Search This Blog

Saturday, April 6, 2019

Aristotelianism

From Wikipedia, the free encyclopedia

Aristotle, by Francesco Hayez

Aristotelianism is a tradition of philosophy that takes its defining inspiration from the work of Aristotle. This school of thought, in the modern sense of philosophy, covers existence, ethics, mind and related subjects. In Aristotle's time, philosophy included natural philosophy, which preceded the advent of modern science during the Scientific Revolution. The works of Aristotle were initially defended by the members of the Peripatetic school and later on by the Neoplatonists, who produced many commentaries on Aristotle's writings. In the Islamic Golden Age, Avicenna and Averroes translated the works of Aristotle into Arabic and under them, along with philosophers such as Al-Kindi and Al-Farabi, Aristotelianism became a major part of early Islamic philosophy

Moses Maimonides adopted Aristotelianism from the Islamic scholars and based his famous Guide for the Perplexed on it and that became the basis of Jewish scholastic philosophy. Although some of Aristotle's logical works were known to western Europe, it was not until the Latin translations of the 12th century that the works of Aristotle and his Arabic commentators became widely available. Scholars such as Albertus Magnus and Thomas Aquinas interpreted and systematized Aristotle's works in accordance with Christian theology.

After retreating under criticism from modern natural philosophers, the distinctively Aristotelian idea of teleology was transmitted through Wolff and Kant to Hegel, who applied it to history as a totality. Although this project was criticized by Trendelenburg and Brentano as non-Aristotelian, Hegel's influence is now often said to be responsible for an important Aristotelian influence upon Marx. Postmodernists, in contrast, reject Aristotelianism's claim to reveal important theoretical truths. In this, they follow Heidegger's critique of Aristotle as the greatest source of the entire tradition of Western philosophy. 

Recent Aristotelian ethical and "practical" philosophy, such as that of Gadamer and McDowell, is often premissed upon a rejection of Aristotelianism's traditional metaphysical or theoretical philosophy. From this viewpoint, the early modern tradition of political republicanism, which views the res publica, public sphere or state as constituted by its citizens' virtuous activity, can appear thoroughly Aristotelian. 

The most famous contemporary Aristotelian philosopher is Alasdair MacIntyre. Especially famous for helping to revive virtue ethics in his book After Virtue, MacIntyre revises Aristotelianism with the argument that the highest temporal goods, which are internal to human beings, are actualized through participation in social practices. He juxtaposes Aristotelianism with the managerial institutions of capitalism and its state, and with rival traditions — including the philosophies of Hume and Nietzsche — that reject Aristotle's idea of essentially human goods and virtues and instead legitimate capitalism. Therefore, on MacIntyre's account, Aristotelianism is not identical with Western philosophy as a whole; rather, it is "the best theory so far, [including] the best theory so far about what makes a particular theory the best one." Politically and socially, it has been characterized as a newly "revolutionary Aristotelianism". This may be contrasted with the more conventional, apolitical and effectively conservative uses of Aristotle by, for example, Gadamer and McDowell. Other important contemporary Aristotelian theorists include Fred D. Miller, Jr. in politics and Rosalind Hursthouse in ethics.

History

Ancient Greek

The original followers of Aristotle were the members of the Peripatetic school. The most prominent members of the school after Aristotle were Theophrastus and Strato of Lampsacus, who both continued Aristotle's researches. During the Roman era the school concentrated on preserving and defending his work. The most important figure in this regard was Alexander of Aphrodisias who commentated on Aristotle's writings. With the rise of Neoplatonism in the 3rd century, Peripateticism as an independent philosophy came to an end, but the Neoplatonists sought to incorporate Aristotle's philosophy within their own system, and produced many commentaries on Aristotle.

Byzantine Empire

Byzantine Aristotelianism emerged in the Byzantine Empire in the form of Aristotelian paraphrase: adaptations in which Aristotle's text is rephrased, reorganized, and pruned, in order to make it more easily understood. This genre was allegedly invented by Themistius in the mid-4th century, revived by Michael Psellos in the mid-11th century, and further developed by Sophonias in the late 13th to early 14th centuries.

Leo the Mathematician was appointed to the chair of philosophy at the Magnaura School in the mid-9th century to teach Aristotelian logic. The 11th and 12th centuries saw the emergence of twelfth-century Byzantine Aristotelianism. Before the 12th century, the whole Byzantine output of Aristotelian commentaries was focused on logic. However, the range of subjects covered by the Aristotelian commentaries produced in the two decades after 1118 is much greater due to the initiative of the princess Anna Comnena who commissioned a number of scholars to write commentaries on previously neglected works of Aristotle.

Islamic world

A medieval Arabic representation of Aristotle teaching a student.
 
In the Abbasid Empire, many foreign works were translated into Arabic, large libraries were constructed, and scholars were welcomed. Under the caliphs Harun al-Rashid and his son Al-Ma'mun, the House of Wisdom in Baghdad flourished. Christian scholar Hunayn ibn Ishaq (809–873) was placed in charge of the translation work by the caliph. In his lifetime, Ishaq translated 116 writings, including works by Plato and Aristotle, into Syriac and Arabic.

With the founding of House of Wisdom, the entire corpus of Aristotelian works that had been preserved (excluding the Eudemian Ethics, Magna Moralia and Politics) became available, along with its Greek commentators; this corpus laid a uniform foundation for Islamic Aristotelianism.

Al-Kindi (801–873) was the first of the Muslim Peripatetic philosophers, and is known for his efforts to introduce Greek and Hellenistic philosophy to the Arab world. He incorporated Aristotelian and Neoplatonist thought into an Islamic philosophical framework. This was an important factor in the introduction and popularization of Greek philosophy in the Muslim intellectual world.

The philosopher Al-Farabi (872–950) had great influence on science and philosophy for several centuries, and in his time was widely thought second only to Aristotle in knowledge (alluded to by his title of "the Second Teacher"). His work, aimed at synthesis of philosophy and Sufism, paved the way for the work of Avicenna (980–1037). Avicenna was one of the main interpreters of Aristotle. The school of thought he founded became known as Avicennism, which was built on ingredients and conceptual building blocks that are largely Aristotelian and Neoplatonist.

At the western end of the Mediterranean Sea, during the reign of Al-Hakam II (961 to 976) in Córdoba, a massive translation effort was undertaken, and many books were translated into Arabic. Averroes (1126–1198), who spent much of his life in Cordoba and Seville, was especially distinguished as a commentator of Aristotle. He often wrote two or three different commentaries on the same work, and some 38 commentaries by Averroes on the works of Aristotle have been identified. Although his writings had only marginal impact in Islamic countries, his works would eventually have a huge impact in the Latin West, and would lead to the school of thought known as Averroism.

Western Europe

Although some knowledge of Aristotle seems to have lingered on in the ecclesiastical centres of western Europe after the fall of the Roman empire, by the ninth century nearly all that was known of Aristotle consisted of Boethius's commentaries on the Organon, and a few abridgments made by Latin authors of the declining empire, Isidore of Seville and Martianus Capella. From that time until the end of the eleventh century, little progress is apparent in Aristotelian knowledge.

The renaissance of the 12th century saw a major search by European scholars for new learning. James of Venice, who probably spent some years in Constantinople, translated Aristotle's Posterior Analytics from Greek into Latin in the mid-twelfth century, thus making the complete Aristotelian logical corpus, the Organon, available in Latin for the first time. Scholars travelled to areas of Europe that once had been under Muslim rule and still had substantial Arabic-speaking populations. From central Spain, which had come under Christian rule in the eleventh century, scholars produced many of the Latin translations of the 12th century. The most productive of these translators was Gerard of Cremona, (c. 1114–1187), who translated 87 books, which included many of the works of Aristotle such as his Posterior Analytics, Physics, On the Heavens, On Generation and Corruption, and Meteorology. Michael Scot (c. 1175–1232) translated Averroes' commentaries on the scientific works of Aristotle.

Aristotle's physical writings began to be discussed openly, and at a time when Aristotle's method was permeating all theology, these treatises were sufficient to cause his prohibition for heterodoxy in the Condemnations of 1210–1277. In the first of these, in Paris in 1210, it was stated that "neither the books of Aristotle on natural philosophy or their commentaries are to be read at Paris in public or secret, and this we forbid under penalty of excommunication." However, despite further attempts to restrict the teaching of Aristotle, by 1270 the ban on Aristotle's natural philosophy was ineffective.

William of Moerbeke (c. 1215–1286) undertook a complete translation of the works of Aristotle or, for some portions, a revision of existing translations. He was the first translator of the Politics (c. 1260) from Greek into Latin. Many copies of Aristotle in Latin then in circulation were assumed to have been influenced by Averroes, who was suspected of being a source of philosophical and theological errors found in the earlier translations of Aristotle. Such claims were without merit, however, as the Alexandrian Aristotelianism of Averroes followed "the strict study of the text of Aristotle, which was introduced by Avicenna, [because] a large amount of traditional Neoplatonism was incorporated with the body of traditional Aristotelianism".

Albertus Magnus (c. 1200–1280) was among the first medieval scholars to apply Aristotle's philosophy to Christian thought. He produced paraphrases of most of the works of Aristotle available to him. He digested, interpreted and systematized the whole of Aristotle's works, gleaned from the Latin translations and notes of the Arabian commentators, in accordance with Church doctrine. His efforts resulted in the formation of a Christian reception of Aristotle in the Western Europe. Magnus did not repudiate Plato. In that, he belonged to the dominant tradition of philosophy that preceded him, namely the "concordist tradition", which sought to harmonize Aristotle with Plato through interpretation (see for example Porphyry's On Plato and Aristotle Being Adherents of the Same School). Magnus famously wrote:
"Scias quod non perficitur homo in philosophia nisi ex scientia duarum philosophiarum: Aristotelis et Platonis." (Metaphysics, I, tr. 5, c. 5)
(Know that a man is not perfected in philosophy if it weren't for the knowledge of the two philosophers, Aristotle and Plato)
Thomas Aquinas (1225–1274), the pupil of Albertus Magnus, wrote a dozen commentaries on the works of Aristotle. Thomas was emphatically Aristotelian, he adopted Aristotle's analysis of physical objects, his view of place, time and motion, his proof of the prime mover, his cosmology, his account of sense perception and intellectual knowledge, and even parts of his moral philosophy. The philosophical school that arose as a legacy of the work of Thomas Aquinas was known as Thomism, and was especially influential among the Dominicans, and later, the Jesuits.

Using Albert's and Thomas's commentaries, as well as Marsilius of Padua's Defensor pacis, 14th-century scholar Nicole Oresme translated Aristotle's moral works into French and wrote extensive comments on them.

Modern era

After retreating under criticism from modern natural philosophers, the distinctively Aristotelian idea of teleology was transmitted through Wolff and Kant to Hegel, who applied it to history as a totality. Although this project was criticized by Trendelenburg and Brentano as un-Aristotelian, Hegel's influence is now often said to be responsible for an important Aristotelian influence upon Marx. Postmodernists, in contrast, reject Aristotelianism's claim to reveal important theoretical truths. In this, they follow Heidegger's critique of Aristotle as the greatest source of the entire tradition of Western philosophy.

Contemporary Aristotelianism

Aristotelianism is understood by its proponents as critically developing Plato's theories. Recent Aristotelian ethical and 'practical' philosophy, such as that of Gadamer and McDowell, is often premised upon a rejection of Aristotelianism's traditional metaphysical or theoretical philosophy. From this viewpoint, the early modern tradition of political republicanism, which views the res publica, public sphere or state as constituted by its citizens' virtuous activity, can appear thoroughly Aristotelian.

The contemporary Aristotelian philosopher Alasdair MacIntyre is specially famous for helping to revive virtue ethics in his book After Virtue. MacIntyre revises Aristotelianism with the argument that the highest temporal goods, which are internal to human beings, are actualized through participation in social practices. He opposes Aristotelianism to the managerial institutions of capitalism and its state, and to rival traditions—including the philosophies of Hume, Kant, Kierkegaard, and Nietzsche—that reject its idea of essentially human goods and virtues and instead legitimize capitalism. Therefore, on MacIntyre's account, Aristotelianism is not identical with Western philosophy as a whole; rather, it is "the best theory so far, [including] the best theory so far about what makes a particular theory the best one." Politically and socially, it has been characterized as a newly 'revolutionary Aristotelianism'. This may be contrasted with the more conventional, apolitical and effectively conservative uses of Aristotle by, for example, Gadamer and McDowell. Other important contemporary Aristotelian theorists include Fred D. Miller, Jr. in politics and Rosalind Hursthouse in ethics.

In metaphysics, an Aristotelian realism about universals is defended by such philosophers as David Malet Armstrong and Stephen Mumford, and is applied to the philosophy of mathematics by James Franklin.

Criticism

Bertrand Russell criticizes Aristotle's logic on the following points:
  1. The Aristotelian system allows formal defects leading to "bad metaphysics". For example, the following syllogism is permitted: "All golden mountains are mountains, all golden mountains are golden, therefore some mountains are golden", which insinuates the existence of at least one golden mountain. Furthermore, according to Russell, a predicate of a predicate can be a predicate of the original subject, which blurs the distinction between names and predicates with disastrous consequences; for example, a class with only one member is erroneously identified with that one member, making impossible to have a correct theory of the number one.
  2. The syllogism is overvalued in comparison to other forms of deduction. For example, syllogisms are not employed in mathematics since they are less convenient.
In addition, Russell ends his review of the Aristotelian logic with these words:
I conclude that the Aristotelian doctrines with which we have been concerned in this chapter are wholly false, with the exception of the formal theory of the syllogism, which is unimportant. Any person in the present day who wishes to learn logic will be wasting his time if he reads Aristotle or any of his disciples. Nonetheless, Aristotle's logical writings show great ability, and would have been useful to mankind if they had appeared at a time when intellectual originality was still active. Unfortunately, they appeared at the very end of the creative period of Greek thought, and therefore came to be accepted as authoritative. By the time that logical originality revived, a reign of two thousand years had made Aristotle very difficult to dethrone. Throughout modern times, practically every advance in science, in logic, or in philosophy has had to be made in the teeth of the opposition from Aristotle's disciples.

European science in the Middle Ages

From Wikipedia, the free encyclopedia
For most medieval scholars, who believed that God created the universe according to geometric and harmonic principles, science – particularly geometry and astronomy – was linked directly to the divine. To seek these principles, therefore, would be to seek God.
 
European science in the Middle Ages comprised the study of nature, mathematics and natural philosophy in medieval Europe. Following the fall of the Western Roman Empire and the decline in knowledge of Greek, Christian Western Europe was cut off from an important source of ancient learning. Although a range of Christian clerics and scholars from Isidore and Bede to Buridan and Oresme maintained the spirit of rational inquiry, Western Europe would see a period of scientific decline during the Early Middle Ages. However, by the time of the High Middle Ages, the region had rallied and was on its way to once more taking the lead in scientific discovery. Scholarship and scientific discoveries of the Late Middle Ages laid the groundwork for the Scientific Revolution of the Early Modern Period

According to Pierre Duhem, who founded the academic study of medieval science as a critique of the Enlightenment-positivist theory of a 17th-century anti-Aristotelian and anticlerical scientific revolution, the various conceptual origins of that alleged revolution lay in the 12th to 14th centuries, in the works of churchmen such as Aquinas and Buridan.

In the context of this article, "Western Europe" refers to the European cultures bound together by the Roman Catholic Church and the Latin language.

Western Europe

As Roman imperial authority effectively ended in the West during the 5th century, Western Europe entered the Middle Ages with great difficulties that affected the continent's intellectual production dramatically. Most classical scientific treatises of classical antiquity written in Greek were unavailable, leaving only simplified summaries and compilations. Nonetheless, Roman and early medieval scientific texts were read and studied, contributing to the understanding of nature as a coherent system functioning under divinely established laws that could be comprehended in the light of reason. This study continued through the Early Middle Ages, and with the Renaissance of the 12th century, interest in this study was revitalized through the translation of Greek and Arabic scientific texts. Scientific study further developed within the emerging medieval universities, where these texts were studied and elaborated, leading to new insights into the phenomena of the universe. These advances are virtually unknown to the lay public of today, partly because most theories advanced in medieval science are today obsolete, and partly because of the caricature of Middle Ages as a supposedly "Dark Age" which placed "the word of religious authorities over personal experience and rational activity."

Early Middle Ages (AD 476–1000)

In the ancient world, Greek had been the primary language of science. Even under the Roman Empire, Latin texts drew extensively on Greek work, some pre-Roman, some contemporary; while advanced scientific research and teaching continued to be carried on in the Hellenistic side of the empire, in Greek. Late Roman attempts to translate Greek writings into Latin had limited success.

As the knowledge of Greek declined during the transition to the Middle Ages, the Latin West found itself cut off from its Greek philosophical and scientific roots. Most scientific inquiry came to be based on information gleaned from sources which were often incomplete and posed serious problems of interpretation. Latin-speakers who wanted to learn about science only had access to books by such Roman writers as Calcidius, Macrobius, Martianus Capella, Boethius, Cassiodorus, and later Latin encyclopedists. Much had to be gleaned from non-scientific sources: Roman surveying manuals were read for what geometry was included.

Ninth century diagram of the observed and computed positions of the seven planets on 18 March 816.
 
De-urbanization reduced the scope of education and by the 6th century teaching and learning moved to monastic and cathedral schools, with the center of education being the study of the Bible. Education of the laity survived modestly in Italy, Spain, and the southern part of Gaul, where Roman influences were most long-lasting. In the 7th century, learning began to emerge in Ireland and the Celtic lands, where Latin was a foreign language and Latin texts were eagerly studied and taught.

The leading scholars of the early centuries were clergymen for whom the study of nature was but a small part of their interest. They lived in an atmosphere which provided little institutional support for the disinterested study of natural phenomena. The study of nature was pursued more for practical reasons than as an abstract inquiry: the need to care for the sick led to the study of medicine and of ancient texts on drugs, the need for monks to determine the proper time to pray led them to study the motion of the stars, the need to compute the date of Easter led them to study and teach rudimentary mathematics and the motions of the Sun and Moon. Modern readers may find it disconcerting that sometimes the same works discuss both the technical details of natural phenomena and their symbolic significance.

Around 800, Charles the Great, assisted by the English monk Alcuin of York, undertook what has become known as the Carolingian Renaissance, a program of cultural revitalization and educational reform. The chief scientific aspect of Charlemagne's educational reform concerned the study and teaching of astronomy, both as a practical art that clerics required to compute the date of Easter and as a theoretical discipline. From the year 787 on, decrees were issued recommending the restoration of old schools and the founding of new ones throughout the empire. Institutionally, these new schools were either under the responsibility of a monastery, a cathedral or a noble court

The scientific work of the period after Charlemagne was not so much concerned with original investigation as it was with the active study and investigation of ancient Roman scientific texts. This investigation paved the way for the later effort of Western scholars to recover and translate ancient Greek texts in philosophy and the sciences.

High Middle Ages (AD 1000–1300)

The translation of Greek and Arabic works allowed the full development of Christian philosophy and the method of scholasticism.
 
Beginning around the year 1050, European scholars built upon their existing knowledge by seeking out ancient learning in Greek and Arabic texts which they translated into Latin. They encountered a wide range of classical Greek texts, some of which had earlier been translated into Arabic, accompanied by commentaries and independent works by Islamic thinkers.

Gerard of Cremona is a good example: an Italian who traveled to Spain to copy a single text, he stayed on to translate some seventy works. His biography describes how he came to Toledo: "He was trained from childhood at centers of philosophical study and had come to a knowledge of all that was known to the Latins; but for love of the Almagest, which he could not find at all among the Latins, he went to Toledo; there, seeing the abundance of books in Arabic on every subject and regretting the poverty of the Latins in these things, he learned the Arabic language, in order to be able to translate." 

Map of medieval universities. They started a new infrastructure which was needed for scientific communities.
 
This period also saw the birth of medieval universities, which benefited materially from the translated texts and provided a new infrastructure for scientific communities. Some of these new universities were registered as an institution of international excellence by the Holy Roman Empire, receiving the title of Studium Generale. Most of the early Studia Generali were found in Italy, France, England, and Spain, and these were considered the most prestigious places of learning in Europe. This list quickly grew as new universities were founded throughout Europe. As early as the 13th century, scholars from a Studium Generale were encouraged to give lecture courses at other institutes across Europe and to share documents, and this led to the current academic culture seen in modern European universities. 

The rediscovery of the works of Aristotle allowed the full development of the new Christian philosophy and the method of scholasticism. By 1200 there were reasonably accurate Latin translations of the main works of Aristotle, Euclid, Ptolemy, Archimedes, and Galen—that is, of all the intellectually crucial ancient authors except Plato. Also, many of the medieval Arabic and Jewish key texts, such as the main works of Avicenna, Averroes and Maimonides now became available in Latin. During the 13th century, scholastics expanded the natural philosophy of these texts by commentaries (associated with teaching in the universities) and independent treatises. Notable among these were the works of Robert Grosseteste, Roger Bacon, John of Sacrobosco, Albertus Magnus, and Duns Scotus

Scholastics believed in empiricism and supporting Roman Catholic doctrines through secular study, reason, and logic. The most famous was Thomas Aquinas (later declared a "Doctor of the Church"), who led the move away from the Platonic and Augustinian and towards Aristotelianism (although natural philosophy was not his main concern). Meanwhile, precursors of the modern scientific method can be seen already in Grosseteste's emphasis on mathematics as a way to understand nature and in the empirical approach admired by Roger Bacon. 

Optical diagram showing light being refracted by a spherical glass container full of water (from Roger Bacon, De multiplicatione specierum).
 
Grosseteste was the founder of the famous Oxford Franciscan school. He built his work on Aristotle's vision of the dual path of scientific reasoning. Concluding from particular observations into a universal law, and then back again: from universal laws to prediction of particulars. Grosseteste called this "resolution and composition". Further, Grosseteste said that both paths should be verified through experimentation in order to verify the principals. These ideas established a tradition that carried forward to Padua and Galileo Galilei in the 17th century. 

Under the tuition of Grosseteste and inspired by the writings of Arab alchemists who had preserved and built upon Aristotle's portrait of induction, Bacon described a repeating cycle of observation, hypothesis, experimentation, and the need for independent verification. He recorded the manner in which he conducted his experiments in precise detail so that others could reproduce and independently test his results - a cornerstone of the scientific method, and a continuation of the work of researchers like Al Battani

Bacon and Grosseteste conducted investigations into optics, although much of it was similar to what was being done at the time by Arab scholars. Bacon did make a major contribution to the development of science in medieval Europe by writing to the Pope to encourage the study of natural science in university courses and compiling several volumes recording the state of scientific knowledge in many fields at the time. He described the possible construction of a telescope, but there is no strong evidence of his having made one.

Late Middle Ages (AD 1300–1500)

The first half of the 14th century saw the scientific work of great thinkers. The logic studies by William of Occam led him to postulate a specific formulation of the principle of parsimony, known today as Occam's razor. This principle is one of the main heuristics used by modern science to select between two or more underdetermined theories, though it is only fair to point out that this principle was employed explicitly by both Aquinas and Aristotle before him. 

As Western scholars became more aware (and more accepting) of controversial scientific treatises of the Byzantine and Islamic Empires these readings sparked new insights and speculation. The works of the early Byzantine scholar John Philoponus inspired Western scholars such as Jean Buridan to question the received wisdom of Aristotle's mechanics. Buridan developed the theory of impetus which was a step towards the modern concept of inertia. Buridan anticipated Isaac Newton when he wrote: 

Galileo's demonstration of the law of the space traversed in case of uniformly varied motion – as Oresme had demonstrated centuries earlier.
. . . after leaving the arm of the thrower, the projectile would be moved by an impetus given to it by the thrower and would continue to be moved as long as the impetus remained stronger than the resistance, and would be of infinite duration were it not diminished and corrupted by a contrary force resisting it or by something inclining it to a contrary motion.
Thomas Bradwardine and his partners, the Oxford Calculators of Merton College, Oxford, distinguished kinematics from dynamics, emphasizing kinematics, and investigating instantaneous velocity. They formulated the mean speed theorem: a body moving with constant velocity travels distance and time equal to an accelerated body whose velocity is half the final speed of the accelerated body. They also demonstrated this theorem—the essence of "The Law of Falling Bodies"—long before Galileo, who has gotten the credit for this.

In his turn, Nicole Oresme showed that the reasons proposed by the physics of Aristotle against the movement of the Earth were not valid and adduced the argument of simplicity for the theory that the Earth moves, and not the heavens. Despite this argument in favor of the Earth's motion, Oresme fell back on the commonly held opinion that "everyone maintains, and I think myself, that the heavens do move and not the earth."

The historian of science Ronald Numbers notes that the modern scientific assumption of methodological naturalism can be also traced back to the work of these medieval thinkers:
By the late Middle Ages the search for natural causes had come to typify the work of Christian natural philosophers. Although characteristically leaving the door open for the possibility of direct divine intervention, they frequently expressed contempt for soft-minded contemporaries who invoked miracles rather than searching for natural explanations. The University of Paris cleric Jean Buridan (a. 1295–ca. 1358), described as "perhaps the most brilliant arts master of the Middle Ages," contrasted the philosopher’s search for "appropriate natural causes" with the common folk’s erroneous habit of attributing unusual astronomical phenomena to the supernatural. In the fourteenth century the natural philosopher Nicole Oresme (ca. 1320–82), who went on to become a Roman Catholic bishop, admonished that, in discussing various marvels of nature, "there is no reason to take recourse to the heavens, the last refuge of the weak, or demons, or to our glorious God as if He would produce these effects directly, more so than those effects whose causes we believe are well known to us."
However, a series of events that would be known as the Crisis of the Late Middle Ages was under its way. When came the Black Death of 1348, it sealed a sudden end to the previous period of massive scientific change. The plague killed a third of the people in Europe, especially in the crowded conditions of the towns, where the heart of innovations lay. Recurrences of the plague and other disasters caused a continuing decline of population for a century.

Renaissance (15th century)

The 15th century saw the beginning of the cultural movement of the Renaissance. The rediscovery of Greek scientific texts, both ancient and medieval, was accelerated as the Byzantine Empire fell to the Ottoman Turks and many Byzantine scholars sought refuge in the West, particularly Italy

Also, the invention of printing was to have great effect on European society: the facilitated dissemination of the printed word democratized learning and allowed a faster propagation of new ideas. 

When the Renaissance moved to Northern Europe that science would be revived, by figures as Copernicus, Francis Bacon, and Descartes (though Descartes is often described as an early Enlightenment thinker, rather than a late Renaissance one).

Byzantine and Islamic influences

Byzantine interactions

Byzantine science played an important role in the transmission of classical knowledge to the Islamic world and to Renaissance Italy, and also in the transmission of medieval Arabic knowledge to Renaissance Italy. Its rich historiographical tradition preserved ancient knowledge upon which splendid art, architecture, literature and technological achievements were built. 

Byzantine scientists preserved and continued the legacy of the great Ancient Greek mathematicians and put mathematics in practice. In early Byzantium (5th to 7th century) the architects and mathematicians Isidore of Miletus and Anthemius of Tralles used complex mathematical formulas to construct the great “Hagia Sophia” temple, a magnificent technological breakthrough for its time and for centuries afterwards due to its striking geometry, bold design and height. In late Byzantium (9th to 12th century) mathematicians like Michael Psellos considered mathematics as a way to interpret the world. 

John Philoponus, a Byzantine scholar in the 500s, was the first person to systematically question Aristotle's teaching of physics. This served as an inspiration for Galileo Galilei ten centuries later as Galileo cited Philoponus substantially in his works when Galileo also argued why Aristotelian physics was flawed during the Scientific Revolution.

Islamic interactions

A Westerner and an Arab learning geometry in the 15th century.
 
The Byzantine Empire initially provided the medieval Islamic world with Ancient Greek texts on astronomy and mathematics for translation into Arabic. Later with the emerging of the Muslim world, Byzantine scientists such as Gregory Choniades translated Arabic texts on Islamic astronomy, mathematics and science into Medieval Greek, including the works of Ja'far ibn Muhammad Abu Ma'shar al-Balkhi, Ibn Yunus, al-Khazini, Muhammad ibn Mūsā al-Khwārizmī, and Nasīr al-Dīn al-Tūsī among others. There were also some Byzantine scientists who used Arabic transliterations to describe certain scientific concepts instead of the equivalent Ancient Greek terms (such as the use of the Arabic talei instead of the Ancient Greek horoscopus). Byzantine science thus played an important role in not only transmitting ancient Greek knowledge to Western Europe and the Islamic world, but in also transmitting Islamic knowledge to Western Europe. Byzantine scientists also became acquainted with Sassanid and Indian astronomy through citations in some Arabic works.

Conservation of energy

From Wikipedia, the free encyclopedia

In physics and chemistry, the law of conservation of energy states that the total energy of an isolated system remains constant; it is said to be conserved over time. This law means that energy can neither be created nor destroyed; rather, it can only be transformed or transferred from one form to another. For instance, chemical energy is converted to kinetic energy when a stick of dynamite explodes. If one adds up all the forms of energy that were released in the explosion, such as the kinetic energy and potential energy of the pieces, as well as heat and sound, one will get the exact decrease of chemical energy in the combustion of the dynamite. Classically, conservation of energy was distinct from conservation of mass; however, special relativity showed that mass is related to energy and vice versa by E = mc2, and science now takes the view that mass–energy is conserved.
 
Conservation of energy can be rigorously proven by Noether's theorem as a consequence of continuous time translation symmetry; that is, from the fact that the laws of physics do not change over time.

A consequence of the law of conservation of energy is that a perpetual motion machine of the first kind cannot exist, that is to say, no system without an external energy supply can deliver an unlimited amount of energy to its surroundings. For systems which do not have time translation symmetry, it may not be possible to define conservation of energy. Examples include curved spacetimes in general relativity or time crystals in condensed matter physics.

History

Ancient philosophers as far back as Thales of Miletus c. 550 BCE had inklings of the conservation of some underlying substance of which everything is made. However, there is no particular reason to identify this with what we know today as "mass-energy" (for example, Thales thought it was water). Empedocles (490–430 BCE) wrote that in this universal system, composed of four roots (earth, air, water, fire), "nothing comes to be or perishes"; instead, these elements suffer continual rearrangement.

In 1605, Simon Stevinus was able to solve a number of problems in statics based on the principle that perpetual motion was impossible.

In 1639, Galileo published his analysis of several situations—including the celebrated "interrupted pendulum"—which can be described (in modern language) as conservatively converting potential energy to kinetic energy and back again. Essentially, he pointed out that the height a moving body rises is equal to the height from which it falls, and used this observation to infer the idea of inertia. The remarkable aspect of this observation is that the height to which a moving body ascends on a frictionless surface does not depend on the shape of the surface. 

In 1669, Christiaan Huygens published his laws of collision. Among the quantities he listed as being invariant before and after the collision of bodies were both the sum of their linear momentums as well as the sum of their kinetic energies. However, the difference between elastic and inelastic collision was not understood at the time. This led to the dispute among later researchers as to which of these conserved quantities was the more fundamental. In his Horologium Oscillatorium, he gave a much clearer statement regarding the height of ascent of a moving body, and connected this idea with the impossibility of a perpetual motion. Huygens' study of the dynamics of pendulum motion was based on a single principle: that the center of gravity of a heavy object cannot lift itself.

The fact that kinetic energy is scalar, unlike linear momentum which is a vector, and hence easier to work with did not escape the attention of Gottfried Wilhelm Leibniz. It was Leibniz during 1676–1689 who first attempted a mathematical formulation of the kind of energy which is connected with motion (kinetic energy). Using Huygens' work on collision, Leibniz noticed that in many mechanical systems (of several masses, mi each with velocity vi),
was conserved so long as the masses did not interact. He called this quantity the vis viva or living force of the system. The principle represents an accurate statement of the approximate conservation of kinetic energy in situations where there is no friction. Many physicists at that time, such as Newton, held that the conservation of momentum, which holds even in systems with friction, as defined by the momentum:
was the conserved vis viva. It was later shown that both quantities are conserved simultaneously, given the proper conditions such as an elastic collision

In 1687, Isaac Newton published his Principia, which was organized around the concept of force and momentum. However, the researchers were quick to recognize that the principles set out in the book, while fine for point masses, were not sufficient to tackle the motions of rigid and fluid bodies. Some other principles were also required. 

The law of conservation of vis viva was championed by the father and son duo, Johann and Daniel Bernoulli. The former enunciated the principle of virtual work as used in statics in its full generality in 1715, while the latter based his Hydrodynamica, published in 1738, on this single conservation principle. Daniel's study of loss of vis viva of flowing water led him to formulate the Bernoulli's principle, which relates the loss to be proportional to the change in hydrodynamic pressure. Daniel also formulated the notion of work and efficiency for hydraulic machines; and he gave a kinetic theory of gases, and linked the kinetic energy of gas molecules with the temperature of the gas.

This focus on the vis viva by the continental physicists eventually led to the discovery of stationarity principles governing mechanics, such as the D'Alembert's principle, Lagrangian, and Hamiltonian formulations of mechanics. 

Émilie du Châtelet (1706 – 1749) proposed and tested the hypothesis of the conservation of total energy, as distinct from momentum. Inspired by the theories of Gottfried Leibniz, she repeated and publicized an experiment originally devised by Willem 's Gravesande in 1722 in which balls were dropped from different heights into a sheet of soft clay. Each ball's kinetic energy - as indicated by the quantity of material displaced - was shown to be proportional to the square of the velocity. The deformation of the clay was found to be directly proportional to the height the balls were dropped from, equal to the initial potential energy. Earlier workers, including Newton and Voltaire, had all believed that "energy" (so far as they understood the concept at all) was not distinct from momentum and therefore proportional to velocity. According to this understanding, the deformation of the clay should have been proportional to the square root of the height from which the balls were dropped from. In classical physics the correct formula is , where is the kinetic energy of an object, its mass and its speed. On this basis, Châtelet proposed that energy must always have the same dimensions in any form, which is necessary to be able to relate it in different forms (kinetic, potential, heat…).

Engineers such as John Smeaton, Peter Ewart, Carl Holtzmann, Gustave-Adolphe Hirn and Marc Seguin recognized that conservation of momentum alone was not adequate for practical calculation and made use of Leibniz's principle. The principle was also championed by some chemists such as William Hyde Wollaston. Academics such as John Playfair were quick to point out that kinetic energy is clearly not conserved. This is obvious to a modern analysis based on the second law of thermodynamics, but in the 18th and 19th centuries the fate of the lost energy was still unknown. 

Gradually it came to be suspected that the heat inevitably generated by motion under friction was another form of vis viva. In 1783, Antoine Lavoisier and Pierre-Simon Laplace reviewed the two competing theories of vis viva and caloric theory. Count Rumford's 1798 observations of heat generation during the boring of cannons added more weight to the view that mechanical motion could be converted into heat, and (as importantly) that the conversion was quantitative and could be predicted (allowing for a universal conversion constant between kinetic energy and heat). Vis viva then started to be known as energy, after the term was first used in that sense by Thomas Young in 1807. 

The recalibration of vis viva to
which can be understood as converting kinetic energy to work, was largely the result of Gaspard-Gustave Coriolis and Jean-Victor Poncelet over the period 1819–1839. The former called the quantity quantité de travail (quantity of work) and the latter, travail mécanique (mechanical work), and both championed its use in engineering calculation. 

In a paper Über die Natur der Wärme(German "On the Nature of Heat/Warmth"), published in the Zeitschrift für Physik in 1837, Karl Friedrich Mohr gave one of the earliest general statements of the doctrine of the conservation of energy in the words: "besides the 54 known chemical elements there is in the physical world one agent only, and this is called Kraft [energy or work]. It may appear, according to circumstances, as motion, chemical affinity, cohesion, electricity, light and magnetism; and from any one of these forms it can be transformed into any of the others."

Mechanical equivalent of heat

A key stage in the development of the modern conservation principle was the demonstration of the mechanical equivalent of heat. The caloric theory maintained that heat could neither be created nor destroyed, whereas conservation of energy entails the contrary principle that heat and mechanical work are interchangeable. 

In the middle of the eighteenth century, Mikhail Lomonosov, a Russian scientist, postulated his corpusculo-kinetic theory of heat, which rejected the idea of a caloric. Through the results of empirical studies, Lomonosov came to the conclusion that heat was not transferred through the particles of the caloric fluid. 

In 1798, Count Rumford (Benjamin Thompson) performed measurements of the frictional heat generated in boring cannons, and developed the idea that heat is a form of kinetic energy; his measurements refuted caloric theory, but were imprecise enough to leave room for doubt. 

The mechanical equivalence principle was first stated in its modern form by the German surgeon Julius Robert von Mayer in 1842. Mayer reached his conclusion on a voyage to the Dutch East Indies, where he found that his patients' blood was a deeper red because they were consuming less oxygen, and therefore less energy, to maintain their body temperature in the hotter climate. He discovered that heat and mechanical work were both forms of energy and in 1845, after improving his knowledge of physics, he published a monograph that stated a quantitative relationship between them.

Joule's apparatus for measuring the mechanical equivalent of heat. A descending weight attached to a string causes a paddle immersed in water to rotate.
 
Meanwhile, in 1843, James Prescott Joule independently discovered the mechanical equivalent in a series of experiments. In the most famous, now called the "Joule apparatus", a descending weight attached to a string caused a paddle immersed in water to rotate. He showed that the gravitational potential energy lost by the weight in descending was equal to the internal energy gained by the water through friction with the paddle. 

Over the period 1840–1843, similar work was carried out by engineer Ludwig A. Colding, although it was little known outside his native Denmark. 

Both Joule's and Mayer's work suffered from resistance and neglect but it was Joule's that eventually drew the wider recognition. 


In 1844, William Robert Grove postulated a relationship between mechanics, heat, light, electricity and magnetism by treating them all as manifestations of a single "force" (energy in modern terms). In 1846, Grove published his theories in his book The Correlation of Physical Forces. In 1847, drawing on the earlier work of Joule, Sadi Carnot and Émile Clapeyron, Hermann von Helmholtz arrived at conclusions similar to Grove's and published his theories in his book Über die Erhaltung der Kraft (On the Conservation of Force, 1847). The general modern acceptance of the principle stems from this publication.

In 1850, William Rankine first used the phrase the law of the conservation of energy for the principle.

In 1877, Peter Guthrie Tait claimed that the principle originated with Sir Isaac Newton, based on a creative reading of propositions 40 and 41 of the Philosophiae Naturalis Principia Mathematica. This is now regarded as an example of Whig history.

Mass–energy equivalence

Matter is composed of such things as atoms, electrons, neutrons, and protons. It has intrinsic or rest mass. In the limited range of recognized experience of the nineteenth century it was found that such rest mass is conserved. Einstein's 1905 theory of special relativity showed that it corresponds to an equivalent amount of rest energy. This means that it can be converted to or from equivalent amounts of other (non-material) forms of energy, for example kinetic energy, potential energy, and electromagnetic radiant energy. When this happens, as recognized in twentieth century experience, rest mass is not conserved, unlike the total mass or total energy. All forms of energy contribute to the total mass and total energy. 

For example, an electron and a positron each have rest mass. They can perish together, converting their combined rest energy into photons having electromagnetic radiant energy, but no rest mass. If this occurs within an isolated system that does not release the photons or their energy into the external surroundings, then neither the total mass nor the total energy of the system will change. The produced electromagnetic radiant energy contributes just as much to the inertia (and to any weight) of the system as did the rest mass of the electron and positron before their demise. Likewise, non-material forms of energy can perish into matter, which has rest mass. 

Thus, conservation of energy (total, including material or rest energy), and conservation of mass (total, not just rest), each still holds as an (equivalent) law. In the 18th century these had appeared as two seemingly-distinct laws.

Conservation of energy in beta decay

The discovery in 1911 that electrons emitted in beta decay have a continuous rather than a discrete spectrum appeared to contradict conservation of energy, under the then-current assumption that beta decay is the simple emission of an electron from a nucleus. This problem was eventually resolved in 1933 by Enrico Fermi who proposed the correct description of beta-decay as the emission of both an electron and an antineutrino, which carries away the apparently missing energy.

First law of thermodynamics

For a closed thermodynamic system, the first law of thermodynamics may be stated as:
, or equivalently,
where is the quantity of energy added to the system by a heating process, is the quantity of energy lost by the system due to work done by the system on its surroundings and is the change in the internal energy of the system.

The δ's before the heat and work terms are used to indicate that they describe an increment of energy which is to be interpreted somewhat differently than the increment of internal energy (see Inexact differential). Work and heat refer to kinds of process which add or subtract energy to or from a system, while the internal energy is a property of a particular state of the system when it is in unchanging thermodynamic equilibrium. Thus the term "heat energy" for means "that amount of energy added as the result of heating" rather than referring to a particular form of energy. Likewise, the term "work energy" for means "that amount of energy lost as the result of work". Thus one can state the amount of internal energy possessed by a thermodynamic system that one knows is presently in a given state, but one cannot tell, just from knowledge of the given present state, how much energy has in the past flowed into or out of the system as a result of its being heated or cooled, nor as the result of work being performed on or by the system. 

Entropy is a function of the state of a system which tells of limitations of the possibility of conversion of heat into work. 

For a simple compressible system, the work performed by the system may be written:
where is the pressure and is a small change in the volume of the system, each of which are system variables. In the fictive case in which the process is idealized and infinitely slow, so as to be called quasi-static, and regarded as reversible, the heat being transferred from a source with temperature infinitesimally above the system temperature, then the heat energy may be written
where is the temperature and is a small change in the entropy of the system. Temperature and entropy are variables of state of a system. 

If an open system (in which mass may be exchanged with the environment) has several walls such that the mass transfer is through rigid walls separate from the heat and work transfers, then the first law may be written:
where is the added mass and is the internal energy per unit mass of the added mass, measured in the surroundings before the process.

Noether's theorem

Emmy Noether (1882-1935) was an influential mathematician known for her groundbreaking contributions to abstract algebra and theoretical physics.
 
The conservation of energy is a common feature in many physical theories. From a mathematical point of view it is understood as a consequence of Noether's theorem, developed by Emmy Noether in 1915 and first published in 1918. The theorem states every continuous symmetry of a physical theory has an associated conserved quantity; if the theory's symmetry is time invariance then the conserved quantity is called "energy". The energy conservation law is a consequence of the shift symmetry of time; energy conservation is implied by the empirical fact that the laws of physics do not change with time itself. Philosophically this can be stated as "nothing depends on time per se". In other words, if the physical system is invariant under the continuous symmetry of time translation then its energy (which is canonical conjugate quantity to time) is conserved. Conversely, systems which are not invariant under shifts in time (an example, systems with time dependent potential energy) do not exhibit conservation of energy – unless we consider them to exchange energy with another, external system so that the theory of the enlarged system becomes time invariant again. Conservation of energy for finite systems is valid in such physical theories as special relativity and quantum theory (including QED) in the flat space-time.

Relativity

With the discovery of special relativity by Henri Poincaré and Albert Einstein, energy was proposed to be one component of an energy-momentum 4-vector. Each of the four components (one of energy and three of momentum) of this vector is separately conserved across time, in any closed system, as seen from any given inertial reference frame. Also conserved is the vector length (Minkowski norm), which is the rest mass for single particles, and the invariant mass for systems of particles (where momenta and energy are separately summed before the length is calculated—see the article on invariant mass). 

The relativistic energy of a single massive particle contains a term related to its rest mass in addition to its kinetic energy of motion. In the limit of zero kinetic energy (or equivalently in the rest frame) of a massive particle, or else in the center of momentum frame for objects or systems which retain kinetic energy, the total energy of particle or object (including internal kinetic energy in systems) is related to its rest mass or its invariant mass via the famous equation

Thus, the rule of conservation of energy over time in special relativity continues to hold, so long as the reference frame of the observer is unchanged. This applies to the total energy of systems, although different observers disagree as to the energy value. Also conserved, and invariant to all observers, is the invariant mass, which is the minimal system mass and energy that can be seen by any observer, and which is defined by the energy–momentum relation

In general relativity, energy–momentum conservation is not well-defined except in certain special cases. Energy-momentum is typically expressed with the aid of a stress–energy–momentum pseudotensor. However, since pseudotensors are not tensors, they do not transform cleanly between reference frames. If the metric under consideration is static (that is, does not change with time) or asymptotically flat (that is, at an infinite distance away spacetime looks empty), then energy conservation holds without major pitfalls. In practice, some metrics such as the Friedmann–Lemaître–Robertson–Walker metric do not satisfy these constraints and energy conservation is not well defined. The theory of general relativity leaves open the question of whether there is a conservation of energy for the entire universe.

Quantum theory

In quantum mechanics, energy of a quantum system is described by a self-adjoint (or Hermitian) operator called the Hamiltonian, which acts on the Hilbert space (or a space of wave functions) of the system. If the Hamiltonian is a time-independent operator, emergence probability of the measurement result does not change in time over the evolution of the system. Thus the expectation value of energy is also time independent. The local energy conservation in quantum field theory is ensured by the quantum Noether's theorem for energy-momentum tensor operator. Note that due to the lack of the (universal) time operator in quantum theory, the uncertainty relations for time and energy are not fundamental in contrast to the position-momentum uncertainty principle, and merely holds in specific cases. Energy at each fixed time can in principle be exactly measured without any trade-off in precision forced by the time-energy uncertainty relations. Thus the conservation of energy in time is a well defined concept even in quantum mechanics.

Agricultural education

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Agr...