Fact-checked by Grok 2 weeks ago

Logarithm

In , a logarithm is the inverse operation to , defined as the exponent x to which a fixed positive b > 0 (where b \neq 1) must be raised to produce a given positive number a, expressed as \log_b a = x such that b^x = a. This concept encapsulates the relationship between and , as logarithmic properties allow products to be converted into sums, simplifying complex computations. Common forms include the (base 10, denoted \log a) and the natural logarithm (base e \approx 2.718, denoted \ln a), each serving distinct roles in and modeling. The invention of logarithms is credited to Scottish mathematician , who published the first tables in his 1614 work Mirifici Logarithmorum Canonis Descriptio, motivated by the need to streamline astronomical calculations involving large multiplications and . Over the next two decades, English mathematician Henry Briggs refined Napier's approach by introducing base-10 logarithms, publishing extensive tables in Arithmetica Logarithmica (1624), while publisher Adrian Vlacq extended these to numbers up to 100,000 by 1628. In the , Leonhard Euler formalized modern notation and emphasized natural logarithms, linking them to the with base e, defined as the limit \lim_{n \to \infty} (1 + 1/n)^n. These developments transformed , enabling tools like slide rules—first described by in 1622—which relied on logarithmic scales for and division until the advent of electronic calculators in the 1970s. Logarithms remain fundamental in mathematics and science, underpinning calculus through their role as antiderivatives of exponential functions and enabling the modeling of exponential growth and decay processes. In applied fields, they quantify vast scales in phenomena such as bacterial population growth in biology, radioactive decay in physics, acidity via pH in chemistry, and earthquake intensity on the Richter scale in geology. Their properties— including \log_b (xy) = \log_b x + \log_b y and \log_b (x/y) = \log_b x - \log_b y—facilitate solving equations and analyzing data across engineering, economics, and social sciences, where they compress wide-ranging values into manageable forms.

Fundamentals

Motivation

Logarithms were originally developed in the early to address the computational challenges faced by astronomers and mathematicians working with large numbers and complex trigonometric calculations. Scottish mathematician spent over two decades devising a method to simplify these tedious multiplications, which were essential for tasks like predicting planetary positions and solving problems in and astronomy. Conceptually, logarithms serve as the inverse operation to , providing a way to transform multiplicative processes into additive ones, which aligns with the aptitude for over repeated . This inversion allows problems involving products of numbers to be reframed as sums of their logarithmic counterparts, thereby streamlining operations that would otherwise require laborious manual computation. In , logarithms enable the compression of datasets spanning vast ranges of magnitudes into a more manageable , facilitating and comparison of trends that would be obscured on linear representations. For instance, values from billions to fractions of a can be mapped to a concise , as seen in applications like modeling or scientific measurements, where the reveals patterns in exponential phenomena. A practical illustration of this utility is the transformation of , such as computing a \times b, into the \log a + \log b, which then corresponds back to the product via , eliminating the need for direct multiplication of potentially enormous figures.

Definition

The logarithm is the inverse operation to , providing a way to solve for the exponent in equations of the form b^c = a. Formally, the logarithm of a positive a to base b is defined as the c such that b^c = a, where b > 0, b \neq 1, and a > 0. This is denoted by \log_b a = c. The is well-defined under these conditions because the with base b is bijective from the real numbers to the positive reals. Common notations vary by base: \log a typically denotes the common logarithm with base 10, \ln a or \log_e a the natural logarithm with base e \approx 2.71828, and \log_2 a or sometimes \lg a the binary logarithm with base 2. These conventions facilitate applications in different fields, such as engineering for base 10 and computer science for base 2. The domain of the logarithm \log_b a consists of all a > 0, with the base b fixed as a positive real not equal to 1; the is all real numbers. For b > 1, the is strictly increasing; for $0 < b < 1, it is strictly decreasing. These properties ensure the logarithm's invertibility with the exponential .

Examples

To illustrate the logarithm, consider its definition as the exponent to which a base b > 0, b \neq 1, must be raised to obtain a positive argument a > 0. Basic computations demonstrate this directly. For instance, \log_{10}(100) = 2 because $10^2 = 100. Similarly, \log_2(8) = 3 since $2^3 = 8, and \ln(e) = 1 as e^1 = e. Algebraic examples highlight the inverse relationship with exponentiation. The property \log_b(b^k) = k holds for any real k, because b^{\log_b(b^k)} = b^k by definition, and the exponential function is one-to-one, so the exponents must be equal. Solving logarithmic equations reverses this: \log_{10}(x) = 3 implies x = 10^3 = 1000. Common pitfalls arise from domain restrictions. Logarithms are undefined for negative arguments, as no real exponentiation with a positive base yields a negative result, and bases of 1 are invalid since $1^k = 1 for all k, failing to produce other values. For visual intuition, view as repeated —such as $2^3 = 2 \times 2 \times 2 = 8—while the logarithm counts the necessary steps of by the to reach the argument, like \log_2(8) = 3.

Properties

Logarithmic Identities

Logarithmic identities form the foundational algebraic rules for simplifying and manipulating expressions involving logarithms. These identities derive from the inverse relationship between logarithmic and functions and are applicable to any valid b > 0, b \neq 1. They enable the transformation of complex logarithmic expressions into simpler forms, facilitating computation and analysis in various mathematical contexts. The states that for x and y, and base b > 0, b \neq 1, \log_b (xy) = \log_b x + \log_b y. To prove this, let u = \log_b x and v = \log_b y, so x = b^u and y = b^v. Then xy = b^u \cdot b^v = b^{u+v}, and thus \log_b (xy) = u + v = \log_b x + \log_b y. This identity holds because the logarithm is the inverse operation to with the same base. Similarly, the quotient rule provides \log_b \left( \frac{x}{y} \right) = \log_b x - \log_b y for x > 0, y > 0. The proof follows analogously: with u = \log_b x and v = \log_b y, \frac{x}{y} = \frac{b^u}{b^v} = b^{u-v}, so \log_b \left( \frac{x}{y} \right) = u - v. This rule leverages the subtraction property of exponents under the same base. The power rule extends this to exponents: \log_b (x^k) = k \log_b x for real k and x > 0. Proof: Let u = \log_b x, so x = b^u and x^k = (b^u)^k = b^{ku}. Therefore, \log_b (x^k) = ku = k \log_b x. A special case is the root rule, where \log_b \sqrt{x} = \log_b (x^{1/k}) = \frac{1}{k} \log_b x for k \geq 2 and x > 0, directly from the power rule. These confirm the scalability of logarithms with respect to powers. The change of base formula allows conversion between different bases: \log_b a = \frac{\log_k a}{\log_k b} for any valid bases b and k, and a > 0. Commonly, k = e () or k = 10 () is used for computational convenience. Proof: Let u = \log_b a, so a = b^u. Taking \log_k of both sides yields \log_k a = \log_k (b^u) = u \log_k b, hence u = \frac{\log_k a}{\log_k b}. This identity underscores the equivalence of logarithmic scales across bases.

Particular Bases

The common logarithm is the logarithm to base 10, widely employed in , , and due to its alignment with the decimal system for measurements and calculations. It is conventionally denoted as \log x without specifying the base or as \log_{10} x for clarity. The natural logarithm employs base e, the approximately equal to 2.71828 that serves as the foundation for and continuous . This base emerges inherently in the definitions of derivatives and integrals, rendering the natural logarithm indispensable in , differential equations, and . It is denoted as \ln x or \log_e x. The uses base 2 and holds particular significance in , where it measures , data storage in bits, and entropy in . It is denoted as \log_2 x or lg x. Relations between these bases facilitate conversions via the change of base formula, yielding useful constants such as \log_{10} e \approx 0.4343, \ln 10 \approx 2.3026, and \log_2 e \approx 1.4427.

History

Early Development

The Scottish mathematician invented logarithms in the early , primarily to simplify complex trigonometric calculations required for astronomical observations. His motivation stemmed from the laborious multiplications and divisions involved in computing sines and tangents for and planetary motion studies, which were essential to contemporary astronomy. In 1614, Napier published Mirifici Logarithmorum Canonis Descriptio, introducing his concept of "point logarithms" as a novel computational aid. Unlike the modern exponential definition, Napier's logarithms were defined geometrically as proportional to the lengths of line segments generated by points moving along parallel lines, where one point travels at constant speed and the other at a speed inversely proportional to its remaining distance to the endpoint. This construction related numbers in geometric progression to corresponding arithmetic progressions, allowing multiplication of sines to be reduced to addition of their logarithms. Following the publication, English mathematician Henry Briggs corresponded with and visited Napier in during the summers of 1615 and 1616 to refine the system. Briggs proposed shifting the logarithms to a base-10 scale, making them more practical for decimal-based arithmetic, and by 1624 he had computed and published extensive tables of these common logarithms in Arithmetica Logarithmica. Napier's work and Briggs's modifications were rapidly disseminated across through printed tables and treatises, gaining adoption among astronomers and navigators for simplifying computations in the 1620s and 1630s. By the mid-17th century, logarithms had become a standard tool in scientific circles, praised by figures like for accelerating astronomical calculations.

Key Advancements and Contributors

Building on John Napier's foundational work on logarithms published in 1614, later mathematicians advanced the concept through independent inventions, reformulations, and theoretical integrations. Swiss mathematician and instrument maker Jost Bürgi independently developed a logarithmic system around to facilitate astronomical computations, distinct from Napier's approach in its proportional scaling method. His Progress Tabulen, printed in 1620 under the influence of , presented tables equivalent to logarithms for numbers up to 10^12, though without the explicit logarithmic nomenclature, emphasizing practical multiplication and division aids. English mathematician Henry Briggs significantly refined the logarithm by proposing a base-10 system during his 1615–1616 visits with Napier, setting log(1) = 0 and log(10) = 1 for alignment with arithmetic. In 1624, he published Arithmetica Logarithmica, the first comprehensive table of common logarithms, listing values for natural numbers from 1 to 20,000 (and 90,000 to 100,000) to 14 places, alongside , revolutionizing computational efficiency in astronomy and . In 1647, Belgian Jesuit Grégoire de Saint-Vincent advanced the theoretical understanding of logarithms through geometric analysis in his Opus geometricum quadraturae circuli et sectionum coni. He demonstrated that the area under the rectangular hyperbola xy = 1 between points where the ratio of abscissas equals that of another interval yields equal areas, effectively characterizing the \int \frac{1}{x} \, dx as a logarithmic function—a property later explicitly linked to logarithms by his student Alphonse Antonio de Sarasa. Leonhard Euler provided a rigorous formalization of the natural logarithm in his 1748 treatise , defining it with base e \approx 2.71828 as the limit \lim_{n \to \infty} (1 + \frac{1}{n})^n and via the series e = \sum_{k=0}^{\infty} \frac{1}{k!}, explicitly connecting it to the inverse of the . This work established e as the fundamental base for , enabling logarithms' seamless integration into differential and integral calculus. Throughout the 18th and 19th centuries, logarithms evolved from mere computational tools to essential analytic functions in , with Euler's notation facilitating their treatment as continuous, differentiable entities, as seen in the \frac{d}{dx} \ln x = \frac{1}{x} and its role in solving equations modeling growth and decay. This transition underscored logarithms' ubiquity in , bridging arithmetic and methods. A culminating 19th-century advancement came in 1882 when proved that e^\alpha is for any nonzero algebraic \alpha, implying the of \pi (via e^{i\pi} = -1) and, more broadly, that \ln a is for any algebraic a > 0, a \neq 1. This result, building on Charles Hermite's 1873 proof of e's , resolved long-standing questions about the of logarithmic values and affirmed the depth of .

Historical Tools

Logarithm Tables

Logarithm tables, essential pre-digital computational aids, were pioneered by Henry Briggs in the early 1620s. In his 1624 work Arithmetica Logarithmica, Briggs compiled decimal logarithms to 14 places for integers from 1 to 20,000 and from 90,000 to 100,000, marking the first extensive base-10 tables designed for practical arithmetic. Adriaan Vlacq advanced this effort in the 1630s by extending Briggs's tables to cover all integers from 1 to 100,000, publishing them to 10 decimal places in his 1628 Arithmetica Logarithmica, which became a foundational reference for subsequent compilations. These revolutionized by transforming and into simpler and operations. Users located the logarithms of the operands in the , applied logarithmic identities to add or subtract the values, and then consulted the antilogarithm section to retrieve the result, drastically reducing the tedium of manual calculations for large numbers. For values not directly tabulated, accuracy was maintained through methods, typically between adjacent entries, which yielded errors small enough for scientific and applications. Higher-order techniques, such as those using first and second differences, were occasionally employed in more precise to minimize rounding errors. The utility of logarithm tables extended significantly to navigation and surveying, where they facilitated complex trigonometric and positional calculations critical for maritime voyages and land mapping during the Age of Exploration and industrial expansion. By the 1970s, the advent of affordable handheld electronic calculators rendered logarithm tables obsolete, as direct computation supplanted the need for manual lookup and interpolation in routine arithmetic.

Slide Rules

The slide rule, an analog computing device, was invented in 1622 by English and Anglican minister , who adapted logarithmic scales originally developed for printed tables into a portable instrument with movable parts. Oughtred's design built upon earlier logarithmic concepts by placing two engraved scales in sliding configuration, allowing users to perform arithmetic operations mechanically without direct numerical computation. The core mechanism of a relies on logarithmic scales marked along rulers, where the physical distance between numbers represents their logarithms, enabling and through simple alignment and reading. To multiply two numbers, for instance, the user aligns the initial value on one scale with the first on the other, then reads the product where the second intersects, as adding the logarithmic distances yields the logarithm of the result. follows a similar process by reversing the alignment, subtracting logarithms via scale opposition. This analog approach extended to other operations like square roots by using doubled-length scales, though it inherently approximated results due to the continuous nature of the markings. The standard Mannheim slide rule, introduced in 1859 by French artillery officer Amédée Mannheim, became the most common type with its layout featuring four primary scales: A and B (halved logarithmic scales for squares and roots, spanning 1 to 100) on the upper fixed part, and C and D (full logarithmic scales for basic arithmetic, spanning 1 to 10) on the lower fixed and sliding parts, respectively. Specialized variants emerged for , incorporating sine (S) and (T) scales on the slide's edges for angle calculations in and , while models added specialized scales like K (for cubes) or L (for common logs) to handle domain-specific computations such as beam stresses or electrical circuits. These types varied in materials—, , or —and length, with 10-inch pocket models being portable for field use and longer 20-inch versions offering marginally higher readability. Slide rules typically provided precision limited to 3 to 4 significant digits, constrained by scale length, engraving quality, and the user's ability to interpolate between marks, making them suitable for estimates rather than exact . The A, B, C, and D scales formed the foundation of most models, with the cursor—a movable indicator—facilitating precise and reading across scales. Slide rules reached their peak usage in the 20th century among engineers, scientists, and pilots, powering calculations for projects like the Apollo moon missions and ballistics, where rapid approximations were essential. Their portability and speed made them indispensable until the mid-1970s, when affordable electronic calculators like the , introduced in 1972, rendered them obsolete by offering greater precision and versatility at lower cost.

Computational Applications

Logarithms played a pivotal role in enabling complex scientific computations during the 17th to 19th centuries by transforming multiplications and divisions into additions and subtractions, thus facilitating the handling of large datasets in fields requiring precise numerical work. In astronomy, John Napier's logarithm tables, published in 1614, were instrumental for calculations essential to and planetary analysis. Napier's analogies, mnemonic devices for solving spherical triangles, streamlined these operations, which were central to astronomical modeling. utilized these logarithms to process Tycho Brahe's observational data, particularly in deriving his third law of planetary motion in 1618 by recognizing a logarithmic relationship between orbital periods and distances, a breakthrough that reduced exhaustive manual verifications. In applications, logarithms accelerated and computations throughout the 17th to 19th centuries. Naval gunners in the English employed logarithm tables to determine cannonball trajectories, powder charges, and firing angles, enhancing accuracy in combat scenarios despite initial resistance due to the low mathematical among crews. For instance, during the , these tables allowed for rapid adjustments in shipboard gunnery, contributing to tactical advantages in fleet engagements. In land , instruments like the sector incorporated logarithmic scales to compute distances and angles from trigonometric measurements, as seen in 18th-century colonial practices where surveyors used such tools to divide territories efficiently without prolonged . Slide rules, based on logarithmic principles, provided a portable alternative for field engineers performing these calculations on-site. Financial computations also benefited significantly, with logarithms enabling the efficient calculation of in the . applied logarithmic methods to approximate and values, as in his 1670 estimation of a 7.43% effective for a specific scenario, which informed early actuarial tables and economic modeling. This approach expedited the generation of growth projections, vital for merchants and investors handling accumulations. The broader impact of logarithms on the lay in accelerating hypothesis testing by minimizing computational drudgery, allowing scientists to iterate models and verify theories more rapidly. later remarked that logarithms effectively "doubled the life of the astronomer" through time savings. A notable example is the 1758 prediction of return, where and collaborators used logarithm tables for months to compute planetary perturbations, refining Edmond Halley's earlier elliptical orbit estimate and confirming the comet's perihelion on March 13, 1759. Such applications underscored logarithms' role in transforming empirical science into predictive endeavors.

Analytic Properties

Existence and Characterization

The existence of the real logarithm function can be established by first considering the \exp: \mathbb{R} \to (0, \infty), defined as the unique solution to the y' = y with y(0) = 1. This function is continuous and strictly increasing on \mathbb{R}, with \lim_{x \to -\infty} \exp(x) = 0 and \lim_{x \to \infty} \exp(x) = \infty. By the applied to the continuous image of [n, n+1] for integers n, which covers intervals in (0, \infty), and the density of such intervals, \exp is surjective onto (0, \infty). Thus, \exp is bijective, and its , the natural logarithm \ln: (0, \infty) \to \mathbb{R}, exists and is unique, continuous, and strictly increasing. More generally, for any b > 0, b \neq [1](/page/1), the logarithm \log_b: (0, \infty) \to \mathbb{R} exists as the inverse of the continuous, strictly increasing x \mapsto b^x, ensuring a unique value for each positive argument. The natural logarithm is the unique f: (0, \infty) \to \mathbb{R} satisfying the f(xy) = f(x) + f(y) for all x, y > 0, f([1](/page/1)) = [0](/page/0), and the normalization f([e](/page/e)) = [1](/page/1). More generally, continuous solutions to the with f([1](/page/1)) = [0](/page/0) are scalar multiples c \ln x for some constant c. To see that the natural logarithm satisfies the equation: \ln(xy) = \int_1^{xy} \frac{1}{t} \, dt = \int_1^x \frac{1}{t} \, dt + \int_x^{xy} \frac{1}{t} \, dt = \ln x + \int_1^y \frac{1}{u} \, du = \ln x + \ln y by the substitution u = t/x in the second integral, and \ln 1 = \int_1^1 \frac{1}{t} \, dt = [0](/page/0). For uniqueness up to scalar, suppose g is another continuous solution. Then g(e^r) = r \cdot g(e) for rational r by on the equation, and by and of , g(x) = c \ln x where c = g(e). Without the continuity assumption, the implies the existence of discontinuous (pathological) solutions to the , constructed using a Hamel basis for \mathbb{R} over \mathbb{Q} in the additive group, which transfer to the multiplicative case via the x \mapsto \ln x. These solutions are highly irregular, non-measurable, and lack practical . Logarithms in different bases are : \log_b a = \frac{\ln a}{\ln b} for b > 0, b \neq 1, a > 0, reflecting that any such function is a constant multiple of the natural logarithm, with the constant determined by the base.

Graph and Representations

The graph of the logarithm function \log_b x for base b > 1 is defined for all x > 0, with a vertical at x = 0 where the function approaches -\infty as x approaches 0 from the right. The curve passes through the point (1, 0), is strictly increasing, and exhibits concave-down behavior, reflecting its role as the of the b^x. For $0 < b < 1, the graph is decreasing and concave up, but the standard case b > 1 shows slower growth compared to linear functions. The natural logarithm, denoted \ln x, admits an integral representation that underscores its foundational properties: for x > 0, \ln x = \int_1^x \frac{1}{t} \, dt. This definition captures the area under the hyperbola $1/t from 1 to x, providing a geometric of . For a general base b > 0, b \neq 1, the logarithm \log_b x relates to the natural logarithm via the change-of-base formula: \log_b x = \frac{\ln x}{\ln b}. This expression allows computation of logarithms in any base by reducing to the natural logarithm, leveraging its unique properties. Asymptotically, the natural logarithm grows slower than any positive power of x; that is, \ln x = o(x^\epsilon) for any \epsilon > 0 as x \to \infty. Moreover, \ln x and \log_{10} x exhibit equivalent growth rates for large x, differing only by the constant factor $1/\ln 10.

Calculus and Transcendence

The derivative of the natural logarithm function is given by \frac{d}{dx} \ln x = \frac{1}{x} for x > 0, which follows directly from its integral definition as \ln x = \int_1^x \frac{1}{t} \, dt by the . For the general logarithm base b > 0, b \neq 1, the derivative is \frac{d}{dx} \log_b x = \frac{1}{x \ln b}, obtained via the change-of-base formula \log_b x = \frac{\ln x}{\ln b}. This simple form underscores the logarithm's role in measuring relative growth rates; the derivative \frac{1}{x} represents the instantaneous relative rate of change, making logarithms essential for analyzing proportional growth in , such as in models where the rate is constant relative to the quantity. The of the natural logarithm is found using : \int \ln x \, dx = x \ln x - x + C, where the substitution u = \ln x, dv = dx yields du = \frac{1}{x} dx, v = x. For the base-b logarithm, it follows as \int \log_b x \, dx = x \log_b x - \frac{x}{\ln b} + C, derived similarly by scaling the natural logarithm . These integrals appear in applications involving accumulated or , where logarithms help quantify total change over intervals proportional to the function's value. Beyond its calculus operations, the natural logarithm exhibits transcendence: for any \alpha > 0 with \alpha \neq 1, \ln \alpha is transcendental. This follows as a of the Lindemann-Weierstrass theorem, which states that if \alpha_1, \dots, \alpha_n are distinct algebraic numbers, then e^{\alpha_1}, \dots, e^{\alpha_n} are algebraically independent over the rationals; assuming \ln \alpha algebraic would imply e^{\ln \alpha} = \alpha contradicts this independence for non-zero algebraic exponents. The transcendence of logarithms has profound implications in and , distinguishing them from algebraic functions and highlighting their non-polynomial nature in integral and differential contexts.

Calculation Methods

Power Series

The power series expansion provides a foundational method for computing logarithms through infinite summation, particularly for the natural logarithm. The , also known as the Newton-Mercator series, expresses the natural logarithm of $1 + x as \ln(1 + x) = \sum_{n=1}^{\infty} (-1)^{n+1} \frac{x^n}{n} = x - \frac{x^2}{2} + \frac{x^3}{3} - \frac{x^4}{4} + \cdots for |x| < 1. This series is derived from the expansion of \ln(1 + x) about x = 0, where the coefficients are the reciprocals of the positive integers with alternating signs. discovered the series around 1665–1666 during his early work on infinite series, while Nicolas Mercator independently developed and first published it in his 1668 book Logarithmotechnia. James Gregory also contributed to the broader development of such series expansions in the , though his work focused more on related . The series converges for -1 < x \leq 1, with conditional convergence at x = 1 yielding \ln 2 = \sum_{n=1}^{\infty} (-1)^{n+1} / n, the alternating series. Outside this interval, the series diverges; for example, at x = -1, it becomes the negative series, which diverges. For practical computation, the is 1, limiting direct application to values of x near 0, but transformations like expressing \ln y for y > 0 as \ln(1 + (y-1)) extend usability to $0 < y < 2. Error bounds for partial sums leverage the alternating series estimation theorem: if the first n terms are used, the absolute error is less than the magnitude of the next term, \frac{|x|^{n+1}}{n+1}, provided the terms decrease in absolute value and approach zero. This bound ensures rapid convergence for small |x|, making the series historically valuable for tabular logarithm computation before electronic calculators. For instance, summing the first few terms approximates \ln(1 + x) with controllable precision, as the error decreases monotonically with additional terms within the convergence interval. To compute logarithms in arbitrary base b > 0, b \neq 1, the change-of-base formula applies: \log_b x = \frac{\ln x}{\ln b}, where \ln x uses the around x = 1 via \ln x = \ln(1 + (x - 1)) for $0 < x < 2, and \ln b is precomputed similarly or exactly if b is a constant like 10 or e. This approach was instrumental in 17th-century numerical methods, allowing extension of the natural logarithm series to common logarithms used in astronomy and .

Iterative Approximations

Iterative methods for approximating logarithms provide efficient ways to compute values through successive refinements, often converging faster than direct series expansions for certain ranges. These techniques are particularly valuable for high-precision calculations and were instrumental in early computational contexts where resources were limited. The arithmetic-geometric mean (AGM) iteration offers a powerful approach to approximate the natural logarithm, especially for values away from 1. Developed by in the early 19th century, the AGM computes the common limit M(a,b) of iteratively defined arithmetic and geometric means starting from initial values a_0 and b_0. For logarithms, adaptations like Borchardt's algorithm or relations to elliptic integrals enable efficient evaluation; for example, to compute \ln x for x > 1, suitable initial values can be chosen to relate the AGM to the logarithm via integral representations, achieving quadratic convergence with few iterations. These methods minimize arithmetic operations and are used in arbitrary-precision libraries for high accuracy. Halley's method, a third-order iterative derived from rational approximations to the logarithmic series, provides another robust framework for root-finding in the context of inverse exponential functions, effectively computing logarithms. For the natural logarithm, it can be applied by solving e^y = z for y = \ln z, using the y_{n+1} = y_n + \frac{2(z - e^{y_n})}{z + e^{y_n}}, which originates from Halley's 1694 work on rational series expansions and achieves cubic convergence. This method excels in scenarios requiring rapid refinement from an initial guess, such as a partial sum, and has been adapted for numerical libraries due to its superior error reduction compared to . For high-precision computation of specific constants like \ln 2, binary splitting emerges as an efficient iterative that recursively divides the evaluation of the series \ln 2 = \sum_{k=1}^\infty \frac{(-1)^{k+1}}{k} into binary subproducts, minimizing intermediate precision loss. Proposed by in 1993 for multiple-precision , it constructs factorials and coefficients through a tree-like : for a [a, b], the partial is computed as S(a, b) = S(a, m) + S(m+1, b) where m = \lfloor (a+b)/2 \rfloor, paired with a denominator product that avoids full recomputation, enabling thousands of digits with near-linear time scaling in precision. This technique proved essential for early computer-based verifications of logarithmic constants and remains a staple in arbitrary-precision software. These iterative approximations were particularly efficient for hand calculations or early electronic computers, where the AGM and splitting could yield 10-20 places with under 20 iterations, far outperforming manual table lookups in speed and accuracy for computations.

Specialized Algorithms

One specialized for computing the natural logarithm leverages the fundamental property that the derivative of \ln x is $1/x, allowing the logarithm to be obtained through of this . Specifically, \ln 2 = \int_1^2 \frac{1}{x} \, dx, which can be approximated using the to estimate the area under the curve of $1/x from to 2. The trapezoidal rule divides the interval into n subintervals of width h = 1/n and approximates the integral as \frac{h}{2} \left( f(1) + 2 \sum_{i=1}^{n-1} f(1 + i h) + f(2) \right), where f(x) = 1/x; for small n, this yields \ln 2 \approx 0.693. This method provides a straightforward way to generate initial approximations without relying on series expansions. Spigot algorithms offer a digit-by-digit computation of logarithms, generating each decimal (or binary) digit independently without requiring the full precision of previous digits, which is particularly useful for high-precision calculations on limited resources. These algorithms typically transform the logarithm into a form amenable to modular arithmetic or series manipulation, such as expressing \ln(1 + x) in a way that allows sequential digit extraction via carry-over operations similar to those in spigot methods for \pi. For instance, a spigot algorithm for \ln 2 can produce digits to arbitrary precision by iteratively refining an array representing the fractional part, avoiding the storage overhead of traditional methods. This approach ensures reliable computation even for large digit counts, as demonstrated in implementations that achieve exact digit isolation through rational approximations. In hardware implementations, logarithms in floating-point units are often computed using a combination of table lookup for the leading bits and a correction term for the mantissa to achieve efficient, high-speed evaluation. The input floating-point number is first normalized, with the exponent providing a scaled logarithm contribution, while the mantissa is used to index a small storing precomputed logarithm values for a reduced range (e.g., [1, 2)). A linear or quadratic correction polynomial is then applied to interpolate the exact value, minimizing table size while maintaining IEEE 754-compliant accuracy. This hybrid method balances latency and resource usage in processors, with table sizes typically around 2-8 for single-precision operations. Post-2000 optimizations for logarithm computation in systems have focused on low-power, high-efficiency approximations that prioritize , often through bit-level manipulations and reduced-precision tables tailored to resource-constrained environments. For example, advanced exponent bit techniques decompose the input into components for rapid estimation, followed by minimal correction steps to ensure bounded and avoid issues in . These methods achieve up to 20-30% latency reductions compared to standard iterative approximations while preserving stability across wide dynamic ranges, as verified in FPGA prototypes for . Such innovations are critical for battery-powered devices, where they enable logarithm operations with errors below 1 ulp without excessive power draw.

Applications

Logarithmic Scales

Logarithmic scales are measurement systems in which the positions or values are determined by the logarithms of the quantities being represented, such that equal intervals on the scale correspond to constant multiplicative factors in the original values. This approach compresses wide-ranging data into a more manageable , making it easier to compare quantities that differ by orders of magnitude. A prominent example is the for earthquake magnitudes, which measures the logarithm base 10 of the maximum of seismic waves recorded by a seismograph, adjusted for distance; each whole-number increase on the scale indicates a tenfold increase in wave and approximately 31.6 times more energy release. In acoustics, the (dB) scale quantifies levels using the formula L_P = 10 \log_{10} \left( \frac{P}{P_0} \right) \ \text{dB}, where P is the measured power and P_0 is a reference power (often $10^{-12} W/m² for sound in air); this logarithmic basis reflects the human ear's nonlinear perception of loudness, with a 10 dB increase corresponding to a tenfold power ratio. Similarly, the pH scale in chemistry measures acidity as \text{pH} = -\log_{10} [\text{H}^+], where [\text{H}^+] is the molar concentration of hydrogen ions; a one-unit decrease in pH represents a tenfold increase in acidity. The primary advantage of logarithmic scales lies in their ability to handle phenomena spanning vast ranges, such as physical sizes from dimensions (around $10^{-10} m) to interstellar distances (up to $10^{21} m), without distorting relative differences at either extreme. This is particularly valuable in fields dealing with or decay, where linear scales would either compress small values or exaggerate large ones, leading to loss of detail. Base-10 logarithms are commonly employed for their alignment with decimal notation, facilitating intuitive interpretation. In graphical representations, semilog plots apply a to one (typically the dependent ) and a to the other, which linearizes trends for easier analysis. Log-log plots, by contrast, use logarithmic scales on both axes, transforming power-law relationships (e.g., y = k x^n) into straight lines with n, aiding in the identification of behaviors across datasets. These plotting techniques enhance the of logarithmic scales by revealing underlying mathematical structures that might be obscured in linear formats.

Natural and Social Sciences

In physics, logarithms are essential for modeling processes, such as or the cooling of objects, where the t_{1/2}, the time for the quantity to reduce by half, is given by t_{1/2} = \frac{\ln 2}{\lambda} with \lambda as the . This arises from the N(t) = N_0 e^{-\lambda t}, precise predictions of over time. In astronomy, the stellar magnitude scale quantifies star logarithmically, where the difference in magnitudes between two stars is m_1 - m_2 = -2.5 \log_{10} \left( \frac{F_1}{F_2} \right), with F representing ; this compresses the vast range of observed intensities into manageable numerical values. Biological growth models often incorporate logarithms to describe exponential phases, as in bacterial where follows N(t) = N_0 e^{rt}, and the growth rate r is derived using natural logs to linearize for . The for earthquakes uses a base-10 logarithm of , such that each whole-number increase represents a tenfold change in and about 31 times more energy release, facilitating comparison of event severities. , which predicts that leading digits in naturally occurring numerical datasets follow a — with digit 1 appearing about 30% of the time— aids in detecting anomalies in biological or ecological compilations. In chemistry, the pH scale defines acidity as \mathrm{pH} = -\log_{10} [\mathrm{H}^+], where [\mathrm{H}^+] is the concentration in moles per liter, allowing a compact representation of concentration spans from 10^{-14} to 10^0. For chemical equilibria, the K relates to the standard free energy change via \ln K = -\frac{\Delta G^\circ}{RT}, with R as the and T as temperature, providing a thermodynamic basis for reaction feasibility. Social sciences apply logarithms to approximate inequality measures like the in lognormal income distributions, where the coefficient G can be estimated as G = 2\Phi\left( \frac{\sigma}{\sqrt{2}} \right) - 1 with \sigma as the standard deviation of log incomes, capturing wealth disparities efficiently. Population growth models in often use semi-logarithmic plots to reveal trends, as the relative growth rate is the slope of \ln P(t) versus time, aiding forecasts of societal expansion. In , the Weber-Fechner law posits that perceived stimulus intensity \psi is proportional to the logarithm of physical intensity I, expressed as \psi = k \log I, explaining why sensory responses diminish relatively as stimuli strengthen, as observed in weight perception or sound loudness.

Information and Complexity

In probability and statistics, the logarithm plays a central role in (MLE), where the is maximized to find parameter estimates that best explain observed data. The is the natural logarithm of the , which simplifies optimization because the logarithm is a monotonic transformation that preserves the location of maxima while converting products into sums. This approach was formalized by Ronald A. Fisher in his seminal 1922 paper, where he demonstrated that MLE provides efficient estimators under certain regularity conditions. The log-normal distribution arises when the logarithm of a random variable follows a normal distribution, making it suitable for modeling positively skewed data such as stock prices or particle sizes, where multiplicative processes dominate. If X is log-normally distributed, then \ln X \sim \mathcal{N}(\mu, \sigma^2), and the mean and variance of X involve exponential terms of \mu and \sigma. This distribution's properties were comprehensively analyzed in early works, with key characterizations appearing in McAlister's 1879 study on variable proportionality, highlighting its applicability to phenomena driven by independent multiplicative factors. In information theory, the logarithm quantifies uncertainty through Shannon entropy, defined as H = -\sum_i p_i \log_2 p_i, where p_i are the probabilities of outcomes in a discrete , measuring the average information content in bits. A positive entropy value indicates randomness, with maximum entropy occurring for uniform distributions; this metric underpins data compression and communication limits. The formula was introduced by in his 1948 paper, establishing the foundation for quantifying information in noisy channels. Computational complexity theory employs logarithms to describe efficient runtimes, particularly the in for operations like binary search, which achieves O(\log n) time by halving the search space at each step. This logarithmic scaling reflects divide-and-conquer strategies in and tree traversals, where the number of iterations grows slowly with input size n. Such analyses are standard in design, as detailed in foundational texts on computational efficiency. In , Lyapunov exponents characterize the rate of divergence of nearby trajectories in dynamical systems, defined as \lambda = \lim_{t \to \infty} \frac{1}{t} \log \frac{|\delta(t)|}{|\delta(0)|}, where \delta measures separation; positive values signal behavior through exponential to initial conditions. The largest exponent determines overall , with sums over all directions relating to phase-space volume contraction. This concept originated in Lyapunov's 1892 stability analysis and was adapted to chaos in modern . For fractals, the extends topological dimension to non-integer values using logarithmic scaling, approximated for self-similar sets as D = \frac{\log N}{\log (1/s)}, where N is the number of scaled copies and s is the similarity ratio. This measures roughness by how measure changes under rescaling, yielding values like \log 2 / \log 3 \approx 0.631 for the . The dimension was introduced by in 1919, providing a rigorous measure for irregular geometric objects.

Arts and Number Theory

In music theory, the perception of pitch is logarithmic, with intervals defined by ratios of frequencies rather than absolute differences, allowing consistent scaling across octaves. An octave corresponds to a frequency doubling, equivalent to a base-2 logarithm where the interval measure is \log_2 (f_2 / f_1), ensuring that equal steps in logarithmic space produce perceptually uniform pitch changes. This logarithmic foundation underpins musical scales, as stacking intervals multiplies frequency ratios, which adds in logarithmic measures. Equal temperament tuning divides the into 12 equal s, each with a ratio of $2^{1/12}, approximating while enabling modulation across keys without retuning. This system, formalized in the and widely adopted by the 18th, uses the of to balance harmonic purity and practicality, with the interval measuring approximately 100 s on a logarithmic where 1 equals $2^{1/1200}. In , logarithmic projections facilitate scaling and perspective by compressing vast ranges into finite representations, mimicking human through exponential transformations. For instance, M.C. Escher's Path of Life III (1962) employs logarithmic spirals governed by r = (5/2)^{-3\pi\theta} to depict interlocking paths, creating a with a of approximately 2.469 that warps an infinite tiling into a circular frame. This logarithmic scaling produces an illusion of radial depth without traditional linear perspective, as the spirals intersect at angles near 98.4°, evoking projective geometry on conical or spherical surfaces. The states that the number of primes up to x, denoted \pi(x), is asymptotically \pi(x) \sim x / \ln x, providing the leading term for prime distribution density. Conjectured by around 1792 based on numerical evidence and independently by in 1808, the theorem was rigorously proved in 1896 by and Charles Jean de la Vallée Poussin using . The natural logarithm here captures the thinning of primes, with \ln x growing slowly to reflect their sparsity. The \zeta(s) = \sum_{n=1}^\infty n^{-s} for \Re(s) > 1, extended analytically, encodes prime through its Euler product \zeta(s) = \prod_p (1 - p^{-s})^{-1}, where the non-trivial zeros in the critical strip influence oscillations in \pi(x) via the explicit formula involving logarithmic derivatives. The locations of these zeros, conjectured by the to lie on \Re(s) = 1/2, refine the error term in the to O(\sqrt{x} \log x), linking prime gaps to logarithmic scales in the . Discrete logarithms arise in as the inverse of in finite fields: given a prime p, g, and h = g^x \mod p, solving for x is computationally hard for large p, underpinning cryptographic protocols. In the Diffie-Hellman (1976), parties agree on a g^{ab} \mod p without transmitting it, relying on the problem's intractability to prevent adversaries from computing x in g^x \equiv h \mod p. This hardness, although subexponential-time algorithms exist, remains computationally infeasible for the large parameters used in cryptographic applications, securing modern public-key systems. Mertens' theorems provide asymptotics for products over primes, including the third theorem: \prod_{p \leq x} (1 - 1/p) \sim e^{-\gamma} / \ln x, where \gamma \approx 0.57721 is the Euler-Mascheroni , quantifying the harmonic series tied to primes. Proved by Franz Mertens in using estimates, this result implies the sum of reciprocals of primes up to x is \sim \ln \ln x + B, with B \approx 0.261497 (), for methods and approximations.

Generalizations

Complex Logarithms

The complex logarithm extends the real logarithm to nonzero complex numbers z \in \mathbb{C} \setminus \{0\}. It is defined as \log(z) = \ln |z| + i \arg(z), where \ln denotes the natural logarithm of a positive real number and \arg(z) is the argument of z. The principal branch, often denoted \operatorname{Log}(z), restricts the argument to the interval (-\pi, \pi], ensuring a single-valued function on the complex plane minus the non-positive real axis. Due to the periodicity of the argument, the complex logarithm is multi-valued: \log(z) = \ln |z| + i (\arg(z) + 2\pi k) for any integer k. To define a single-valued branch, a branch cut is introduced, typically along the negative real axis from 0 to -\infty, where the function experiences a discontinuity of $2\pi i. This cut connects the branch points at z = 0 and z = \infty, preventing closed paths that encircle these points and cause the argument to wind indefinitely. As the inverse of the complex exponential, the principal logarithm satisfies \exp(\operatorname{Log}(z)) = z for all z in its domain. However, the reverse identity holds only modulo $2\pi i: \operatorname{Log}(\exp(w)) = w + 2\pi i k for some k depending on the imaginary part of w, reflecting the multi-valued nature. In , the is essential for defining powers of complex numbers via z^w = \exp(w \log(z)), which yields multiple values corresponding to the branches of the logarithm. This construction facilitates solving equations like finding all roots or evaluating non-integer exponents in the . Deformed exponentials in nonextensive statistical mechanics, such as the q-exponential e_q(z) = [1 + (1-q)z]^{1/(1-q)} for q \neq 1, have as their inverse the q-logarithm \ln_q(x) = \frac{x^{1-q} - 1}{1-q}, which recovers the natural logarithm in the limit q \to 1. These functions deform the standard exponential structure to accommodate q-deformations in algebraic settings like quantum algebras. Separately, the Jackson q-exponential, defined by e_q(z) = \sum_{k=0}^{\infty} \frac{z^k}{(q;q)_k} where (q;q)_k is the q-Pochhammer symbol, arises in the representation theory of quantum groups as a q-analogue of classical special functions, but has a different series-based inverse. In , the p-adic logarithm is defined on principal units of p-adic fields, specifically for x \in K with |x - 1|_p < 1, as the power series \log_p(1 + s) = \sum_{n=1}^{\infty} (-1)^{n+1} \frac{s^n}{n}, which converges due to the non-archimedean valuation. This logarithm maps the of 1-units isomorphically onto an additive subgroup of the p-adic field, facilitating the study of local and over p-adic completions. Unlike the real logarithm, it satisfies \log_p(xy) = \log_p(x) + \log_p(y) only within the domain of convergence, and its kernel includes roots of unity of order coprime to p. Related concepts include the , denoted \log^* n, which counts the number of iterated applications of the logarithm (base 2 or e) to n until the result is at most 1, and appears in the amortized of -find data structures for dynamic graph connectivity in and algorithms. Specifically, with path compression and by , the find operation achieves O(\alpha(n)) time, where \alpha(n) is the inverse , asymptotically equivalent to \log^* n and bounding the height growth in disjoint-set forests. Another extension is the superlogarithm, the inverse of {}^y a = a \uparrow\uparrow y, defined such that \operatorname{slog}_a(x) = y {}^y a = x, providing a way to solve for the height in sequences beyond . Generalized logarithms encompass the q-logarithm in Tsallis entropy, where the nonextensive entropy S_q = k \frac{1 - \sum_i p_i^q}{q-1} = -k \sum_i p_i^q \ln_q (p_i) uses the q-log to measure information in systems with long-range correlations, differing from Shannon entropy by its non-additivity: S_q(p \oplus q) \neq S_q(p) + S_q(q) for non-independent systems. Similarly, the log*-function in , as the , quantifies the slow growth in algorithmic bounds for problems like minimum spanning trees or , where it replaces polylogarithmic factors in nearly linear-time solutions. These variants highlight non-additivity as a key departure from standard logarithms, where properties like \log(ab) = \log a + \log b fail under deformations, reflecting altered algebraic structures in their domains.

References

  1. [1]
    Intro to Logarithms (article) - Khan Academy
    What is a logarithm? ... Logarithms are another way of thinking about exponents. For example, we know that 2 ‍ raised to the 4 th ‍ power equals 16 ‍ . This is ...
  2. [2]
    1.6 Exponential and Logarithm Functions - Dartmouth Mathematics
    The definition of a general exponential, including its graph. The definition of a general logarithm, including its graph. Laws of exponents. Laws of logarithms.<|control11|><|separator|>
  3. [3]
    Math Skills - Logarithms - Chemistry
    The power to which a base of 10 must be raised to obtain a number is called the common logarithm (log) of the number.Missing: definition | Show results with:definition
  4. [4]
    [PDF] Chopping Logs: A Look at the History and Uses of Logarithms
    Logarithms have been a part of mathematics for several centuries, but the concept of a logarithm has changed notably over the years.
  5. [5]
    [PDF] NHTI Learning Center Math Lab G-25 Rules for Logs
    Logarithms are used in chemistry and soil science (pH levels), and in biomathematics and ecology (growth rates) and geology (Richter scale).
  6. [6]
    How do I use logarithms? Logarithms (logs) in the Earth sciences
    Aug 16, 2024 · Logarithms quantify large data ranges, like in pH and earthquake scales. They are used in a 'plug and chug' approach, often with a calculator.
  7. [7]
    John Napier | Biography, Invention, Logarithms, Bones, & Facts
    Sep 26, 2025 · Invented in the 17th century to speed up calculations, logarithms vastly reduced the time required for multiplying numbers with many digits.
  8. [8]
    Inverse Properties of Logarithmic Functions | CK-12 Foundation
    By the definition of a logarithm, it is the inverse of an exponent. Therefore, a logarithmic function is the inverse of an exponential function.Missing: motivation | Show results with:motivation
  9. [9]
  10. [10]
    Logarithm -- from Wolfram MathWorld
    The logarithm log_bx for a base b and a number x is defined to be the inverse function of taking b to the power x, ie, b^x.
  11. [11]
    Common Logarithm -- from Wolfram MathWorld
    The common logarithm is the logarithm to base 10. The notation logx is used by physicists, engineers, and calculator keypads to denote the common logarithm.
  12. [12]
    Natural Logarithm -- from Wolfram MathWorld
    The natural logarithm lnx is the logarithm having base e, where e=2.718281828. This function can be defined lnx=int_1^x(dt)/t for x>0.
  13. [13]
    Binary Logarithm -- from Wolfram MathWorld
    The binary logarithm log_2x is the logarithm to base 2. The notation lgx is sometimes used to denote this function in number theoretic literature.Missing: definition | Show results with:definition<|control11|><|separator|>
  14. [14]
    DLMF: §4.2 Definitions ‣ Logarithm, Exponential, Powers ‣ Chapter ...
    The general logarithm function Ln ⁡ z is defined by where the integration path does not intersect the origin. This is a multivalued function of z with branch ...
  15. [15]
    logarithm - PlanetMath
    Mar 22, 2013 · There are a number of basic algebraic identities involving logarithms. logx(yz) log x ⁡ ( y ⁢ z ), =logx(y)+logx(z) = log x ⁡ ( y ) + log x ⁡ ( ...
  16. [16]
    Intro to logarithms (video) - Khan Academy
    Jul 19, 2019 · Yup! In math, log usually means base 10, as that is the base system we normally use. ln is another button you'd find on a scientific ...
  17. [17]
    4.5: Logarithmic Properties - Mathematics LibreTexts
    Oct 28, 2022 · Some important properties of logarithms are given here. First, the following properties are easy to prove.Using the Product Rule for... · Expanding Logarithmic...
  18. [18]
    Log rules: Justifying the logarithm properties (article) | Khan Academy
    In this lesson, we will prove three logarithm properties: the product rule, the quotient rule, and the power rule.
  19. [19]
    Proofs of Logarithm Properties or Rules - ChiliMath
    Use the exponent rules to prove logarithmic properties like Product Property, Quotient Property and Power Property. Learn the justification of these ...
  20. [20]
    Proofs of Logarithm Properties - OnlineMathLearning.com
    In these lessons, we will look at the four properties of logarithms and their proofs. They are the product rule, quotient rule, power rule and change of base ...
  21. [21]
    Algebra - Logarithm Functions - Pauls Online Math Notes
    Nov 16, 2022 · So, the common logarithm is simply the log base 10, except we drop the “base 10” part of the notation. Similarly, the natural logarithm is ...
  22. [22]
    3.12 Mathematical Constants ‣ Areas ‣ Chapter 3 Numerical Methods
    The constants discussed are π (3.14159...), e (2.71828...), and Euler's constant γ (0.57721...).
  23. [23]
    Calculus I - Logarithm Functions - Pauls Online Math Notes
    Jul 10, 2024 · In the natural logarithm the base e is the same number as in the natural exponential function that we saw in the last section. Here is a ...
  24. [24]
    Analysis of Algorithms: Lecture 4 - Texas Computer Science
    The logarithm base two comes up often in divide-and-conquer algorithms ... For example, the number of bits required to store a number n in binary is log2n.
  25. [25]
    Natural Logarithm of 2 -- from Wolfram MathWorld
    The natural logarithm of 2 is a transcendental quantity that arises often in decay problems, especially when half-lives are being converted to decay constants.
  26. [26]
    Logarithms: The Early History of a Familiar Function - John Napier ...
    He coined a term from the two ancient Greek terms logos, meaning proportion, and arithmos, meaning number; compounding them to produce the word “logarithm.” ...
  27. [27]
    [PDF] Napier's ideal construction of the logarithms - LOCOMAT
    Sep 12, 2012 · A precise definition of the logarithm was given by Napier as follows.8 He considered two lines (figure 1), with two points moving from left to ...
  28. [28]
    The making of the logarithm | plus.maths.org
    Jan 14, 2014 · Napier took the length of the line segment from A to B to be very large: 10 , 000 , 000 = 10 7 units. He did this to ensure accuracy and ...
  29. [29]
    [PDF] A reconstruction of the tables of Briggs' Arithmetica logarithmica (1624)
    Jan 11, 2011 · After Napier's publication, Briggs went to visit him in Scotland in the summers of 1615 and 1616 and they agreed on the need to reformulate the ...
  30. [30]
    Mathematical Treasure: Arithmetica Logarithmica of Henry Briggs
    Briggs published his first extensive table of these logarithms in 1624 in Arithmetica Logarithmica. His table contained the logarithms of 30,000 natural numbers ...
  31. [31]
    [PDF] JOHN NAPIER (1550 - University of St Andrews
    There is no doubt that the introduction of logarithms has spurred the development of science. JOHANNES KEPLER recognized the tremendous advantage of his ...
  32. [32]
    John Napier - Biography - MacTutor - University of St Andrews
    John Napier was a Scottish scholar who is best known for his invention of logarithms, but other mathematical contributions include a mnemonic for formulas used ...
  33. [33]
    Jost Bürgi - Biography - MacTutor - University of St Andrews
    Jost Bürgi was a Swiss mathematician who discovered logarithms independently of the Scottish mathematician Napier. ... 1620. Bürgi's method is different ...
  34. [34]
    [PDF] Jost B¨urgi and the discovery of the logarithms - ETH Zürich
    It seems to be fair, therefore, to consider Jost Bürgi and John Napier as the two simultaneous and independent discoverers of the logarithms, see also Clark and ...
  35. [35]
    Henry Briggs (1561 - 1630) - Biography - MacTutor
    Briggs's mathematical treatise Arithmetica Logarithmica was published in 1624. This gave the logarithms of the natural numbers from 1 to 20,000 and 90,000 to ...
  36. [36]
    Gregory of Saint-Vincent (1584 - 1667) - Biography - MacTutor
    Saint-Vincent worked on mathematics and developed methods which were important in setting the scene for the invention of the differential and integral calculus.
  37. [37]
    Gregory of St. Vincent and the Rectangular Hyperbola - ResearchGate
    Aug 5, 2025 · The calculation of the area under the hyperbola y = 1/t was carried out by Gregory of St. Vincent in 1647 [24], showing that The logarithmic ...
  38. [38]
    [PDF] An Appreciation of Euler's Formula - Rose-Hulman Scholar
    In the book Introductio in Analysin Infinitorum published in 1748, Euler gives us a lot of the notation crucial to mathematics. Before this publication, there ...<|separator|>
  39. [39]
    The History of the Derivation of Euler's Number
    Leonhard Euler Goes on to Make Various Discoveries Regarding e—1748. Leonhard Euler had made several discoveries of the properties of e in the years that ...
  40. [40]
    History of the Exponential and Logarithmic Concepts - jstor
    At this stage the logarithmic function is introduced to great advantage through the relation zmIn = em/n log z. Here the n-valued expression zmIn becomes a one- ...
  41. [41]
    Von Lindemann's Proof Made Accessible to Today's Undergraduates
    Feb 13, 2015 · The proof that pi is a transcendental number, first provided by Carl Louis Ferdinand von Lindemann in 1882, was and remains one of the most celebrated results ...
  42. [42]
    Analytic proof of the Lindemann theorem - MSP
    This discussion is preceded by a proof that e is transcendental, which con- tains the central idea of the proof of the Lindemann theorem, and yet is not.
  43. [43]
    Adriaan Vlacq - Biography - MacTutor - University of St Andrews
    In 1628 Vlacq republished the 10 decimal place logarithm tables as Arithmetica logarithma sive logarithmorum chiliades tentum, pro numeris naturali serie ...Missing: 1630s | Show results with:1630s
  44. [44]
    The logarithmic tables of Edward Sang and his daughters
    Edward Sang (1805–1890), aided only by his daughters Flora and Jane, compiled vast logarithmic and other mathematical tables.
  45. [45]
    Using a table of logarithms - Applied Mathematics Consulting
    Jun 3, 2024 · Interpolation is a kind of compression. A&S would be 100 billion times larger if it tabulated functions at 15 figure inputs. Instead, it ...
  46. [46]
    Mathematical Treasure: Babbage's Tables of Logarithms
    Maritime navigation required accurate logarithmic tables. Babbage undertook this task and published his tables in 1827. The images shown are from the 1872 ...Missing: surveying | Show results with:surveying
  47. [47]
    Slide Rule History - The Oughtred Society
    Dec 27, 2021 · William Oughtred, an Anglican minister, today recognized as the inventor of the slide rule, places two such scales side by side and slides them to read the ...
  48. [48]
    The Slide Rule - IEEE Pulse
    May 28, 2019 · This design, first created by William Oughtred in 1633, saw widespread use well into the 19th century. In 1722, John Warner, a London instrument ...
  49. [49]
    Historical Notes: Oughtred and the Slide Rule - IMA
    Feb 11, 2022 · William Oughtred, the inventor of the slide rule, was born at Eton, Berkshire, circa 1574. His father appears to have been a junior master at Eton College.<|control11|><|separator|>
  50. [50]
    Linear Slide Rules | Smithsonian Institution
    Between 1614 and 1622, John Napier discovered logarithms, Edmund Gunter devised a scale on which numerals could be multiplied and divided by measuring the ...Missing: sources | Show results with:sources
  51. [51]
    3 Slide Rule ABC's and D's
    Slide rules consist of sets of logarithmic scales that are used to add logarithms of numbers in order to perform multiplication and division, as well as other ...Missing: mechanism | Show results with:mechanism
  52. [52]
    How and Why a Slide Rule Works - Sphere Research Corporation
    Noting how the scales on a slide rule simply repeat themselves, one can put the 10 instead of the 1 under the number by which one wishes to multiply to see the ...<|control11|><|separator|>
  53. [53]
    Eric's Types of Slide Rules and their Scales
    Sep 12, 2002 · All slide rules consist of logarithmic scales that can be moved in relation to each other in order to do basic mathematical calculations.
  54. [54]
    Trigonometry on the Slide Rule - by Mike Syphers - Substack
    Nov 30, 2024 · So by the time Mannheim slide rules were first developed in the mid-1800s, the inclusion of sine and tangent scales was almost immediate.Missing: specialized | Show results with:specialized
  55. [55]
    Slide Rules - CHM Revolution - Computer History Museum
    Metal slide rule. A rule for calculating surveying angles and distances. Slide rules have been made from wood, steel, aluminum, bamboo, and plastic.
  56. [56]
    2.972 How A Slide Rule Works - MIT
    The scales (A-D) are labeled on the left-hand side of the slide rule. The number of scales on a slide rule vary depending on the number of mathematical ...
  57. [57]
    [PDF] with a, b, c, d, ci, k, s, t, l scales - International Slide Rule Museum
    This slide rule has scales A, B, C, D, CI, K, S, T, and L. It solves multiplication, division, and other math problems by comparing scales.
  58. [58]
    An old-school tool revisited - College of Engineering at Carnegie ...
    Aug 26, 2025 · Before handheld calculators, engineers relied on slide rules. These mechanical analog computers helped get us go to the moon.Missing: peak 20th century 1970s
  59. [59]
    Slide Rule Nostalgia | American Scientist
    By the late 1970s the cost of the scientific calculators had fallen to the $20–$30 range, and we were soon able to convince him that it would be all right to ...
  60. [60]
  61. [61]
    8.1 Kepler, Napier, and the Third Law - MathPages
    By the 18th of May, 1618, Kepler had fully grasped the logarithmic pattern in the planetary orbits: Now, because 18 months ago the first dawn, three months ...
  62. [62]
    How Mathematical Logarithms Aided the Royal Navy - News
    Aug 15, 2013 · Gunners in the English navy applied logarithms to simplify their calculations and improve their aim and effectiveness.
  63. [63]
    How to Survey Land with 18th Century Tools — Using the Sector
    Jan 15, 2022 · For example, linear, trigonometric, and logarithmic scales were engraved to the two bars. Computations work by solving problems with similar ...
  64. [64]
    Critical developments in land surveying - Coordinates
    Computational methods were enhanced with the use of logarithmic tables introduced by Napier in the 17th century, and supplemented by hand-powered rotational ...
  65. [65]
  66. [66]
    Remarkable numerical calculations before electronic computers
    Mar 22, 2019 · Halley's original prediction had been for 1758, based on his assessment that the comet ... Both of them calculated huge logarithm tables by hand:.
  67. [67]
    [PDF] The Fundamental Theorem of Calculus and integration methods
    Oct 12, 2016 · In fact, log is differentiable, hence continuous, so by the Intermediate. Value Theorem, it suffices to check that log x is unbounded.
  68. [68]
    [PDF] Chapter 8 Logarithms and Exponentials: logx and e
    Here, we are going to use our knowledge of the Fun- damental Theorem of Calculus and the Inverse Function Theorem to develop the properties of the Logarithm ...<|separator|>
  69. [69]
    5.3 Properties and Graphs of Logarithmic Functions
    5.3 Properties and Graphs of Logarithmic Functions. In Section 5.2, we saw exponential functions f(x) = b^x are one-to-one which means they are invertible.
  70. [70]
    4.2 - Logarithmic Functions and Their Graphs
    The inverse of an exponential function is a logarithmic function. Remember that the inverse of a function is obtained by switching the x and y coordinates.
  71. [71]
    7.1: The Logarithm Defined as an Integral
    ### Summary of Natural Logarithm Definition and Properties
  72. [72]
    4.3 - Properties of Logarithms
    Logarithms are exponents, and when you multiply, you're going to add the logarithms. The log of a product is the sum of the logs.Missing: notation | Show results with:notation
  73. [73]
    [PDF] 2 Exponents and logarithms - Penn Math
    Exponentials grow at different rates and every exponential grows faster than every power. 3. Logarithms grow so slowly that any power of ln x is less than any ...
  74. [74]
    Change of Base formula - Ximera - Xronos
    This leads us to the 'change of base' formula for logarithms. Write the logarithm as a log with a base of We want to rewrite the log above as something.
  75. [75]
    [PDF] Introducing ex - MIT Mathematics
    The whole purpose of calculus is to understand change. It is wonderful to ... Johann Bernoulli connected logarithms to exponential series in 1697 Œ1Ќ: And by 1751 ...
  76. [76]
    [PDF] ln x dx - MIT OpenCourseWare
    If we choose v = x then v = 1 and: . ln x dx = uv dx. The formula for integration by parts is: uv dx = uv - u v dx. So by plugging in u = ln x and v = x ...
  77. [77]
    proof of Lindemann-Weierstrass theorem and that e and π - π
    Mar 22, 2013 · ▫. Corollary 6. If α>0 α > 0 is algebraic with α≠1 α ≠ 1 , then lnα ln ⁡ α is transcendental. Proof. If β=lnα β = ln ⁡ α , then eβ=α e β = α .
  78. [78]
    [PDF] Trapezoid Rule Approximation of / 2 - MIT
    We can use a calculator to find that this value is approximately 0.693147. Numerical methods will allow us to estimate this with accuracy about equal.
  79. [79]
    Richard Feynman and The Connection Machine
    ### Summary of Richard Feynman's Method for Computing Logarithms
  80. [80]
    Spigot Algorithm and Reliable Computation of Natural Logarithm
    A SPIGOT ALGORITHM FOR THE DIGITS OF PI · Mathematics · 1995. It is remarkable that the algorithm illustrated in Table 1, which uses no floating-point ...
  81. [81]
    [PDF] Hardware Implementation of the Logarithm Function - LTH/EIT
    This thesis presents a hardware design using Improved Parabolic Synthesis to approximate the fractional part of the base two logarithm function.
  82. [82]
    A fast hardware approach for approximate, efficient logarithm and ...
    In this paper, we present an approach to compute log() and antilog() in hardware. Our approach is based on a table lookup, followed by an interpolation step.Missing: post | Show results with:post<|control11|><|separator|>
  83. [83]
    A New Fast Logarithm Algorithm Using Advanced Exponent Bit ...
    Dec 30, 2022 · In this paper, we present a new method to efficiently implement a logarithm operation based on exponent bit extraction.Missing: post | Show results with:post
  84. [84]
    [PDF] Computing floating-point logarithms with fixed-point operations - HAL
    Nov 12, 2015 · This paper explores computing floating-point logarithms using fixed-point arithmetic, re-evaluating its relevance, and achieving performance on ...Missing: lookup correction
  85. [85]
    [PDF] low precision logarithmic number systems: beyond base-2 - arXiv
    Feb 12, 2021 · We optimize hardware for low-precision LNS arithmetic by selecting a base with favourable add and subtract truth tables. We achieve large ...
  86. [86]
    [PDF] Logarithmic Scales - Math 141
    A logarithmic scale represents numbers by their logarithms, useful when physical quantities have a very large variance.
  87. [87]
    Earthquake Magnitude, Energy Release, and Shaking Intensity
    The Richter magnitude of an earthquake is determined from the logarithm of the amplitude of waves recorded by seismographs. Adjustments are included for the ...
  88. [88]
    NIST Guide to the SI, Chapter 8
    Jan 28, 2016 · NIST Guide to the SI, Appendix A: Definitions of the SI Base Units ... decibel, symbol dB (1 dB = 0.1 B). Level-of-a-field-quantity is ...
  89. [89]
    pH Scale | U.S. Geological Survey - USGS.gov
    Jun 19, 2019 · pH is reported in "logarithmic units". Each number represents a 10-fold change in the acidity/basicness of the water. Water with a pH of five is ...
  90. [90]
    A refresher on logs
    Logarithmic scales allow one to examine values that span many orders of magnitude without losing information on the smaller scales.
  91. [91]
    6. Slopes on logarithmic graph paper.
    Semi-log paper has a logarithmic scale on one axis and a linear scale on the other; log-log paper has logarithmic scales on both axes. The logarithmic scale has ...
  92. [92]
    0.2 Exponential and Logarithmic Functions
    The half-life of an exponentially decaying quantity is the time required for the quantity to be reduced by a factor of one half.
  93. [93]
  94. [94]
    5: The Magnitude Scale - Physics LibreTexts
    Feb 18, 2025 · As you will see below, the magnitude system is logarithmic, which turns the huge range in brightness ratios into a much smaller range in ...
  95. [95]
    [PDF] Modeling with Exponential and Logarithmic Functions - Mathematics
    Goal: Many processes that occur in nature, such as population growth, radioactive decay, heat diffusion, can be modeled using exponential functions. Logarithmic ...
  96. [96]
    6.2: Logarithmic Scales in Natural Sciences - Mathematics LibreTexts
    Sep 12, 2020 · The Richter Scale is a base-ten logarithmic scale. In other words, an earthquake of magnitude 8 is not twice as great as an earthquake of magnitude 4.Converting from Logarithmic to... · Using Common Logarithms · Try it Now 5
  97. [97]
    A concise proof of Benford's law - PMC - NIH
    In many natural and human phenomena, the distribution of the first significant digit of a random number follows a logarithm-type law.
  98. [98]
    A primer on pH - NOAA/PMEL
    Because the pH scale is logarithmic (pH = -log[H+]), a change of one pH unit corresponds to a ten-fold change in hydrogen ion concentration (Figure 1).
  99. [99]
    [PDF] Calculation of Equilibrium Constants for Isotopic Exchange Reactions
    The natural logarithm of the equilibrium con- stant, K, at any temperature T for any chemical reaction is the difference in the standard free energies of ...
  100. [100]
    Estimation of the Gini coefficient for the lognormal distribution of ...
    Jul 28, 2016 · Gini coefficient as a measure of inequality is widely used in various contexts such as energy, credit availability, income, health care and ...
  101. [101]
    Human population growth - Graphs and maps - Ined
    The graph is thus semi-logarithmic, i.e. linear along the x-axis and logarithmic along the y-axis. It shows the relative rate of population growth. A population ...
  102. [102]
    [PDF] Weber's Law and Fechner's Law Introduction
    Weber's law expresses a general relationship between a quantity or intensity of something and how much more needs to be added for us to be able to tell that ...
  103. [103]
    [PDF] A Mathematical Theory of Communication
    N(t) = N(t -t1) +N(t -t2) +•••+N(t -tn). 0 where X0 is the largest real solution of the characteristic equation: X,t1 +X,t2 +•••+X,tn = 1 3 Page 4 and ...
  104. [104]
    [PDF] The physics of musical scales: Theory and experiment
    Sep 26, 2015 · We hear pitch logarithmically, meaning that relationships between notes are defined by frequency ratios rather than frequency differences. For ...
  105. [105]
    Logarithmic Interval Measures - Stichting Huygens-Fokker
    Because stacking two intervals involves multiplication of frequency ratios, this is equivalent to addition of the logarithmic measures of these ratios. See ...
  106. [106]
    [PDF] PHY 103: Scales and Musical Temperament
    Sep 22, 2015 · 12 Tone Equal Temperament​​ ‣Better: use a multiplicative factor such that fn = an/12f ‣For f12 = 2f (one octave) we need a = 2. Therefore,
  107. [107]
    [PDF] Logarithmic Spirals and Projective Geometry in M.C. Escher's "Path ...
    Jan 1, 2012 · The two diagonals meet closer to the center of the woodcut than they would if the image were a perspective projection of a regularly tiled ...<|separator|>
  108. [108]
    Prime Number Theorem -- from Wolfram MathWorld
    The prime number theorem gives an asymptotic form for the prime counting function pi(n), which counts the number of primes less than some integer n.Missing: original | Show results with:original
  109. [109]
    [PDF] 2.3 Diffie–Hellman key exchange - Brown Math
    The Diffie–Hellman key exchange algorithm solves the following dilemma. Alice and Bob want to share a secret key for use in a symmetric cipher, but.
  110. [110]
    Mertens' theorems | What's new
    Dec 11, 2013 · Mertens' theorems are a set of classical estimates concerning the asymptotic distribution of the prime numbers.
  111. [111]
  112. [112]
    [PDF] The complex logarithm, exponential and power functions
    In order to define the complex logarithm, one must solve the complex equation: ... logarithm and complex power function are defined by their principal values.
  113. [113]
    [PDF] Branch Points and Branch Cuts (18.04, MIT). - MIT Mathematics
    Oct 11, 1999 · Figure 1.6: Cut Complex Plane. The value of log(z) at A (a point infinitesi- mally close to and above the positive x-axis) ...
  114. [114]
    [PDF] q-Special functions, an overview - arXiv
    Aug 5, 2023 · Abstract. This article gives a brief introduction to q-special functions, i.e., q-analogues of the clas- sical special functions.
  115. [115]
    [PDF] A possible deformed algebra and calculus inspired in nonextensive ...
    Aug 14, 2004 · We present a deformed algebra related to the q-exponential and the q-logarithm functions that emerge from nonextensive statistical mechanics ...
  116. [116]
    [PDF] arXiv:q-alg/9610002v1 1 Oct 1996
    Different generators of a deformed oscillator algebra give rise to one- parameter families of q-exponential functions and q-Hermite polyno-.
  117. [117]
    p-adic exponential and p-adic logarithm - PlanetMath.org
    Mar 22, 2013 · 1. If expp(s) exp p ⁡ ( s ) and expp(t) exp p ⁡ ( t ) are defined then expp(s+t)=expp(s)expp(t) exp p ⁡ ( s + t ) = exp p ⁡ ( s ) ⁢ exp p ⁡ ( t ) ...
  118. [118]
    [PDF] Infinite series in p-adic fields - Keith Conrad
    In these notes we will develop the theory of power series over complete nonarchimedean fields. Let K be a field that is complete with respect to a nontrivial ...
  119. [119]
    Disjoint Set Union - Algorithms for Competitive Programming
    Oct 12, 2024 · This article discusses the data structure Disjoint Set Union or DSU. Often it is also called Union Find because of its two main operations.
  120. [120]
    None
    ### Summary of Definition of Super-Logarithm as Inverse of Tetration