The Method of Fluxions is Isaac Newton's foundational mathematical framework for what is now known as calculus, developed during his annus mirabilis in 1665–1666, which conceptualizes continuously varying quantities as "fluents" and their instantaneous rates of change as "fluxions," enabling the solution of problems in tangents, areas, and motion.[1][2]Newton first documented the method in a manuscript dated May 28, 1665, while isolated at his family home in Woolsthorpe during the Great Plague, building on his earlier work in infinite series and geometry.[2] The core ideas revolve around "moments," which are infinitesimally small increments of fluents over time, allowing for the approximation and exact resolution of geometric and physical problems without relying on traditional algebraic methods.[1] Fluxions, denoted by a dot over the variable (e.g., ẋ for the fluxion of x), represent derivatives, while the inverse process of finding fluents corresponds to integration, applied to tasks such as quadrature (area under curves) and rectification (arc lengths).[2]Although Newton circulated drafts among contemporaries starting in 1669, the method remained unpublished for decades due to his reluctance and the priority dispute with Gottfried Wilhelm Leibniz over the invention of calculus.[1] The earliest public account appeared in 1693 through letters to John Wallis, followed by a more detailed exposition in the 1704 appendix De Quadratura Curvarum to Newton's Opticks, and a posthumous English translation of his 1671 manuscript De Methodis Fluxionum et Serierum Infinitorum in 1736 by John Colson.[2] Newton employed the method extensively in his Philosophiæ Naturalis Principia Mathematica (1687), particularly in analyzing orbital motion and forces, though he disguised it in geometric language to avoid controversy over infinitesimals.[1]The approach faced philosophical scrutiny, notably from George Berkeley in his 1734 critique The Analyst, which derided fluxions as "the ghosts of departed quantities" for their reliance on vanishingly small quantities without rigorous foundation.[1] Despite such challenges, the Method of Fluxions laid essential groundwork for modern analysis, influencing subsequent mathematicians like the Bernoulli brothers and Euler, and underscoring Newton's profound impact on mathematical physics.[2]
Historical Context
Newton's Early Mathematical Influences
During his undergraduate studies at Trinity College, Cambridge, beginning in 1661, Isaac Newton encountered key mathematical works that laid the groundwork for his later innovations in analysis. He engaged deeply with René Descartes' La Géométrie (1637), which introduced algebraic methods to solve geometric problems, emphasizing coordinate geometry and the representation of curves through equations. This exposure prompted Newton to annotate and extend Cartesian techniques, particularly in resolving loci and tangents, as evidenced by his early notebooks from the mid-1660s.[3] Complementing this, Newton studied John Wallis's Arithmetica Infinitorum (1656), a treatise on the arithmetic of infinites that interpolated areas under curves using sequences of fractions, effectively bridging finite arithmetic with infinite processes. Wallis's approach to quadratures through infinite series profoundly influenced Newton's emerging interest in continuous change, as Newton later referenced it in his own interpolations of binomial expansions.[4]Newton also drew inspiration from earlier continental methods that anticipated infinitesimal reasoning. Bonaventura Cavalieri's Geometria indivisibilibus continuorum (1635) introduced the method of indivisibles, treating plane figures as sums of infinitely many lines and solids as stacks of planes to compute areas and volumes without rigorous limits. Although Cavalieri's technique faced philosophical critiques for its ambiguity regarding indivisibles, it provided Newton with a practical tool for handling "ultimate ratios" in curved figures during his initial forays into motion and variation.[5] The Scottish mathematician James Gregory's work on infinite series, detailed in Geometriae pars universalis (1668) and earlier manuscripts, offered convergent expansions for arcs, sines, and hyperbolas, demonstrating how infinite sums could resolve transcendental problems. Gregory's series methods, circulated in academic correspondence, were analogous to expansions Newton pursued independently.[5]In 1665–1666, amid the Great Plague that closed Cambridge, Newton returned to his family home in Woolsthorpe, where he developed the generalized binomial theorem with fractional exponents, enabling infinite series expansions for various functions. He applied these techniques to approximate circular arcs and facilitate quadratures. These innovations stemmed directly from his synthesis of Wallis's interpolations and other techniques.[6] In 1669, Newton derived power series expansions for trigonometric functions such as sine and cosine, such as \sin y = y - \frac{y^3}{3!} + \frac{y^5}{5!} - \cdots. Earlier, in his "Waste Book"—a notebook of jottings from 1663 to 1669—Newton recorded preliminary ideas on moments and velocities as of October 1664, exploring instantaneous rates in projectile motion through geometric proportions akin to emerging infinitesimal methods. These entries, focusing on small increments in circular paths, reveal his nascent conceptualization of variable quantities in physical contexts.[7]This personal trajectory aligned with the 17th-century European preoccupation with motion problems, such as planetary orbits and falling bodies, which demanded tools for analyzing changing velocities.[8]
In the early 17th century, mathematicians began addressing problems of tangents, extrema, and areas under curves using algebraic and geometric techniques that foreshadowed calculus. These efforts were driven by the need to analyze planetary motion, optimize designs, and compute volumes, building on Renaissance recoveries of ancient Greek methods while incorporating emerging algebraic tools. Key figures like René Descartes and Pierre de Fermat shifted focus from static geometry to dynamic algebraic representations, setting the stage for conceptualizing change over time.[9]René Descartes introduced analytic geometry in his 1637 treatise La Géométrie, part of Discours de la méthode, where he represented curves as equations in coordinates, enabling algebraic manipulation of geometric problems. This innovation allowed for the study of tangents and intersections through solving polynomials, though Descartes himself avoided infinitesimals and focused on finite differences. His approach unified algebra and geometry, providing a framework for later rate-of-change calculations essential to Newton's fluxions.[10]Pierre de Fermat advanced tangent and extrema methods in correspondence and notes from the 1630s, using his "method of adequality" to approximate instantaneous slopes by setting an equation equal to itself with a small increment and then discarding higher-order terms. For instance, to find the tangent to y = x^2 at x = a, Fermat equated (a + e)^2 = y and eliminated the e^2 term, yielding the slope $2a. This technique, applied to optimization and area problems, prefigured differentiation without explicit limits.[9]Bonaventura Cavalieri's Geometria indivisibilibus continuorum (1635) proposed the method of indivisibles, treating plane figures as sums of infinitely many parallel lines and solids as stacks of planes, to compute areas and volumes. For example, he showed the area under a parabola equals four-thirds the inscribed triangle by comparing indivisible sums. This heuristic, refined against critics like Paul Guldin, offered a practical precursor to integration, influencing quadrature techniques that Newton later formalized.[10]Mid-century, John Wallis's Arithmetica infinitorum (1656) interpolated areas under power curves using infinite series patterns, such as expressing \int_0^1 x^{n} \, dx = \frac{1}{n+1} for integer n via recursive fractions. Wallis extended this to non-integers and derived the infinite product \frac{\pi}{2} = \prod_{n=1}^\infty \frac{(2n)^2}{(2n-1)(2n+1)}, blending arithmetic progression with geometric summation. His work on infinite processes provided Newton with tools for series expansions in fluxional computations.[9]Isaac Barrow, in Lectiones geometricae (lectures delivered 1664–1666, published 1670), geometrically linked tangent construction to area quadrature, showing that the tangent at a point relates inversely to the area accumulation rate—essentially the fundamental theorem of calculus in curve-sketching form. Barrow used inscribed and circumscribed polygons to bound areas and derived tangent rules from circle properties. As Newton's teacher at Cambridge, Barrow's synthesis directly inspired the method of fluxions, where flowing quantities (fluents) and their rates (fluxions) extended these static relations to motion.[10]
Development of the Method
Initial Formulation in the 1660s
During the Great Plague of 1665–1666, which forced the closure of Cambridge University, Isaac Newton retreated to his family estate at Woolsthorpe Manor, entering a period of intense intellectual productivity known as his annus mirabilis.[11] Isolated from formal academic life, Newton turned his attention to mathematics and natural philosophy, laying the groundwork for what would become the method of fluxions. This isolation provided the uninterrupted focus that enabled rapid advancements, including early explorations of infinite series and rates of change, culminating in the composition of key manuscripts based on ideas developed during these years.[11]A central conceptual breakthrough emerged from Newton's consideration of changing quantities, or "fluents," and their instantaneous rates of change, termed fluxions. To circumvent the philosophical issues associated with explicit infinitesimals, Newton introduced the notion of "ultimate ratios" of nascent moments— the limiting ratios of infinitesimal increments as they approach zero without ever reaching it.[11] These nascent moments represented the incipient stages of generated quantities, allowing Newton to rigorously define tangents, areas, and velocities through geometric limits rather than undefined infinities. This approach, rooted in his 1665–1666 reflections, provided a foundation for differential calculus while aligning with his commitment to mathematical certainty.[11]Newton's fluxional ideas found early application in analyzing celestial motion, particularly in deriving the inverse square law of gravitation. By 1666, using nascent principles akin to fluxions, he demonstrated that Kepler's laws of planetary motion implied a centripetal force varying inversely with the square of the distance from the sun, explaining elliptical orbits around a focus.[11] This insight connected mathematical rates of change to physical dynamics, foreshadowing the synthesis in his later Principia.The first explicit manuscript incorporating these developments was De Analysi per aequationes numero terminorum infinitas, composed around mid-1669 from ideas originating in the plague years and circulated privately among scholars like Isaac Barrow and John Collins.[12] In this work, fluxions appear implicitly through the use of incremental moments to revert infinite series—transforming power series expressions for solving equations and computing areas under curves.[12] For instance, Newton employed rules for series expansion and reversion to extract roots and quadratures, effectively applying fluxional increments to resolve indeterminate forms without formal notation.[12] This implicit framework marked the initial written articulation of fluxional methods for infinite series analysis.[12]
Refinements and Private Manuscripts
Following the foundational ideas developed in the 1660s, Newton continued to refine his method of fluxions through private correspondence and manuscripts in the ensuing years. In 1671, Newton composed the treatiseDe methodis fluxionum et serierum infinitarum, which outlined the application of fluxions to solving differential equations through geometric constructions, demonstrating how flowing quantities could be used to determine tangents, areas, and other curve properties without explicit algebraic notation.[13] This manuscript, circulated among a small circle of trusted scholars like Collins, represented an early systematic exposition but remained unpublished during Newton's lifetime.[13]Over the subsequent decades, particularly in the 1690s, Newton revisited and expanded the 1671 tract known as the "Method of Fluxions," incorporating treatments of higher-order fluxions to handle more complex rates of change in curved motions. These revisions, added intermittently to the original manuscript, aimed to integrate infinite series expansions more deeply with fluxional computations, enhancing the method's versatility for advanced geometric problems.[14] Despite these improvements, the full tract stayed in private hands, shared selectively with figures such as David Gregory and Nicolas Fatio de Duillier, rather than being prepared for broader dissemination.[15]Newton's decision to withhold publication stemmed from his apprehension of sparking mathematical controversies, as experienced earlier with his optical theories, and his philosophical preference for presenting the method in a geometric framework—emphasizing limits and ratios over purely analytic fluxional symbols—to align with classical Euclidean traditions.[16] In 1711, William Jones published some of Newton's early mathematical manuscripts, including De analysi per aequationes numero terminorum infinitas (1669), in Analysis per quantitatum series, fluxiones, ac differentias, as part of efforts in the calculus priority dispute.[17] The full 1671 treatise on fluxions remained unpublished until 1736.
Core Concepts and Notation
Fluxions and Fluents Defined
In Isaac Newton's method of fluxions, a fluent is defined as a flowing or varying quantity that increases or decreases over time, such as the position of a moving point or the area under a curve, conceptualized as generated continuously rather than composed of discrete parts.[18] The fluxion, in turn, represents the instantaneous rate of change or velocity of this fluent, akin to its momentary increment at any given instant.[18] For instance, if the fluent denotes distance traveled as a function of time, its fluxion corresponds to the speed at that moment.[19]Newton deliberately avoided the use of actual infinitesimals, instead treating fluxions through the limits of ratios of vanishing quantities or moments—small increments that approach zero without ever being zero in a finite sense.[15] These moments, denoted momentarily before they vanish, allow computation of rates without invoking indivisible atoms of magnitude. This approach, using evanescent quantities or moments that vanish instantaneously, allowed computation of rates without assuming the existence of actual infinitesimals or indivisible magnitudes, grounding the method in the limits of finite increments while emphasizing their transient nature.[19]Newton introduced a distinctive notation for fluxions, using a dot accent above the variable to indicate the first-order rate of change, such as \dot{x} for the fluxion of the fluent x.[20] Higher-order fluxions followed similarly, with multiple dots (e.g., \ddot{x} for the second fluxion), while the fluent itself was recovered as the integral of its fluxion, though Newton often expressed this inversely through summation of moments.[21] This dotted notation, developed in the early 1690s, facilitated algebraic manipulation of rates in his manuscripts.[20]Conceptually, Newton viewed fluxions not as mere symbolic abstractions but as arising from a genuine physical cause, such as the uniform motion generating time or a body's continuous trajectory, thereby grounding the method in dynamic generation rather than static division.[18] This perspective distinguished his framework from purely algebraic treatments, insisting that quantities "flow" through perpetual motion, with fluxions capturing the inherent velocities produced thereby.[15]
Basic Rules for Computation
Newton's method of fluxions provided a systematic framework for calculating the rates of change of mathematical quantities, with basic rules that parallel modern differentiation techniques. These rules were derived through geometric and algebraic manipulations, often involving increments or moments proportional to the fluxions themselves. Central to the operations is the handling of products, quotients, composite expressions, and powers of fluents.The fluxion of the product of two fluents u and v is computed as\dot{(uv)} = u \dot{v} + v \dot{u},a rule explicitly formulated by considering the simultaneous changes in u and v. This product rule allows for the differentiation of multiplicative expressions by distributing the fluxions across the factors, forming the basis for more complex computations in the method.For quotients, Newton derived the corresponding rule through inversion or by applying the product rule to u and the reciprocal of v, yielding\dot{\left(\frac{u}{v}\right)} = \frac{v \dot{u} - u \dot{v}}{v^2}.This formula accounts for the relative rates in division, ensuring consistency in handling fractional fluents without direct division of fluxions.In cases of composite fluents, where one fluent y depends on another x, the chain rule analogue expresses the fluxion as\dot{y} = \dot{x} \cdot \frac{dy}{dx},with \frac{dy}{dx} treated as an intermediate fluent derived from the functional relationship. This relational approach enables the propagation of fluxions through nested dependencies, akin to composing rates of change.The fluxion of powers follows a binomial expansion pattern, where for a fluent x raised to the power n (positive, negative, integer, or fractional), the first fluxion is\dot{(x^n)} = n x^{n-1} \dot{x}.Successive higher fluxions of x^n produce terms that diminish the power and incorporate factorial-like coefficients with powers of \dot{x}, serving as a direct precursor to the general Leibniz rule for the k-th derivative (x^n)^{(k)}. These power rules underpin the expansion of series and polynomial manipulations in fluxional calculus.
Mathematical Examples and Applications
Finding Tangents and Areas
One of the primary applications of Newton's method of fluxions was the determination of tangents to curves, addressing a classical problem in geometry that had challenged mathematicians for centuries. In this approach, Newton conceptualized curves as generated by flowing quantities, or fluents, such as x and y, with their rates of change denoted by fluxions \dot{x} and \dot{y}. The direction of the tangent at any point is given by the ratio of these fluxions, \frac{\dot{y}}{\dot{x}}, which corresponds to the modern derivative \frac{dy}{dx}. This ratio determines the slope of the tangent line, allowing for the construction of the tangent through geometric interpretation of the infinitesimal increments. For instance, Newton applied this method to the cycloid, a curve traced by a point on a rolling circle, deriving the fluxions from its parametric equations and computing the tangent's inclination at arbitrary points along the arch.[22] Similarly, for the conchoid of Nicomedes, defined by the relation x y = b + y \sqrt{c^2 - y^2} (with constants b and c), Newton differentiated the equation using fluxional rules to obtain \dot{y} in terms of \dot{x}, yielding the tangent's direction and enabling precise geometric drawings.[22] These examples illustrate how fluxions provided a systematic algebraic tool for tangent-finding, surpassing earlier synthetic methods limited to specific conic sections.[23]The method's companion problem involved quadrature, or finding the area under a curve—another longstanding geometric challenge. Newton treated areas as fluents accumulating via their bounding ordinates' fluxions, effectively inverting the differentiation process through integration. A key technique was the reversion of power series for cases where relations were implicit, transforming transcendental relations into manageable series. This reversion method extended to more complex curves.[22]A foundational quadrature example in fluxions is the area under y = x^n for positive integer n \neq -1. Applying the basic integration rule, the fluxion \dot{z} = x^n implies z = \frac{x^{n+1}}{n+1} + c, where the constant c is determined by initial conditions, such as zero area at x=0. This yields the familiar antiderivative, demonstrating fluxions' power for polynomial areas and foreshadowing the general binomial theorem's role in non-integer cases.[22]Newton's innovations culminated in his 1669 treatise De analysi per aequationes numero terminorum infinitas, where he explicitly solved quadratures unsolvable by finite algebraic equations, such as the area under \sqrt{\frac{a^2 - x^2}{a^2 + x^2}}, by expanding the integrand into an infinite series and term-by-term integration. This approach resolved problems previously deemed intractable, like certain transcendental curves, by leveraging series expansions derived from the binomial theorem, thus broadening geometry's scope beyond algebraic solvability.[24]
Higher-Order Fluxions
Newton extended the method of fluxions beyond first-order rates of change to higher orders, enabling analysis of acceleration and more complex dynamic behaviors. The second fluxion, denoted \ddot{x}, is the fluxion of the first fluxion \dot{x}, representing the rate of change of velocity, or acceleration, in the context of flowing quantities or motion. As Newton described, these higher fluxions are "velocities of velocities," essential for resolving equations involving varying speeds and for applications in geometry and physics.[25]In dynamics, second fluxions proved crucial for modeling forces in curvilinear motion. For centripetal acceleration in polar coordinates, Newton formulated the radial equation as \ddot{r} - r \dot{\theta}^2 = -\frac{f}{m}, where r is the radial distance, \theta the angular position, \dot{\theta} the angular velocity, f the centripetal force, and m the mass; this expression captures the inward acceleration required to maintain an orbital trajectory around a center of force. This application appears in his treatment of planetary motion, linking fluxional calculus directly to gravitational attraction.[26]To compute higher-order fluxions systematically, Newton provided rules that extended the basic operations, particularly for products of fluents. The fluxion of a product uv follows \dot{(uv)} = u \dot{v} + v \dot{u}, and for the nth fluxion, a recursive application yields the general Leibniz rule: the nth fluxion of uv is \sum_{k=0}^{n} \binom{n}{k} u^{(k)} v^{(n-k)}, where u^{(k)} denotes the kth fluxion of u. This recursive structure allowed Newton to handle complex expressions in series expansions and equation resolutions without explicit limits.[25]Higher-order fluxions also facilitated the analysis of curve properties, such as curvature in trajectories. For the parabola y^2 = ax, Newton derived the second fluxions to find the acceleration along the curve, yielding $2 y \ddot{y} + 2 \dot{y}^2 = a \ddot{x}, which relates to the radius of curvature given by DH = y + \frac{a^2}{9x}. Extending to third fluxions, \dddot{x}, enables identification of inflection points on a trajectory, where the curvature changes sign—occurring when the second fluxion vanishes and the third fluxion determines the transition from concave to convex behavior.[25]
Rivalry with Leibniz
Timeline of Independent Inventions
Isaac Newton developed the foundational ideas of his method of fluxions during his annus mirabilis in 1665–1666, while isolated at his family estate in Woolsthorpe due to the Great Plague, where he explored infinite series and methods for finding tangents and areas under curves.[27] These early formulations remained private, with no publication or wide dissemination at the time, allowing Newton's work to precede similar developments elsewhere by nearly a decade. In a 1671 letter to mathematician John Collins, Newton outlined aspects of his fluxional calculus, including rules for computing fluxions of products and quotients, but this correspondence was confined to a small circle of British mathematicians and not shared beyond.[28]Gottfried Wilhelm Leibniz, working independently on the continent, began formulating his differential calculus in 1675 while in Paris, where he introduced notation for infinitesimals and rules for maxima, minima, and tangents in unpublished manuscripts dated October 29 of that year.[29] Although Leibniz visited London in early 1673 as part of a diplomatic mission and met Henry Oldenburg, the secretary of the Royal Society, any exposure to Newtonian ideas through Oldenburg remains unproven and did not influence his core insights, as Leibniz's mathematical interests at the time focused on other topics like series expansions.[30] No direct correspondence between Newton and Leibniz on their respective methods occurred until the mid-1670s, when mediated exchanges via Oldenburg in 1676 shared some optical and series results but omitted details of fluxions or differentials, thereby preserving the independence of their inventions.[30]Leibniz's first public account of his differential method appeared in 1684 in the journal Acta Eruditorum, where he presented the "Nova Methodus" for solving problems of maxima and minima using differentials, marking the formal introduction of his notation to the broader mathematical community.[31] This timeline underscores the parallel yet isolated paths: Newton's private maturation from 1665 onward and Leibniz's breakthrough in 1675, with neither having access to the other's systematic approach until well after their initial discoveries.[32]
Accusations of Plagiarism
The initial public accusations of plagiarism against Gottfried Wilhelm Leibniz regarding the invention of calculus emerged in 1699, when Nicolas Fatio de Duillier, a Swissmathematician and associate of Isaac Newton, published the book Lineæ brevissimæ descensus investigatio geometrica duplex.[33] In this piece, Fatio claimed that Leibniz had derived his differential methods from Newton's unpublished work on fluxions, citing correspondence from the 1670s between English mathematician John Collins and Leibniz as evidence of Leibniz's access to Newton's ideas.[32]The controversy intensified in the early 1710s amid growing tensions over priority, which both Newton and Leibniz had independently developed around the same time in the late 1660s and 1670s. In 1710, John Keill, a supporter of Newton, published a paper in the Philosophical Transactions of the Royal Society (vol. 26) explicitly charging Leibniz with plagiarizing Newton's fluxional calculus. Newton himself amplified these allegations in 1715 through an anonymous review in the same journal, portraying Leibniz's contributions as derivative and reinforcing the narrative of theft based on the earlier letters.[34]Leibniz swiftly countered these claims in a 1711 article in Acta Eruditorum, where he defended his independent discovery of the infinitesimal calculus and asserted priority through references to his private manuscript notes dating back to 1675, predating his awareness of Newton's work. He dismissed the accusations as unfounded, emphasizing that his notations and methods, such as the differential dx, were original developments without reliance on fluxions.[30]A pivotal event contributing to Newton's suspicions occurred in 1690, when Jacob Bernoulli posed a challenge problem in Acta Eruditorum to determine the equation of the catenary curve formed by a hanging chain. Leibniz responded with a solution employing his nascent calculus techniques, which Newton later interpreted as suspiciously akin to his own fluxional methods for handling rates of change, thereby heightening perceptions of plagiarism among Newton's circle.[35]
Priority Dispute and Resolution
Evidence Presented by Newton
In the priority dispute over the invention of calculus, Isaac Newton presented key evidence through the Commercium Epistolicum (1712), a publication by the Royal Society that he largely authored himself, which included selectively edited excerpts from the 1672–1676 correspondence between Newton, Henry Oldenburg, and Gottfried Wilhelm Leibniz to imply that Leibniz had prior knowledge of the method of fluxions.[32] This editing aimed to demonstrate that Leibniz's differential notation derived from exposure to Newton's ideas during his 1676 visit to London, though the correspondence primarily discussed infinite series rather than fluxions explicitly.[36] Newton supplemented this with references to his unpublished manuscripts, notably De Analysi per Aequationes Numero Terminorum Infinitas (written in 1669 and circulated privately via Isaac Barrow to John Collins), which outlined methods for expanding functions into infinite series and finding tangents and areas—elements he argued constituted the foundational principles of fluxions predating Leibniz's work by several years.[37]Newton further contended that his approach via fluxions and "ultimate ratios"—the limiting ratios of vanishing quantities as time intervals approach zero—offered a more rigorous foundation than Leibniz's use of infinitesimals, which he viewed as fictive and less grounded in geometry.[38] In his Account of the Method of Fluxions and Infinite Series (published 1715 but drafted earlier), Newton emphasized that ultimate ratios avoided the paradoxes of actual infinitesimals by treating increments as generated quantities whose ratios were evaluated in the limit, thereby providing a kinematic interpretation superior to Leibniz's static differentials.[19]Historians have since criticized some of Newton's exhibits for containing errors and possible alterations, such as the inaccurate claim in the Commercium Epistolicum that Leibniz accessed a 1672 letter from Newton to Collins detailing fluxional methods, which was not the case, highlighting selective presentation to bolster his priority claim.[32] These manipulations, including the anonymous authorship of the Commercium under the guise of an impartial committee report, have been noted as biasing the evidence in Newton's favor.[39]
Role of the Royal Society
In late 1711, following accusations of plagiarism leveled by John Keill in the Philosophical Transactions of the Royal Society, Gottfried Wilhelm Leibniz appealed to the Society's secretary, Hans Sloane, demanding a public recantation from Keill and an investigation into the claims against him.[32] As president of the Royal Society since 1703, Isaac Newton, who had been privately directing Keill's actions, responded by appointing a committee on 6 March 1712 to examine the priority dispute over the invention of calculus.[40] The committee, comprising Newton's close allies including Keill himself, Edmund Halley, and John Machin, was widely criticized for its lack of impartiality, as its members were predisposed to support Newton's claims.[41]Under Newton's secretive guidance, the committee produced the Commercium Epistolicum D. Johannis Collins et aliorum de Analysi Promota, published by the Royal Society in April 1712. This report, authored anonymously by Newton himself, compiled correspondence and evidence—such as Newton's early manuscripts from the 1660s shared with contemporaries like John Collins—to assert Newton's priority in developing the method of fluxions before Leibniz's publication of his differential calculus in 1684.[32] The document ruled in Newton's favor, concluding that Leibniz had concealed his knowledge of Newton's prior work during his 1676 visit to England, though it stopped short of a direct plagiarism charge.[40]Leibniz vehemently protested the report's bias in an anonymous 1713 pamphlet, Charta Volans, but his subsequent appeal to the Royal Society for reconsideration was ignored, with no further official response from the body.[32] The dispute's direct involvement of the Society effectively concluded with Leibniz's death on 14 November 1716, leaving the Commercium as the institution's final pronouncement affirming Newton's precedence.[41]
Legacy and Modern Interpretation
Influence on British Mathematics
The posthumous English translation of Newton's Method of Fluxions and Infinite Series by John Colson in 1736 played a crucial role in disseminating the work within Britain, including over 200 pages of explanatory commentary that made the kinematical approach accessible to a wider audience of mathematicians and educators.[42] Priced affordably at 15 shillings, this edition reinforced fluxions as a cornerstone of British mathematical instruction, particularly in universities like Cambridge and Edinburgh, where it supplanted earlier manuscript circulations.[42]Brook Taylor was among the earliest adopters, integrating fluxions into his Methodus Incrementorum Directa et Inversa (1715), where he employed the dot notation—such as \dot{x} for the first fluxion and \ddot{x} for the second—to analyze finite differences and series expansions in mechanics.[42] Taylor favored limits over infinitesimals in line with Newtonian rigor, which helped establish fluxions as a tool for curve analysis and integration without relying on Leibnizian symbolism.[42] Colin Maclaurin built on this foundation with his comprehensive Treatise of Fluxions (1742), a systematic exposition that treated fluxions analytically while grounding them in kinematics to counter critiques like George Berkeley's The Analyst (1734); this text became a standard reference, influencing subsequent fluxional calculus treatises and teaching practices across British institutions.[42]The dot notation persisted as a hallmark of British mathematical works well into the 19th century, appearing prominently in texts by figures like Thomas Simpson (Doctrine and Application of Fluxions, 1750) and John Landen (Residual Analysis, 1764), where it symbolized a commitment to Newton's original framework over the more algebraic Leibnizian differentials.[42] This preference delayed the widespread adoption of symbolic analysis from the Continent, as British mathematicians prioritized the geometric and kinematical interpretations of fluxions in their treatises.[42]Applications of fluxions extended to practical fields, with Edmond Halley employing the method in celestial mechanics to compute comet orbits and planetary perturbations, as seen in his contributions to the Philosophical Transactions.[42] These uses underscored how the resolution of the priority dispute in Newton's favor bolstered fluxions' dominance in British science during the 18th century.[42]
Relation to Contemporary Calculus
Newton's method of fluxions prefigures the modern concept of the derivative as the limit of a difference quotient, where the fluxion \dot{x} of a quantity x is understood as \dot{x} = \lim_{\Delta t \to 0} \frac{\Delta x}{\Delta t}, representing the instantaneous rate of change. This intuitive approach, grounded in geometric interpretations of motion and infinitesimals, aligns closely with contemporary calculus by capturing the essence of differentiation without the formal epsilon-delta rigor developed later. Historians of mathematics note that fluxions effectively solved problems of tangents and areas through this limiting process, making them a foundational precursor to the analytic tools used today.Despite their ingenuity, fluxions lacked the rigorous foundations essential for modern calculus, relying on informal notions of "ultimate ratios" that avoided explicit infinitesimals but did not fully resolve paradoxes like those in Zeno's arguments. This approach proved problematic for higher-order fluxions, particularly with non-analytic functions where infinite series expansions might not converge, highlighting vulnerabilities exposed in the 19th century. The work of Augustin-Louis Cauchy and Karl Weierstrass in the 1820s and 1870s provided the epsilon-delta limits and real analysis framework that elevated calculus to a precise discipline, rendering Newton's methods obsolete in formal proofs while preserving their heuristic value.In the modern view, fluxions are seen as geometrically oriented and intuitive, emphasizing visualization over algebraic manipulation, in contrast to Gottfried Wilhelm Leibniz's differential notation, which facilitated symbolic computation and broader applicability in analysis. This geometric focus limited fluxions' integration into emerging fields like partial differential equations, though their conceptual clarity influenced pedagogical approaches. Higher-order fluxions briefly served as precursors to Taylor series expansions in early approximations.The 19th-century shift in Britain toward Leibniz's "d/dx" notation, championed by Augustus De Morgan in his 1836 textbook Elements of Algebra, marked the obsolescence of fluxions, aligning British mathematics with continental standards and enabling more efficient symbolic reasoning in advanced studies. This transition underscored fluxions' historical role as a bridge from classical geometry to analytic calculus, now viewed through the lens of real analysis.