Search This Blog

Sunday, July 3, 2022

Mathematical physics

From Wikipedia, the free encyclopedia
 
An example of mathematical physics: solutions of Schrödinger's equation for quantum harmonic oscillators (left) with their amplitudes (right).

Mathematical physics refers to the development of mathematical methods for application to problems in physics. The Journal of Mathematical Physics defines the field as "the application of mathematics to problems in physics and the development of mathematical methods suitable for such applications and for the formulation of physical theories". An alternative definition would also include those mathematics that are inspired by physics (also known as physical mathematics).

Scope

There are several distinct branches of mathematical physics, and these roughly correspond to particular historical periods.

Classical mechanics

The rigorous, abstract and advanced reformulation of Newtonian mechanics adopting the Lagrangian mechanics and the Hamiltonian mechanics even in the presence of constraints. Both formulations are embodied in analytical mechanics and lead to understanding the deep interplay of the notions of symmetry and conserved quantities during the dynamical evolution, as embodied within the most elementary formulation of Noether's theorem. These approaches and ideas have been extended to other areas of physics as statistical mechanics, continuum mechanics, classical field theory and quantum field theory. Moreover, they have provided several examples and ideas in differential geometry (e.g. several notions in symplectic geometry and vector bundle).

Partial differential equations

Following mathematics: the theory of partial differential equation, variational calculus, Fourier analysis, potential theory, and vector analysis are perhaps most closely associated with mathematical physics. These were developed intensively from the second half of the 18th century (by, for example, D'Alembert, Euler, and Lagrange) until the 1930s. Physical applications of these developments include hydrodynamics, celestial mechanics, continuum mechanics, elasticity theory, acoustics, thermodynamics, electricity, magnetism, and aerodynamics.

Quantum theory

The theory of atomic spectra (and, later, quantum mechanics) developed almost concurrently with some parts of the mathematical fields of linear algebra, the spectral theory of operators, operator algebras and more broadly, functional analysis. Nonrelativistic quantum mechanics includes Schrödinger operators, and it has connections to atomic and molecular physics. Quantum information theory is another subspecialty.

Relativity and quantum relativistic theories

The special and general theories of relativity require a rather different type of mathematics. This was group theory, which played an important role in both quantum field theory and differential geometry. This was, however, gradually supplemented by topology and functional analysis in the mathematical description of cosmological as well as quantum field theory phenomena. In the mathematical description of these physical areas, some concepts in homological algebra and category theory are also important.

Statistical mechanics

Statistical mechanics forms a separate field, which includes the theory of phase transitions. It relies upon the Hamiltonian mechanics (or its quantum version) and it is closely related with the more mathematical ergodic theory and some parts of probability theory. There are increasing interactions between combinatorics and physics, in particular statistical physics.

Usage

Relationship between mathematics and physics

The usage of the term "mathematical physics" is sometimes idiosyncratic. Certain parts of mathematics that initially arose from the development of physics are not, in fact, considered parts of mathematical physics, while other closely related fields are. For example, ordinary differential equations and symplectic geometry are generally viewed as purely mathematical disciplines, whereas dynamical systems and Hamiltonian mechanics belong to mathematical physics. John Herapath used the term for the title of his 1847 text on "mathematical principles of natural philosophy"; the scope at that time being

"the causes of heat, gaseous elasticity, gravitation, and other great phenomena of nature".

Mathematical vs. theoretical physics

The term "mathematical physics" is sometimes used to denote research aimed at studying and solving problems in physics or thought experiments within a mathematically rigorous framework. In this sense, mathematical physics covers a very broad academic realm distinguished only by the blending of some mathematical aspect and physics theoretical aspect. Although related to theoretical physics, mathematical physics in this sense emphasizes the mathematical rigour of the similar type as found in mathematics.

On the other hand, theoretical physics emphasizes the links to observations and experimental physics, which often requires theoretical physicists (and mathematical physicists in the more general sense) to use heuristic, intuitive, and approximate arguments. Such arguments are not considered rigorous by mathematicians.

Such mathematical physicists primarily expand and elucidate physical theories. Because of the required level of mathematical rigour, these researchers often deal with questions that theoretical physicists have considered to be already solved. However, they can sometimes show that the previous solution was incomplete, incorrect, or simply too naïve. Issues about attempts to infer the second law of thermodynamics from statistical mechanics are examples. Other examples concern the subtleties involved with synchronisation procedures in special and general relativity (Sagnac effect and Einstein synchronisation).

The effort to put physical theories on a mathematically rigorous footing not only developed physics but also has influenced developments of some mathematical areas. For example, the development of quantum mechanics and some aspects of functional analysis parallel each other in many ways. The mathematical study of quantum mechanics, quantum field theory, and quantum statistical mechanics has motivated results in operator algebras. The attempt to construct a rigorous mathematical formulation of quantum field theory has also brought about some progress in fields such as representation theory.

Prominent mathematical physicists

Before Newton

There is a tradition of mathematical analysis of nature that goes back to the ancient Greeks; examples include Euclid (Optics), Archimedes (On the Equilibrium of Planes, On Floating Bodies), and Ptolemy (Optics, Harmonics). Later, Islamic and Byzantine scholars built on these works, and these ultimately were reintroduced or became available to the West in the 12th century and during the Renaissance.

In the first decade of the 16th century, amateur astronomer Nicolaus Copernicus proposed heliocentrism, and published a treatise on it in 1543. He retained the Ptolemaic idea of epicycles, and merely sought to simplify astronomy by constructing simpler sets of epicyclic orbits. Epicycles consist of circles upon circles. According to Aristotelian physics, the circle was the perfect form of motion, and was the intrinsic motion of Aristotle's fifth element—the quintessence or universal essence known in Greek as aether for the English pure air—that was the pure substance beyond the sublunary sphere, and thus was celestial entities' pure composition. The German Johannes Kepler [1571–1630], Tycho Brahe's assistant, modified Copernican orbits to ellipses, formalized in the equations of Kepler's laws of planetary motion.

An enthusiastic atomist, Galileo Galilei in his 1623 book The Assayer asserted that the "book of nature is written in mathematics". His 1632 book, about his telescopic observations, supported heliocentrism. Having introduced experimentation, Galileo then refuted geocentric cosmology by refuting Aristotelian physics itself. Galileo's 1638 book Discourse on Two New Sciences established the law of equal free fall as well as the principles of inertial motion, founding the central concepts of what would become today's classical mechanics. By the Galilean law of inertia as well as the principle of Galilean invariance, also called Galilean relativity, for any object experiencing inertia, there is empirical justification for knowing only that it is at relative rest or relative motion—rest or motion with respect to another object.

René Descartes famously developed a complete system of heliocentric cosmology anchored on the principle of vortex motion, Cartesian physics, whose widespread acceptance brought the demise of Aristotelian physics. Descartes sought to formalize mathematical reasoning in science, and developed Cartesian coordinates for geometrically plotting locations in 3D space and marking their progressions along the flow of time.

An older contemporary of Newton, Christiaan Huygens, was the first to idealize a physical problem by a set of parameters and the first to fully mathematize a mechanistic explanation of unobservable physical phenomena, and for these reasons Huygens is considered the first theoretical physicist and one of the founders of modern mathematical physics.

Newtonian and post Newtonian

In this era, important concepts in calculus such as the fundamental theorem of calculus (proved in 1668 by Scottish mathematician James Gregory) and finding extrema and minima of functions via differentiation using Fermat's theorem (by French mathematician Pierre de Fermat) were already known before Leibniz and Newton. Isaac Newton (1642–1727) developed some concepts in calculus (although Gottfried Wilhelm Leibniz developed similar concepts outside the context of physics) and Newton's method to solve problems in physics. He was extremely successful in his application of calculus to the theory of motion. Newton's theory of motion, shown in his Mathematical Principles of Natural Philosophy, published in 1687, modeled three Galilean laws of motion along with Newton's law of universal gravitation on a framework of absolute space—hypothesized by Newton as a physically real entity of Euclidean geometric structure extending infinitely in all directions—while presuming absolute time, supposedly justifying knowledge of absolute motion, the object's motion with respect to absolute space. The principle of Galilean invariance/relativity was merely implicit in Newton's theory of motion. Having ostensibly reduced the Keplerian celestial laws of motion as well as Galilean terrestrial laws of motion to a unifying force, Newton achieved great mathematical rigor, but with theoretical laxity.

In the 18th century, the Swiss Daniel Bernoulli (1700–1782) made contributions to fluid dynamics, and vibrating strings. The Swiss Leonhard Euler (1707–1783) did special work in variational calculus, dynamics, fluid dynamics, and other areas. Also notable was the Italian-born Frenchman, Joseph-Louis Lagrange (1736–1813) for work in analytical mechanics: he formulated Lagrangian mechanics) and variational methods. A major contribution to the formulation of Analytical Dynamics called Hamiltonian dynamics was also made by the Irish physicist, astronomer and mathematician, William Rowan Hamilton (1805-1865). Hamiltonian dynamics had played an important role in the formulation of modern theories in physics, including field theory and quantum mechanics. The French mathematical physicist Joseph Fourier (1768 – 1830) introduced the notion of Fourier series to solve the heat equation, giving rise to a new approach to solving partial differential equations by means of integral transforms.

Into the early 19th century, following mathematicians in France, Germany and England had contributed to mathematical physics. The French Pierre-Simon Laplace (1749–1827) made paramount contributions to mathematical astronomy, potential theory. Siméon Denis Poisson (1781–1840) worked in analytical mechanics and potential theory. In Germany, Carl Friedrich Gauss (1777–1855) made key contributions to the theoretical foundations of electricity, magnetism, mechanics, and fluid dynamics. In England, George Green (1793-1841) published An Essay on the Application of Mathematical Analysis to the Theories of Electricity and Magnetism in 1828, which in addition to its significant contributions to mathematics made early progress towards laying down the mathematical foundations of electricity and magnetism.

A couple of decades ahead of Newton's publication of a particle theory of light, the Dutch Christiaan Huygens (1629–1695) developed the wave theory of light, published in 1690. By 1804, Thomas Young's double-slit experiment revealed an interference pattern, as though light were a wave, and thus Huygens's wave theory of light, as well as Huygens's inference that light waves were vibrations of the luminiferous aether, was accepted. Jean-Augustin Fresnel modeled hypothetical behavior of the aether. The English physicist Michael Faraday introduced the theoretical concept of a field—not action at a distance. Mid-19th century, the Scottish James Clerk Maxwell (1831–1879) reduced electricity and magnetism to Maxwell's electromagnetic field theory, whittled down by others to the four Maxwell's equations. Initially, optics was found consequent of[clarification needed] Maxwell's field. Later, radiation and then today's known electromagnetic spectrum were found also consequent of this electromagnetic field.

The English physicist Lord Rayleigh [1842–1919] worked on sound. The Irishmen William Rowan Hamilton (1805–1865), George Gabriel Stokes (1819–1903) and Lord Kelvin (1824–1907) produced several major works: Stokes was a leader in optics and fluid dynamics; Kelvin made substantial discoveries in thermodynamics; Hamilton did notable work on analytical mechanics, discovering a new and powerful approach nowadays known as Hamiltonian mechanics. Very relevant contributions to this approach are due to his German colleague mathematician Carl Gustav Jacobi (1804–1851) in particular referring to canonical transformations. The German Hermann von Helmholtz (1821–1894) made substantial contributions in the fields of electromagnetism, waves, fluids, and sound. In the United States, the pioneering work of Josiah Willard Gibbs (1839–1903) became the basis for statistical mechanics. Fundamental theoretical results in this area were achieved by the German Ludwig Boltzmann (1844-1906). Together, these individuals laid the foundations of electromagnetic theory, fluid dynamics, and statistical mechanics.

Relativistic

By the 1880s, there was a prominent paradox that an observer within Maxwell's electromagnetic field measured it at approximately constant speed, regardless of the observer's speed relative to other objects within the electromagnetic field. Thus, although the observer's speed was continually lost relative to the electromagnetic field, it was preserved relative to other objects in the electromagnetic field. And yet no violation of Galilean invariance within physical interactions among objects was detected. As Maxwell's electromagnetic field was modeled as oscillations of the aether, physicists inferred that motion within the aether resulted in aether drift, shifting the electromagnetic field, explaining the observer's missing speed relative to it. The Galilean transformation had been the mathematical process used to translate the positions in one reference frame to predictions of positions in another reference frame, all plotted on Cartesian coordinates, but this process was replaced by Lorentz transformation, modeled by the Dutch Hendrik Lorentz [1853–1928].

In 1887, experimentalists Michelson and Morley failed to detect aether drift, however. It was hypothesized that motion into the aether prompted aether's shortening, too, as modeled in the Lorentz contraction. It was hypothesized that the aether thus kept Maxwell's electromagnetic field aligned with the principle of Galilean invariance across all inertial frames of reference, while Newton's theory of motion was spared.

Austrian theoretical physicist and philosopher Ernst Mach criticized Newton's postulated absolute space. Mathematician Jules-Henri Poincaré (1854–1912) questioned even absolute time. In 1905, Pierre Duhem published a devastating criticism of the foundation of Newton's theory of motion. Also in 1905, Albert Einstein (1879–1955) published his special theory of relativity, newly explaining both the electromagnetic field's invariance and Galilean invariance by discarding all hypotheses concerning aether, including the existence of aether itself. Refuting the framework of Newton's theory—absolute space and absolute time—special relativity refers to relative space and relative time, whereby length contracts and time dilates along the travel pathway of an object.

In 1908, Einstein's former mathematics professor Hermann Minkowski modeled 3D space together with the 1D axis of time by treating the temporal axis like a fourth spatial dimension—altogether 4D spacetime—and declared the imminent demise of the separation of space and time. Einstein initially called this "superfluous learnedness", but later used Minkowski spacetime with great elegance in his general theory of relativity, extending invariance to all reference frames—whether perceived as inertial or as accelerated—and credited this to Minkowski, by then deceased. General relativity replaces Cartesian coordinates with Gaussian coordinates, and replaces Newton's claimed empty yet Euclidean space traversed instantly by Newton's vector of hypothetical gravitational force—an instant action at a distance—with a gravitational field. The gravitational field is Minkowski spacetime itself, the 4D topology of Einstein aether modeled on a Lorentzian manifold that "curves" geometrically, according to the Riemann curvature tensor. The concept of Newton's gravity: "two masses attract each other" replaced by the geometrical argument: "mass transform curvatures of spacetime and free falling particles with mass move along a geodesic curve in the spacetime" (Riemannian geometry already existed before the 1850s, by mathematicians Carl Friedrich Gauss and Bernhard Riemann in search for intrinsic geometry and non-Euclidean geometry.), in the vicinity of either mass or energy. (Under special relativity—a special case of general relativity—even massless energy exerts gravitational effect by its mass equivalence locally "curving" the geometry of the four, unified dimensions of space and time.)

Quantum

Another revolutionary development of the 20th century was quantum theory, which emerged from the seminal contributions of Max Planck (1856–1947) (on black-body radiation) and Einstein's work on the photoelectric effect. In 1912, a mathematician Henri Poincare published Sur la théorie des quanta. He introduced the first non-naïve definition of quantization in this paper. The development of early quantum physics followed by a heuristic framework devised by Arnold Sommerfeld (1868–1951) and Niels Bohr (1885–1962), but this was soon replaced by the quantum mechanics developed by Max Born (1882–1970), Werner Heisenberg (1901–1976), Paul Dirac (1902–1984), Erwin Schrödinger (1887–1961), Satyendra Nath Bose (1894–1974), and Wolfgang Pauli (1900–1958). This revolutionary theoretical framework is based on a probabilistic interpretation of states, and evolution and measurements in terms of self-adjoint operators on an infinite-dimensional vector space. That is called Hilbert space (introduced by mathematicians David Hilbert (1862–1943), Erhard Schmidt(1876-1959) and Frigyes Riesz (1880-1956) in search of generalization of Euclidean space and study of integral equations), and rigorously defined within the axiomatic modern version by John von Neumann in his celebrated book Mathematical Foundations of Quantum Mechanics, where he built up a relevant part of modern functional analysis on Hilbert spaces, the spectral theory (introduced by David Hilbert who investigated quadratic forms with infinitely many variables. Many years later, it had been revealed that his spectral theory is associated with the spectrum of the hydrogen atom. He was surprised by this application.) in particular. Paul Dirac used algebraic constructions to produce a relativistic model for the electron, predicting its magnetic moment and the existence of its antiparticle, the positron.

List of prominent contributors to mathematical physics in the 20th century

Prominent contributors to the 20th century's mathematical physics include, (ordered by birth date) William Thomson (Lord Kelvin) [1824–1907], Oliver Heaviside [1850–1925], Jules Henri Poincaré [1854–1912] , David Hilbert [1862–1943], Arnold Sommerfeld [1868–1951], Constantin Carathéodory [1873–1950], Albert Einstein [1879–1955], Max Born [1882–1970], George David Birkhoff [1884-1944], Hermann Weyl [1885–1955], Satyendra Nath Bose [1894-1974], Norbert Wiener [1894–1964], John Lighton Synge [1897–1995], Wolfgang Pauli [1900–1958], Paul Dirac [1902–1984], Eugene Wigner [1902–1995], Andrey Kolmogorov [1903-1987], Lars Onsager [1903-1976], John von Neumann [1903–1957], Sin-Itiro Tomonaga [1906–1979], Hideki Yukawa [1907–1981], Nikolay Nikolayevich Bogolyubov [1909–1992], Subrahmanyan Chandrasekhar [1910-1995], Mark Kac [1914–1984], Julian Schwinger [1918–1994], Richard Phillips Feynman [1918–1988], Irving Ezra Segal [1918–1998], Ryogo Kubo [1920–1995], Arthur Strong Wightman [1922–2013], Chen-Ning Yang [1922– ], Rudolf Haag [1922–2016], Freeman John Dyson [1923–2020], Martin Gutzwiller [1925–2014], Abdus Salam [1926–1996], Jürgen Moser [1928–1999], Michael Francis Atiyah [1929–2019], Joel Louis Lebowitz [1930– ], Roger Penrose [1931– ], Elliott Hershel Lieb [1932– ], Sheldon Glashow [1932– ], Steven Weinberg [1933–2021], Ludvig Dmitrievich Faddeev [1934–2017], David Ruelle [1935– ], Yakov Grigorevich Sinai [1935– ], Vladimir Igorevich Arnold [1937–2010], Arthur Michael Jaffe [1937–], Roman Wladimir Jackiw [1939– ], Leonard Susskind [1940– ], Rodney James Baxter [1940– ], Michael Victor Berry [1941- ], Giovanni Gallavotti [1941- ], Stephen William Hawking [1942–2018], Jerrold Eldon Marsden [1942–2010], Michael C. Reed [1942– ], Israel Michael Sigal [1945], Alexander Markovich Polyakov [1945– ], Barry Simon [1946– ], Herbert Spohn [1946– ], John Lawrence Cardy [1947– ], Giorgio Parisi [1948– ], Edward Witten [1951– ], Ashoke Sen [1956-] and Juan Martín Maldacena [1968– ].

Stanford–Binet Intelligence Scales

From Wikipedia, the free encyclopedia
 
Stanford–Binet Intelligence scales
ICD-9-CM94.01

The Stanford–Binet Intelligence Scales (or more commonly the Stanford–Binet) is an individually administered intelligence test that was revised from the original Binet–Simon Scale by Alfred Binet and Theodore Simon. The Stanford–Binet Intelligence Scale is now in its fifth edition (SB5) and was released in 2003. It is a cognitive ability and intelligence test that is used to diagnose developmental or intellectual deficiencies in young children. The test measures five weighted factors and consists of both verbal and nonverbal subtests. The five factors being tested are knowledge, quantitative reasoning, visual-spatial processing, working memory, and fluid reasoning.

The development of the Stanford–Binet initiated the modern field of intelligence testing and was one of the first examples of an adaptive test. The test originated in France, then was revised in the United States. It was initially created by the French psychologist Alfred Binet, who, following the introduction of a law mandating universal education by the French government, began developing a method of identifying "slow" children, so that they could be placed in special education programs, instead of labelled sick and sent to the asylum. As Binet indicated, case studies might be more detailed and helpful, but the time required to test many people would be excessive. In 1916, at Stanford University, the psychologist Lewis Terman released a revised examination that became known as the Stanford–Binet test.

Development

As discussed by Fancher & Rutherford in 2012, the Stanford–Binet is a modified version of the Binet-Simon Intelligence scale. The Binet-Simon scale was created by the French psychologist Alfred Binet and his student Theodore Simon. Due to changing education laws of the time, Binet had been requested by a government commission to come up with a way to detect children who were falling behind developmentally and in need of help. Binet believed that intelligence is malleable and that intelligence tests would help target children in need of extra attention to advance their intelligence.

To create their test, Binet and Simon first created a baseline of intelligence. A wide range of children were tested on a broad spectrum of measures in an effort to discover a clear indicator of intelligence. Failing to find a single identifier of intelligence, Binet and Simon instead compared children in each category by age. The children's highest levels of achievement were sorted by age and common levels of achievement considered the normal level for that age. Because this testing method merely compares a person's ability to the common ability level of others their age, the general practices of the test can easily be transferred to test different populations, even if the measures used are changed.

Reproduction of an item from the 1908 Binet–Simon intelligence scale, that shows three pairs of pictures, and asks the tested child, "Which of these two faces is the prettier?" Reproduced from the article "A Practical Guide for Administering the Binet-Simon Scale for Measuring Intelligence" by J. W. Wallace Wallin in the December 1911 issue of the journal The Psychological Clinic (volume 5 number 7), public domain

One of the first intelligence tests, the Binet-Simon test quickly gained support in the psychological community, many of whom further spread it to the public. Lewis M. Terman, a psychologist at Stanford University, was one of the first to create a version of the test for people in the United States, naming the localized version the Stanford–Binet Intelligence Scale. Terman used the test not only to help identify children with learning difficulties but also to find children and adults who had above average levels of intelligence. In creating his version, Terman also tested additional methods for his Stanford revision, publishing his first official version as The Measurement of Intelligence: An Explanation of and a Complete Guide for the Use of the Stanford Revision and Extension of the Binet-Simon Intelligence Scale (Fancher & Rutherford, 2012) (Becker, 2003).

The original tests in the 1905 form include:

  1. "Le Regard"
  2. Prehension Provoked by a Tactile Stimulus
  3. Prehension Provoked by a Visual Perception
  4. Recognition of Food
  5. Quest of Food Complicated by a Slight Mechanical Difficulty
  6. Execution of Simple Commands and Imitation of Simple Gestures
  7. Verbal Knowledge of Objects
  8. Verbal Knowledge of Pictures
  9. Naming of Designated Objects
  10. Immediate Comparison of Two Lines of Unequal Lengths
  11. Repetition of Three Figures
  12. Comparison of Two Weights
  13. Suggestibility
  14. Verbal Definition of Known Objects
  15. Repetition of Sentences of Fifteen Words
  16. Comparison of Known Objects from Memory
  17. Exercise of Memory on Pictures
  18. Drawing a Design from Memory
  19. Immediate Repetition of Figures
  20. Resemblances of Several Known Objects Given from Memory
  21. Comparison of Lengths
  22. Five Weights to be Placed in Order
  23. Gap in Weights
  24. Exercise upon Rhymes
  25. Verbal Gaps to be Filled
  26. Synthesis of Three Words in One Sentence
  27. Reply to an Abstract Question
  28. Reversal of the Hands of a Clock
  29. Paper Cutting
  30. Definitions of Abstract Terms

Historical use

One hindrance to widespread understanding of the test is its use of a variety of different measures. In an effort to simplify the information gained from the Binet-Simon test into a more comprehensible and easier to understand form, German psychologist William Stern created the now well known Intelligence Quotient (IQ). By comparing the mental age a child scored at to their biological age, a ratio is created to show the rate of their mental progress as IQ. Terman quickly grasped the idea for his Stanford revision with the adjustment of multiplying the ratios by 100 to make them easier to read.

As also discussed by Leslie, in 2000, Terman was another of the main forces in spreading intelligence testing in the United States (Becker, 2003). Terman quickly promoted the use of the Stanford–Binet for schools across the United States where it saw a high rate of acceptance. Terman's work also had the attention of the U.S. government, who recruited him to apply the ideas from his Stanford–Binet test for military recruitment near the start of World War I. With over 1.7 million military recruits taking a version of the test and the acceptance of the test by the government, the Stanford–Binet saw an increase in awareness and acceptance (Fancher & Rutherford, 2012).

Given the perceived importance of intelligence and with new ways to measure intelligence, many influential individuals, including Terman, began promoting controversial ideas to increase the nation's overall intelligence. These ideas included things such as discouraging individuals with low IQ from having children and granting important positions based on high IQ scores. While there was significant opposition, many institutions proceeded to adjust students' education based on their IQ scores, often with a heavy influence on future career possibilities (Leslie, 2000).

Revisions of the Stanford–Binet Intelligence Scale

Maud Merrill

Since the first publication in 1916, there have been four additional revised editions of the Stanford–Binet Intelligence Scales, the first of which was developed by Lewis Terman. Over twenty years later, Maud Merrill was accepted into Stanford's education program shortly before Terman became the head of the psychology department. She completed both her Masters Degree and Ph.D. under Terman and quickly became a colleague of his as they started the revisions of the second edition together. There were 3,200 examinees, aged one and a half to eighteen years, ranging in different geographic regions as well as socioeconomic levels in attempts to comprise a broader normative sample (Roid & Barram, 2004). This edition incorporated more objectified scoring methods, while placing less emphasis on recall memory and including a greater range of nonverbal abilities (Roid & Barram, 2004) compared to the 1916 edition.

When Terman died in 1956, the revisions for the third edition were well underway, and Merrill was able to publish the final revision in 1960 (Roid & Barram, 2004). The use of deviation IQ made its first appearance in third edition, however the use of the mental age scale and ratio IQ were not eliminated. Terman and Merrill attempted to calculate IQs with a uniform standard deviation while still maintaining the use of the mental age scale by including a formula in the manual to convert the ratio IQs with means varying between age ranges and nonuniform standard deviations to IQs with a mean of 100 and a uniform standard deviation of 16. However, it was later demonstrated that very high scores occurred with much greater frequency than what would be predicted by the normal curve with a standard deviation of 16, and scores in the gifted range were much higher than those yielded by essentially every other major test, so it was deemed that the ratio IQs modified to have a uniform mean and standard deviation, referred to as "deviation IQs" in the manual of the third edition of the Stanford–Binet (Terman & Merrill, 1960), could not be directly compared to scores on "true" deviation IQ tests, such as the Wechsler Intelligence Scales, and the later versions of the Stanford–Binet, as those tests compare the performance of examinees to their own age group on a normal distribution (Ruf, 2003). While new features were added, there were no newly created items included in this revision. Instead, any items from the 1937 form that showed no substantial change in difficulty from the 1930s to the 1950s were either eliminated or adjusted (Roid & Barram, 2004).

Robert Thorndike was asked to take over after Merrill's retirement. With the help of Elizabeth Hagen and Jerome Sattler, Thorndike produced the fourth edition of the Stanford–Binet Intelligence Scale in 1986. This edition covers the ages two through twenty-three and has some considerable changes compared to its predecessors (Graham & Naglieri, 2003). This edition was the first to use the fifteen subtests with point scales in place of using the previous age scale format. In an attempt to broaden cognitive ability, the subtests were grouped and resulted in four area scores, which improved flexibility for administration and interpretation (Youngstrom, Glutting, & Watkins, 2003). The fourth edition is known for assessing children that may be referred for gifted programs. This edition includes a broad range of abilities, which provides more challenging items for those in their early adolescent years, whereas other intelligence tests of the time did not provide difficult enough items for the older children (Laurent, Swerdlik, & Ryburn, 1992).

Gale Roid published the most recent edition of the Stanford–Binet Intelligence Scale. Roid attended Harvard University where he was a research assistant to David McClelland. McClelland is well known for his studies on the need for achievement. While the fifth edition incorporates some of the classical traditions of these scales, there were several significant changes made.

Timeline

  • April 1905: Development of Binet-Simon Test announced at a conference in Rome
  • June 1905: Binet-Simon Intelligence Test introduced
  • 1908 and 1911: New Versions of Binet-Simon Intelligence Test
  • 1916: Stanford–Binet First Edition by Terman
  • 1937: Second Edition by Terman and Merrill
  • 1960: L-M modified second edition by Merrill
  • 1973: Third Edition by Merrill (1937 re-normed)
  • 1986: Fourth Edition by Thorndike, Hagen, and Sattler
  • 2003: Fifth Edition by Roid

Stanford–Binet Intelligence Scale: Fifth Edition

Just as it was used when Binet first developed the IQ test, the Stanford–Binet Intelligence Scale: Fifth Edition (SB5) is based in the schooling process to assess intelligence. It continuously and efficiently assesses all levels of ability in individuals with a broader range in age. It is also capable of measuring multiple dimensions of abilities (Ruf, 2003).

The SB5 can be administered to individuals as early as two years of age. There are ten subsets included in this revision including both verbal and nonverbal domains. Five factors are also incorporated in this scale, which are directly related to Cattell-Horn-Carroll (CHC) hierarchical model of cognitive abilities. These factors include fluid reasoning, knowledge, quantitative reasoning, visual-spatial processing, and working memory (Bain & Allin, 2005). Many of the familiar picture absurdities, vocabulary, memory for sentences, and verbal absurdities still remain from the previous editions (Janzen, Obrzut, & Marusiak, 2003), however with more modern artwork and item content for the revised fifth edition.

For every verbal subtest that is used, there is a nonverbal counterpart across all factors. These nonverbal tasks consist of making movement responses such as pointing or assembling manipulatives (Bain & Allin, 2005). These counterparts have been included to address language-reduced assessments in multicultural societies. Depending on age and ability, administration can range from fifteen minutes to an hour and fifteen minutes.

The fifth edition incorporated a new scoring system, which can provide a wide range of information such as four intelligence score composites, five factor indices, and ten subtest scores. Additional scoring information includes percentile ranks, age equivalents, and a change-sensitive score (Janzen, Obrzut, & Marusiak, 2003). Extended IQ scores and gifted composite scores are available with the SB5 in order to optimize the assessment for gifted programs (Ruf, 2003). To reduce errors and increase diagnostic precision, scores are obtained electronically through the use of computers now.

The standardization sample for the SB5 included 4,800 participants varying in age, sex, race/ethnicity, geographic region, and socioeconomic level (Bain & Allin, 2005).

Reliability

Several reliability tests have been performed on the SB5 including split-half reliability, standard error of measurement, plotting of test information curves, test-retest stability, and inter-scorer agreement. On average, IQ scores for this scale have been found quite stable across time (Janzen, Obrzut, & Marusiak, 2003). Internal consistency was tested by split-half reliability and was reported to be substantial and comparable to other cognitive batteries (Bain & Allin, 2005). The median interscorer correlation was .90 on average (Janzen, Obrzut, & Marusiak, 2003). The SB5 has also been found to have great precision at advanced levels of performance meaning that the test is especially useful in testing children for giftedness (Bain & Allin, 2005). There have only been a small amount of practice effects and familiarity of testing procedures with retest reliability; however, these have proven to be insignificant. Readministration of the SB5 can occur in a six-month interval rather than one year due to the small mean differences in reliability (Bain & Allin, 2005).

Validity

Content validity has been found based on the professional judgments Roid received concerning fairness of items and item content as well as items concerning the assessment of giftedness (Bain & Allin, 2005). With an examination of age trends, construct validity was supported along with empirical justification of a more substantial g loading for the SB5 compared to previous editions. The potential for a variety of comparisons, especially for within or across factors and verbal/nonverbal domains, has been appreciated with the scores received from the SB5 (Bain & Allin, 2005).

Score classification

The test publisher includes suggested score classifications in the test manual.

Stanford–Binet Fifth Edition (SB5) classification
IQ Range ("deviation IQ") IQ Classification
145–160 Very gifted or highly advanced
130–144 Gifted or very advanced
120–129 Superior
110–119 High average
90–109 Average
80–89 Low average
70–79 Borderline impaired or delayed
55–69 Mildly impaired or delayed
40–54 Moderately impaired or delayed

The classifications of scores used in the Fifth Edition differ from those used in earlier versions of the test.

Subtests and factors

Fluid reasoning Knowledge Quantitative reasoning Visual-spatial processing Working memory
Early reasoning Vocabulary Non-verbal quantitative reasoning (non-verbal) Form board and form patterns

(non-verbal)

Delayed response (non-verbal)
Verbal absurdities Procedural knowledge (non-verbal) Verbal quantitative reasoning Position and direction Block span (non-verbal)
Verbal analogies Picture absurdities (non-verbal)

Memory for sentences
Object series matrices (non-verbal)


Last word

Present use

Since its inception, the Stanford–Binet has been revised several times. Currently, the test is in its fifth edition, which is called the Stanford–Binet Intelligence Scales, Fifth Edition, or SB5. According to the publisher's website, "The SB5 was normed on a stratified random sample of 4,800 individuals that matches the 2000 U.S. Census". By administering the Stanford–Binet test to large numbers of individuals selected at random from different parts of the United States, it has been found that the scores approximate a normal distribution. The revised edition of the Stanford–Binet over time has devised substantial changes in the way the tests are presented. The test has improved when looking at the introduction of a more parallel form and more demonstrative standards. For one, a non-verbal IQ component is included in the present day tests whereas in the past, there was only a verbal component. In fact, it now has equal balance of verbal and non-verbal content in the tests. It is also more animated than the other tests, providing the test-takers with more colourful artwork, toys and manipulatives. This allows the test to have a higher range in the age of the test takers. This test is purportedly useful in assessing the intellectual capabilities of people ranging from young children all the way to young adults. However, the test has come under criticism for not being able to compare people of different age categories, since each category gets a different set of tests. Furthermore, very young children tend to do poorly on the test due to the fact that they lack the ability to concentrate long enough to finish it.

Current uses for the test include clinical and neuropsychological assessment, educational placement, compensation evaluations, career assessment, adult neuropsychological treatment, forensics, and research on aptitude. Various high-IQ societies also accept this test for admission into their ranks; for example, the Triple Nine Society accepts a minimum qualifying score of 151 for Form L or M, 149 for Form L-M if taken in 1986 or earlier, 149 for SB-IV, and 146 for SB-V; in all cases the applicant must have been at least 16 years old at the date of the test. Intertel accepts a score of 135 on SB5 and 137 on Form L-M.

Moons of Jupiter

From Wikipedia, the free encyclopedia

A montage of Jupiter and its four largest moons (distance and sizes not to scale)

There are 80 known moons of Jupiter, not counting a number of moonlets likely shed from the inner moons. All together, they form a satellite system which is called the Jovian system. The most massive of the moons are the four Galilean moons: Io, Europa, Ganymede, and Callisto, which were independently discovered in 1610 by Galileo Galilei and Simon Marius and were the first objects found to orbit a body that was neither Earth nor the Sun. Much more recently, beginning in 1892, dozens of far smaller Jovian moons have been detected and have received the names of lovers (or other sexual partners) or daughters of the Roman god Jupiter or his Greek equivalent Zeus. The Galilean moons are by far the largest and most massive objects to orbit Jupiter, with the remaining 76 known moons and the rings together composing just 0.003% of the total orbiting mass.

Of Jupiter's moons, eight are regular satellites with prograde and nearly circular orbits that are not greatly inclined with respect to Jupiter's equatorial plane. The Galilean satellites are nearly spherical in shape due to their planetary mass, and so would be considered at least dwarf planets if they were in direct orbit around the Sun. The other four regular satellites are much smaller and closer to Jupiter; these serve as sources of the dust that makes up Jupiter's rings. The remainder of Jupiter's moons are irregular satellites whose prograde and retrograde orbits are much farther from Jupiter and have high inclinations and eccentricities. These moons were probably captured by Jupiter from solar orbits. Twenty-three of the irregular satellites have not yet been officially named.

Characteristics

The Galilean moons. From left to right, in order of increasing distance from Jupiter: Io; Europa; Ganymede; Callisto.

The physical and orbital characteristics of the moons vary widely. The four Galileans are all over 3,100 kilometres (1,900 mi) in diameter; the largest Galilean, Ganymede, is the ninth largest object in the Solar System, after the Sun and seven of the planets, Ganymede being larger than Mercury. All other Jovian moons are less than 250 kilometres (160 mi) in diameter, with most barely exceeding 5 kilometres (3.1 mi). Their orbital shapes range from nearly perfectly circular to highly eccentric and inclined, and many revolve in the direction opposite to Jupiter's rotation (retrograde motion). Orbital periods range from seven hours (taking less time than Jupiter does to rotate around its axis), to some three thousand times more (almost three Earth years).

Origin and evolution

The relative masses of the Jovian moons. Those smaller than Europa are not visible at this scale, and combined would only be visible at 100× magnification.

Jupiter's regular satellites are believed to have formed from a circumplanetary disk, a ring of accreting gas and solid debris analogous to a protoplanetary disk. They may be the remnants of a score of Galilean-mass satellites that formed early in Jupiter's history.

Simulations suggest that, while the disk had a relatively high mass at any given moment, over time a substantial fraction (several tens of a percent) of the mass of Jupiter captured from the solar nebula was passed through it. However, only 2% of the proto-disk mass of Jupiter is required to explain the existing satellites. Thus, several generations of Galilean-mass satellites may have been in Jupiter's early history. Each generation of moons might have spiraled into Jupiter, because of drag from the disk, with new moons then forming from the new debris captured from the solar nebula. By the time the present (possibly fifth) generation formed, the disk had thinned so that it no longer greatly interfered with the moons' orbits. The current Galilean moons were still affected, falling into and being partially protected by an orbital resonance with each other, which still exists for Io, Europa, and Ganymede. Ganymede's larger mass means that it would have migrated inward at a faster rate than Europa or Io.

The outer, irregular moons are thought to have originated from captured asteroids, whereas the protolunar disk was still massive enough to absorb much of their momentum and thus capture them into orbit. Many are believed to have broken up by mechanical stresses during capture, or afterward by collisions with other small bodies, producing the moons we see today.

Discovery

Jupiter and the Galilean moons through a 25 cm (10 in) Meade LX200 telescope.
 
The number of moons known for each of the four outer planets up to October 2019. Jupiter currently has 80 known satellites.
 

Chinese historian Xi Zezong claimed that the earliest record of a Jovian moon (Ganymede or Callisto) was a note by Chinese astronomer Gan De of an observation around 364 BC regarding a "reddish star". However, the first certain observations of Jupiter's satellites were those of Galileo Galilei in 1609. By January 1610, he had sighted the four massive Galilean moons with his 20× magnification telescope, and he published his results in March 1610.

Simon Marius had independently discovered the moons one day after Galileo, although he did not publish his book on the subject until 1614. Even so, the names Marius assigned are used today: Ganymede, Callisto, Io, and Europa. No additional satellites were discovered until E. E. Barnard observed Amalthea in 1892.

With the aid of telescopic photography, further discoveries followed quickly over the course of the 20th century. Himalia was discovered in 1904, Elara in 1905, Pasiphae in 1908, Sinope in 1914, Lysithea and Carme in 1938, Ananke in 1951, and Leda in 1974. By the time that the Voyager space probes reached Jupiter, around 1979, 13 moons had been discovered, not including Themisto, which had been observed in 1975, but was lost until 2000 due to insufficient initial observation data. The Voyager spacecraft discovered an additional three inner moons in 1979: Metis, Adrastea, and Thebe.

No additional moons were discovered for two decades, but between October 1999 and February 2003, researchers found another 34 moons using sensitive ground-based detectors. These are tiny moons, in long, eccentric, generally retrograde orbits, and averaging 3 km (1.9 mi) in diameter, with the largest being just 9 km (5.6 mi) across. All of these moons are thought to have been captured asteroidal or perhaps comet bodies, possibly fragmented into several pieces.

By 2015, a total of 15 additional moons were discovered. Two more were discovered in 2017 by the team led by Scott S. Sheppard at the Carnegie Institution for Science, bringing the total to 69. On 17 July 2018, the International Astronomical Union confirmed that Sheppard's team had discovered ten more moons around Jupiter, bringing the total number to 79. Among these is Valetudo, which has a prograde orbit, but crosses paths with several moons that have retrograde orbits, making an eventual collision—at some point on a billions-of-years timescale—likely.

In September 2020, researchers from the University of British Columbia identified 45 candidate moons from an analysis of archival images taken in 2010 by the Canada-France-Hawaii Telescope. These candidates were mainly small and faint, down to a magnitude of 25.7 or over 800 m (0.50 mi) in diameter. From the number of candidate moons detected within a sky area of one square degree, the team extrapolated that the population of retrograde Jovian moons brighter than magnitude 25.7 is around 600, within a factor of 2. Although the team considers their characterised candidates to be likely moons of Jupiter, they all remain unconfirmed due to their insufficient observation data for determining reliable orbits for each of them.

Naming

Galilean moons around Jupiter   Jupiter ·   Io ·   Europa ·   Ganymede ·   Callisto
 
Orbits of Jupiter's inner moons within its rings

The Galilean moons of Jupiter (Io, Europa, Ganymede, and Callisto) were named by Simon Marius soon after their discovery in 1610. However, these names fell out of favor until the 20th century. The astronomical literature instead simply referred to "Jupiter I", "Jupiter II", etc., or "the first satellite of Jupiter", "Jupiter's second satellite", and so on. The names Io, Europa, Ganymede, and Callisto became popular in the mid-20th century, whereas the rest of the moons remained unnamed and were usually numbered in Roman numerals V (5) to XII (12). Jupiter V was discovered in 1892 and given the name Amalthea by a popular though unofficial convention, a name first used by French astronomer Camille Flammarion.

The other moons were simply labeled by their Roman numeral (e.g. Jupiter IX) in the majority of astronomical literature until the 1970s. Several different suggestions were made for names of Jupiter's outer satellites, but none were universally accepted until 1975 when the International Astronomical Union's (IAU) Task Group for Outer Solar System Nomenclature granted names to satellites V–XIII, and provided for a formal naming process for future satellites still to be discovered. The practice was to name newly discovered moons of Jupiter after lovers and favorites of the god Jupiter (Zeus) and, since 2004, also after their descendants. All of Jupiter's satellites from XXXIV (Euporie) onward are named after descendants of Jupiter or Zeus, except LIII (Dia), named after a lover of Jupiter. Names ending with "a" or "o" are used for prograde irregular satellites (the latter for highly inclined satellites), and names ending with "e" are used for retrograde irregulars. With the discovery of smaller, kilometre-sized moons around Jupiter, the IAU has established an additional convention to limit the naming of small moons with absolute magnitudes greater than 18 or diameters smaller than 1 km (0.62 mi). Some of the most recently confirmed moons have not received names.

Some asteroids share the same names as moons of Jupiter: 9 Metis, 38 Leda, 52 Europa, 85 Io, 113 Amalthea, 239 Adrastea. Two more asteroids previously shared the names of Jovian moons until spelling differences were made permanent by the IAU: Ganymede and asteroid 1036 Ganymed; and Callisto and asteroid 204 Kallisto.

Groups

The orbits of Jupiter's irregular satellites, and how they cluster into groups: by semi-major axis (the horizontal axis in Gm); by orbital inclination (the vertical axis); and orbital eccentricity (the yellow lines). The relative sizes are indicated by the circles.

Regular satellites

These have prograde and nearly circular orbits of low inclination and are split into two groups:

  • Inner satellites or Amalthea group: Metis, Adrastea, Amalthea, and Thebe. These orbit very close to Jupiter; the innermost two orbit in less than a Jovian day. The latter two are respectively the fifth and seventh largest moons in the Jovian system. Observations suggest that at least the largest member, Amalthea, did not form on its present orbit, but farther from the planet, or that it is a captured Solar System body. These moons, along with a number of seen and as-yet-unseen inner moonlets (see Amalthea moonlets), replenish and maintain Jupiter's faint ring system. Metis and Adrastea help to maintain Jupiter's main ring, whereas Amalthea and Thebe each maintain their own faint outer rings.
  • Main group or Galilean moons: Io, Europa, Ganymede and Callisto. They are some of the largest objects in the Solar System outside the Sun and the eight planets in terms of mass, larger than any known dwarf planet. Ganymede exceeds (and Callisto nearly equals) even the planet Mercury in diameter, though they are less massive. They are respectively the fourth-, sixth-, first-, and third-largest natural satellites in the Solar System, containing approximately 99.997% of the total mass in orbit around Jupiter, while Jupiter is almost 5,000 times more massive than the Galilean moons. The inner moons are in a 1:2:4 orbital resonance. Models suggest that they formed by slow accretion in the low-density Jovian subnebula—a disc of the gas and dust that existed around Jupiter after its formation—which lasted up to 10 million years in the case of Callisto. Europa, Ganymede, and Callisto are suspected of having subsurface water oceans, and Io may have a subsurface magma ocean.

Irregular satellites

Orbits and positions of Jupiter's irregular satellites as of 1 January 2021. Prograde orbits are colored blue while retrograde orbits are colored red.
 
Inclinations (°) vs. eccentricities of Jupiter's irregular satellites, with the major groups identified. Data as of 2021.
 

The irregular satellites are substantially smaller objects with more distant and eccentric orbits. They form families with shared similarities in orbit (semi-major axis, inclination, eccentricity) and composition; it is believed that these are at least partially collisional families that were created when larger (but still small) parent bodies were shattered by impacts from asteroids captured by Jupiter's gravitational field. These families bear the names of their largest members. The identification of satellite families is tentative, but the following are typically listed:

  • Prograde satellites:
    • Themisto is the innermost irregular moon and is not part of a known family.
    • The Himalia group is spread over barely 1.4 Gm in semi-major axes, 1.6° in inclination (27.5 ± 0.8°), and eccentricities between 0.11 and 0.25. It has been suggested that the group could be a remnant of the break-up of an asteroid from the asteroid belt.
    • Carpo is another prograde moon and is not part of a known family. It has the highest inclination of all of the prograde moons.
    • Valetudo is the outermost prograde moon and is not part of a known family. Its prograde orbit crosses paths with several moons that have retrograde orbits and may in the future collide with them.
  • Retrograde satellites:
    • The Carme group is spread over only 1.2 Gm in semi-major axis, 1.6° in inclination (165.7 ± 0.8°), and eccentricities between 0.23 and 0.27. It is very homogeneous in color (light red) and is believed to have originated from a D-type asteroid progenitor, possibly a Jupiter trojan.
    • The Ananke group has a relatively wider spread than the previous groups, over 2.4 Gm in semi-major axis, 8.1° in inclination (between 145.7° and 154.8°), and eccentricities between 0.02 and 0.28. Most of the members appear gray, and are believed to have formed from the breakup of a captured asteroid.
    • The Pasiphae group is quite dispersed, with a spread over 1.3 Gm, inclinations between 144.5° and 158.3°, and eccentricities between 0.25 and 0.43. The colors also vary significantly, from red to grey, which might be the result of multiple collisions. Sinope, sometimes included in the Pasiphae group, is red and, given the difference in inclination, it could have been captured independently; Pasiphae and Sinope are also trapped in secular resonances with Jupiter.

List

The moons of Jupiter are listed below by orbital period. Moons massive enough for their surfaces to have collapsed into a spheroid are highlighted in bold. These are the four Galilean moons, which are comparable in size to the Moon. The other moons are much smaller, with the least massive Galilean moon being more than 7,000 times more massive than the most massive of the other moons. The irregular captured moons are shaded light gray when prograde and dark gray when retrograde. The orbits and mean distances of the irregular moons are strongly variable over short timescales due to frequent planetary and solar perturbations, therefore the listed orbital elements of all irregular moons are averaged over a 400-year numerical integration. Their orbital elements are all based on the epoch of 1 January 2000. As of 2021, S/2003 J 10 is the only moon of Jupiter considered lost due to its uncertain orbit. A number of other moons have only been observed for a year or two, but have decent enough orbits to be easily measurable at present.

Exploration

The orbit and motion of the Galilean moons around Jupiter, as captured by JunoCam aboard the Juno spacecraft.

Nine spacecraft have visited Jupiter. The first were Pioneer 10 in 1973, and Pioneer 11 a year later, taking low-resolution images of the four Galilean moons and returning data on their atmospheres and radiation belts. The Voyager 1 and Voyager 2 probes visited Jupiter in 1979, discovering the volcanic activity on Io and the presence of water ice on the surface of Europa. Ulysses further studied Jupiter's magnetosphere in 1992 and then again in 2000.

The Galileo spacecraft was the first to enter orbit around Jupiter, arriving in 1995 and studying it until 2003. During this period, Galileo gathered a large amount of information about the Jovian system, making close approaches to all of the Galilean moons and finding evidence for thin atmospheres on three of them, as well as the possibility of liquid water beneath the surfaces of Europa, Ganymede, and Callisto. It also discovered a magnetic field around Ganymede.

Then the Cassini probe to Saturn flew by Jupiter in 2000 and collected data on interactions of the Galilean moons with Jupiter's extended atmosphere. The New Horizons spacecraft flew by Jupiter in 2007 and made improved measurements of its satellites' orbital parameters.

Ganymede taken by Juno during its 34th perijove.

In 2016, the Juno spacecraft imaged the Galilean moons from above their orbital plane as it approached Jupiter orbit insertion, creating a time-lapse movie of their motion.

Computer-aided software engineering

From Wikipedia, the free encyclopedia ...