularity in an idealized space-time universe was predicted from the general relativity theory by the German astronomer Karl Schwarzschild in 1916. Its role in stellar evolution was later described by the American physicists J. Robert Oppenheimer and John Wheeler. During the 1980s, possible black holes were thought to have been located in X-ray sources and at the centre of certain galaxies.

Solar-system (solar system) astronomy
      This area of investigation, which lay relatively dormant through the first half of the 20th century, was revived in the 1960s under the stimulus of the Soviet and American space programs. Missions to the Moon and planets yielded a wealth of complex information which has yet to be completely assimilated by scientists. A single example of the resulting change in ideas about the history of the solar system will have to suffice here. Before the first manned lunar landing in 1969, there were three competing hypotheses about the origin of the Moon: (1) formation in its present orbit simultaneously with the Earth, as described in the nebular hypothesis; (2) formation elsewhere and subsequent capture by the Earth; and (3) ejection from the Earth by fission (popularly known theory that the Moon emanated from what is now the Pacific Ocean Basin). Following the analysis of lunar samples and theoretical criticism of these hypotheses, lunar scientists came to the conclusion that none of them was satisfactory. Photographs of the surface of Mercury taken by the U.S. Mariner 10 spacecraft in 1974, however, showed that it is heavily cratered like the Moon's surface. This finding, together with theoretical calculations by V.S. Safronov of the Soviet Union and George W. Wetherill of the United States on the formation of planets by accumulation (accretion or aggregation) of smaller solid bodies, suggested that the Earth was also probably subject to heavy bombardment soon after its formation. In line with this, a theory proposed by the American astronomers William K. Hartmann and A.G.W. Cameron has become the most popular. According to their theory, the Earth was struck by a Mars-sized object, and the force of the impact vaporized the outer parts of both bodies. The vapour thus produced remained in orbit around the Earth and eventually condensed to form the Moon. Like the hypothesis proposed by Luis Alvarez that attributes the extinction of the dinosaurs to an asteroid impact, the Hartmann–Cameron theory seemed so bizarre that it could not have been taken seriously until compelling evidence became available.

      During the years 1896–1932 the foundations of physics changed so radically that many observers describe this period as a scientific revolution comparable in depth, if not in scope, to the one that took place during the 16th and 17th centuries. The 20th-century revolution changed many of the ideas about space, time, mass, energy, atoms, light, force, determinism, and causality that had apparently been firmly established by Newtonian physics during the 18th and 19th centuries. Moreover, according to some interpretations, the new theories demolished the basic metaphysical assumption of earlier science that the entire physical world has a real existence and objective properties independent of human observation.

      Closer examination of 19th-century physics shows that Newtonian ideas were already being undermined in many areas and that the program of mechanical explanation was openly challenged by several influential physicists toward the end of the century. Yet, there was no agreement as to what the foundations of a new physics might be. Modern textbook writers and popularizers often try to identify specific paradoxes or puzzling experimental results—e.g., the failure to detect the Earth's absolute motion in the Michelson–Morley experiment—as anomalies that led physicists to propose new fundamental theories such as relativity. Historians of science have shown, however, that most of these anomalies did not directly cause the introduction of the theories that later resolved them. As with Copernicus' introduction of heliocentric astronomy, the motivation seems to have been a desire to satisfy aesthetic principles of theory structure rooted in earlier views of the world rather than a need to account for the latest experiment or calculation.

radioactivity and the transmutation of elements
      The discovery of radioactivity by the French physicist Henri Becquerel in 1896 is generally taken to mark the beginning of 20th-century physics. The successful isolation of radium and other intensely radioactive substances by Marie and Pierre Curie focused the attention of scientists and the public on this remarkable phenomenon and promoted a wide range of experiments.

      Ernest Rutherford (Rutherford, Ernest, Baron Rutherford of Nelson, of Cambridge) soon took the lead in studying the nature of radioactivity. He found that there are two distinct kinds of radiation emitted in radioactivity called alpha (alpha particle) and beta rays. The alpha rays proved to be positively charged particles identical to ionized helium atoms. Beta rays are much less massive negatively charged particles; they were shown to be the same as the electrons discovered by J.J. Thomson in cathode rays in 1897. A third kind of ray, designated gamma, consists of high-frequency electromagnetic radiation.

      Rutherford proposed that radioactivity involves a transmutation of one element into another. This proposal called into question one of the basic assumptions of 19th-century chemistry: that the elements consist of qualitatively different substances—92 of them by the end of the century. It implied a return to the ideas of Prout and the ancient atomists—namely, that everything in the world is composed of only one or a few basic substances.

      Transmutation, according to Rutherford and his colleagues, was governed by certain empirical rules. For example, in alpha decay the atomic number of the “daughter” element is two less than that of the “mother” element, and its atomic weight is four less; this seems consistent with the fact that the alpha (alpha decay) ray, identified as helium, has atomic number 2 and atomic weight 4, so that total atomic number and total atomic weight are conserved in the decay reaction.

      Using these rules, Rutherford and his colleagues could determine the atomic numbers and atomic weights of many substances formed by radioactive decay, even though the substances decayed so quickly into others that these properties could not be measured directly. The atomic number of an element determines its place in Mendeleyev's periodic table (and thus its chemical properties; see above). It was found that substances of different atomic weight could have the same atomic number; such substances were called isotopes of an element.

      Although the products of radioactive decay are determined by simple rules, the decay process itself seems to occur at random. All one can say is that there is a certain probability that an atom of a radioactive substance will decay during a certain time interval, or, equivalently, that half of the atoms of the sample will have decayed after a certain time—i.e., the half life of the material.

The nucleus
      At the University of Manchester (England), Rutherford led a group that rapidly developed new ideas about atomic structure. On the basis of an experiment conducted by Hans Geiger and Ernest Marsden in which alpha particles were scattered by a thin film of metal, Rutherford (Rutherford atomic model) proposed a nuclear model of the atom (1911). In this model, the atom consists mostly of empty space, with a tiny, positively charged nucleus that contains most of the mass, surrounded by one or more negatively charged electrons. Henry G.J. Moseley, an English physicist, showed by an analysis of X-ray spectra that the electric charge on the nucleus is simply proportional to the atomic number of the element.

      During the 1920s physicists thought that the nucleus was composed of two particles: the proton (the positively charged nucleus of hydrogen) and the electron. In 1932 the English physicist James Chadwick discovered the neutron, a particle with about the same mass as the proton but no electric charge. Since there were technical difficulties with the proton–electron model of the nucleus, physicists were willing to accept Heisenberg's hypothesis that it consists instead of protons and neutrons. The atomic number is then simply the number of protons in the nucleus, while the mass number, the integer closest to the atomic weight, is equal to the total number of neutrons and protons. As mentioned above, this simple model of nuclear structure provided the basis for Hans Bethe's theory of the formation of elements from hydrogen in stars.

      In 1938 the German physicists Otto Hahn and Fritz Strassmann found that, when uranium is bombarded by neutrons, lighter elements such as barium and krypton are produced. This phenomenon was interpreted by Lise Meitner and her nephew Otto Frisch as a breakup, or fission, of the uranium nucleus into smaller nuclei. Other physicists soon realized that since fission produces more neutrons, a chain reaction could result in a powerful explosion. World War II was about to begin, and physicists who had emigrated from Germany, Italy, and Hungary to the United States and Great Britain feared that Germany might develop an atomic bomb (nuclear weapon) that could determine the outcome of the war. They persuaded the U.S. and British governments to undertake a major project to develop such a weapon first. The U.S. Manhattan Project did eventually produce atomic bombs based on the fission of uranium or of plutonium, a new artificially created element, and these were used against Japan in August 1945. Later, an even more powerful bomb based on the fusion of hydrogen atoms was developed and tested by both the United States and the Soviet Union. Thus nuclear physics began to play a major role in world history.

Einstein's 1905 trilogy
      In a few months during the years 1665–66, Newton discovered the composite nature of light, analyzed the action of gravity, and invented the mathematical technique now known as calculus—or so he recalled in his old age. The only person who has ever matched Newton's amazing burst of scientific creativity—three revolutionary discoveries within a year—was Albert Einstein (Einstein, Albert), who in 1905 published the special theory of relativity, the quantum theory of radiation, and a theory of Brownian movement that led directly to the final acceptance of the atomic structure of matter.

      Relativity theory has already been mentioned several times in this article, an indication of its close connection with several areas of physical science. There is no room here to discuss the subtle line of reasoning that Einstein followed in arriving at his amazing conclusions; a brief summary of his starting point and some of the consequences will have to suffice.

      In his 1905 paper on the electrodynamics of moving bodies, Einstein called attention to an apparent inconsistency in the usual presentation of Maxwell's electromagnetic theory as applied to the reciprocal action of a magnet and a conductor. The equations are different depending on which is “at rest” and which is “moving,” yet the results must be the same. Einstein located the difficulty in the assumption that absolute space exists; he postulated instead that the laws of nature are the same for observers in any inertial frame of reference (reference frame) and that the speed of light is the same for all such observers.

      From these postulates Einstein inferred: (1) an observer in one frame would find from his own measurements that lengths of objects in another frame are contracted by an amount given by the Lorentz–FitzGerald formula; (2) each observer would find that clocks in the other frame run more slowly; (3) there is no absolute time—events that are simultaneous in one frame of reference may not be so in another; and (4) the observable mass of any object increases as it goes faster.

      Closely connected with the mass-increase effect is Einstein's (Einstein's mass-energy relation) famous formula E = mc2: mass and energy are no longer conserved but can be interconverted. The explosive power of the atomic and hydrogen bombs derives from the conversion of mass to energy.

      In a paper on the creation and conversion of light (usually called the “photoelectric effect paper”), published earlier in 1905, Einstein proposed the hypothesis that electromagnetic radiation consists of discrete energy quanta that can be absorbed or emitted only as a whole. Although this hypothesis would not replace the wave theory (wave-particle duality) of light, which gives a perfectly satisfactory description of the phenomena of diffraction, reflection, refraction, and dispersion, it would supplement it by also ascribing particle properties to light.

      Until recently the invention of the quantum theory of radiation was generally credited to another German physicist, Max Planck (Planck, Max), who in 1900 discussed the statistical distribution of radiation energy in connection with the theory of blackbody radiation. Although Planck did propose the basic hypothesis that the energy of a quantum of radiation is proportional to its frequency of vibration, it is not clear whether he used this hypothesis merely for mathematical convenience or intended it to have a broader physical significance. In any case, he did not explicitly advocate a particle theory of light before 1905. Historians of physics still disagree on whether Planck or Einstein should be considered the originator of the quantum theory.

      Einstein's paper on Brownian movement (Brownian motion) seems less revolutionary than the other 1905 papers because most modern readers assume that the atomic structure of matter was well established at that time. Such was not the case, however. In spite of the development of the chemical atomic theory and of the kinetic theory of gases in the 19th century, which allowed quantitative estimates of such atomic properties as mass and diameter, it was still fashionable in 1900 to question the reality of atoms. This skepticism, which does not seem to have been particularly helpful to the progress of science, was promoted by the empiricist, or “positivist,” philosophy advocated by Auguste Comte, Ernst Mach, Wilhelm Ostwald, Pierre Duhem, Henri Poincaré, and others. It was the French physicist Jean Perrin who, using Einstein's theory of Brownian movement, finally convinced the scientific community to accept the atom as a valid scientific concept.

      The Danish physicist Niels Bohr (Bohr, Niels) pioneered the use of the quantum hypothesis in developing a successful theory of atomic structure. Adopting Rutherford's nuclear model, he proposed in 1913 that the atom is like a miniature solar system, with the electrons moving in orbits around the nucleus just as the planets move around the Sun. Although the electrical attraction between the electrons and nucleus is mathematically similar to the gravitational attraction between the planets and the Sun, the quantum hypothesis is needed to restrict the electrons to certain orbits and to forbid them from radiating energy except when jumping from one orbit to another.

      Bohr's model provided a good description of the spectra and other properties of atoms containing only one electron—neutral hydrogen and singly ionized helium—but could not be satisfactorily extended to multi-electron atoms or molecules. It relied on an inconsistent mixture of old and new physical principles, hinting but not clearly specifying how a more adequate general theory might be constructed.

      The nature of light was still puzzling to those who demanded that it should behave either like waves or like particles. Two experiments performed by American physicists seemed to favour the particle theory: Robert A. Millikan's confirmation of the quantum theory of the photoelectric effect proposed by Einstein; and Arthur H. Compton's experimental demonstration that X rays behave like particles when they collide with electrons. The findings of these experiments had to be considered along with the unquestioned fact that electromagnetic radiation also exhibits wave properties such as interference and diffraction.

      Louis de Broglie, a French physicist, proposed a way out of the dilemma: accept the wave–particle dualism as a description not only of light but also of electrons and other entities previously assumed to be particles. In 1926 the Austrian physicist Erwin Schrödinger (Schrödinger, Erwin) constructed a mathematical “wave mechanics” based on this proposal. His theory tells how to write down an equation for the wave function of any physical system in terms of the masses and charges of its components. From the wave function, one may compute the energy levels and other observable properties of the system.

      Schrödinger's equation (Schrödinger equation), the most convenient form of a more general theory called quantum mechanics to which the German physicists Werner Heisenberg and Max Born also contributed, was brilliantly successful. Not only did it yield the properties of the hydrogen atom but it also allowed the use of simple approximating methods for more complicated systems even though the equation could not be solved exactly. The application of quantum mechanics to the properties of atoms, molecules, and metals occupied physicists for the next several decades.

      The founders of quantum mechanics did not agree on the philosophical significance of the new theory. Born proposed that the wave function determines only the probability distribution of the electron's position or path; it does not have a well-defined instantaneous position and velocity. Heisenberg made this view explicit in his indeterminacy principle (uncertainty principle): the more accurately one determines the position, the less accurately the velocity is fixed; the converse is also true. Heisenberg's principle is often called the uncertainty principle, but this is somewhat misleading. It tends to suggest incorrectly that the electron really has a definite position and velocity and that they simply have not been determined.

      Einstein objected to the randomness implied by quantum mechanics in his famous statement that God “does not play dice.” He also was disturbed by the apparent denial of the objective reality of the atomic world: Somehow the electron's position or velocity comes into existence only when it is measured. Niels Bohr expressed this aspect of the quantum worldview in his complementarity principle, building on de Broglie's resolution of the wave–particle dichotomy: A system can have such properties as wave or particle behaviour that would be considered incompatible in Newtonian physics but that are actually complementary; light exhibits either wave behaviour or particle behaviour, depending on whether one chooses to measure the one property or the other. To say that it is really one or the other, or to say that the electron really has both a definite position and momentum at the same time, is to go beyond the limits of science.

      Bohr's viewpoint, which became known as the Copenhagen Interpretation of quantum mechanics, was that reality can be ascribed only to a measurement. Einstein argued that the physical world must have real properties whether or not one measures them; he and Schrödinger published a number of thought experiments designed to show that things can exist beyond what is described by quantum mechanics. During the 1970s and 1980s, advanced technology made it possible to actually perform some of these experiments, and quantum mechanics was vindicated in every case.

      The long-standing problem of the nature of the force that holds atoms together in molecules was finally solved by the application of quantum mechanics. Although it is often stated that chemistry has been “reduced to physics” in this way, it should be pointed out that one of the most important postulates of quantum mechanics was introduced primarily for the purpose of explaining chemical facts and did not originally have any other physical justification. This was the so-called exclusion principle (Pauli exclusion principle) put forth by the Austrian physicist Wolfgang Pauli, which forbids more than one electron occupying a given quantum state in an atom. The state of an electron includes its spin, a property introduced by the Dutch-born American physicists George E. Uhlenbeck and Samuel A. Goudsmit. Using that principle and the assumption that the quantum states in a multi-electron atom are essentially the same as those in the hydrogen atom, one can postulate a series of “shells (shell atomic model)” of electrons and explain the chemical valence of an element in terms of the loss, gain, or sharing of electrons in the outer shell.

      Some of the outstanding problems to be solved by quantum chemistry were: (1) The “saturation” of chemical forces. If attractive forces hold atoms together to form molecules, why is there a limit on how many atoms can stick together (generally only two of the same kind)? (2) Stereochemistry—the three-dimensional structure of molecules, in particular the spatial directionality of bonds as in the tetrahedral carbon atom. (3) Bond (chemical bonding) length—i.e., there seems to be a well-defined equilibrium distance between atoms in a molecule that can be determined accurately by experiment. (4) Why some atoms (e.g., helium) normally form no bonds with other atoms, while others form one or more. (These are the empirical rules of valence.)

      Soon after J.J. Thomson's discovery of the electron in 1897, there were several attempts to develop theories of chemical bonds based on electrons. The most successful was that proposed in the United States by G.N. Lewis (Lewis theory) in 1916 and Irving Langmuir in 1919. They emphasized shared pairs of electrons and treated the atom as a static arrangement of charges. While the Lewis–Langmuir model as a whole was inconsistent with quantum theory, several of its specific features continued to be useful.

      The key to the nature of the chemical bond was found to be the quantum-mechanical exchange effect, first described by Heisenberg in 1926–27. Resonance is related to the requirement that the wave function for two or more identical particles must have definite symmetry properties with respect to the coordinates of those particles—it must have plus or minus the same value (symmetric or antisymmetric, respectively) when those particles are interchanged. Particles such as electrons and protons, according to a hypothesis proposed by Enrico Fermi and P.A.M. Dirac, must have antisymmetric wave functions. Exchange may be imagined as a continual jumping back and forth or interchange of the electrons between two possible states. In 1927 the German physicists Walter Heitler and Fritz London used this idea to obtain an approximate wave function for two interacting hydrogen atoms. They found that with an antisymmetric wave function (including spin) there is an attractive force, while with a symmetric one there is a repulsive force. Thus two hydrogen atoms can form a molecule if their electron spins are opposite, but not if they are the same.

      The Heitler–London approach to the theory of chemical bonds was rapidly developed by John C. Slater and Linus C. Pauling (Pauling, Linus) in the United States. Slater proposed a simple general method for constructing multiple-electron wave functions that would automatically satisfy the Pauli exclusion principle. Pauling introduced a valence-bond method, picking out one electron in each of the two combining atoms and constructing a wave function representing a paired-electron bond between them. Pauling and Slater were able to explain the tetrahedral carbon structure in terms of a particular mixture of wave functions that has a lower energy than the original wave functions, so that the molecule tends to go into that state.

      About the same time another American scientist, Robert S. Mulliken, was developing an alternative theory of molecular structure based on what he called molecular orbitals. (The idea had been used under a different name by John E. Lennard-Jones of England in 1929 and by Erich Hückel of Germany in 1931.) Here, the electron is not considered to be localized in a particular atom or two-atom bond, but rather it is treated as occupying a quantum state (anorbital”) that is spread over the entire molecule.

      In treating the benzene molecule by the valence-bond method in 1933, Pauling and George W. Wheland constructed a wave function that was a linear combination of five possible structures—i.e., five possible arrangements of double and single bonds. Two of them are the structures that had been proposed by the German chemist August Kekulé (later Kekule von Stradonitz) in 1865, with alternating single and double bonds between adjacent carbon atoms in the six-carbon ring. The other three (now called Dewar structures for the British chemist and physicist James Dewar, though they were first suggested by H. Wichelhaus in 1869) have one longer bond going across the ring. Pauling and Dewar described their model as involving resonance between the five structures. According to quantum mechanics, this does not mean that the molecule is sometimes “really” in one state and at other times in another, but rather that it is always in a composite state.

      The valence-bond method, with its emphasis on resonance between different structures as a means of analyzing aromatic molecules, dominated quantum chemistry during the 1930s. The method was comprehensively presented and applied in Pauling's classic treatise The Nature of the Chemical Bond (1939), the most important work on theoretical chemistry in the 20th century. One reason for its popularity was that ideas similar to resonance had been developed by organic chemists, notably F.G. Arndt in Germany and Christopher K. Ingold in England, independently of quantum theory during the late 1920s.

      After World War II there was a strong movement away from the valence-bond method toward the molecular-orbital method, led by Mulliken in the United States and by Charles Coulson, Lennard-Jones, H.C. Longuet-Higgins, and Michael J.S. Dewar in England. The advocates of the molecular-orbital method argued that their approach was simpler and easier to apply to complicated molecules, since it allowed one to visualize a definite charge distribution for each electron.

Stephen G. Brush

Additional Reading
A. Rupert Hall and Marie Boas Hall, A Brief History of Science (1964, reprinted 1988), provides a good introduction to the subject; A.E.E. McKenzie, The Major Achievements of Science, 2 vol. (1960, reprinted 1988), concentrates on developments from the 16th century, with brief extracts from original sources; and Cecil J. Schneer, The Search for Order: The Development of the Major Ideas in the Physical Sciences from the Earliest Times to the Present (1960, reissued 1984 as The Evolution of Physical Science), accounts for the developments from the 17th through the 19th centuries. Comprehensive surveys include Stephen F. Mason, A History of the Sciences, new rev. ed. (1962); and Stephen Toulmin and June Goodfield, The Architecture of Matter (1962, reissued 1982). Thomas S. Kuhn, The Structure of Scientific Revolutions, 2nd enl. ed. (1970), presents the “paradigm” theory of science, based on historical examples; Gerald Holton, Thematic Origins of Scientific Thought: Kepler to Einstein, rev. ed. (1988), offers a new interpretation of the history of science, with case studies of the work of Einstein and others; and I. Bernard Cohen, Revolution in Science (1985), is a comparative study. See also Stephen G. Brush, The History of Modern Science: A Guide to the Second Scientific Revolution, 1800–1950 (1988). Charles Coulston Gillispie (ed.), Dictionary of Scientific Biography, 16 vol. (1970–80), is an excellent source for authoritative biographical data. For references to the scholarly literature in the history of science, consult the “Critical Bibliography of the History of Science and Its Cultural Influences,” an annual feature in Isis, an international review of the history of science. Cumulations of this bibliography appeared as Magda Whitrow (ed.), Isis Cumulative Bibliography: 1913–65, 6 vol. (1971–84); and John Neu (ed.), Isis Cumulative Bibliography 1966–1975, 2 vol. (1980–85).

* * *


Universalium. 2010.

Игры ⚽ Нужна курсовая?

Look at other dictionaries:

  • Physical science — is an encompassing term for the branches of natural science and science that study non living systems, in contrast to the biological sciences. However, the term physical creates an unintended, somewhat arbitrary distinction, since many branches… …   Wikipedia

  • physical science — Science Sci ence, n. [F., fr. L. scientia, fr. sciens, entis, p. pr. of scire to know. Cf. {Conscience}, {Conscious}, {Nice}.] 1. Knowledge; knowledge of principles and causes; ascertained truth of facts. [1913 Webster] If we conceive God s sight …   The Collaborative International Dictionary of English

  • physical science — n any of the natural sciences (as physics, chemistry, and astronomy) that deal primarily with nonliving materials physical scientist n …   Medical dictionary

  • physical science — n [U] also the physical sciences [plural] the sciences, for example ↑chemistry and ↑physics, that are concerned with studying things that are not living …   Dictionary of contemporary English

  • physical science — noun count sciences such as geography and physics that deal with things that are not alive. Sciences such as biology that deal with things that are alive are called life sciences …   Usage of the words and phrases in modern English

  • physical science — n. any of the sciences that deal with inanimate matter or energy, as physics, chemistry, geology, astronomy, etc …   English World dictionary

  • physical science — noun the physical properties, phenomena, and laws of something (Freq. 3) he studied the physics of radiation • Syn: ↑physics • Hypernyms: ↑natural science • Hyponyms: ↑acoustics * * * …   Useful english dictionary

  • physical science — UK / US noun [countable] Word forms physical science : singular physical science plural physical sciences sciences such as geography and physics that deal with things that are not alive. Sciences such as biology that deal with things that are… …   English dictionary

  • Physical Science for Christian Schools — (1974) is written by Emmett L. Williams and George Mulfinger, Jr. and was the first text book published by Bob Jones University Press. This book has been described as follows [in 1974] a full scale, Bible science textbook rolls off the Bob Jones… …   Wikipedia

  • physical science — physical sciences N COUNT: usu pl The physical sciences are branches of science such as physics, chemistry, and geology that are concerned with natural forces and with things that do not have life. ...the rapid growth of interest in both the… …   English dictionary

Share the article and excerpts

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