Religion And Science Research Paper

Academic Writing Service

Sample Religion And Science Research Paper. Browse other  research paper examples and check the list of research paper topics for more inspiration. If you need a religion research paper written according to all the academic standards, you can always turn to our experienced writers for help. This is how your paper can get an A! Feel free to contact our research paper writing service for professional assistance. We offer high-quality assignments for reasonable rates.

The intersection of religion and sciences has been both more complex and less conflictive than is frequently depicted by those who cite the trial of Galileo, the dispute over the theological implications of Charles Darwin’s theory of natural selection, and the Scopes Trial as evidence of the historical incompatibility of religious beliefs and scientific endeavor. This research paper will first examine the conflict thesis, then survey the history of the relations between religion and science in the Western tradition from the fifth century BC to the end of the twentieth century with special reference to Christianity.

Academic Writing, Editing, Proofreading, And Problem Solving Services

Get 10% OFF with 24START discount code


1. The Historiography Of Religion And Sciences

The thesis that Christianity has been historically hostile to science was popularized by two influential works that appeared in the late nineteenth century: John William Draper’s (1811–1882) History of the Conflict Between Religion and Science (1874) and Andrew Dickson White’s (1832–1918) A History of the Warfare of Science with Theology in Christendom (1896). White, the first president of Cornell University, argued that Christianity had a long history of opposing scientific progress in the interest of dogmatic theology. Wedding a triumphalist view of science to a dismissive view of theology, his thesis struck a responsive chord in American thought, which was at the turn of the century increasingly committed both to a secular outlook and to recognizing the central role that the sciences played in modern society. The Draper–White thesis became enormously influential in America both among academics and in the popular mind. It came to dominate the historical interpretation of the relations of religion to science during much of the twentieth century. Reflecting a positivist outlook, it viewed science as continually progressing and overcoming the entrenched antagonism of religious opinions, which invariably retreated before its advance.

Some historians regarded the Draper–White thesis as oversimplifying and distorting a complicated relationship. Thus, Burtt, in The Metaphysical Foundations of Modern Physical Science (1932), suggested that, far from being hindered by religion, science was in fact indebted to theological ideas. Alfred North Whitehead, in Science and the Modern World (1926), rejected the view that in controversies between science and religion, religion was wrong, and … science was always right. The true facts of the case are very much more complex, and refuse to be summarized in these simple terms. A number of important studies, by Robert K. Merton, Herbert Butterfield, and Alexandre Koyre, among others, also challenged the conflict thesis.




A smaller number of historians of science went even further to argue that specific components of Christian theology were responsible for the scientific revolution. Reijer Hooykaas contended that Calvinism’s accommodationist understanding of Scripture and its voluntaristic theology encouraged scientific endeavor. Similarly Stanley Jaki, in Science and Creation (1974), maintained that Christian theism, the doctrine of Creation, and the belief that the universe is governed by a rational and benevolent deity were necessary preconditions of the development of modern science. In the last three decades of the twentieth century several scholars have provided a more systematic reevaluation of the conflict thesis. Two major contributions to the literature came from Lindberg and Numbers (1986) and Brooke (1991). In the introduction to their composite volume, Lindberg and Numbers argued for an historical picture that recognized ‘a complex and diverse interaction that defies reduction to simple ‘‘conflict’’ or ‘‘harmony’’’ to replace the largely polemical one drawn by Draper and White. Brooke objected to imparting essentialist understandings to words like ‘science’ and ‘religion,’ which had many different meanings over time and could not be reduced to simple static concepts. ‘Natural philosophy,’ for example, denoted something very different in the seventeenth century from ‘science’ in the twentieth. The growing acknowledgment that in their relations religion and the sciences have exhibited a multiplicity of attitudes, reflecting local conditions and particular historical circumstances, led Brooke to suggest that a ‘complexity thesis’ is a more accurate model than the familiar conflict thesis. The views of Brooke, and Lindberg and Numbers have gained increasing acceptance among professional historians of science.

2. Greece And Rome

Western science as an explanatory enterprise began with the Greeks. It was in the sixth century BC that the Pre-Socratic philosophers like Thales (fl. 585 BC), Anaxagoras (c. 500–428 BC), and Anaximander (c. 610–546 BC) challenged the mythopoeic picture of the natural world. Most Pre-Socratics understood the world in nonmythological terms, viewing it as the product of natural forces. In the late fifth century BC they experienced some popular opposition in Athens. In c. 437–436 BC Anaxagoras was prosecuted on a charge of impiety and his books publicly burned.

By the end of the fifth century BC natural philosophy (physical science) had come to be recognized by the Greeks as offering the most suitable means of understanding the world of nature and religious issues and hereafter were to play a limited role in either hindering or encouraging scientific endeavor in the classical world. This does not mean that religious ideas were of no account in formulating world views for the Greeks or for the Romans, who took over the philosophical and scientific culture of the Greeks. The Epicureans explicitly repudiated a religious view of the universe while for the Stoics, science (especially cosmology) and religious thought were related intimately. Moreover, the Greek view of the world was predominantly an organicist or vitalist one. Philosophical naturalism, which was espoused by the Pre-Socratic philosophers, the sophists, Aristotle (384–322 BC), and the Epicureans, always remained a minority opinion in Graeco–Roman thought. In general the Greeks conceived of nature as a divinity. Plato (c. 427–347 BC) believed that a world soul (psyche) animated the universe, in which the sun, the moon, the stars, and the planets were gods. A vitalist world picture, stripped of its pagan elements, was transmitted by the Romans to Western Europeans, who continued to hold it well into the Middle Ages.

3. Late Antiquity And The Middle Ages

Christianity, founded by Jesus of Nazareth (c. 4 BC to 29 AD), spread rapidly throughout the Roman Empire owing to its extensive missionary activity. In the second century AD Christian apologists (theologians who defended their faith philosophically against pagan critics) began the process of harmonizing Christian revelation with Graeco–Roman philosophy. Apologists like Justin Martyr (c. 100–165), Tertullian (c. 160–220), and Irenaeus (c. 130–200) were intellectuals who had received their education in classical culture and were influenced strongly by it. Their opinion of natural philosophy, which was a component of Greek philosophy and inseparable from it, was ambiguous. While they often denounced it for its pagan religious or anti-Christian elements, they employed its methods, thought forms, and vocabulary—and drew on it when it was useful in the rational defense of Christianity. These church Fathers hellenized Christianity by taking over elements of classical culture and incorporating them into the Christian world view. Tertullian, for example, who asked, ‘What indeed has Athens to do with Jerusalem? What harmony is there between the Academy and the church?’ is sometimes cited as an example of early Christian anti-intellectualism. Yet he often borrowed from Greek natural science and medicine to buttress his theological arguments. Not all Christian apologists accepted Greek science, however. A few, like Tatian (second century), who were influenced by ascetic ideals or concerns about potential demonic influences, were either guarded in their acceptance or hostile in their approach to Greek pagan natural philosophy.

The most influential of the church Fathers was Augustine of Hippo in North Africa (354–430). Augustine shared the reservations of some of the Fathers about the value of natural philosophy but he recognized its utility. Like other Fathers he rejected curiositas, the desire to pursue knowledge for the sake of knowing, which he regarded as a characteristic feature of pagan learning. But Augustine believed that the study of natural philosophy was legitimated when it served higher purposes, such as aiding in the understanding of Scripture or lending support to Christian theology. By becoming the handmaiden of Christian theology natural philosophy was assured a place in the emerging Christian culture, one that was to last through the Middle Ages and well into the early-modern period.

The early Middle Ages (c. 500–1000) was a period of disintegration that followed the collapse of the Roman Empire in the West under repeated attacks and occupation by Germanic barbarians. As towns became villages, public services ceased to exist, and roads became impassable, the cultural continuity with the classical world was nearly severed. Educational institutions disappeared and learning remained alive only in Christian monasteries. A few distinguished figures stand out from the intellectual gloom of the period for their interest in natural philosophy, much of it derived from Neoplatonic sources. They include Cassiodorus (c. 485–580), Isidore, bishop of Seville (c. 560–636), the Venerable Bede (c. 673–735), and John Scotus Erigena (c. 810–877). These men could do little more than preserve as much as possible of the accumulated classical heritage, including natural philosophy, but they did so as clerics, within a monastic framework, and with the confidence that scientific knowledge, as a handmaiden to theology ( philosophia ancilla theologiae), was worthy of being transmitted to future generations.

The eleventh and twelfth centuries experienced the revival of trade between Western Europe and the eastern Mediterranean, together with the concomitant renewal of urban life, the growth of manufacturing, and the beginning of the gradual transformation of Europe from a rural to an urban economy. Cathedral schools sprang up in cities and gradually replaced monastic schools. In the twelfth century natural philosophers avidly studied Plato’s Timaeus and attempted to harmonize its cosmology with the Creation account in Genesis. The Arabs had preserved Greek learning, and works by Greek authors on natural philosophy were translated into Latin from Arabic, thus making available much new scientific material. The most important were the extant works of Aristotle, which by the thirteenth century had been translated in their entirety, replacing the influence of Plato and creating a major intellectual transformation of Western European universities. Muslim contributions (for example, in optics) were significant as well.

There was much in Aristotle’s writings that was useful to European Christians and that could be harmonized with Christian theology. But there were also areas of potential conflict, such as Aristotle’s belief that the world is eternal, that the soul is not immortal, and that the universe can be accounted for by naturalistic and even deterministic explanations apart from divine providence. These and other incompatible doctrines caused the teaching of Aristotle’s natural philosophy to be banned from the University of Paris (site of the preeminent theological faculty in Europe) in 1210 and 1215, but by 1255 his writings had gained official sanction and they came to be widely studied. A number of distinguished scholars, including Roger Bacon (1213–1291), Albert the Great (1200– 1280), and his student Thomas Aquinas (c. 1224– 1274), attempted to create a synthesis of Aristotelian philosophy and Christian theology that would reconcile the Christian faith and classical reason. The existence of a more radical faction at Paris, headed by Siger of Brabant (c. 1240–1284), led the bishop of Paris, Etienne Tempier, to condemn, at the request of traditional theologians, 13 philosophical propositions espoused by Siger. In 1277 he increased the number of propositions condemned to 219. The condemnations represented a conservative backlash that placed limits on speculation in natural philosophy. In the long run, however, they provided arts masters at Paris with incentives to explore other scientific options that lay outside Aristotelian natural philosophy, particularly by way of ‘thought experiments’ that entertained hypothetical possibilities, such as the existence of other worlds or of a vacuum. In fact, freedom of scientific speculation was relatively broad in late medieval universities, which were supported by the church.

4. The Copernican Revolution And Galileo

When joined with Aristotelian physics, the geocentric theory of the solar system, as refined by Ptolemy of Alexandria (second century AD), was almost universally accepted in the West and taken to be compatible with their theologies by Jews, Muslims, and Christians until the seventeenth century. The first major challenge to the Ptolemaic theory came with the publication of Nicolaus Copernicus’s (1473–1543) De revolutionibus orbium coelestium (On the Revolutions of the Heavenly Bodies), which appeared after the author’s death in

  1. Copernicus advocated a heliocentric model of the solar system. How he came to espouse it is unknown but he might never have published it had it not been for the encouragement of a young Lutheran astronomer, Georg Rheticus (1514–1574) of Wittenberg. Although they initially rejected its heliocentric cosmology, Lutherans at the University of Wittenberg, including the Protestant reformer Philip Melanchthon (1497–1560), were attracted to the mathematical convenience of the new theory. Elsewhere, however, European natural philosophers rejected it for a variety of reasons, many of which were scientific rather than theological. In the year 1600 there were probably no more than 10 Copernicans in all Europe. One scientist who was quick to embrace the new theory was Galileo Galilei (1564–1642), a young professor of mathematics at the University of Padua. Although secretly a Copernican by 1597, Galileo did not publicize his views until 1613, when in Letters on Sunspots he declared his acceptance of the heliocentric system. It was, however, his Letter to the Grand Duchess Christina (1615), in which Galileo argued that Scripture did not intend to impart scientific truth but accommodated itself to popular language in describing the natural world, that provoked the Catholic Church’s response. Although a layman, Galileo had undertaken to interpret the Bible. In the following year (1616) he appeared before the Congregation of the Holy Office (the Inquisition). Cardinal Bellarmine stated his willingness to accept the Copernican theory if it could be proven (which it could not be at the time); otherwise Scripture ought to be read literally as teaching geocentricity. The Church declared the theory ‘false and contrary to Holy Scripture’ and recommended that De revolutionibus be ‘suspended until corrected.’ Galileo was warned not to teach or defend heliocentrism except as a hypothesis until it could be proven. In 1632, after maintaining silence for 16 years, Galileo published his Dialogue on the Two Principal World Systems, an ostensibly even-handed debate between proponents of the Ptolemaic and Copernican theories in which the arguments for the Ptolemaic theory were made to look feeble. Galileo was summoned to Rome, where he stood trial in 1633. Compelled to abjure his belief that the earth moved, he was confined to a country estate near Florence, where he continued writing until his death in 1642.

The traditional picture of Galileo as a martyr to intellectual freedom and a victim of the Catholic Church’s hostility to science is little more than a caricature. The trial of Galileo must be understood against the backdrop of his own times. A vain and ambitious man, he did not suffer fools gladly and he had a penchant for making enemies in high places. Among those he alienated was Pope Urban VIII (1623–1644), who had earlier encouraged Galileo’s researches on the motion of the earth, but came to believe that Galileo had betrayed him in deliberately violating an injunction not to teach Copernicanism. Moreover, strong reaction against Copernicanism (and Galileo’s espousal of it) came from the Aristotelians of the scientific establishment, some of whom were determined to entangle Galileo with the Church. In particular, Galileo aroused the ire of a loosely organized movement of traditional scientists called the Liga, which succeeded in introducing theological issues into scientific discourse and attacked Galileo for his views.

In contrast with the relative freedom of intellectual debate that natural philosophers in European universities had enjoyed in the late Middle Ages, the Protestant Reformation had made the Catholic Church more cautious. In the question of whether Scripture ought to be read as teaching a geocentric universe, the Church was anxious to prove its good faith in interpreting Scripture literally in the face of Protestant attacks on its allegorizing tendencies in biblical exegesis. Aristotelianism became, in its Thomistic form, a rigid dogma and the papal court became, after the Council of Trent (1545–1563), more authoritarian. The Inquisition and the Index of Prohibited Books, introduced into Italy by forces of the Counter Reformation, were employed against Galileo. The prosecution of Galileo represents not the cause celebre in the alleged warfare of science with theology, but a dispute that involved multiple issues, including powerful personalities and vested professional and ecclesiastical interests, which complicated Galileo’s championing of a revolutionary change in astronomical models.

5. The Mechanical Philosophy And Natural Theology

By the mid-seventeenth century, the Copernican theory had come to be accepted widely throughout Europe, owing largely to Galileo’s promotion of heliocentrism. The seventeenth century saw, too, the formulation of a mechanical model of the universe, which attempted to account for all natural phenomena in terms of matter and motion without reference to action at a distance. The mechanical philosophy had its roots in corpuscular materialism, which explained physical phenomena by the interaction of tiny particles of matter called atoms. While elements of the model had been adopted by several of the leading figures of the scientific revolution, its most influential advocates were Rene Descartes (1596–1650) and Pierre Gassendi (1592–1655), who sought to devise a new philosophy that would replace Aristotelianism. While Descartes and Gassendi grounded their mechanical philosophies in theological considerations, critics feared that their views would lead to materialism, as they did in the case of Thomas Hobbes (1588–1679), or to deism, as they did in the case of Voltaire (1694–1778).

The mechanical philosophy received its greatest synthesis at the hands of Isaac Newton (1642–1727) in his Principia Mathematica Naturalis Philosophiae (Mathematical Principles of Natural Philosophy (1687)). By his discovery of the principle of universal gravitation, Newton succeeded in harmonizing celestial and terrestrial mechanics, bringing for the first time the whole of nature into a precise rational interpretation. Newton’s picture of a universe governed by uniform laws that could be described without recourse to primary causes resulted in the widespread belief that nature was self-governing and aided the ‘natural religion’ of deism and its idea of a ‘watchmaker’ God. It is ironic that Newton, himself a Christian (although an Arian), inadvertently laid the groundwork for the rationalism of the eighteenth-century Enlightenment. On the basis of his synthesis of the mechanical philosophy David Hume (1711–1776) developed a materialistic philosophy in England, while in France philosophers like Julien Offray de La Mettrie (1709–1751) and Baron d’Holbach (1723–1789) embraced atheistic materialism.

The popularity of the mechanical philosophy led in the latter half of the seventeenth century to the creation of a natural theology that continued into the nineteenth century. Natural theology describes the existence and attributes of God by means of knowledge derived from natural reason rather than from special revelation. There were several grounds for its attraction: the wonders of scientific discoveries, many of which could be seen for the first time with the invention of the microscope; the desire to secure an adequate foundation for a theology of divine providence that would confront the perceived dangers of materialistic atheism; and the enormous influence of Newton’s mechanical picture of the world. The argument from design became the cornerstone of a natural theology, in which God was described as a craftsman, a geometer, or an architect. Not all exponents of natural theology were deists; many theists, especially those who were repelled by religious enthusiasm, were involved in the enterprise of demonstrating the majesty of the Creator from the intricacies embedded in his creation. In William Paley’s (1743–1805) influential Natural Theology, or Evidences of the Existence and Attributes of the Deity (1802) natural theology reached its apex.

6. Geology And Darwinian Evolution

The influence of the Enlightenment, with its exaltation of reason and rejection of special revelation, can be observed in the developments that were occurring in science, particularly in the increasing tendency to seek naturalistic explanations and to divorce science altogether from theological considerations. In the seventeenth and eighteenth centuries theories of the earth had sought to weave together geology and theology in what was termed ‘physico-theology.’ Many of these theories calculated the age of the earth by using biblical chronologies in a way that could be harmonized with physical principles. Invariably they postulated a young earth. While most attributed organic fossils to the Noachian Flood, by the eighteenth century it was becoming increasingly difficult to explain the global distribution of fossils by a single event (the Deluge) that had taken place a few thousand years ago. The French naturalist Georges Louis Leclerc, Comte de Buffon (1707–1788), a deist, proposed a much older earth (publicly of 75,000 years and privately perhaps as old as three million years) and suggested that each of the six biblical days of Creation might correspond to an immense length of time. Georges Cuvier (1769–1832), a devout Protestant, demonstrated the existence of fossil mammals unlike any in his own day and without any apparent intermediate forms. He argued that a series of geologic catastrophes had eliminated species in past ages. Cuvier’s catastrophes were not global and in his case did not derive from his religious convictions. But they were later read as such and were appropriated in English natural theology to support catastrophism, which became a popular form of physico-theology that attributed geological changes to great catastrophes of the past, the most recent being the biblical Flood.

The Scottish geologist Charles Lyell (1797–1875) rejected altogether providential interventions in geologic history. In his Principles of Geology (3 vols. 1830–1833) he succeeded in severing geology from theology. Lyell rejected diluvialism (catastrophism) for what became known as uniformitarianism: the theory that geological changes on the earth’s surface have occurred slowly over long periods of time and that in explaining change we refer to processes that we currently observe. Previous proponents of diluvialism, like William Buckland (1784–1856) of Oxford and Adam Sedgwick (1785–1873) of Cambridge, quickly abandoned catastrophism and adopted the new geology, which Christians harmonized with the Genesis account of Creation in a relatively painless fashion. A group of laymen, who called themselves Mosaic or Scriptural geologists, continued to take their geology from Genesis but they were marginalized since nearly all professional geologists abandoned the Noachian Flood as an organizing idea in accounting for changes in the earth’s history.

Charles Darwin (1809–82) had, as a young man, been influenced by Paley’s natural theology. During his voyage aboard the H.M.S. Beagle, however, he read Lyell’s work and began to think that slow changes had produced the geological phenomena that he observed. His theory of evolution by natural selection, published in Origin of Species (1859), made use of Lyell’s theories that the earth was very old and the fossil record imperfect. Evolutionary ideas had preceded Darwin. Originating in the Enlightenment, they had already been proposed in the eighteenth century by Charles’s grandfather, Erasmus Darwin (1731–1802), and by Jean-Baptiste Lamarck (1744– 1829), the latter proposing the theory that acquired characteristics were inherited. By the mid-nineteenth century the belief that transmutationism (evolution) represented a divine plan was widespread. In advancing natural selection as a mechanism for evolution, Darwin argued that in the ‘struggle for existence’ those individuals with even a slight advantage would survive and pass their advantage on to their offspring. Herbert Spencer (1820–1903) referred to it as the principle of the ‘survival of the fittest.’

Darwin’s theory met with a mixed response that cut across conventional professional and religious boundaries. Some scientists initially rejected it, while a number of clergymen accepted it. Conservative anti-Darwinists, including several leading naturalists like Louis Agassiz of Harvard (1807–1873), geologists like Canadian William Dawson (1820–1899) of McGill, and theologians like Charles Hodge (1797–1878) of Princeton Seminary rejected it, fearing the break with final causes. A second group of Christians accepted evolutionary theory as compatible with a biblical faith. They included the Scot James McCosh (1811– 1894), president of Princeton University, and the American botanist Asa Gray (1810–1888) of Harvard. A third group came to believe that evolution had rendered supernatural religion untenable and opted (most often) for some form of philosophical materialism.

By the 1870s, however, most scientists and a large body of the educated public had accepted some form of evolution (though not necessarily natural selection). Full acceptance of Darwin’s views on human nature, as set forth in his Descent of Man (1871), was more difficult for Christians. Darwin maintained that moral values were rationalizations of social instincts that had developed by natural selection, thus asserting a full-blown materialism. Alfred Russell Wallace (1823– 1913), the codiscoverer of the theory of evolution by natural selection, rejected materialism, maintaining that the creation of the human mind required nonmaterial forces.

In the 1920s evolution was one of several theological issues that divided the major Protestant denominations of North America in the modernist–fundamentalist controversy. While many Protestants (as well as some Roman Catholics and Jews) had come to believe in an old earth and to harmonize belief in evolution with a nonliteral interpretation of the Genesis account of Creation, conservatives emphasized in their rejection the atheistic and materialistic dangers inherent in a naturalistic theory of human origins. But the Scopes trial, held in Dayton, TN in 1925, symbolized the inability of biblical literalists to withstand the force of Darwinism, which was taken for granted by the intellectual elites of Europe and America. Antievolutionism was pilloried in the American press as a rearguard reaction to scientific progress and it was associated with nativism and rural and uneducated religious fundamentalism. Although biblical catastrophism was revived in the 1920s, it enjoyed favor only in conservative and sectarian Protestant circles in North America before it exploded into the popular creationist movement (called ‘Flood Geology’) in the 1960s.

7. Twentieth-Century Physics And Cosmology

While religious issues had been increasingly excluded from scientific discourse in geology and natural history (the study of natural objects, plants, minerals, and animals) in the nineteenth century, they had continued to be discussed in natural philosophy. From about 1895 modern physics began to deal with issues that went beyond those of classical physics, including quantum mechanics and special and general relativity. Several discoveries in modern physics challenged the determinism of the Newtonian mechanical model. Foremost was the indeterminacy principle of Werner Heisenberg (1901–1976), which demonstrated that one could not know perfectly both the position and momentum of a subatomic particle. Heisenberg’s principle gave rise to much discussion of its wider philosophical and theological implications, particularly relating to the possibility of intelligent design of the universe and human free will.

In cosmology too discoveries were made that suggested theological implications. Edwin Hubble’s (1889–1953) discovery that the universe was expanding at a tremendous rate led the Roman Catholic priest and physicist Georges Lemaıtre (1894–1966) to propose what became known as the big-bang theory: that the universe had begun as a dense ‘primeval atom’ that had exploded in the distant past. It was challenged by the steady-state theory, developed in England and championed by Fred Hoyle (1915– ), according to which the universe is eternal but matter is being continuously created in space. In the late 1960s, several observational discoveries led to the almost universal acceptance of the big-bang theory, which was argued by some researchers, like Paul Davies (1946– ) and Frank Tipler (1947– ), to strengthen belief in God’s creation of the universe.

Another cosmological theory that engendered a good deal of theological discussion was the Anthropic Principle (a term coined by Brandon Carter in 1974), which asserts that the emergence of human beings in the universe is not an accidental by product of purposeless cosmic evolution. According to this theory the universe was so finely tuned during its earliest moments that if any one of several physical parameters had been slightly different it would have been unsuitable for the survival of human life. Hence, it is argued that the inhabited universe is anthropocentric, having been designed for human beings. The Anthropic Principle reintroduced into cosmological speculation at the end of the twentieth century issues that were long thought to have been laid to rest, including questions of teleology, intelligent design, and divine providence.

8. Conclusion

By the 1940s the modern synthesis of genetics and Darwin’s theory of natural selection, which accounted for random variation by genetic mutation, made it possible for biological scientists to adopt a fully naturalistic position in both method and metaphysics. Naturalistic evolution, it was alleged, was able offer a complete explanation for the origin and development of the universe and for all forms of life, as well as for social factors like language, religion, and ethics. While scientists like George Gaylord Simpson (1902–1984) considered materialism to be the only philosophical position capable of scientific support, it was far from universally accepted within the scientific community. The last decade of the twentieth century, moreover, saw a resurgence of interest in the relationship of religion to the sciences. This interest arose from a variety of sources both within and outside the scientific community, including those of Christian and other theistic traditions (who most often asserted the complementarity of religion and the sciences), process philosophers, proponents of New Age and pantheistic spiritualism, and non-Western religious traditions. Perhaps it reflected the widespread recognition that, for all its achievements, science does not provide ultimate answers to questions regarding the meaning of the universe or of life itself.

Bibliography:

  1. Barrow J D, Tipler F J 1986 The Anthropic Cosmological Principle. Oxford University Press, Oxford, UK
  2. Barbour I G 1997 Religion and Science: Historical and Con- temporary Issues. 1st rev. edn. Harper, San-Francisco
  3. Biagoli M 1993 Galileo Courtier: The Practice of Science in the Culture of Absolutism. University of Chicago Press, Chicago
  4. Brooke J H 1991 Science and Religion: Some Historical Perspectives. Cambridge University Press, Cambridge, UK
  5. Brooke J H, Cantor G 2000 Reconstructing Nature: The Engagement of Science and Religion. Oxford University Press, New York
  6. Burtt E A 1932 The Metaphysical Foundations of Modern Physical Science, rev. edn. Harcourt Brace, New York
  7. Durant J (ed.) 1985 Darwinism and Divinity. Blackwell, Oxford, UK
  8. Fantoli A 1996 Galileo: For Copernicanism and For the Church. Vatican Observatory Publications, Vatican City, Italy
  9. Ferngren G (ed.) 2000 The History of Science and Religion in the Western Tradition: An Encyclopedia. Garland, New York
  10. Harrison P 1998 The Bible, Protestantism, and the Rise of Natural Science. Cambridge University Press, Cambridge, UK
  11. Jaki S L 1974 Science and Creation: From Eternal Cycles to an Oscillating Universe. Science History Publications, New York
  12. Koyre A 1957 From the Closed World to the Infinite Universe. Johns Hopkins Press, Baltimore, MA
  13. Larson E J 1997 Summer for the Gods: The Scopes Trial and America’s Continuing Debate over Science and Religion. Basic Books, New York
  14. Lindberg D C, Numbers R L (eds.) 1986 God and Nature: Historical Essays on the Encounter Between Christianity and Science. University of California Press, Berkeley, CA
  15. Lindberg D C, Numbers R L 2001 Science and the Christian Tradition: Twelve Case Histories. University of Chicago Press, Chicago
  16. Moore J R 1979 The Post-Darwinian Controversies: A Study of the Protestant Struggle to Come to Terms with Darwin in Great Britain and America, 1870–1900. Cambridge University Press, Cambridge, UK
  17. Russell R J, Stoeger W R, Coyne G V (eds.) 1988 Physics, Philosophy, and Theology: A Common Quest for Understanding. Vatican Observatory Publications, Vatican City, Italy
  18. White A D 1896 A History of the Warfare of Science with Theology in Christendom. Appleton, New York

 

Religion And Youth Research Paper
Religion And Politics In The United States Research Paper

ORDER HIGH QUALITY CUSTOM PAPER


Always on-time

Plagiarism-Free

100% Confidentiality
Special offer! Get 10% off with the 24START discount code!