algebra

algebra
/al"jeuh breuh/, n.
1. the branch of mathematics that deals with general statements of relations, utilizing letters and other symbols to represent specific sets of numbers, values, vectors, etc., in the description of such relations.
2. any of several algebraic systems, esp. a ring in which elements can be multiplied by real or complex numbers (linear algebra) as well as by other elements of the ring.
3. any special system of notation adapted to the study of a special system of relationship: algebra of classes.
[1535-45; < ML < Ar al-jabr lit., restoration]

* * *

(as used in expressions)

* * *

Introduction

      branch of mathematics in which arithmetical operations and formal manipulations are applied to abstract symbols rather than specific numbers. The notion that there exists such a distinct subdiscipline of mathematics, as well as the term algebra to denote it, resulted from a slow historical development. This article presents that history, tracing the evolution over time of the concept of the equation, number systems, symbols for conveying and manipulating mathematical statements, and the modern abstract structural view of algebra. For information on specific branches of algebra, see elementary algebra (algebra, elementary), linear algebra (algebra, linear), and modern algebra (algebra, modern).

Emergence of formal equations
      Perhaps the most basic notion in mathematics is the equation, a formal statement that two sides of a mathematical expression are equal—as in the simple equation x + 3 = 5—and that both sides of the equation can be simultaneously manipulated (by adding, dividing, taking roots, and so on to both sides) in order to “solve” the equation. Yet, as simple and natural as such a notion may appear today, its acceptance first required the development of numerous mathematical ideas, each of which took time to mature. In fact, it took until the late 16th century to consolidate the modern concept of an equation as a single mathematical entity.

      Three main threads in the process leading to this consolidation deserve special attention:
● Attempts to solve equations involving one or more unknown quantities. In describing the early history of algebra, the word equation is frequently used out of convenience to describe these operations, although early mathematicians would not have been aware of such a concept.
● The evolution of the notion of exactly what qualifies as a legitimate number. Over time this notion expanded to include broader domains (rational numbers, irrational numbers (irrational number), negative numbers, and complex numbers (complex number)) that were flexible enough to support the abstract structure of symbolic algebra.
● The gradual refinement of a symbolic language suitable for devising and conveying generalized algorithms (algorithm), or step-by-step procedures for solving entire categories of mathematical problems.

      These three threads are traced in this section, particularly as they developed in the ancient Middle East and Greece, the Islamic era, and the European Renaissance.

Problem solving in Egypt (Egypt, ancient) and Babylon
      The earliest extant mathematical text from Egypt is the Rhind papyrus (c. 1650 BC). It and other texts attest to the ability of the ancient Egyptians to solve linear equations (linear equation) in one unknown. A linear equation is a first-degree equation, or one in which all the variables are only to the first power. (In today's notation, such an equation in one unknown would be 7x + 3x = 10.) Evidence from about 300 BC indicates that the Egyptians also knew how to solve problems involving a system of two equations in two unknown quantities, including quadratic (second-degree, or squared unknowns) equations. For example, given that the perimeter of a rectangular plot of land is 100 units and its area is 600 square units, the ancient Egyptians could solve for the field's length l and width w. (In modern notation, they could solve the pair of simultaneous equations 2w + 2l =100 and wl = 600.) However, throughout this period there was no use of symbols—problems were stated and solved verbally. The following problem is typical:

● Method of calculating a quantity,
● multiplied by 1 1/2 added 4 it has come to 10.
● What is the quantity that says it?
● First you calculate the difference of this 10 to this 4. Then 6 results.
● Then you divide 1 by 1 1/2. Then 2/3 results.
● Then you calculate 2/3 of this 6. Then 4 results.
● Behold, it is 4, the quantity that said it.
● What has been found by you is correct.

      Note that except for 2/3, for which a special symbol existed, the Egyptians expressed all fractional quantities using only unit fractions, that is, fractions bearing the numerator 1. For example, 3/4 would be written as 1/2 + 1/4.

      Babylonian (Babylonia) mathematics dates from as early as 1800 BC, as indicated by cuneiform texts preserved in clay tablets. Babylonian arithmetic was based on a well-elaborated, positional sexagesimal system—that is, a system of base 60, as opposed to the modern decimal system (decimal number system), which is based on units of 10. The Babylonians, however, made no consistent use of zero. A great deal of their mathematics consisted of tables, such as for multiplication, reciprocals, squares (but not cubes), and square and cube roots.

      In addition to tables, many Babylonian tablets contained problems that asked for the solution of some unknown number. Such problems explained a procedure to be followed for solving a specific problem, rather than proposing a general algorithm for solving similar problems. The starting point for a problem could be relations involving specific numbers and the unknown, or its square, or systems of such relations. The number sought could be the square root of a given number, the weight of a stone, or the length of the side of a triangle. Many of the questions were phrased in terms of concrete situations—such as partitioning a field among three pairs of brothers under certain constraints. Still, their artificial character made it clear that they were constructed for didactical purposes.

Greece (ancient Greek civilization) and the limits of geometric expression
The Pythagoreans (Pythagoreanism) and Euclid
 A major milestone of Greek mathematics was the discovery by the Pythagoreans (Pythagoreanism) around 430 BC that not all lengths are commensurable, that is, measurable by a common unit. This surprising fact became clear while investigating what appeared to be the most elementary ratio between geometric magnitudes, namely, the ratio between the side and the diagonal of a square. The Pythagoreans knew that for a unit square (that is, a square whose sides have a length of 1), the length of the diagonal must be √2—owing to the Pythagorean theorem, which states that the square on the diagonal of a triangle must equal the sum of the squares on the other two sides (a2 + b2 = c2). The ratio between the two magnitudes thus deduced, 1 and √2, had the confounding property of not corresponding to the ratio of any two whole, or counting, numbers (1, 2, 3,…). This discovery of incommensurable quantities contradicted the basic metaphysics of Pythagoreanism, which asserted that all of reality was based on the whole numbers.

      Attempts to deal with incommensurables eventually led to the creation of an innovative concept of proportion by Eudoxus of Cnidus (c. 400–350 BC), which Euclid preserved in his Elements (c. 300 BC). The theory of proportions remained an important component of mathematics well into the 17th century, by allowing the comparison of ratios of pairs of magnitudes of the same kind. Greek proportions, however, were very different from modern equalities, and no concept of equation could be based on it. For instance, a proportion could establish that the ratio between two line segments, say A and B, is the same as the ratio between two areas, say R and S. The Greeks would state this in strictly verbal fashion, since symbolic expressions, such as the much later A:B::R:S (read, A is to B as R is to S), did not appear in Greek texts. The theory of proportions enabled significant mathematical results, yet it could not lead to the kind of results derived with modern equations. Thus, from A:B::R:S the Greeks could deduce that (in modern terms) A + B:A − B::R + S:R − S, but they could not deduce in the same way that A:R::B:S. In fact, it did not even make sense to the Greeks to speak of a ratio between a line and an area since only like, or homogeneous, magnitudes were comparable. Their fundamental demand for homogeneity was strictly preserved in all Western mathematics until the 17th century.

      When some of the Greek geometric constructions, such as those that appear in Euclid's Elements, are suitably translated into modern algebraic language, they establish algebraic identities, solve quadratic equations, and produce related results. However, not only were symbols of this kind never used in classical Greek works but such a translation would be completely alien to their spirit. Indeed, the Greeks not only lacked an abstract language for performing general symbolic manipulations but they even lacked the concept of an equation to support such an algebraic interpretation of their geometric constructions.

      For the classical Greeks, especially as shown in Books VII–XI of the Elements, a number was a collection of units, and hence they were limited to the counting numbers. Negative numbers were obviously out of this picture, and zero could not even start to be considered. In fact, even the status of 1 was ambiguous in certain texts, since it did not really constitute a collection as stipulated by Euclid. Such a numerical limitation, coupled with the strong geometric orientation of Greek mathematics, slowed the development and full acceptance of more elaborate and flexible ideas of number in the West.

Diophantus
      A somewhat different, and idiosyncratic, orientation to solving mathematical problems can be found in the work of a later Greek, Diophantus of Alexandria (fl. c. AD 250), who developed original methods for solving problems that, in retrospect, may be seen as linear or quadratic equations. Yet even Diophantus, in line with the basic Greek conception of mathematics, considered only positive rational solutions; he called a problem “absurd” whose only solutions were negative numbers. Diophantus solved specific problems using ad hoc methods convenient for the problem at hand, but he did not provide general solutions. The problems that he solved sometimes had more than one (and in some cases even infinitely many) solutions, yet he always stopped after finding the first one. In problems involving quadratic equations, he never suggested that such equations might have two solutions.

      On the other hand, Diophantus was the first to introduce some kind of systematic symbolism (symbol) for polynomial equations. A polynomial equation is composed of a sum of terms, in which each term is the product of some constant and a nonnegative power of the variable or variables. Because of their great generality, polynomial equations can express a large proportion of the mathematical relationships that occur in nature—for example, problems involving area, volume, mixture, and motion. In modern notation, polynomial equations in one variable take the form

anxn + an−1xn−1 + … + a2x2 + a1x + a0 = 0,
where the ai are known as coefficients and the highest power of n is known as the degree of the equation (for example, 2 for a quadractic, 3 for a cubic, 4 for a quartic, 5 for a quintic, and so on). Diophantus's symbolism was a kind of shorthand, though, rather than a set of freely manipulable symbols. A typical case was:
ΔνΔβζδΜβΚνβανγ
(meaning: 2x4 − x3 − 3x2 + 4x + 2). Here M represents units, ζ the unknown quantity, Kν its square, and so forth. Since there were no negative coefficients, the terms that corresponded to the unknown and its third power appeared to the right of the special symbol . This symbol did not function like the equals sign of a modern equation, however; there was nothing like the idea of moving terms from one side of the symbol to the other. Also, since all of the Greek letters were used to represent specific numbers, there was no simple and unambiguous method of representing abstract coefficients in an equation.

      A typical Diophantine problem would be: “Find two numbers such that each, after receiving from the other a given number, will bear to the remainder a given relation.” In modern terms, this problem would be stated

(x + a)/(y − a) = r, (y + b)/(x − b) = s.
Diophantus always worked with a single unknown quantity ζ. In order to solve this specific problem, he assumed as given certain values that allowed him a smooth solution: a = 30, r = 2, b = 50, s = 3. Now the two numbers sought were ζ + 30 (for y) and 2ζ − 30 (for x), so that the first ratio was an identity, /ζ = 2, that was fulfilled for any nonzero value of ζ. For the modern reader, substituting these values in the second ratio would result in (ζ + 80)/(2ζ − 80) = 3. By applying his solution techniques, Diophantus was led to ζ = 64. The two required numbers were therefore 98 and 94.

The equation in India and China
      Indian mathematicians, such as Brahmagupta (AD 598–670) and Bhaskara II (Bhāskara II) (AD 1114–1185), developed nonsymbolic, yet very precise, procedures for solving first- and second-degree equations and equations with more than one variable. However, the main contribution of Indian mathematicians was the elaboration of the decimal (decimal number system), positional numeral system. A full-fledged decimal, positional system certainly existed in India by the 9th century, yet many of its central ideas had been transmitted well before that time to China and the Islamic world. Indian arithmetic, moreover, developed consistent and correct rules for operating with positive and negative numbers and for treating zero like any other number, even in problematic contexts such as division. Several hundred years passed before European mathematicians fully integrated such ideas into the developing discipline of algebra.

      Chinese mathematicians during the period parallel to the European Middle Ages developed their own methods for classifying and solving quadratic equations by radicals—solutions that contain only combinations of the most tractable operations: addition, subtraction, multiplication, division, and taking roots. They were unsuccessful, however, in their attempts to obtain exact solutions to higher-degree equations. Instead, they developed approximation methods of high accuracy, such as those described in Yang Hui's Yang Hui suanfa (1275; “Yang Hui's Mathematical Methods”). The calculational advantages afforded by their expertise with the abacus may help explain why Chinese mathematicians gravitated to numerical analysis methods.

Islamic contributions
 Islamic (Islāmic world) contributions to mathematics began around AD 825, when the Baghdad mathematician Muḥammad ibn Mūsā al-Khwārizmī (Khwārizmī, al-) wrote his famous treatise al-Kitāb al-mukhtaṣar fī ḥisāb al-jabr wa'l-muqābala (translated into Latin in the 12th century as Algebra et Almucabal, from which the modern term algebra is derived). By the end of the 9th century a significant Greek mathematical corpus, including works of Euclid, Archimedes (c. 285–212/211 BC), Apollonius of Perga (c. 262–190 BC), Ptolemy (fl. AD 127–145), and Diophantus, had been translated into Arabic. Similarly, ancient Babylonian and Indian mathematics, as well as more recent contributions by Jewish sages, were available to Islamic scholars. This unique background allowed the creation of a whole new kind of mathematics that was much more than a mere amalgamation of these earlier traditions. A systematic study of methods for solving quadratic equations constituted a central concern of Islamic mathematicians. A no less central contribution was related to the Islamic reception and transmission of ideas related to the Indian system of numeration, to which they added decimal fractions (fractions such as 0.125, or 1/8).

      Al-Khwārizmī's algebraic work embodied much of what was central to Islamic contributions. He declared that his book was intended to be of “practical” value, yet this definition hardly applies to its contents. In the first part of his book, al-Khwārizmī presented the procedures for solving six types of equations: squares equal roots, squares equal numbers, roots equal numbers, squares and roots equal numbers, squares and numbers equal roots, and roots and numbers equal squares. In modern notation, these equations would be stated ax2 = bx, ax2 = c, bx = c, ax2 + bx = c, ax2 + c = bx, and bx + c = ax2, respectively. Only positive numbers were considered legitimate coefficients or solutions to equations. Moreover, neither symbolic representation nor abstract symbol manipulation appeared in these problems—even the quantities were written in words rather than in symbols. In fact, all procedures were described verbally. This is nicely illustrated by the following typical problem (recognizable as the modern method of completing the square):

What must be the square which, when increased by 10 of its own roots, amounts to 39? The solution is this: You halve the number of roots, which in the present instance yields 5. This you multiply by itself; the product is 25. Add this to 39; the sum is 64. Now take the root of this, which is 8, and subtract from it half the number of the roots, which is 5; the remainder is 3. This is the root of the square which you sought.

      In the second part of his book, al-Khwārizmī used propositions taken from Book II of Euclid's Elements in order to provide geometric justifications for his procedures. As remarked above, in their original context these were purely geometric propositions. Al-Khwārizmī directly connected them for the first time, however, to the solution of quadratic equations. His method was a hallmark of the Islamic approach to solving equations—systematize all cases and then provide a geometric justification, based on Greek sources. Typical of this approach was the Persian mathematician and poet Omar Khayyam's Risālah fiʾl-barāhīn ʿalā masāʾil al-jabr waʾl-muqābalah (c. 1070; “Treatise on Demonstration of Problems of Algebra”), in which Greek knowledge concerning conic sections (conic section) (ellipses (ellipse), parabolas (parabola), and hyperbolas (hyperbola)) was applied to questions involving cubic equations.

      The use of Greek-style geometric arguments in this context also led to a gradual loosening of certain traditional Greek constraints. In particular, Islamic mathematics allowed, and indeed encouraged, the unrestricted combination of commensurable and incommensurable magnitudes within the same framework, as well as the simultaneous manipulation of magnitudes of different dimensions as part of the solution of a problem. For example, the Egyptian mathematician Abu Kāmil (c. 850–930) treated the solution of a quadratic equation as a number rather than as a line segment or an area. Combined with the decimal system, this approach was fundamental in developing a more abstract and general conception of number, which was essential for the eventual creation of a full-fledged abstract idea of an equation.

Commerce and abacists in the European Renaissance
      Greek and Islamic mathematics were basically “academic” enterprises, having little interaction with day-to-day matters involving building, transportation, and commerce. This situation first began to change in Italy in the 13th and 14th centuries. In particular, the rise of Italian mercantile companies and their use of modern financial instruments for trade with the East, such as letters of credit, bills of exchange, promissory notes, and interest calculations, led to a need for improved methods of bookkeeping.

       Leonardo Pisano, known to history as Fibonacci, studied the works of Kāmil and other Arabic mathematicians as a boy while accompanying his father's trade mission to North Africa on behalf of the merchants of Pisa. In 1202, soon after his return to Italy, Fibonacci wrote Liber Abbaci (“Book of the Abacus”). Although it contained no specific innovations, and although it strictly followed the Islamic tradition of formulating and solving problems in purely rhetorical fashion, it was instrumental in communicating the Hindu-Arabic numerals to a wider audience in the Latin world. Early adopters of the “new” numerals became known as abacists, regardless of whether they used the numerals for calculating and recording transactions or employed an abacus for doing the actual calculations. Soon numerous abacist schools sprang up to teach the sons of Italian merchants the “new math.”

      The abacists first began to introduce abbreviations for unknowns in the 14th century—another important milestone toward the full-fledged manipulation of abstract symbols. For instance, c stood for cossa (“thing”), ce for censo (“square”), cu for cubo (“cube”), and R for Radice (“root”). Even combinations of these symbols were introduced for obtaining higher powers. This trend eventually led to works such as the first French algebra text, Nicolas Chuquet's Triparty en la science des nombres (1484; “The Science of Numbers in Three Parts”). As part of a discussion on how to use the Hindu-Arabic numerals, Triparty contained relatively complicated symbolic expressions, such as

R214pR2180
(meaning: ).

      Chuquet also introduced a more flexible way of denoting powers of the unknown—i.e., 122 (for 12 squares) and even m12m (to indicate −12x−2). This was, in fact, the first time that negative numbers were explicitly used in European mathematics. Chuquet could now write an equation as follows:

.3.2p.12 egaulx a .9.1
(meaning: 3x2 + 12 = 9x).

      Following the ancient tradition, coefficients were always positive, and thus the above was only one of several possible equations involving an unknown and squares of it. Indeed, Chuquet would say that the above was an impossible equation, since its solution would involve the square root of −63. This illustrates the difficulties involved in reaching a more general and flexible concept of number: the same mathematician would allow negative numbers in a certain context and even introduce a useful notation for dealing with them, but he would completely avoid their use in a different, albeit closely connected, context.

      In the 15th century, the German-speaking countries developed their own version of the abacist tradition: the Cossists, including mathematicians such as Michal Stiffel, Johannes Scheubel, and Christoff Rudolff. There one finds the first use of specific symbols for the arithmetic operations, equality, roots, and so forth. The subsequent process of standardizing symbols was, nevertheless, lengthy and involved.

Cardano and the solving of cubic and quartic equations
      Girolamo Cardano (Cardano, Girolamo) was a famous Italian physician, an avid gambler, and a prolific writer with a lifelong interest in mathematics. His widely read Ars Magna (1545; “Great Work”) contains the Renaissance era's most systematic and comprehensive account of solving cubic and quartic equations. Cardano's presentation followed the Islamic tradition of solving one instance of every possible case and then giving geometric justifications for his procedures, based on propositions from Euclid's Elements. He also followed the Islamic tradition of expressing all coefficients as positive numbers, and his presentation was fully rhetorical, with no real symbolic manipulation. Nevertheless, he did expand the use of symbols as a kind of shorthand for stating problems and describing solutions. Thus, the Greek geometric perspective still dominated—for instance, the solution of an equation was always a line segment, and the cube was the cube built on such a segment. Still, Cardano could write a cubic equation to be solved as

cup p: 6 reb aequalis 20
(meaning: x3 + 6x = 20) and present the solution as
R.V: cu.R. 108 p: 10 m: R.V: cu. R. 108m: 10,
meaning
 x = .

      Because Cardano refused to view negative numbers as possible coefficients in equations, he could not develop a notion of a general third-degree equation. This meant that he had to consider 13 “different” third-degree equations. Similarly, he considered 20 different cases for fourth-degree equations, following procedures developed by his student Ludovico Ferrari. However, Cardano was sometimes willing to consider the possibility of negative (or “false”) solutions. This allowed him to formulate some general rules, such as that in an equation with three real roots (including even negative roots), the sum of the roots must, except for sign, equal the coefficient of the square's term.

      In spite of his basic acceptance of traditional views on numbers, the solution of certain problems led Cardano to consider more radical ideas. For instance, he demonstrated that 10 could be divided into two parts whose product was 40. The answer, 5 +  √(−15) and 5 −  √(−15) , however, required the use of imaginary, or complex numbers (complex number), that is, numbers involving the square root of a negative number. Such a solution made Cardano uneasy, but he finally accepted it, declaring it to be “as refined as it is useless.”

      The first serious and systematic treatment of complex numbers had to await the Italian mathematician Rafael Bombelli, particularly the first three volumes of his unfinished L'Algebra (1572). Nevertheless, the notion of a number whose square is a negative number left most mathematicians uncomfortable. Where, exactly, in nature could one point to the existence of a negative or imaginary quantity? Thus the acceptance of numbers beyond the positive rational numbers was slow and reluctant.

Viète and the formal equation
      It is in the work of the French mathematician François Viète (Viète, François, Seigneur De La Bigotiere) that the first consistent, coherent, and systematic conception of an algebraic equation in the modern sense appeared. A main innovation of Viète's In artem analyticam isagoge (1591; “Introduction to the Analytic Art”) was its use of well-chosen symbols of one kind (vowels) for unknowns and of another kind (consonants) for known quantities. This allowed not only flexibility and generality in solving linear and quadratic equations but also something absent from all his predecessors' work, namely, a clear analysis of the relationship between the forms of the solutions and the values of the coefficients of the original equation. Viète saw his contribution as developing a “systematic way of thinking” leading to general solutions, rather than just a “bag of tricks” to solve specific problems.

      By combining existing usage with his own innovations, Viète was able to formulate equations clearly and to provide rules for transposing factors from one side of an equation to the other in order to find solutions. An example of an equation would be:

A cubus + C plano in A aequatus D solido
(meaning: x3 + cx = d).

      Note that each of the terms involved was one-dimensional, that is, after canceling powers, the remaining terms on each side of the equation are to the first power. Thus, on the left-hand side, the two-dimensional magnitude Z plano (a square) was divided by the one-dimensional variable G, leaving one dimension. On the right-hand side, a sum of two three-dimensional magnitudes (a third power) was divided by a product of two one-dimensional variables (which make a square), leaving one dimension. Thus, Viète did not break the important Greek tradition whereby the terms equated must always be of the same dimension. Nevertheless, for the first time it became possible, in the framework of an equation, to multiply or divide both sides by a certain magnitude. The result was a new equation, homogeneous in itself yet not homogeneous with the original one.

      Viète showed how to transform given equations into others, already known. For example, in modern notation, he could transform x3 + ax2 = b2x into x2 + ax = b2. He thus reduced the number of cases of cubic equations from the 13 given by Cardano and Bombelli. Nevertheless, since he still did not use negative or zero coefficients, he could not reduce all the possible cases to just one.

      Viète applied his methods to solve, in a general, abstract-symbolic fashion, problems similar to those in the Diophantine tradition. However, very often he also rephrased his answers in plain words—as if to reassure his contemporaries, and perhaps even himself, of the validity of his new methods.

The concept of numbers
      The work of Viète, described above, contained a clear, systematic, and coherent conception of the notion of equation that served as a broadly accepted starting point for later developments. No similar single reference point exists for the general conception of number, however. Some significant milestones may nevertheless be mentioned, and prominent among them was De Thiende (Disme: The Art of Tenths), an influential booklet published in 1585 by the Flemish mathematician Simon Stevin (Stevin, Simon). De Thiende was intended as a practical manual aimed at teaching the essentials of operating with decimal fractions, but it also contained many conceptual innovations. It was the first mathematical text where the all-important distinction between number and magnitude, going back to the ancient Greeks, was explicitly and totally abolished. Likewise, Stevin declared that 1 is a number just like any other and that the root of a number is a number as well. Stevin also showed how one single idea of number, expressed as decimal fractions, could be used equally in such separate contexts as land surveying, volume measurement, and astronomical and financial computations. The very need for an explanation of this kind illuminates how far Stevin's contemporaries and predecessors were from the modern notion of numbers.

      Indeed, throughout the 17th century, lively debates continued among mathematicians over the legitimacy of using various numbers. For example, concerning the irrationals, some prominent mathematicians, such as the Frenchman Blaise Pascal (Pascal, Blaise) and the Britons Isaac Barrow (Barrow, Isaac) and Isaac Newton (Newton, Sir Isaac), were willing only to grant them legitimacy as geometric magnitudes. The negative numbers were sometimes seen as even more problematic, and in many cases negative solutions of equations were still considered by many to be “absurd” or “devoid of interest.” Finally, the complex numbers were still ignored by many mathematicians, even though Bombelli had given precise rules for working with them.

      All these discussions dwindled away as the 18th century approached. A new phase in the development of the concept of number began, involving a systematization and search for adequate foundations for the various systems. This new phase is described in the next section of this article.

Classical algebra
      François Viète's work at the close of the 16th century, described in the section Viète and the formal equation (algebra), marks the start of the classical discipline of algebra. Further developments included several related trends, among which the following deserve special mention: the quest for systematic solutions of higher order equations, including approximation techniques; the rise of polynomials and their study as autonomous mathematical entities; and the increased adoption of the algebraic perspective in other mathematical disciplines, such as geometry, analysis, and logic. During this same period, new mathematical objects arose that eventually replaced polynomials as the main focus of algebraic study.

      The creation of what came to be known as analytic geometry can be attributed to two great 17th-century French thinkers: Pierre de Fermat (Fermat, Pierre de) and René Descartes (Descartes, René). Using algebraic techniques developed by Viète and Girolamo Cardano, as described earlier in this article, Fermat and Descartes tackled geometric problems that had remained unsolved since the time of the classical Greeks. The new kind of organic connection that they established between algebra and geometry was a major breakthrough, without which the subsequent development of mathematics in general, and geometry and calculus in particular, would be unthinkable.

 In his famous book La Géométrie (1637), Descartes established equivalences between algebraic operations and geometric constructions. In order to do so, he introduced a unit length that served as a reference for all other lengths and for all operations among them. For example, suppose that Descartes was given a segment AB and was asked to find its square root. He would draw the straight line DB (see the figure—>), where DA was defined as the unit length. Then he would bisect DB at C, draw the semicircle on the diameter DB with centre C, and finally draw the perpendicular from A to E on the semicircle. Elementary properties of the circle imply that ∠DEB = 90 °, which in turn implies that ∠ADE = ∠AEB and ∠DEA = ∠EBA. Thus, △DEA is similar to △EBA, or in other words, the ratio of corresponding sides is equal. Substituting x, 1, and y for AB, DA, and AE, respectively, one obtains x/y = y/1. Simplifying, x = y2, or y is the square root of x. Thus, in what might appear to be an ordinary application of classical Greek techniques, Descartes demonstrated that he could find the square root of any given number, as represented by a line segment. The key step in his construction was the introduction of the unit length DA. This seemingly trivial move, or anything similar to it, had never been done before, and it had enormous repercussions for what could thereafter be done by applying algebraic reasoning to geometry.

      Descartes also introduced a notation that allowed great flexibility in symbolic manipulation. For instance, he would write

to denote the cubic root of this algebraic expression. This was a direct continuation (with some improvement) of techniques and notations introduced by Viète. Descartes also introduced a new idea with truly far-reaching consequences when he explicitly eliminated the demand for homogeneity among the terms in an equation—although for convenience he tried to stick to homogeneity wherever possible.

      Descartes's program was based on the idea that certain geometric loci (straight lines, circles, and conic sections) could be characterized in terms of specific kinds of equations involving magnitudes that were taken to represent line segments. However, he did not envision the equally important, reciprocal idea of finding the curve that corresponded to an arbitrary algebraic expression. Descartes was aware that much information about the properties of a curve—such as its tangents and enclosed areas—could be derived from its equation, but he did not elaborate.

      On the other hand, Descartes was the first to discuss separately and systematically the algebraic properties of polynomial equations. This included his observations on the correspondence between the degree of an equation and the number of its roots, the factorization of a polynomial with known roots into linear factors, the rule for counting the number of positive and negative roots of an equation, and the method for obtaining a new equation whose roots were equal to those of a given equation, though increased or diminished by a given quantity.

      Descartes's work was the start of the transformation of polynomials into an autonomous object of intrinsic mathematical interest. To a large extent, algebra became identified with the theory of polynomials. A clear notion of a polynomial equation, together with existing techniques for solving some of them, allowed coherent and systematic reformulations of many questions that had previously been dealt with in a haphazard fashion. High on the agenda remained the problem of finding general algebraic solutions for equations of degree higher than four. Closely related to this was the question of the kinds of numbers that should count as legitimate solutions, or roots, of equations. Attempts to deal with these two important problems forced mathematicians to realize the centrality of another pressing question, namely, the number of solutions for a given polynomial equation.

      The answer to this question is given by the fundamental theorem of algebra, first suggested by the French-born mathematician Albert Girard in 1629, and which asserts that every polynomial with real number coefficients could be expressed as the product of linear and quadratic real number factors or, alternatively, that every polynomial equation of degree n with complex coefficients had n complex roots. For example, x3 + 2x2 − x − 2 can be decomposed into the quadratic factor x2 − 1 and the linear factor x + 2, that is, x3 + 2x2 − x − 2 = (x2-1)(x+2). The mathematical beauty of having n solutions for n-degree equations overcame most of the remaining reluctance to consider complex numbers as legitimate.

      Although every single polynomial equation had been shown to satisfy the theorem, the essence of mathematics since the time of the ancient Greeks has been to establish universal principles. Therefore, leading mathematicians throughout the 18th century sought the honour of being the first to prove the theorem. The flaws in their proofs were generally related to the lack of rigorous foundations for polynomials and the various number systems. Indeed, the process of criticism and revision that accompanied successive attempts to formulate and prove some correct version of the theorem contributed to a deeper understanding of both.

      The first complete proof of the theorem was given by the German mathematician Carl Friedrich Gauss (Gauss, Carl Friedrich) in his doctoral dissertation of 1799. Subsequently, Gauss provided three additional proofs. A remarkable feature of all these proofs was that they were based on methods and ideas from calculus and geometry, rather than algebra. The theorem was fundamental in that it established the most basic concept around which the discipline as a whole was built. The theorem was also fundamental from the historical point of view, since it contributed to the consolidation of the discipline, its main tools, and its main concepts.

Impasse with radical methods
      A major breakthrough in the algebraic solution of higher-degree equations was achieved by the Italian-French mathematician Joseph-Louis Lagrange (Lagrange, Joseph-Louis, comte de l'Empire) in 1770. Rather than trying to find a general solution for quintic equations directly, Lagrange attempted to clarify first why all attempts to do so had failed by investigating the known solutions of third- and fourth-degree equations. In particular, he noticed how certain algebraic expressions connected with those solutions remained invariant when the coefficients of the equations were permuted (exchanged) with one another. Lagrange was certain that a deeper analysis of this invariance would provide the key to extending existing solutions to higher-degree equations.

      Using ideas developed by Lagrange, in 1799 the Italian mathematician Paolo Ruffini (Ruffini, Paolo) was the first to assert the impossibility of obtaining a radical solution for general equations beyond the fourth degree. He adumbrated in his work the notion of a group of permutations (permutations and combinations) of the roots of an equation and worked out some basic properties. Ruffini's proofs, however, contained several significant gaps.

      Between 1796 and 1801, in the framework of his seminal number-theoretical investigations, Gauss systematically dealt with cyclotomic equations: xp − 1 = 0 (p > 2 and prime). Although his new methods did not solve the general case, Gauss did demonstrate solutions for these particular higher-degree equations.

      In 1824 the Norwegian mathematician Niels Henrik Abel (Abel, Niels Henrik) provided the first valid proof of the impossibility of obtaining radical solutions for general equations beyond the fourth degree. However, this did not end polynomial research; rather, it opened an entirely new field of research since, as Gauss's example showed, some equations were indeed solvable. In 1828 Abel suggested two main points for research in this regard: to find all equations of a given degree solvable by radicals, and to decide if a given equation can be solved by radicals. His early death in complete poverty, two days before receiving an announcement that he had been appointed professor in Berlin, prevented Abel from undertaking this program.

Galois theory
      Rather than establishing whether specific equations can or cannot be solved by radicals, as Abel had suggested, the French mathematician Évariste Galois (Galois, Évariste) (1811–32) pursued the somewhat more general problem of defining necessary and sufficient conditions for the solvability of any given equation. Although Galois's life was short and exceptionally turbulent—he was arrested several times for supporting Republican causes, and he died the day before his 21st birthday from wounds incurred in a duel—his work reshaped the discipline of algebra.

Galois's work on permutations
      Prominent among Galois's seminal ideas was the clear realization of how to formulate precise solvability conditions for a polynomial in terms of the properties of its group of permutations (permutations and combinations). A permutation of a set, say the elements a, b, and c, is any re-ordering of the elements, and it is usually denoted as follows:

This particular permutation takes a to c, b to a, and c to b. For three elements, as here, there are six different possible permutations. In general, for n elements there are n! permutations to choose from. (Where n! = n(n − 1)(n − 2)⋯2∙1.) Furthermore, two permutations can be combined to produce a third permutation in an operation known as composition. (The set of permutations are closed under the operation of composition.) For example,

      Here a goes first to c (in the first permutation) and then from c to b (in the second permutation), which is equivalent to a going directly to b, as given by the permutation to the right of the equation. Composition is associative—given three permutations P, Q, and R, then (P * Q) * R = P * (Q * R). Also, there exists an identity permutation that leaves the elements unchanged:

Finally, for each permutation there exists another permutation, known as its inverse, such that their composition results in the identity permutation. The set of permutations for n elements is known as the symmetric group Sn.

      The concept of an abstract group developed somewhat later. It consisted of a set of abstract elements with an operation defined on them such that the conditions given above were satisfied: closure, associativity, an identity element, and an inverse element for each element in the set.

      This abstract notion is not fully present in Galois's work. Like some of his predecessors, Galois focused on the permutation group of the roots of an equation. Through some beautiful and highly original mathematical ideas, Galois showed that a general polynomial equation was solvable by radicals if and only if its associated symmetric group was “soluble.” Galois's result, it must be stressed, referred to conditions for a solution to exist; it did not provide a way to calculate radical solutions in those cases where they existed.

Acceptance of Galois theory
      Galois's work was both the culmination of a main line of algebra—solving equations by radical methods—and the beginning of a new line—the study of abstract structures. Work on permutations, started by Lagrange and Ruffini, received further impetus in 1815 from the leading French mathematician, Augustin-Louis Cauchy (Cauchy, Augustin-Louis, Baron). In a later work of 1844, Cauchy systematized much of this knowledge and introduced basic concepts. For instance, the permutation

was denoted by Cauchy in cycle notation as (ab)(ced), meaning that the permutation was obtained by the disjoint cycles a to b (and back to a) and c to e to d (and back to c).

      A series of unusual and unfortunate events involving the most important contemporary French mathematicians prevented Galois's ideas from being published for a long time. It was not until 1846 that Joseph Liouville (Liouville, Joseph) edited and published for the first time, in his prestigious Journal de Mathématiques Pures et Appliquées, the important memoire in which Galois had presented his main ideas and that the Paris Academy had turned down in 1831. In Germany, Leopold Kronecker (Kronecker, Leopold) applied some of these ideas to number theory in 1853, and Richard Dedekind (Dedekind, Richard) lectured on Galois theory in 1856. At this time, however, the impact of the theory was still minimal.

      A major turning point came with the publication of Traité des substitutions et des équations algebriques (1870; “Treatise on Substitutions and Algebraic Equations”) by the French mathematician Camille Jordan (Jordan, Camille). In his book and papers, Jordan elaborated an abstract theory of permutation groups, with algebraic equations merely serving as an illustrative application of the theory. In particular, Jordan's treatise was the first group theory book and it served as the foundation for the conception of Galois theory as the study of the interconnections between extensions of fields and the related Galois groups of equations—a conception that proved fundamental for developing a completely new abstract approach to algebra in the 1920s. Major contributions to the development of this point of view for Galois theory came variously from Enrico Betti (Betti, Enrico) (1823–92) in Italy and from Dedekind, Henrich Weber (1842–1913), and Emil Artin (Artin, Emil) (1898–1962) in Germany.

Applications of group theory
      Galois theory arose in direct connection with the study of polynomials, and thus the notion of a group developed from within the mainstream of classical algebra. However, it also found important applications in other mathematical disciplines throughout the 19th century, particularly geometry and number theory.

      In 1872 Felix Klein (Klein, Felix) suggested in his inaugural lecture at the University of Erlangen, Germany, that group theoretical ideas might be fruitfully put to use in the context of geometry. Since the beginning of the 19th century, the study of projective geometry had attained renewed impetus, and later on non-Euclidean geometries (non-Euclidean geometry) were introduced and increasingly investigated. This proliferation of geometries raised pressing questions concerning both the interrelations among them and their relationship with the empirical world. Klein suggested that these geometries could be classified and ordered within a conceptual hierarchy. For instance, projective geometry seemed particularly fundamental because its properties were also relevant in Euclidean geometry, while the main concepts of the latter, such as length and angle, had no significance in the former.

      A geometric hierarchy may be expressed in terms of which transformations leave the most relevant properties of a particular geometry unchanged. It turned out that these sets of transformations were best understood as forming a group. Klein's idea was that the hierarchy of geometries might be reflected in a hierarchy of groups whose properties would be easier to understand. An example from Euclidean geometry illustrates the basic idea. The set of rotations in the plane has closure: if rotation I rotates a figure by an angle α, and rotation J by an angle β, then rotation I*J rotates it by an angle α + β. The rotation operation is obviously associative, α + (β + γ) = (α + β) + γ. The identity element is the rotation through an angle of 0 degrees, and the inverse of the rotation through angle α is the angle −α. Thus the set of rotations of the plane is a group of invariant transformations for Euclidean geometry. The groups associated with other kinds of geometries is somewhat more involved, but the idea remains the same.

      In the 1880s and '90s, Klein's friend, the Norwegian Sophus Lie (Lie, Sophus), undertook the enormous task of classifying all possible continuous groups of geometric transformations, a task that eventually evolved into the modern theory of Lie groups and Lie algebras. At roughly the same time, the French mathematician Henri Poincaré (Poincaré, Henri) studied the groups of motions of rigid bodies, a work that helped to establish group theory as one of the main tools in modern geometry.

      The notion of a group also started to appear prominently in number theory in the 19th century, especially in Gauss's work on modular arithmetic. In this context, he proved results that were later reformulated in the abstract theory of groups—for instance (in modern terms), that in a cyclic group (all elements generated by repeating the group operation on one element) there always exists a subgroup of every order (number of elements) dividing the order of the group.

      In 1854 Arthur Cayley (Cayley, Arthur), one of the most prominent British mathematicians of his time, was the first explicitly to realize that a group could be defined abstractly—without any reference to the nature of its elements and only by specifying the properties of the operation defined on them. Generalizing on Galois's ideas, Cayley took a set of meaningless symbols 1, α, β,… with an operation defined on them as shown in the table below.

Cayley demanded only that the operation be closed with respect to the elements on which it was defined, while he assumed implicitly that it was associative and that each element had an inverse. He correctly deduced some basic properties of the group, such as that if the group has n elements, then θn = 1 for each element θ. Nevertheless, in 1854 the idea of permutation groups was rather new, and Cayley's work had little immediate impact.

Fundamental concepts of modern algebra
Prime factorization
      Some other fundamental concepts of modern algebra also had their origin in 19th-century work on number theory, particularly in connection with attempts to generalize the theorem of (unique) prime factorization (fundamental theorem of arithmetic) beyond the natural numbers. This theorem asserted that every natural number could be written as a product of its prime factors in a unique way, except perhaps for order (e.g., 24 = 2∙2∙2∙3). This property of the natural numbers was known, at least implicitly, since the time of Euclid. In the 19th century, mathematicians sought to extend some version of this theorem to the complex numbers (complex number).

      One should not be surprised, then, to find the name of Gauss (Gauss, Carl Friedrich) in this context. In his classical investigations on arithmetic Gauss was led to the factorization properties of numbers of the type a + ib (a and b integers and i =  √(−1) ), sometimes called Gaussian integers. In doing so, Gauss not only used complex numbers to solve a problem involving ordinary integers, a fact remarkable in itself, but he also opened the way to the detailed investigation of special subdomains of the complex numbers.

      In 1832 Gauss proved that the Gaussian integers satisfied a generalized version of the factorization theorem where the prime factors had to be especially defined in this domain. In the 1840s the German mathematician Ernst Eduard Kummer (Kummer, Ernst Eduard) extended these results to other, even more general domains of complex numbers, such as numbers of the form a + θb, where θ2 = n for n a fixed integer, or numbers of the form a + ρb, where ρn = 1, ρ ≠ 1, and n > 2. Although Kummer did prove interesting results, it finally turned out that the prime factorization theorem was not valid in such general domains. The following example illustrates the problem.

      Consider the domain of numbers of the form a + b √(−5) and, in particular, the number 21 = 21 + 0 √(−5) . 21 can be factored as both 3∙7 and as (4 +  √(−5) )(4 −  √(−5) ). It can be shown that none of the numbers 3, 7, 4 ± √(−5) could be further decomposed as a product of two different numbers in this domain. Thus, in one sense they were prime. However, at the same time they violated a property of prime numbers known from the time of Euclid: if a prime number p divides a product ab, then it either divides a or b. In this instance, 3 divides 21 but neither of the factors 4 +  √(−5) or 4 −  √(−5) .

      This situation led to the concept of indecomposable numbers. In classical arithmetic any indecomposable number is a prime (and vice versa), but in more general domains a number may be indecomposable, such as 3 here, yet not prime in the earlier sense. The question thus remained open which domains the prime factorization theorem was valid in and how properly to formulate a generalized version of it. This problem was undertaken by Dedekind (Dedekind, Richard) in a series of works spanning over 30 years, starting in 1871. Dedekind's general methodological approach promoted the introduction of new concepts around which entire theories could be built. Specific problems were then solved as instances of the general theory.

      A main question pursued by Dedekind was the precise identification of those subsets of the complex numbers for which some generalized version of the theorem made sense. The first step toward answering this question was the concept of a field, defined as any subset of the complex numbers that was closed under the four basic arithmetic operations (except division by zero). The largest of these fields was the whole system of complex numbers, whereas the smallest field was the rational numbers. Using the concept of field and some other derivative ideas, Dedekind identified the precise subset of the complex numbers for which the theorem could be extended. He named that subset the algebraic integers.

Ideals
      Finally, Dedekind introduced the concept of an ideal. A main methodological trait of Dedekind's innovative approach to algebra was to translate ordinary arithmetic properties into properties of sets of numbers. In this case, he focused on the set I of multiples of any given integer and pointed out two of its main properties:
● If n and m are two numbers in I, then their difference is also in I.
● If n is a number in I and a is any integer, then their product is also in I.

      As he did in many other contexts, Dedekind took these properties and turned them into definitions. He defined a collection of algebraic integers that satisfied these properties as an ideal in the complex numbers. This was the concept that allowed him to generalize the prime factorization theorem in distinctly set-theoretical terms.

      In ordinary arithmetic, the ideal generated by the product of two numbers equals the intersection of the ideals generated by each of them. For instance, the set of multiples of 6 (the ideal generated by 6) is the intersection of the ideal generated by 2 and the ideal generated by 3. Dedekind's generalized versions of the theorem were phrased precisely in these terms for general fields of complex numbers and their related ideals. He distinguished among different types of ideals and different types of decompositions, but the generalizations were all-inclusive and precise. More important, he reformulated what were originally results on numbers, their factors, and their products as far more general and abstract results on special domains, special subsets of numbers, and their intersections.

      Dedekind's results were important not only for a deeper understanding of factorization. He also introduced the set-theoretical approach into algebraic research, and he defined some of the most basic concepts of modern algebra that became the main focus of algebraic research throughout the 20th century. Moreover, Dedekind's ideal-theoretical approach was soon successfully applied to the factorization of polynomials as well, thus connecting itself once again to the main focus of classical algebra.

Systems of equations
      In spite of the many novel algebraic ideas that arose in the 19th century, solving equations and studying properties of polynomial forms continued to be the main focus of algebra. The study of systems of equations led to the notion of a determinant and to matrix theory.

Determinants (determinant)
      Given a system of n linear equations in n unknowns, its determinant was defined as the result of a certain combination of multiplication and addition of the coefficients of the equations that allowed the values of the unknowns to be calculated directly. For example, given the system

a1x + b1y = c1
a2x + b2y = c2
the determinant Δ of the system is the number Δ = a1b2 − a2b1, and the values of the unknowns are given by
x = (c1b2 − c2b1)/Δ
y = (a1c2 − a2c1)/Δ.

      Historians agree that the 17th-century Japanese mathematician Seki Kōwa (Seki Takakazu) was the earliest to use methods of this kind systematically. In Europe, credit is usually given to his contemporary, the German coinventor of calculus, Gottfried Wilhelm Leibniz (Leibniz, Gottfried Wilhelm).

      In 1815 Cauchy (Cauchy, Augustin-Louis, Baron) published the first truly systematic and comprehensive study of determinants, and he was the one who coined the name. He introduced the notation (al, n) for the system of coefficients of the system and demonstrated a general method for calculating the determinant.

      Closely related to the concept of a determinant was the idea of a matrix as an arrangement of numbers in lines and columns. That such an arrangement could be taken as an autonomous mathematical object, subject to special rules that allow for manipulation like ordinary numbers, was first conceived in the 1850s by Cayley (Cayley, Arthur) and his good friend the attorney and mathematician James Joseph Sylvester (Sylvester, James Joseph). Determinants were a main, direct source for this idea, but so were ideas contained in previous work on number theory by Gauss and by the German mathematician Ferdinand Gotthold Max Eisenstein (Eisenstein, Ferdinand Gotthold Max) (1823–52).

      Given a system of linear equations:

ξ = αx + βy + γz + …
η = α′x + β′y + γ′z + …
ζ = α″x + β″y + γ″z + …
… = … + … + … + …
Cayley represented it with a matrix as follows:

      The solution could then be written as:

The matrix bearing the −1 exponent was called the inverse matrix, and it held the key to solving the original system of equations. Cayley showed how to obtain the inverse matrix using the determinant of the original matrix. Once this matrix is calculated, the arithmetic of matrices allowed him to solve the system of equations by a simple analogy with linear equations: ax = b {long right arrow} x = a−1b.

      Cayley was joined by other mathematicians, such as the Irish William Rowan Hamilton (Hamilton, Sir William Rowan), the German Georg Frobenius (Frobenius, Georg), and Jordan, in developing the theory of matrices, which soon became a fundamental tool in analysis, geometry, and especially in the emerging discipline of linear algebra (algebra, linear). A further important point was that matrices enlarged the range of algebraic notions. In particular, matrices embodied a new, mathematically significant instance of a system with a well-elaborated arithmetic, whose rules departed from traditional number systems in the important sense that multiplication was not generally commutative.

      In fact, matrix theory was naturally connected after 1830 with a central trend in British mathematics developed by George Peacock and Augustus De Morgan (De Morgan, Augustus), among others. In trying to overcome the last reservations about the legitimacy of the negative and complex numbers, these mathematicians suggested that algebra be conceived as a purely formal, symbolic language, irrespective of the nature of the objects whose laws of combination it stipulated. In principle, this view allowed for new, different kinds of arithmetic, such as matrix arithmetic. The British tradition of symbolic algebra was instrumental in shifting the focus of algebra from the direct study of objects (numbers, polynomials, and the like) to the study of operations among abstract objects. Still, in most respects, Peacock and De Morgan strove to gain a deeper understanding of the objects of classical algebra rather than to launch a new discipline.

      Another important development in Britain concerned the elaboration of an algebra of logic. De Morgan and George Boole (Boole, George), and somewhat later Ernst Schröder in Germany, were instrumental in transforming logic from a purely metaphysical into a mathematical discipline. They also added to the growing realization of the immense potential of algebraic thinking, freed from its narrow conception as the discipline of polynomial equations and number systems.

Quaternions and vectors
      Remaining doubts about the legitimacy of complex numbers were finally dispelled when their geometric interpretation became widespread among mathematicians. This interpretation, initially and independently conceived by the Norwegian surveyor Caspar Wessel and the French bookkeeper Jean-Robert Argand (see Argand diagram), was made known to a larger audience of mathematicians mainly through its explicit use by Gauss in his 1848 proof of the fundamental theorem of algebra. Under this interpretation, every complex number appeared as a directed segment on the plane, characterized by its length and its angle of inclination with respect to the x-axis. The number i thus corresponded to the segment of length 1 that was perpendicular to the x-axis. Once a proper arithmetic was defined on these numbers, it turned out that i2 = −1, as expected.

      An alternative interpretation, very much within the spirit of the British school of symbolic algebra, was published in 1837 by Hamilton (Hamilton, Sir William Rowan). Hamilton defined a complex number a + bi as a pair (ab) of real numbers and gave the laws of arithmetic for such pairs. For example, he defined multiplication as:

(ab)(cd) = (ac − bdbc + ad).

      In Hamilton's notation i = (0, 1) and by the above definition of complex multiplication (0, 1)(0, 1) = (−1, 0)—that is, i2 = −1 as desired. This formal interpretation obviated the need to give any essentialist definition of complex numbers.

      Starting in 1830, Hamilton pursued intensely, and unsuccessfully, a scheme to extend his idea to triplets (abc), which he expected to be of great utility in mathematical physics. His difficulty lay in defining a consistent multiplication for such a system, which in hindsight is known to be impossible. In 1843 Hamilton finally realized that the generalization he was looking for had to be found in the system of quadruplets (abcd), which he named quaternions. He wrote them, in analogy with the complex numbers, as a + bi + cj + dk, and his new arithmetic was based on the rules: i2 = j2 = k2 = ijk = −1 and ij = k, ji = −k, jk = i, kj = −i, ki = j, and ik = −j. This was the first example of a coherent, significant mathematical system that preserved all of the laws of ordinary arithmetic, with the exception of commutativity.

      In spite of Hamilton's initial hopes, quaternions never really caught on among physicists, who generally preferred vector notation when it was introduced later. Nevertheless, his ideas had an enormous influence on the gradual introduction and use of vectors in physics. Hamilton used the name scalar for the real part a of the quaternion, and the term vector for the imaginary part bi + cj + dk, and defined what are now known as the scalar (or dot) and vector (or cross) products. It was through successive work in the 19th century of the Britons Peter Guthrie Tait (Tait, Peter Guthrie), James Clerk Maxwell (Maxwell, James Clerk), and Oliver Heaviside (Heaviside, Oliver) and the American Josiah Willard Gibbs (Gibbs, J Willard) that an autonomous theory of vectors was first established while developing on Hamilton's initial ideas. In spite of physicists' general lack of interest in quaternions, they remained important inside mathematics, although mainly as an example of an alternate algebraic system.

The close of the classical age
      The last major algebra textbook in the classical tradition was Heinrich Weber's Lehrbuch der Algebra (1895; “Textbook of Algebra”), which codified the achievements and current dominant views of the subject and remained highly influential for several decades. At its centre was a well-elaborated, systematic conception of the various systems of numbers, built as a rigorous hierarchy from the natural numbers up to the complex numbers. Its primary focus was the study of polynomials, polynomial equations, and polynomial forms, and all relevant results and methods derived in the book directly depended on the properties of the systems of numbers. Radical methods for solving equations received a great deal of attention, but so did approximation methods, which are now typically covered instead in analysis and numerical analysis textbooks. Recently developed concepts, such as groups and fields, as well as methods derived from Galois's work, were treated in Weber's textbook, but only as useful tools to help deal with the main topic of polynomial equations.

      To a large extent, Weber's textbook was a very fine culmination of a long process that started in antiquity. Fortunately, rather than bring this process to a conclusion, it served as a catalyst for the next stage of algebra.

Structural algebra
      At the turn of the 20th century, algebra reflected a very clear conceptual hierarchy based on a systematically elaborated arithmetic, with a theory of polynomial equations built on top of it. Finally, a well-developed set of conceptual tools, most prominently the idea of groups, offered a comprehensive means of investigating algebraic properties. Then in 1930 a textbook was published that presented a totally new image of the discipline. This was Moderne Algebra, by the Dutch mathematician Bartel van der Waerden, who since 1924 had attended lectures in Germany by Emmy Noether (Noether, Emmy) at Göttingen and by Emil Artin at Hamburg. Van der Waerden's new image of the discipline inverted the conceptual hierarchy of classical algebra. Groups, fields, rings, and other related concepts became the main focus, based on the implicit realization that all of these concepts were, in fact, instances of a more general, underlying idea: the idea of an algebraic structure. Thus, the main task of algebra became the elucidation of the properties of each of these structures and of the relationships among them. Similar questions were now asked about all these concepts, and similar concepts and techniques were used where possible. The main tasks of classical algebra became ancillary. The systems of real numbers, rational numbers, and polynomials were studied as particular instances of certain algebraic structures; the properties of these systems depended on what was known about the general structures of which they were instances, rather than the other way round.

Precursors to the structural approach
      Van der Waerden's book did not contain many new results or concepts. Its innovation lay in the unified picture it presented of the discipline of algebra. Van der Waerden brought together, in a surprisingly illuminating manner, algebraic research that had taken place over the previous three decades and in doing so he combined the contributions of several leading German algebraists from the beginning of the 20th century.

Hilbert and Steinitz
      Of these German mathematicians, few were more important than David Hilbert (Hilbert, David). Among his important contributions, his work in the 1890s on the theory of algebraic number fields was decisive in establishing the conceptual approach promoted by Dedekind as dominant for several decades. As the undisputed leader of mathematics at Göttingen, then the world's premiere research institution, Hilbert's influence propagated through the 68 doctoral dissertations he directed as well as through the many students and mathematicians who attended his lectures. To a significant extent, the structural view of algebra was the product of some of Hilbert's innovations, yet he basically remained a representative of the classical discipline of algebra. It is likely that the kind of algebra that developed under the influence of van der Waerden's book had no direct appeal for Hilbert.

      In 1910 Ernst Steinitz published an influential article on the abstract theory of fields that was an important milestone on the road to the structural image of algebra. His work was highly structural in that he first established the simplest kinds of subfields that any field contains and established a classification system. He then investigated how properties were passed from a field to any extension of it or to any of its subfields. In this way, he was able to characterize all possible fields abstractly. To a great extent, van der Waerden extended to the whole discipline of algebra what Steinitz accomplished for the more restricted domain of fields.

Noether and Artin
      The greatest influence behind the consolidation of the structural image of algebra was no doubt Noether (Noether, Emmy), who became the most prominent figure in Göttingen in the 1920s. Noether synthesized the ideas of Dedekind, Hilbert, Steinitz, and others in a series of articles in which the theory of factorization of algebraic numbers and of polynomials was masterly and succinctly subsumed under a single theory of abstract rings. She also contributed important papers to the theory of hypercomplex systems (extensions, such as the quaternions, of complex numbers to higher dimensions) that followed a similar approach, further demonstrating the potential of the structural approach.

      The last significant influence on van der Waerden's structural image of algebra was by Artin (Artin, Emil), above all for the latter's reformulation of Galois theory. Rather than speaking of the Galois group of a polynomial equation with coefficients in a particular field, Artin focused on the group of automorphisms (automorphism) of the coefficients' splitting field (the smallest extension of the field such that the polynomial could be factored into linear terms). Galois theory could then be seen as the study of the interrelations between the extensions of a field and the possible subgroups of the Galois group of the original field. In this typical structural reformulation of a classical 19th-century theory of algebra, the problem of solvability of equations by radicals appeared as a particular application of an abstract general theory.

The structural approach dominates
      After the late 1930s it was clear that algebra, and in particular the structural approach within it, had become one of the most dynamic areas of research in mathematics. Structural methods, results, and concepts were actively pursued by algebraists in Germany, France, the United States, Japan, and elsewhere. The structural approach was also successfully applied to redefine other mathematical disciplines. An important early example of this was the thorough reformulation of algebraic geometry in the hands of van der Waerden, André Weil (Weil, André) in France, and the Russian-born Oscar Zariski in Italy and the United States. In particular, they used the concepts and approach developed in ring theory by Noether and her successors. Another important example was the work of the American Marshall Stone, who in the late 1930s defined Boolean algebras (Boolean algebra), bringing under a purely algebraic framework ideas stemming from logic, topology, and algebra itself.

      Over the following decades, algebra textbooks appeared around the world along the lines established by van der Waerden. Prominent among these was A Survey of Modern Algebra (1941) by Saunders Mac Lane (Mac Lane, Saunders) and Garret Birkhoff, a book that was fundamental for the next several generations of mathematicians in the United States. Neverthless, it must be stressed that not all algebraists felt, at least initially, that the new direction implied by Moderne Algebra was paramount. More classically oriented research was still being carried out well beyond the 1930s. The research of Frobenius and his former student Issai Schur, who were the most outstanding representatives of the Berlin mathematical school at the beginning of the 20th century, and of Hermann Weyl, one of Hilbert's most prominent students, merit special mention.

Algebraic superstructures
      Although the structural approach had become prominent in many mathematical disciplines, the notion of structure remained more a regulative, informal principle than a real mathematical concept for independent investigation. It was only natural that sooner or later the question would arise how to define structures in such a way that the concept could be investigated. For example, Noether brought new and important insights into certain rings (algebraic numbers and polynomials) previously investigated under separate frameworks by studying their underlying structures. Similarly, it was expected that a general metatheory of structures, or superstructures, would prove fruitful for studying other related concepts.

Bourbaki
      Attempts to develop such a metatheory were undertaken starting in the 1940s. The first one came from a group of young French mathematicians working under the common pseudonym of Nicolas Bourbaki (Bourbaki, Nicolas). The founders of the group included Weil, Jean Dieudonné (Dieudonné, Jean), and Henri Cartan (Cartan, Henri). Over the next few decades, the group published a collection of extremely influential textbooks, Eléments de mathématique, that covered several central mathematical disciplines, particularly from a structural perspective. Yet, to the extent that Bourbaki's mathematics was structural, it was so in a general, informal way. As van der Waerden extended to all of algebra the structural approach that Steinitz introduced in the theory of fields, so Bourbaki's Eléments extended this approach to a truly broad range of mathematical disciplines. Although Bourbaki did define a formal concept of structure in the first book of the collection, their concept turned out to be quite cumbersome and was not pursued further.

Category theory
      The second attempt to formalize the notion of structure developed within category theory. The first paper on the subject was published in the United States in 1942 by Mac Lane (Mac Lane, Saunders) and Samuel Eilenberg. The idea behind their approach was that the essential features of any particular mathematical domain (a category) could be identified by focusing on the interrelations among its elements, rather than looking at the behaviour of each element in isolation. For example, what characterized the category of groups were the properties of its homomorphisms (homomorphism) (mappings between groups that preserve algebraic operations) and comparisons with morphisms for other categories, such as homeomorphisms (homeomorphism) for topological spaces. Another important concept of Mac Lane and Eilenberg was their formulation of “functors,” a generalization of the idea of function that enabled them to connect different categories. For example, in algebraic topology functors associated topological spaces with certain groups such that their topological properties could be expressed as algebraic properties of the groups—a process that enabled powerful algebraic tools to be used on previously intractable problems.

      Although category theory did not become a universal language for all of mathematics, it did become the standard formulation for algebraic topology and homology. Category theory also led to new approaches in the study of the foundations of mathematics by means of Topos theory (mathematics, foundations of). Some of these developments were further enhanced between 1956 and 1970 through the intensive work of Alexandre Grothendieck (Grothendieck, Alexandre) and his collaborators in France, using still more general concepts based on categories.

New challenges and perspectives
      The enormous productivity of research in algebra over the second half of the 20th century precludes any complete synopsis. Nevertheless, two main issues deserve some comment. The first was a trend toward abstraction and generalization as embodied in the structural approach. This trend was not exclusive, however. Researchers moved back and forth, studying general structures as well as classical entities such as the real and rational numbers. The second issue was the introduction of new kinds of proofs and techniques. The following examples are illustrative.

      A subgroup H of a group G is called a normal group if for every element g in G and h in H, g−1hg is an element of H. A group with no normal subgroups is known as a simple group. Simple groups are the basic components of group theory, and since Galois's time it was known that the general quintic was unsolvable by radicals because its Galois group was simple. However, a full characterization of simple groups remained unattainable until a major breakthrough in 1963 by two Americans, Walter Feit and John G. Thomson, who proved an old conjecture of the British mathematician William Burnside, namely, that the order of noncommutative finite simple groups is always even. Their proof was long and involved, but it reinforced the belief that a full classification of finite simple groups might, after all, be possible. The completion of the task was announced in 1983 by the American mathematician Daniel Gorenstein, following the contributions of hundreds of individuals over thousands of pages. Although this classification seems comprehensive, it is anything but clear-cut and systematic, since simple groups appear in all kinds of situations and under many guises. Thus, there seems to be no single individual who can boast of knowing all of its details. This kind of very large, collective theorem is certainly a novel mathematical phenomenon.

      Another example concerns the complex and involved question of the use of computers in proving and even formulating new theorems. This now incipient trend will certainly receive increased attention in the 21st century.

      Finally, probabilistic methods of proof in algebra, and in particular for solving difficult, open problems in group theory, have been introduced. This trend began with a series of papers by the Hungarian mathematicians Paul Erdös (Erdös, Paul) and Paul Turán, both of whom introduced probabilistic methods into many other branches of mathematics as well.

Leo Corry

Additional Reading

General history
B.L. van der Waerden, A History of Algebra: From al-Khwārizmī to Emmy Noether (1985), is a highly respected classic. Two works that contain selections from original mathematical texts, including many that are directly relevant to the history of algebra, are David Eugene Smith, A Source Book in Mathematics (1929, reissued 1959); and John Fauvel and Jeremy Gray (eds.), The History of Mathematics: A Reader (1987, reissued 1990).

Ancient and Greek algebra
Among the books on ancient mathematics, including sections on algebra, the reader may consult O. Neugebauer, The Exact Sciences in Antiquity, 2nd ed. (1969, reissued 1993); Richard J. Gillings, Mathematics in the Time of the Pharaohs (1972, reprinted 1982); and Jens Hoyrup, Lengths, Widths, Surfaces: A Portrait of Old Babylonian Algebra and Its Kin (2002).The degree to which algebraic ideas do or do not appear in Greek geometric texts has been widely discussed by historians. A comprehensive summary of work on this controversial question appears in Michael N. Fried and Sabetai Unguru, Apollonius of Perga's Conica: Text, Context, Subtext (2001).

Indian and Chinese algebra
Among the few English-language books on the history of algebra in India and China, the following are recommended: C.N. Srinivasiengar, The History of Ancient Indian Mathematics (1967); and Ulrich Libbrecht, Chinese Mathematics in the Thirteenth Century: The Shu-Shu Chiu-Chang of Ch'in, Chiu-Shao (1973).

Islamic algebra
Research on Islamic mathematics has vigorously developed in recent years. Two important works are Roshdi Rashed (Rushdi Rashid), The Development of Arabic Mathematics: Between Arithmetic and Algebra (1994; originally published in French, 1984); and J.L. Berggren, Episodes in the Mathematics of Medieval Islam (1986).

Algebra in Renaissance Europe
Jacob Klein, Greek Mathematical Thought and the Origin of Algebra, trans. from German (1968, reprinted 1992), is one of the most important accounts of the evolution of the concept of number from the ancient Greeks to the 17th century. Paul Lawrence Rose, The Italian Renaissance of Mathematics: Studies on Humanists and Mathematicians from Petrarch to Galileo (1975), is a highly respected history of Renaissance mathematics.

Modern algebra
Leo Corry, Modern Algebra and the Rise of Mathematical Structures (1996), traces the emergence of the structural approach, as well as efforts to develop a metatheory of structures.Leo Corry

* * *


Universalium. 2010.

Игры ⚽ Нужно сделать НИР?
Synonyms:
(by means of letters and other symbols), ,


Look at other dictionaries:

  • Algebra — Algebra …   Deutsch Wörterbuch

  • algebră — ALGÉBRĂ s.f. 1. Teorie a operaţiilor privind numerele reale (pozitive ori negative) sau complexe şi rezolvarea ecuaţiilor prin substituirea prin litere a valorilor numerice şi a formulei generale de calcul numeric particular. ♦ Manual şcolar care …   Dicționar Român

  • Algebra — (fra Arabisk al djebr ) er en gren af matematikken der kan beskrives som en genralisering og udvidelse af aritmetikken. Man kan lave en grov inddeling af algebra i disse felter: 10 Elementær algebra hvor man ser på egenskaberne ved de reelle tal …   Danske encyklopædi

  • Algebra — Sf Lehre von den mathematischen Gleichungen (usw.) erw. fach. (15. Jh.) Entlehnung. Entlehnt aus ml. algebra, das seinerseits auf arab. al ǧabr zurückgeht. Dieses ist Teil des Titels eines Lehrbuchs des arabischen Mathematikers Al Ḫwārizmī (9. Jh …   Etymologisches Wörterbuch der deutschen sprache

  • Algĕbra — (v. arab., Math.), der Theil der Arithmetik, der sich mit der Lösung solcher Aufgaben beschäftigt, in denen der Werth einer od. mehrerer Zahlen gefunden werden soll, nachdem das Resultat bekannt ist, welches aus irgend welchen… …   Pierer's Universal-Lexikon

  • Algebra — »Buchstabenrechnung, Lehre von den mathematischen Gleichungen«: Der Fachausdruck der Mathematik wurde im 15. Jh. aus gleichbed. mlat. algebra, evtl. auch durch roman. Vermittlung (beachte entsprechend span., port. álgebra, it. algebra, frz.… …   Das Herkunftswörterbuch

  • Algebra — Al ge*bra, n. [LL. algebra, fr. Ar. al jebr reduction of parts to a whole, or fractions to whole numbers, fr. jabara to bind together, consolidate; al jebr w almuq[=a]balah reduction and comparison (by equations): cf. F. alg[ e]bre, It. & Sp.… …   The Collaborative International Dictionary of English

  • Algebra — statusas T sritis automatika atitikmenys: angl. algebra vok. Algebra, f rus. алгебра, f pranc. algèbre, f …   Automatikos terminų žodynas

  • algebra — statusas T sritis automatika atitikmenys: angl. algebra vok. Algebra, f rus. алгебра, f pranc. algèbre, f …   Automatikos terminų žodynas

  • Algebra — statusas T sritis fizika atitikmenys: angl. algebra vok. Algebra, f rus. алгебра, f pranc. algèbre, f …   Fizikos terminų žodynas

  • algebra — statusas T sritis fizika atitikmenys: angl. algebra vok. Algebra, f rus. алгебра, f pranc. algèbre, f …   Fizikos terminų žodynas

Share the article and excerpts

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