Fact-checked by Grok 2 weeks ago

Logistic function

The logistic function, also known as the , is a that produces an S-shaped , real-valued inputs to outputs between 0 and , and is commonly defined in its standard form as \sigma(x) = \frac{[1](/page/1)}{[1](/page/1) + e^{-x}}. This function initially grows slowly, then accelerates, and finally approaches an upper , making it ideal for modeling bounded growth processes. It satisfies the autonomous f'(x) = f(x)([1](/page/1) - f(x)), which describes self-limiting dynamics where the rate of change is proportional to both the current value and the remaining distance to the . The general form extends this to f(x) = \frac{L}{[1](/page/1) + e^{-k(x - x_0)}}, incorporating parameters for the maximum value L, steepness k, and x_0. Originally developed by Belgian mathematician Pierre-François Verhulst in 1838 as a model for population dynamics, the logistic function addressed limitations of exponential growth by incorporating density-dependent factors that limit expansion toward a carrying capacity K. Verhulst's logistic equation, \frac{dP}{dt} = rP\left(1 - \frac{P}{K}\right), yields the function as its solution, predicting an inflection point at P = K/2 where growth rate peaks. Rediscovered independently in 1920 by American biologists Raymond Pearl and Lowell Reed, who applied it to human population data from U.S. censuses, the model gained prominence for fitting empirical S-curves in demographics and ecology. Beyond biology and , the logistic function has broad applications across disciplines, including where it serves as an in neural networks and the core of for by estimating probabilities. In , it models technology adoption and market saturation; in chemistry, reaction kinetics; and in , the spread of infectious diseases under limited susceptible populations. Its properties—such as differentiability, monotonicity, and bounded range—make it a foundational tool in optimization and statistical modeling, with extensions like the influencing fields from physics to social sciences.

Definition and Basic Form

Standard Form

The standard logistic function is defined by the equation f(x) = \frac{L}{1 + e^{-k(x - x_0)}} where L > 0 represents the curve's maximum value, also known as the carrying capacity; k > 0 is the growth rate, determining the steepness of the curve; and x_0 is the x-coordinate of the midpoint, where f(x_0) = L/2. This function produces a characteristic S-shaped, or sigmoidal, that starts near zero for negative x, rises steeply around x_0, and flattens toward L for large positive x. The is bounded between 0 and L, with an at x = x_0 where the concavity changes and the growth rate is maximized. Asymptotically, f(x) approaches 0 as x \to -\infty and L as x \to \infty, reflecting bounded growth that levels off at the maximum. This form arises as the solution to the for , originally proposed by Pierre-François Verhulst in 1838 (detailed further in the History section).

Parameter Interpretations

In the general form of the logistic function, given by f(x) = \frac{L}{1 + e^{-k (x - x_0)}}, the parameter L represents the curve's upper asymptote, interpreted as the in population growth models, which denotes the maximum sustainable population size limited by environmental resources such as food and habitat. This concept originated with Pierre-François Verhulst's model for human , and in his 1845 analysis for , L (denoted as K) was estimated at approximately 6.584 million based on historical data fitting. The parameter k > 0 governs the steepness of the transition from the lower (near 0) to the upper (L), reflecting the intrinsic growth rate in biological contexts; a larger k indicates faster approach to the , as seen in Verhulst's equation where k (denoted as r) was fitted to 2.62% per year for Belgium's population. In statistical applications, such as for , k effectively scales the input features through the linear predictor, controlling the of the probability output to changes in x. The parameter x_0 specifies the or , where the function value is L/2 and the is maximized at kL/4, often representing the time or input value of maximum in ecological models. Biologically, this point signals the transition from accelerating to decelerating as the population nears half the . A common variation uses the standardized notation f(x) = \frac{1}{1 + e^{-x}}, where L = 1, k = 1, and x_0 = 0, simplifying analysis in and by mapping inputs to probabilities between 0 and 1 without explicit scaling parameters. In diverse domains, parameters adapt via units: in , x is time (e.g., years), L is count, and k has units of inverse time; in , the function is often dimensionless, with f(x) as a probability and parameters absorbed into the model's coefficients for probabilistic interpretation.

History

Early Developments

The logistic function originated in the work of Belgian mathematician Pierre François Verhulst, who introduced the model in 1838 as a model for bounded and named the curve "logistique" in 1845, terming the approach "logistic growth" to reflect its S-shaped trajectory contrasting with unchecked exponential increase. Verhulst derived the formula by extending principles from Thomas Malthus's ideas on population limits, positing that growth rates diminish proportionally as populations approach a maximum sustainable level, leading to a that integrates these constraints for realistic forecasting. Verhulst first presented his model in a brief note titled Notice sur la loi que la population suit dans son accroissement, published in the journal Correspondance Mathématique et Physique in . He elaborated on the derivation and applied it to empirical data from , the , and in two subsequent memoirs: Recherches mathématiques sur la loi d'accroissement de la population (1845) and Deuxième mémoire sur la loi d'accroissement de la population (1847), both appearing in the proceedings of the Royal Academy of Sciences, Letters and Fine Arts of . Despite these foundational publications, Verhulst's logistic model received scant attention and limited adoption in the late 19th and early 20th centuries, remaining largely overlooked until its independent rediscovery by Raymond Pearl and Lowell Reed in the 1920s for analyzing U.S. population trends.

Key Contributors and Milestones

The logistic function experienced a significant rediscovery in 1920 when American biostatisticians Raymond Pearl and Lowell J. Reed independently derived it while analyzing U.S. data from 1790 to 1910, fitting the curve to predict future trends and demonstrating its S-shaped pattern as a limit to . Their work, published in the Proceedings of the , popularized the model in and , though they initially overlooked Pierre-François Verhulst's earlier formulation. Following Pearl and Reed's publications, Verhulst's 19th-century contributions gained full recognition in the , with subsequent studies crediting him as the original developer of the equation in 1838 for modeling bounded . This period also marked a terminological shift: while Verhulst had coined "logistic" (from the Greek logistikós, relating to calculation, or possibly logis for habitation), Pearl and Reed's adoption of "logistic " standardized the name in English-language literature, supplanting references to the "Verhulst equation." In the 1930s and 1940s, the logistic function spread widely in , particularly through the work of Russian biologist Georgii F. Gause, whose 1934 book The Struggle for Existence applied it to model single-species and interspecies competition under resource constraints, influencing experimental designs in laboratory populations like paramecia. This era solidified its role as a foundational tool for understanding density-dependent growth in natural systems. By the 1950s, the function integrated into statistical methodology, notably through David R. Cox's 1958 paper introducing for analyzing binary outcomes, which provided a framework for estimating probabilities via maximum likelihood and spurred its adoption in fields like and . In the 1960s, it found applications in early computing and simulations, including population models run on digital computers to forecast growth limits and chaotic behaviors, as seen in Edward Lorenz's meteorological simulations that foreshadowed the logistic map's use in dynamical systems. In the 2020s, the logistic function saw renewed interest in , particularly as the activation in neural networks for tasks like , though this represents an application of established mathematics rather than a novel mathematical milestone.

Mathematical Properties

Symmetries and Transformations

The logistic function, in its general form f(x) = \frac{L}{1 + e^{-k(x - x_0)}}, where L > 0 is the curve's maximum value, k > 0 is the growth rate, and x_0 is the x-value of the midpoint (), possesses a point about the point (x_0, L/2). This symmetry arises because f(x_0 + u) + f(x_0 - u) = L for any deviation u, meaning the function values equidistant from x_0 sum to the upper L. Equivalently, the shifted and scaled variant g(u) = \frac{2}{L} \left( f(x_0 + u) - \frac{L}{2} \right) satisfies g(-u) = -g(u), rendering it an centered at the . This oddness after transformation highlights the function's antisymmetric behavior around its midpoint, which can be expressed as g(u) = \tanh\left( \frac{k u}{2} \right), connecting it to the hyperbolic (detailed further in subsequent sections). For the logistic \sigma(z) = \frac{1}{1 + e^{-z}} (with L = 1, k = 1, x_0 = 0), the symmetry simplifies to \sigma(z) + \sigma(-z) = 1, or $1 - \sigma(z) = \sigma(-z), confirming rotational symmetry about (0, 1/2). Any general logistic function can be reparameterized to this standard form via an of the input variable, such as z = k(x - x_0), followed by vertical scaling and shifting to match L. The S-shaped (sigmoidal) profile of the logistic function is preserved under affine of both the input and output. Specifically, applying a linear transformation to the argument, x \mapsto a x + b with a \neq 0, and to the output, f(x) \mapsto c f(x) + d with c > 0, yields another logistic function with adjusted parameters k' = |a| k, x_0' = x_0 - b/a, and L' = c L. Reflections over the line y = L/2 (vertical flip) also maintain the form, as L - f(x) = \frac{L e^{-k(x - x_0)}}{1 + e^{-k(x - x_0)}}, which is logistic with negated growth rate k \to -k and unchanged . Horizontal scaling via k compresses or expands the transition region without altering the overall . A key transformation linking the logistic to linear models is the logit function, defined as \ell(p) = \ln\left( \frac{p}{1 - p} \right) for $0 < p < 1. Applying the logit to the logistic output yields \ell(f(x)) = k(x - x_0), a linear relationship that facilitates parameter estimation and interpretation in applications like growth modeling. For the standard sigmoid, this reduces to \ell(\sigma(z)) = z, underscoring the logit as the natural inverse that linearizes the .

Inverse Function

The inverse of the logistic function f(x) = \frac{L}{1 + e^{-k(x - x_0)}} is derived by solving for x in terms of y, where y = f(x). Rearranging gives: y = \frac{L}{1 + e^{-k(x - x_0)}} $1 + e^{-k(x - x_0)} = \frac{L}{y} e^{-k(x - x_0)} = \frac{L}{y} - 1 = \frac{L - y}{y} Taking the natural logarithm on both sides yields: -k(x - x_0) = \ln\left( \frac{L - y}{y} \right) x - x_0 = -\frac{1}{k} \ln\left( \frac{L - y}{y} \right) = \frac{1}{k} \ln\left( \frac{y}{L - y} \right) Thus, the inverse function is: x = x_0 + \frac{1}{k} \ln\left( \frac{y}{L - y} \right) This expression holds for the generalized logistic form with parameters L > 0, k > 0, and real x_0. The inverse is defined for y \in (0, L), corresponding to the open range of the logistic function, which approaches but never reaches 0 or L over the real line. The output x spans all real numbers, reflecting the bijection between \mathbb{R} and (0, L). In the standardized case where L = 1, x_0 = 0, and k = 1, the logistic function simplifies to the sigmoid \sigma(z) = \frac{1}{1 + e^{-z}}, and its inverse is the logit function: \text{logit}(p) = \ln\left( \frac{p}{1 - p} \right), \quad p \in (0, 1). This transformation maps probabilities in (0, 1) to the unbounded real line and is fundamental in statistical modeling. Numerical computation of the can pose challenges, particularly when y approaches 0 or L, as the ratio \frac{y}{L - y} becomes very small or large, risking underflow, , or precision loss in floating-point representations. Stable implementations typically clip y to [\epsilon, L - \epsilon] for small \epsilon > 0 (e.g., $10^{-15}) before computing \frac{1}{k} \ln\left( \frac{y}{L - y} \right).

Relation to Hyperbolic Functions

The logistic function admits an equivalent expression in terms of the hyperbolic tangent function, providing an alternative representation that highlights its connection to and trigonometric identities. For the general logistic function f(x) = \frac{L}{1 + e^{-k(x - x_0)}}, where L > 0 is the curve's maximum value, k > 0 is the growth rate, and x_0 is the (value of x at which f(x_0) = L/2), this equivalence takes the form f(x) = \frac{L}{2} \left( 1 + \tanh\left( \frac{k (x - x_0)}{2} \right) \right). This relation arises directly from the hyperbolic tangent, \tanh(z) = \frac{e^{2z} - 1}{e^{2z} + 1}, which is derived from the ratio of hyperbolic sine and cosine functions. To see the connection, substitute z = \frac{k (x - x_0)}{2} into the expression for (1 + \tanh(z))/2: \frac{1 + \tanh(z)}{2} = \frac{1 + \frac{e^{2z} - 1}{e^{2z} + 1}}{2} = \frac{e^{2z}}{e^{2z} + 1} = \frac{1}{1 + e^{-2z}}. With $2z = k (x - x_0), the right-hand side matches the standard logistic form scaled by L/2 and shifted appropriately, confirming the equivalence. This closed-form identity underscores the logistic function's membership in the family of sigmoid curves, sharing the characteristic S-shape and bounded asymptotes with \tanh. In numerical computations, the hyperbolic tangent formulation offers practical advantages over the exponential-based logistic form, particularly in terms of stability when evaluating large arguments. The \tanh function can be implemented using techniques that rewrite it to avoid direct computation of extremely large or small exponentials—for instance, for large positive z, \tanh(z) \approx 1 - 2e^{-2z}, and for large negative z, \tanh(z) \approx -1 + 2e^{2z}—reducing the risk of overflow or underflow in . This relation to exponentials facilitates more robust evaluations in applications such as optimization algorithms and simulations, where the logistic function appears frequently.

Derivatives

The first derivative of the logistic function f(x) = \frac{L}{1 + e^{-k(x - x_0)}} is obtained via the chain rule and , yielding f'(x) = \frac{L k e^{-k(x - x_0)}}{(1 + e^{-k(x - x_0)})^2} = k f(x) \left(1 - \frac{f(x)}{L}\right). This form highlights the 's dependence on the value itself, scaled by the growth k and modulated by the proximity to the carrying capacity L. The derivative reaches its maximum value at the inflection point x = x_0, where f(x_0) = L/2, giving f'(x_0) = \frac{L k}{4}. This maximum slope represents the steepest rate of change in the S-shaped curve, occurring midway between the lower asymptote (0) and upper asymptote (L). The second derivative is f''(x) = k^2 f(x) \left(1 - \frac{f(x)}{L}\right) \left(1 - \frac{2 f(x)}{L}\right), which factors to show that f''(x) = 0 at x = x_0 (where f(x_0) = L/2) and also at the asymptotes. For x < x_0, f''(x) > 0 (concave up), and for x > x_0, f''(x) < 0 (concave down), marking the transition from accelerating to decelerating growth at the inflection point. In population dynamics and growth models, the first derivative f'(x) interprets as the instantaneous growth rate, initially increasing exponentially before tapering due to resource limits, as captured by the logistic differential equation \frac{df}{dx} = k f (1 - f/L). This property underscores the function's utility in modeling bounded growth processes, such as biological populations approaching carrying capacity.

Integrals

The indefinite integral of the logistic function in its standard form f(x) = \frac{L}{1 + e^{-k(x - x_0)}} can be obtained through substitution. Setting z = k(x - x_0), the integral becomes \int f(x) \, dx = \frac{L}{k} \int \frac{1}{1 + e^{-z}} \, dz. The antiderivative of the unit \frac{1}{1 + e^{-z}} is \ln(1 + e^z) + C, yielding \int f(x) \, dx = \frac{L}{k} \ln \left( 1 + e^{k(x - x_0)} \right) + C. An equivalent expression is \frac{L}{k} \left[ k(x - x_0) + \ln \left( 1 + e^{-k(x - x_0)} \right) \right] + C, which follows from the identity \ln(1 + e^{z}) = z + \ln(1 + e^{-z}). This form highlights the linear growth component for large positive arguments, consistent with the function's asymptotic behavior approaching L. The definite integral of the logistic function over the entire real line diverges, as f(x) approaches L for x \to \infty and 0 for x \to -\infty, resulting in infinite area under the curve. However, for the associated probability density function of the , f'(x) = \frac{k L e^{-k(x - x_0)}}{[1 + e^{-k(x - x_0)}]^2}, the integral from -\infty to \infty equals 1 by definition. In generalized logistic models, such as the generalized logistic distribution or logistic-normal mixtures, closed-form integrals often do not exist, necessitating numerical methods. Gauss-Hermite quadrature, Monte Carlo integration, and quasi-Monte Carlo methods using Sobol' or Halton sequences have been shown to effectively approximate these integrals, with quasi-Monte Carlo offering superior performance for high-dimensional cases in logistic-normal models. These integrals find application in cumulative population models in ecology, where the antiderivative represents total accumulated growth over time.

Series Expansions

The Taylor series expansion of the general logistic function f(x) = \frac{L}{1 + e^{-k(x - x_0)}} around the inflection point x = x_0 is given by f(x) = \frac{L}{2} + \frac{L k}{4} (x - x_0) - \frac{L k^3}{48} (x - x_0)^3 + \frac{L k^5}{480} (x - x_0)^5 - \frac{17 L k^7}{80640} (x - x_0)^7 + \cdots, where the coefficients arise from successive derivatives evaluated at x_0, with the first few terms matching the scaled and shifted standard sigmoid series. The general term of this series for the standard logistic sigmoid \sigma(z) = \frac{1}{1 + e^{-z}} (corresponding to L = 1, k = 1, x_0 = 0) involves B_n and can be expressed as \sigma(z) = \sum_{n=0}^{\infty} \frac{(-1)^{n+1} (2^{n+1} - 1) B_{n+1}}{(n+1) n!} z^n, derived from the higher-order derivatives of \sigma(z), which relate to derivatives of \sech^2(z/2) since \sigma'(z) = \frac{1}{4} \sech^2(z/2); this form stems from the connection to the Taylor series of the hyperbolic tangent function, \sigma(z) = \frac{1}{2} \bigl(1 + \tanh(z/2)\bigr), whose expansion incorporates Bernoulli numbers in its odd-powered terms. For large |k(x - x_0)|, asymptotic expansions capture the exponential approach to the limiting values f(x) \to L as x \to +\infty and f(x) \to 0 as x \to -\infty. Specifically, for large positive z = k(x - x_0), f(x) = L \sum_{m=0}^{\infty} (-1)^m e^{-(m+1) z}, a convergent geometric series expansion reflecting the rapid decay away from the transition region; an analogous expansion holds for large negative z by symmetry, f(x) = L \sum_{m=1}^{\infty} (-1)^{m+1} e^{m z}. The power series expansion around any real point has an infinite radius of convergence along the real line, allowing analytic continuation to represent the function globally on \mathbb{R}, though in the complex plane, convergence is limited by poles at distances \pi / k from the real axis.

Connection to Differential Equations

The logistic differential equation is an autonomous first-order ordinary differential equation of the form \frac{dy}{dt} = k y \left(1 - \frac{y}{L}\right), where k > 0 is the intrinsic growth rate and L > 0 is the , representing the maximum value that y(t) approaches as t \to \infty. This equation arises in models of processes exhibiting initially that slows and saturates due to limitations, and its general solution is the logistic function y(t) = \frac{L}{1 + e^{-k(t - x_0)}}, where x_0 is a determining the . To derive the solution, separate variables to obtain \frac{dy}{y \left(1 - \frac{y}{L}\right)} = k \, dt. The left-hand integral is evaluated using : \frac{1}{y \left(1 - \frac{y}{L}\right)} = \frac{1}{y} + \frac{1/L}{1 - y/L} = \frac{L}{y(L - y)}, yielding \int \frac{L}{y(L - y)} \, dy = k t + C. Integrating gives \ln |y| - \ln |L - y| = k t + C, or \ln \left| \frac{y}{L - y} \right| = k t + C. Exponentiating and solving for y produces y(t) = \frac{L e^{k t + C}}{L + e^{k t + C}} = \frac{L}{1 + A e^{-k t}}, where A = e^C > 0 is a constant determined by initial conditions. The initial condition y(0) = y_0 (with $0 < y_0 < L) fixes the constant A = \frac{L - y_0}{y_0}, ensuring the solution starts at the specified value and approaches L asymptotically. The parameter x_0 in the standard logistic form emerges from this, as it shifts the curve so that y(x_0) = L/2, the point of maximum growth rate; specifically, x_0 = \frac{1}{k} \ln \left( \frac{L/y_0 - 1}{1} \right), linking the initial value directly to the curve's centering. This parameterization highlights how the logistic function encapsulates the DE's behavior, with x_0 encoding the timing of the transition from slow to rapid growth relative to the initial state. For stability analysis, the equilibrium points are found by setting \frac{dy}{dt} = 0, giving y = 0 and y = L. The point y = 0 is unstable, as perturbations away from it (positive y) lead to growth toward L, while y = L is stable, attracting solutions from below. This is confirmed by the phase line: for $0 < y < L, \frac{dy}{dt} > 0 (increasing), and the eigenvalue of the linearized at each (the of the right-hand side) is positive at y=0 (unstable) and negative at y=L (stable).

Probabilistic Interpretations

As a Cumulative Distribution Function

The logistic function serves as the (CDF) of the , a continuous commonly used in statistical modeling. The (PDF) of the , parameterized by a \mu \in \mathbb{R} and a s > 0, is f(x) = \frac{e^{-(x - \mu)/s}}{s \left(1 + e^{-(x - \mu)/s}\right)^2}, \quad x \in \mathbb{R}. Its corresponding CDF is the standard logistic sigmoid: F(x) = \frac{1}{1 + e^{-(x - \mu)/s}}. These parameters \mu and s determine the center and spread of the distribution, respectively. In the more general form of the logistic function \sigma(x) = \frac{L}{1 + e^{-k(x - x_0)}}, the location \mu aligns with the midpoint x_0, while the growth rate k relates inversely to the scale as k = 1/s. The is symmetric about its \mu, with a variance of \pi^2 s^2 / 3. Compared to the normal distribution, it exhibits heavier tails and higher , making it suitable for applications where outliers or extreme events occur with greater probability than under Gaussian assumptions. This property arises from the exponential form in its PDF, which decays more slowly in the tails than the normal's decay. The gained prominence in statistics through its application to quantal response models in bioassays, where it describes the probability of a outcome (e.g., response or no response) as a of dose or . Joseph Berkson introduced this usage in , demonstrating the logistic function's advantages over the normal integral for fitting dose-response data in experimental settings. This historical tie underscores its role in early statistical modeling of threshold phenomena.

Role in Probability Distributions

The logistic distribution is closely related to the extreme value distribution of type I, also known as the , through the property that the difference of two independent Gumbel-distributed random variables follows a . Specifically, if X and Y are independent Gumbel random variables with location parameters \mu_X and \mu_Y, and common \beta > 0, then Z = X - Y has a with location \mu_X - \mu_Y and scale \beta. This connection arises in , where the logistic distribution serves as the CDF of the log-Fisher-Tippett type I distribution under certain transformations, highlighting its role in modeling differences in maxima. The logistic distribution can also be represented as a scale mixture of distributions, providing a hierarchical useful in Bayesian modeling and robustness studies. In this representation, a logistic X is expressed as X = \mu + \sigma Z / \sqrt{V}, where Z \sim \mathcal{N}(0,1) is , independent of V, and V follows a mixing distribution with density g(v) = \frac{2}{v} \exp\left(-\frac{2}{v} - 2\sqrt{\frac{2}{v}}\right) for v > 0. This mixture structure, first derived by Stefanski, explains the logistic's heavier tails compared to the and has been extended in recent work to derive limit laws for randomly indexed maxima and sums. Such representations facilitate computational inference and underscore the logistic's position among location-scale families with mixtures. The (MGF) of the provides a tool for deriving its moments and characterizing its properties. For a with location \mu and scale s > 0, the MGF is M(t) = e^{\mu t} \Gamma(1 + s t) \Gamma(1 - s t) for |t| < 1/s, where \Gamma denotes the gamma function. This form, expressible via the beta function for the standard case (M(t) = B(1 + t, 1 - t) when \mu = 0, s = 1), confirms the existence of all moments and yields the mean \mu and variance \frac{\pi^2 s^2}{3} upon differentiation. Simulation of logistic random variables commonly employs the inverse cumulative distribution function (CDF) method, leveraging the closed-form inverse of the logistic CDF. If U \sim \text{Uniform}(0,1), then X = \mu + s \log\left(\frac{U}{1 - U}\right) follows a with parameters \mu and s. This approach, standard for continuous distributions with invertible CDFs, enables efficient generation for Monte Carlo studies and builds on the logistic's role as a CDF in probabilistic modeling.

Generalizations

Multivariate Extensions

The multinomial logistic function, commonly referred to as the , provides a multidimensional extension of the univariate logistic sigmoid by mapping a vector of real numbers to a probability distribution over multiple categories. It is defined as \sigma(\mathbf{z})_i = \frac{e^{z_i}}{\sum_{j=1}^K e^{z_j}} for each component i = 1, \dots, K, where \mathbf{z} \in \mathbb{R}^K is the input vector. This formulation ensures the outputs sum to 1 and lie in [0,1], making it suitable for modeling categorical outcomes in discrete choice scenarios. The , foundational to this extension, was introduced by in the context of qualitative choice behavior, deriving the probability of selecting alternative i as the ratio of exponentials of utility differences. The term "" and its application in probabilistic neural network outputs were later formalized by , emphasizing maximum mutual information estimation for classification tasks. The multivariate logistic distribution generalizes the univariate logistic to joint distributions over random vectors, preserving logistic marginals while allowing for dependence structures. Under independence assumptions, the joint cumulative distribution function (CDF) is the product of the univariate CDFs, F(\mathbf{x}) = \prod_{i=1}^d F(x_i), where each F(x_i) = \frac{1}{1 + e^{-x_i}} for \mathbf{x} \in \mathbb{R}^d. Gumbel's Type I bivariate logistic distribution, with joint CDF F(x,y) = \frac{1}{1 + e^{-x} + e^{-y}}, provides an early example of a dependent bivariate form that maintains logistic marginals. Malik and Abraham proposed broader families of multivariate logistic distributions, such as F(\mathbf{x}) = \left( \prod_{i=1}^d (1 + e^{-x_i}) \right)^{-1/\alpha} for \alpha > 0, ensuring marginal logistic distributions and positive dependence for \alpha < 1. In dynamical systems, the logistic map extends to higher dimensions through matrix or multidimensional formulations to study chaos in vector-valued iterations. The standard one-dimensional logistic map x_{n+1} = r x_n (1 - x_n) generalizes to a matrix form \mathbf{X}_{n+1} = r \mathbf{X}_n (I - \mathbf{X}_n), where \mathbf{X}_n is a stochastic matrix and I is the identity, exhibiting fractal spectral properties and chaotic transfer across dimensions for certain r. Multidimensional discrete chaotic maps, including logistic variants, preserve bifurcation routes from periodic to chaotic regimes, with invariance in critical parameters across dimensions, as analyzed in generalizations of the . Multivariate extensions of the logistic function, such as the , have applications in reinforcement learning for handling categorical actions, including in multi-agent settings.

Time-Dependent Variants

Time-dependent variants of the logistic function extend the standard model by allowing parameters such as the carrying capacity L or growth rate k to vary with time, enabling better representation of dynamic systems where environmental or external factors evolve. A common generalization takes the form f(t) = \frac{L(t)}{1 + e^{-k(t - x_0)}}, where L(t) and k(t) are functions of time, often specified piecewise or through auxiliary differential equations to capture non-stationary growth patterns. This approach has been applied to model infectious disease outbreaks, such as , where the carrying capacity K(t) evolves according to \frac{dK(t)}{dt} = v (K(t) - K_1) \left[1 - \left(\frac{K(t) - K_1}{K_2}\right)^\mu \right], yielding a logistic-like solution for K(t) that adjusts to changing infection limits. Similarly, the growth rate can be time-dependent, as in the stochastic form \frac{dN}{dt} = r(t) N^\alpha \left[1 - p(t) N \right], where variability in r(t) and p(t) accounts for fluctuating transmission dynamics. The Richards curve represents an asymmetric variant of the logistic function, introducing a shape parameter \nu to allow for uneven growth acceleration and deceleration phases, which is useful for biological processes with temporal asymmetries. Its explicit form is f(t) = \frac{L}{\left(1 + \nu e^{-k(t - x_0)}\right)^{1/\nu}}, reducing to the standard symmetric logistic when \nu = 1. This generalization, originally proposed for empirical plant growth data, provides flexibility in fitting observed S-shaped curves that deviate from perfect symmetry around the inflection point. These time-dependent extensions often arise as solutions to modified differential equations where parameters vary, such as the generalized logistic equation \frac{dy}{dt} = k(t) y \left(1 - \frac{y}{L(t)}\right), which relaxes the constant-rate assumption of the classic to accommodate slowly changing environmental limits. Analytical solutions are typically unavailable in closed form, necessitating numerical integration, but approximations exist for slowly varying L(t) or k(t), preserving the sigmoidal trajectory while adapting to temporal shifts. In recent climate modeling, time-varying carrying capacities have been incorporated into logistic frameworks to simulate adaptive responses to environmental changes, such as productivity shifts in fisheries under warming scenarios. For instance, simulations adjusting L(t) downward by 20-50% due to climate-induced habitat loss reveal unintended overharvesting risks when management targets lag behind dynamic capacities. This approach highlights the 's utility in projecting non-stationary ecological thresholds through 2050.

Applications in Natural Sciences

Ecology and Population Dynamics

In ecology, the logistic function provides a foundational model for describing population growth that approaches an environmental carrying capacity, capturing the transition from exponential increase to stabilization due to resource limitations. Pierre-François Verhulst introduced this model in 1838 to predict bounded population dynamics, where growth slows as the population nears the maximum sustainable size. The Verhulst model expresses population size P(t) at time t as P(t) = \frac{K}{1 + \left( \frac{K}{P_0} - 1 \right) e^{-r t}}, with r denoting the intrinsic growth rate and K the carrying capacity, representing the equilibrium population level under constant environmental conditions. This closed-form solution arises from integrating the underlying differential equation \frac{dP}{dt} = r P \left(1 - \frac{P}{K}\right), which quantifies the total population trajectory over time by separating variables and solving the separable ordinary differential equation. The model's integral form highlights how initial population P_0 influences the approach to K, with early growth approximating exponential rates before density-dependent factors like competition for resources dominate. However, the Verhulst model assumes a static environment with unchanging K and uniform individual contributions to growth, which limits its applicability to real-world scenarios involving fluctuating resources or external perturbations. To address these constraints, extensions incorporate stochastic elements, such as random environmental variability or demographic noise, transforming the deterministic equation into a probabilistic framework that better predicts fluctuations in small populations. A seminal empirical validation of the logistic model came from Georgii Gause's 1934 experiments on yeast and protozoan cultures, where measured growth curves closely matched the S-shaped trajectory, demonstrating resource-limited saturation in controlled microcosms. These laboratory studies underscored the model's utility for microbial ecology, influencing subsequent applications to larger-scale population forecasts while highlighting the need for adaptations in dynamic habitats, such as those with time-varying carrying capacities.

Medicine and Growth Modeling

In medicine, the logistic function models tumor growth particularly well during early phases, where cell proliferation competes for limited resources such as nutrients or space, leading to an S-shaped curve that transitions from exponential to saturating growth. The pure logistic model for tumor volume N(t) is given by N(t) = \frac{N_0 K e^{r t}}{K + N_0 (e^{r t} - 1)}, where N_0 is the initial volume, K is the carrying capacity (maximum tumor size), and r is the intrinsic growth rate. This formulation captures the initial rapid expansion before density-dependent inhibition slows growth, as observed in preclinical studies of breast and lung cancers. For more advanced stages, hybrid models combining logistic and Gompertz functions have been developed to account for asymmetric slowing of growth due to factors like necrosis or vascular limitations, improving predictions of tumor progression in clinical oncology. The logistic function also plays a key role in epidemic modeling within medicine, often integrated into extensions of the SIR (Susceptible-Infectious-Recovered) framework via logistic incidence rates that reflect saturation in transmission as the susceptible population depletes. In such models, the incidence term \beta S I / (1 + \alpha I) incorporates logistic-like density dependence, enabling forecasts of outbreak trajectories analogous to bounded population growth. For instance, early analyses of the 2020 COVID-19 outbreak in Wuhan, China, fitted cumulative cases to a using nonlinear least squares, achieving high accuracy with goodness-of-fit metrics indicating strong alignment (e.g., near-linear relationships in transformed data). The model's inflection point, occurring at the midpoint x_0 = \frac{1}{r} \ln \left( \frac{K - N_0}{N_0} \right) + t_0, precisely predicts the peak infection rate, aiding in resource allocation for peak cases. Recent applications extend this to ongoing outbreaks, such as the 2024 mpox resurgence. A modified logistic growth model, incorporating human-to-human and zoonotic transmission rates, was fitted to U.S. case data, forecasting outbreak peaks and demonstrating that combined interventions could reduce cases by up to 95%. Similarly, for the 2025 avian influenza (H5N1) threats, logistic growth models with multi-delay dynamics have been used to simulate virus spread in avian and human populations, predicting oscillatory outbreaks and evaluating culling strategies to prevent escalation. These medical uses parallel ecological population models but focus on pathogen-host interactions in human health contexts.

Chemistry and Reaction Kinetics

In chemistry, the logistic function arises prominently in modeling autocatalytic reactions, where a product of the reaction catalyzes its own formation, leading to an initial slow phase followed by rapid acceleration and eventual saturation. A classic example is the bimolecular autocatalytic process represented as A + B \to 2B, with the rate law r = -\frac{d[A]}{dt} = k [A][B]. Assuming a constant total concentration L = [A] + [B], this simplifies to \frac{d[A]}{dt} = -k [A] (L - [A]), or equivalently for the product concentration [B], \frac{d[B]}{dt} = k [B] (L - [B]). The solution to this differential equation yields the logistic function for the concentration of the autocatalyst B over time: [B](t) = \frac{L}{1 + \left( \frac{L - [B]_0}{[B]_0} \right) e^{-k L t}}, where [B]_0 is the initial concentration. This S-shaped curve captures the sigmoidal progression typical of such reactions, starting slowly, accelerating due to positive feedback, and leveling off as reactants deplete. This form parallels the connection to differential equations discussed elsewhere, providing a deterministic framework for non-linear kinetics. In enzyme kinetics, the logistic function serves as an approximation for saturation phenomena in certain , particularly when accounting for cooperative effects or probabilistic interpretations of substrate binding. The standard v = \frac{V_{\max} [S]}{K_m + [S]} describes hyperbolic saturation, but generalizations to logistic forms incorporate sigmoidal behavior for allosteric enzymes or when initial rates are influenced by autocatalytic-like substrate activation. For instance, a logistic modification can model the velocity as v = \frac{V_{\max}}{1 + e^{-k([S] - S_0)}}, better fitting data where binding exhibits threshold effects. This approach enhances predictive accuracy in vitro for complex enzymatic systems. The logistic function also models conversion rates during phase transitions in polymerization reactions, where monomer addition accelerates autocatalytically until chain saturation occurs. In free-radical or step-growth polymerizations, the degree of conversion \alpha(t) often follows a logistic curve, reflecting an induction period, rapid propagation, and termination plateau. For example, turbidity measurements in controlled polymerizations fit to logistic functions reveal rate constants that quantify the steepness of the sigmoidal profile, aiding in optimizing reaction conditions for uniform polymer structures. A notable application appears in early models of chemical oscillators, such as the Belousov-Zhabotinsky reaction (observed in the 1950s), where autocatalytic steps drive oscillatory behavior that can exhibit logistic-like growth phases for intermediate concentrations. These models highlight the logistic function's role in capturing transient autocatalysis within cyclic kinetics.

Physics and Statistical Mechanics

In statistical mechanics, the logistic function fundamentally describes the Fermi-Dirac distribution for the average occupation number of fermions in a single-particle state with energy \epsilon, at absolute temperature T and chemical potential \mu. This distribution is expressed as f(\epsilon) = \frac{1}{e^{(\epsilon - \mu)/kT} + 1}, where k is Boltzmann's constant; this form is identical to the standard unit-height logistic function \frac{1}{1 + e^{-x}} upon rescaling the argument x = (\epsilon - \mu)/kT. The logistic shape captures the sharp transition from occupied states below the Fermi level to unoccupied states above it at low temperatures, reflecting the Pauli exclusion principle in fermionic systems. In atmospheric physics, mirages arise from refraction due to temperature gradients affecting the refractive index of air. Laboratory experiments simulate these effects through ray-tracing in varying density profiles. For diffusion processes in spin systems, mean-field approximations employ the to describe probabilistic transitions in models like the . The conditional probability that a spin \sigma_i = +1 given the local field h_i from neighboring spins follows P(\sigma_i = +1 \mid \{\sigma_j\}) = \frac{1}{1 + \exp(-2 \beta h_i)}, a logistic sigmoid that governs flip rates in and facilitates mean-field solutions for magnetization and correlation diffusion. This formulation simplifies the analysis of cooperative behavior and phase transitions in lattice spin configurations under thermal fluctuations. Recent applications in quantum computing leverage simulations of the logistic map on quantum hardware to explore chaotic dynamics and quantum analogs of classical chaos. In 2022, such simulations demonstrated the logistic map's utility in probing sensitivity to initial conditions and bifurcation structures on noisy intermediate-scale quantum devices, highlighting potential for studying quantum information scrambling. The Fermi-Dirac distribution, as a logistic cumulative distribution function, further bridges these physical interpretations to probabilistic frameworks in quantum statistical mechanics.

Applications in Social and Technical Sciences

Statistics and Machine Learning

In statistics, is a fundamental method for modeling binary outcomes, where the probability p of a positive class is given by the logistic function applied to a linear combination of predictors:
p = \frac{1}{1 + e^{-\boldsymbol{\beta}^T \mathbf{x}}},
with \boldsymbol{\beta} as the coefficient vector and \mathbf{x} as the feature vector. This approach, introduced by , transforms the linear predictor into a bounded probability between 0 and 1, making it suitable for classification tasks where outcomes are dichotomous, such as success/failure or presence/absence.
Parameters in logistic regression are typically estimated using maximum likelihood estimation (MLE), where the logit link function— the inverse of the —connects the linear predictor to the binomial distribution in the framework of generalized linear models (GLMs). This estimation maximizes the log-likelihood of observed data, yielding coefficients that best fit the probabilistic model, and has become a cornerstone for interpretable predictive modeling in statistical analysis. In machine learning, the logistic function appears as the sigmoid activation \sigma(z) = \frac{1}{1 + e^{-z}}, commonly used in the hidden layers of early to introduce nonlinearity and enable gradient-based learning through . This activation squashes inputs to [0,1], facilitating the modeling of complex decision boundaries in , though modern networks often favor alternatives like for deeper architectures due to vanishing gradient issues with sigmoid. Extensions of logistic regression address multi-class problems via multinomial logistic regression, which generalizes the binary case using the softmax function derived from the multinomial distribution for probabilistic outputs across categories; this is particularly useful in supervised learning for tasks like image or text classification. To mitigate overfitting, especially in high-dimensional settings, regularization techniques such as L1 (Lasso) and L2 (Ridge) penalties are incorporated into the MLE objective, promoting sparse or stable models as implemented in efficient coordinate descent algorithms. Advancements in 2024 have integrated logistic regression with large language models (LLMs) by applying it to embeddings from small LLMs for binary classification, achieving performance comparable to larger models in few-shot settings while enhancing explainability through interpretable coefficients.

Economics and Innovation Diffusion

In economics, the logistic function underpins models of innovation diffusion, particularly through the Bass model, which describes the adoption of new products or technologies in a population over time. Developed by Frank M. Bass in 1969, this model captures the interplay between innovative adopters, who embrace the innovation independently, and imitators, who follow due to social influence or word-of-mouth. The differential equation governing the rate of adoption N(t) is given by: \frac{dN(t)}{dt} = p (M - N(t)) + q \frac{N(t)}{M} (M - N(t)), where M represents the market potential (total potential adopters), p is the coefficient of innovation (rate of independent adoption), and q is the coefficient of imitation (rate of adoption influenced by others). This equation yields a closed-form solution for cumulative adoption that exhibits an S-shaped curve akin to the logistic function, reflecting slow initial growth, rapid acceleration, and eventual saturation. The Bass model's parameters allow economists to forecast market penetration and assess the relative roles of external marketing efforts (via p) and internal social dynamics (via q). For instance, in applications to consumer durables, higher q values indicate strong network effects, leading to faster diffusion once a critical mass is reached. This framework has been widely adopted in marketing and economic forecasting, enabling predictions of sales trajectories for durable goods by fitting historical data to estimate p, q, and M. The model's logistic-like form parallels population growth models in ecology, where limited resources constrain expansion, but here it highlights bounded market sizes and social contagion. Empirical applications demonstrate the model's utility in real-world economic contexts. For hybrid vehicles, studies using U.S. sales data from the early 2000s show an S-curve adoption pattern, with initial slow uptake among innovators (low p) accelerating via imitation (higher q) as environmental awareness and fuel prices influenced consumer behavior, projecting market saturation around 20-30% by the 2020s. Similarly, smartphone adoption in markets like China during the 2000s-2010s followed a Bass-predicted trajectory, with rapid imitation-driven growth (q > p) leading to approximately 62% by 2015, as amplified word-of-mouth effects. These examples underscore how the model quantifies economic impacts, such as and for green technologies. Beyond products, the Bass model extends to sociological processes, modeling the of norms and behaviors through social networks. In the context of social media penetration, extensions incorporating network data reveal how platforms like achieved global adoption, with imitation coefficients capturing among peers, resulting in S-curves for user growth from niche to mainstream acceptance in the . This application highlights the model's role in understanding cultural shifts, such as the normalization of online sharing, where drives widespread behavioral change akin to economic innovation uptake.

Linguistics and Language Evolution

In historical linguistics, the logistic function models the S-curve pattern observed in the diffusion of linguistic innovations over time, where the frequency of a new form starts slowly, accelerates, and then asymptotes toward completion. This approach captures how sound changes, such as vowel shifts in English, propagate through speech communities, with lexical diffusion showing gradual adoption rather than abrupt uniformity. For instance, analyses of the (c. 1400–1700) have fitted logistic curves to the progressive replacement of diphthongs and long vowels in dated texts, illustrating how regional variants spread from urban centers outward. Labovian models in apply the to describe the probabilistic adoption of variable rules, where the probability of using an innovative variant increases sigmoidally with social and temporal factors. The core is f(t) = \frac{1}{1 + e^{-k(t - t_0)}} where f(t) represents the proportion of the innovative form at time t, k is the growth rate, and t_0 is the . This logistic-incrementation framework, developed by , explains how changes increment steadily across generations, aligning with apparent-time data in which younger speakers lead the shift. It emphasizes social embedding, with adolescents often driving acceleration before stabilization in adulthood. Empirical data from the Northern Cities Shift, a chain shift affecting short vowels in urban Inland North American English dialects, demonstrate this fit across the late 20th century. Sociophonetic studies from the 1970s to 2000s, including Labov's , reveal S-curve trajectories for features like the raising and fronting of /æ/ (as in "cat") and lowering of /ʌ/ (as in "bus"), with logistic regression confirming age-graded adoption rates peaking around 30–50% usage before nearing completion in leading groups. These patterns highlight the shift's ongoing diffusion from cities like and , captured through real- and apparent-time sampling. Recent applications extend logistic modeling to digital contexts, such as the of on . A 2023 study analyzed growth curves of neologisms on (now X), fitting extended logistic equations to track how terms like internet memes or abbreviations surge in frequency, often following S-shaped adoption driven by network effects before plateauing. This parallels broader innovation models in , where social connectivity accelerates linguistic variants akin to economic adoptions.

Agriculture and Crop Yield Modeling

In agriculture, the logistic function is applied to model crop yield responses to fertilizer inputs, capturing the characteristic S-shaped curve that reflects initial slow growth due to nutrient limitations or soil immobilization, followed by rapid yield increases and eventual plateauing at a maximum potential. This contrasts with the more commonly used Mitscherlich equation, an exponential approximation given by Y = A (1 - e^{-b x}), where Y is yield, A is the maximum yield, x is the fertilizer input, and b describes the response rate; while the Mitscherlich model assumes immediate diminishing returns without an initial lag, the logistic function better fits scenarios with delayed responses, such as nitrogen immobilization by soil microbes early in the season. A variant of the logistic function, known as the Hill-logistic form, is particularly useful for dose-response modeling in applications: Y(x) = \frac{L x^k}{c^k + x^k}, where L is the maximum plateau, x is the input level, c is the input at half-maximum (inflection-related ), and k controls the curve's steepness, allowing flexibility for varying response sensitivities across types or crops. This form accommodates plateauing responses observed in field trials, where excessive beyond the inflection yields minimal gains due to . In trials with corn (Zea mays L.), S-curves fitted via logistic models have demonstrated increases from low rates (e.g., 0–50 kg N/ showing limited response due to initial deficiencies) to rapid gains around 100–150 kg N/, plateauing near 200 kg N/ at of 10–12 t/, highlighting efficient . To determine optimal fertilizer levels, techniques fit the logistic model to trial data, identifying the where the yield response is steepest—typically near 50% of maximum yield—and recommending inputs achieving 90–95% of the plateau to balance and environmental risks. For instance, in corn studies, regression-estimated inflection points around 120 kg N/ha have guided recommendations, avoiding over-application that could lead to without yield benefits. These methods emphasize site-specific calibration using field experiments to estimate parameters like L and c, ensuring predictions align with local and variability.

References

  1. [1]
    [PDF] Logistic Regression
    Aug 14, 2017 · The function σ(z) = 1. 1+e−z is called the logistic function (or sigmoid function); it looks like this: σ(z) z. An important quality of this ...
  2. [2]
    [PDF] Unit 27: Sigmoid function
    Apr 8, 2024 · Problem 27.4: The sigmoid function F(x) is also called the standard logistic function because it satisfies the logistic equation F′(x) = F(x)(1 ...Missing: formula | Show results with:formula
  3. [3]
    The Logistic Equation - Department of Mathematics at UTSA
    Oct 29, 2021 · The standard logistic function is the solution of the simple first-order non-linear ordinary differential equation ( 1 − f ( x ) )
  4. [4]
    [PDF] Logistic growth - UNL Math
    Oct 30, 2013 · A logistic function is P(t) = L 1 + Ce-kt, where L is the carrying capacity. The point of diminishing returns is at L/2.Missing: mathematics | Show results with:mathematics
  5. [5]
    Carrying Capacity and Logistic Growth Rate - UTSA
    Oct 24, 2021 · A typical application of the logistic equation is a common model of population growth, originally due to Pierre-François Verhulst in 1838.
  6. [6]
    Logistic Growth Model
    The logistic growth model, also called the Verhulst model, accounts for growth rate declining to 0 by including a factor of 1 - P/K.
  7. [7]
    [PDF] Chapter 4 Logistic Regression
    Raymond Pearl and Lowell Reed, who were studying population growth at Johns Hopkins University, reinvented the logistic function in 1920. They also tried to ...
  8. [8]
    Logistic Growth Model - Department of Mathematics at UTSA
    Oct 24, 2021 · Verhulst derived his logistic equation to describe the self-limiting growth of a biological population. The equation was rediscovered in 1911 by ...
  9. [9]
  10. [10]
    Logistic Equation -- from Wolfram MathWorld
    The logistic equation (sometimes called the Verhulst model or logistic growth curve) is a model of population growth first published by Pierre Verhulst (1845, ...
  11. [11]
    Recherches mathématiques sur la loi d'accroissement de ... - EuDML
    Verhulst, P.F.. "Recherches mathématiques sur la loi d'accroissement de la population.." Nouveaux mémoires de l'Académie Royale des Sciences et Belles ...
  12. [12]
    [PDF] Chapter 6 - Verhulst and the logistic equation (1838)
    He first turned back to Malthus' remark according to which the population of ... Verhulst gave up the logistic equation and chose instead a differential equation ...
  13. [13]
    The Sigmoid Function: A Key Component in Data Science - DataCamp
    May 28, 2025 · The sigmoid function is a logistic function that maps any input values to a range of probabilities between 0 and 1. It is commonly used in ...Mathematical Properties of the... · Sigmoid's Role in Logistic...
  14. [14]
    Verhulst and the logistic equation (1838) - ResearchGate
    In 1838 the Belgian mathematician Verhulst introduced the logistic equation, which is a kind of generalization of the equation for exponential growth.
  15. [15]
    Verhulst, P.F. (1838) Notice sur la loi que la population suit dans son ...
    Verhulst, P.F. (1838) Notice sur la loi que la population suit dans son accroissement. Correspondence Mathematique et Physique (Ghent), 10, 113-121.
  16. [16]
    Recherches mathématiques sur la loi d'accroissement de ... - Persée
    Verhulst Pierre-François. Recherches mathématiques sur la loi d'accroissement de la population. In: Nouveaux mémoires de l'Académie royale des sciences et ...
  17. [17]
    Pearl-Verhulst logistic process - Encyclopedia of Mathematics
    Jun 6, 2020 · Verhulst [a1] introduced the logistic curve in 1838, its use was virtually ignored until R. Pearl and L.J. Reed [a2] rediscovered it empirically ...
  18. [18]
    [PDF] The Origins of Logistic Regression
    Abstract. This paper describes the origins of the logistic function, its adop@ tion in bio@assay, and its wider acceptance in statistics. Its roots.Missing: original | Show results with:original
  19. [19]
    Verhulst and the logistic equation (1838) - SpringerLink
    Verhulst, P.-F.: Recherches mathématiques sur la loi d'accroissement de la population. Nouv. Mém. Acad. R. Sci. B.-lett. Brux. 18, 1–45 (1845). gdz.sub.uni ...
  20. [20]
    Pierre Verhulst (1804 - 1849) - Biography - MacTutor
    Another possible root of the term logistic could have been the French word "logis" (place to live) which was of course related to the limited resources for ...
  21. [21]
    The Logistic Curve and the History of Population Ecology - jstor
    The logistic curve was introduced by Raymond Pearl and Lowell Reed in 1920 and was heavi- ly promoted as a description of human and animal population growth.
  22. [22]
    The Regression Analysis of Binary Sequences - jstor
    Dr. Cox's paper seems likely to result in a much wider acceptance of the logistic function as a regression model. I have never been a partisan in the probit ...Missing: Bernard | Show results with:Bernard
  23. [23]
    Logistic map - Wikipedia
    The map was initially utilized by Edward Lorenz in the 1960s to showcase properties of irregular solutions in climate systems. It was popularized in a 1976 ...
  24. [24]
    The early origins of the logit model - ScienceDirect.com
    This paper describes the origins of the logistic function and its history up to its adoption in bio-assay and the beginning of its wider acceptance in ...Missing: original | Show results with:original
  25. [25]
    None
    ### Summary of Logistic Function: Symmetries, Transformations, Reparameterizations, and Mathematical Properties
  26. [26]
    Logit Transformation -- from Wolfram MathWorld
    The function z=f(x)=ln(x/(1-x)). (1) This function has an inflection point at x=1/2, where f^('')(x)=(2x-1)/(x^2(x-1)^2)=0. (2) Applying the logit ...
  27. [27]
    [PDF] Logistic Regression (1/24/13) 1 Introduction 2 Exponential Family
    logistic function, which is the inverse of the logit function as shown below: η = log. µ. 1 − µ. ⇒ µ = 1. 1 + exp{−η}. , the logistic function. 4 Logistic ...
  28. [28]
    [PDF] Lecture 26 — Logistic regression 26.1 The logistic regression model
    The logistic regression model assumes each response Yi is an independent random variable with distribution Bernoulli(pi), where the log-odds corresponding to pi ...<|control11|><|separator|>
  29. [29]
    Sigmoid Function -- from Wolfram MathWorld
    ### Relation Between Logistic Function and Hyperbolic Tangent
  30. [30]
    [PDF] A Primer on Logistic Growth and Substitution
    The logistic model is symmetric around the midpoint tm. Other models describe non-symmetric or skewed growth, e.g., the. Gompertz. We will not consider non ...
  31. [31]
    What integrals are known that include the logistic function?
    Oct 21, 2020 · The logistic function is a common choice of activation function in neural nets. So is the antiderivative∫x−∞dt1+e−t=∫x−∞etdt1+et=ln(1+ex),.
  32. [32]
    Logistic Function Equation - BYJU'S
    The standard logistic function is a logistic function with parameters k = 1, x0 = 0, L = 1.
  33. [33]
    Integral of Logistic Distribution - calculus - Math Stack Exchange
    Nov 8, 2013 · The integral represents the expected value of the Logistic Distribution which is supposed to be zero.
  34. [34]
    Numerical integration in logistic-normal models - ScienceDirect.com
    Dec 1, 2006 · The main aim of this paper is to compare the performance of three numerical integration methods: GH, MC and QMC, for the integrals appearing in ...
  35. [35]
    [PDF] A generalized logistic function and its applications - EconStor
    The integral curve u = u(t) of equation (3) meeting the condition umin < u(t) < umax is called a generalized logistic function. When the initial level umin = 0, ...
  36. [36]
    [PDF] Non-linear Time Series and Artificial Neural Network of Red Hat ...
    Jun 4, 2018 · The test is based on a Taylor series expansion of the general LSTAR model. We take the third order Taylor approximation of the Logistic function.
  37. [37]
    CRC standard curves and surfaces with Mathematica - Internet Archive
    Jul 8, 2023 · CRC standard curves and surfaces with Mathematica. by: Von Seggern, David H. (David Henry). Publication date: 2007. Topics: Curves on surfaces ...
  38. [38]
    4.4 The Logistic Equation - Calculus Volume 2 | OpenStax
    Mar 30, 2016 · The logistic equation was first published by Pierre Verhulst in 1845 . ... Find the equation and parameter r r that best fit the data for the ...
  39. [39]
    Logistic Differential Equations | Brilliant Math & Science Wiki
    A logistic differential equation is an ordinary differential equation whose solution is a logistic function, modeling bounded growth.
  40. [40]
    Logistic Distribution - MATLAB & Simulink - MathWorks
    The logistic distribution is used for growth models and in logistic regression. It has longer tails and a higher kurtosis than the normal distribution.Missing: formula | Show results with:formula
  41. [41]
    Application of the Logistic Function to Bio-Assay
    (1944). Application of the Logistic Function to Bio-Assay. Journal of the American Statistical Association: Vol. 39, No. 227, pp. 357-365.
  42. [42]
    [PDF] Extreme Value Distributions : Theory and Applications - Minerva
    A rather unexpected relation holds between the logistic and type 1 distributions. ... a standard normal variable, Fisher and Tippett (1928) showed empirically ...
  43. [43]
    The Standard Logistic Distribution - Random Services
    By definition, Hence g ( z ) = h ( y ) d y d z = 1 ( e z + 1 ) 2 e z , z ∈ R which is the PDF of the standard logistic distribution.
  44. [44]
    [PDF] Logistic distribution
    The population mean and variance are. E[X] = −lnλ. V[X] = π2. 3κ2 . The population skewness and kurtosis of X are more complicated expressions. APPL ...
  45. [45]
    Generalized logistic model with time-varying parameters to analyze ...
    Jun 3, 2024 · This work uses time-dependent parameters, so we used a piece-wise function. We believe that the use of more complex functions can improve model ...
  46. [46]
    Logistic equation and COVID-19 - PMC - PubMed Central
    Thus, the generalized logistic equation can be considered to be a stochastic one with time-dependent coefficients. d N d t = r ( t ) N α [ 1 − p ( t ) N ] ...
  47. [47]
    The logistic population model with slowly varying carrying capacity
    Aug 7, 2025 · ... The time-varying carrying capacity K(t) in equation 1 is modelled with the following equation, based on equation 17 in Shepherd & Stojkov ( ...
  48. [48]
    Unintended consequences of climate‐adaptive fisheries ...
    (2023) , who simulated increases and decreases in productivity (by simulating a change in carrying capacity) with logistic models fit to the stocks in the RAM ...<|control11|><|separator|>
  49. [49]
    [PDF] WHAT IS THE LOGISTIC MODEL? | OSU Math
    Logistic model was proposed by Pierre-François Verhulst in 1838, where the rate of reproduction is proportional to both the existing population and the ...Missing: function | Show results with:function<|control11|><|separator|>
  50. [50]
    9 | Population Growth and Regulation - Utexas
    The Verhulst-Pearl logistic equation are three assumptions: (1) that all individuals are equivalent -- that is, the addition of every new individual reduces ...
  51. [51]
    Stochastic dynamics and logistic population growth - NASA ADS
    The Verhulst model is probably the best known macroscopic rate equation in population ecology. It depends on two parameters, the intrinsic growth rate and ...
  52. [52]
    Fitting the Logistic Growth Model - Yeast - Joseph M. Mahaffy
    Aug 10, 2001 · Gause [1] ran a series of experiments that validated the logistic growth equation for a collection of microorganisms. His experiments were ...Missing: ecology | Show results with:ecology
  53. [53]
    [PDF] 6 Applications of Continuous Models to Population Dynamics
    The logistic growth model is perhaps the simplest extension of equation (1a) ... In his book, The Struggle for Existence, Gause (1934) describes a series of ...
  54. [54]
    Autocatalytic – Knowledge and References - Taylor & Francis
    A bimolecular autocatalysis reaction is represented as A + B → 2B. The rate expression is r = −d[A]/dt = k[A][B]. Let us define the progress of the reaction by ...
  55. [55]
    [PDF] Introducing logistic enzyme kinetics
    For treatment of in vitro enzyme kinetics the Michaelis-Menten equation is generalized to a logistic form. From the new probabilistic viewpoint the classical ...
  56. [56]
    Gaining Structural Control by Modification of Polymerization Rate in ...
    Aug 26, 2022 · Fitting the data to the logistic function gives kturbidity of 70.2 and 0.98 h–1 for the fast and slow polymerization systems, respectively.
  57. [57]
    Belousov-Zhabotinsky reaction - Scholarpedia
    Sep 11, 2007 · The BZ reaction makes it possible to observe development of complex patterns in time and space by naked eye on a very convenient human time ...Missing: logistic 1950s<|control11|><|separator|>
  58. [58]
    Physical interpretation of the Rasch model: analogy to Fermi–Dirac ...
    Sep 16, 2025 · ... Fermi–Dirac distribution in statistical physics. Both models share the same mathematical form, with the logistic function determining the ...
  59. [59]
    On Polymer Statistical Mechanics: From Gaussian Distribution to ...
    Aug 22, 2023 · Generalized logistic function or Fermi-Dirac distribution function was able to understand many polymer mechanics problems such as ...
  60. [60]
    (PDF) Mirages in a bottle - ResearchGate
    Aug 9, 2025 · The mirage's image, as seen by the eye or the camera lens, can be used to analyse the deflection and inversion of light rays.Missing: american logistic
  61. [61]
    Generalized Linear Models - Nelder - 1972 - Royal Statistical Society
    These generalized linear models are illustrated by examples relating to four distributions; the Normal, Binomial (probit analysis, etc.), Poisson (contingency ...
  62. [62]
    [PDF] Conditional Logit Analysis of Qualitative Choice Behavior
    Conditional logit analysis of qualitative choice behavior. DANIEL MCFADDEN'. UNIVERSITY OF CALIFORNIA AT BERKELEY. BERKELEY, CALIFORNIA. 1. Preferences and ...
  63. [63]
    [PDF] Consumer Learning and Hybrid Vehicle Adoption
    The purpose of this paper is to study the diffusion of hybrid cars among consumers, and in particular to estimate the effects of learning on consumers' ...
  64. [64]
    [PDF] Diffusion Model for the Adoption of Smartphone Brands under ...
    The model extends the Bass diffusion model to capture dynamic adoption and competitive pricing of smartphone brands, analyzing the influence of average pricing.
  65. [65]
    (PDF) A New Bass Model Utilizing Social Network Data
    May 26, 2017 · The Bass model, however, assumes all potential customers are linked to all other customers in a social network. In this research, we develop the ...Missing: norms | Show results with:norms
  66. [66]
    Descriptive adequacy of the S-curve model in diachronic studies of ...
    The S-shaped curve is typically used to model the diffusion of linguistic change across time. This paper looks at the degree to which the model applies to ...
  67. [67]
    [PDF] 1 Lexical Diffusion and Neogrammarian Regularity Mieko Ogura
    The data in Table 1 are fitted to the logistic function, which transforms the curve into ... Papers from the 8th International Congress on English Historical ...
  68. [68]
    Adolescence, Incrementation, and Language Change - Project MUSE
    To summarize, the logistic-incrementation model proposed by Labov (2001) not only provides a mathematical model that is consistent with the S-curve of ...
  69. [69]
    The initiation and incrementation of sound change
    Dec 23, 2020 · This article presents a theory of the initiation and incrementation mechanisms whereby individual phonetic innovations become community-wide sound changes.1 Introduction · 5 Social Meaning · 5.4 The Curvilinear Pattern
  70. [70]
    1. THE NORTHERN CITIES SHIFT - Duke University Press
    The Northern Cities Shift (NCS) is a chain shift affecting vowel systems in American English, mainly in cities like Chicago, Detroit, and Buffalo.Missing: logistic | Show results with:logistic
  71. [71]
    [PDF] The Northern Cities Shift in Real Time: Evidence from Chicago - CORE
    Mar 21, 2010 · Labov, Yaeger, and Steiner (1972) first described the Northern Cities Shift as a series of movements of five vowels; Eckert (1988) adds the ...Missing: logistic | Show results with:logistic
  72. [72]
    Minor extensions of the logistic equation for growth curves of word ...
    Jun 12, 2023 · The model can consistently reproduce various patterns of actual growth curves, such as the logistic function, linear growth, and finite-time divergence.<|control11|><|separator|>
  73. [73]
    Extracting information from S-curves of language change - Journals
    Dec 6, 2014 · In this paper, we investigate the shape and significance of S-curves in models of adoption of innovations and in data of language change. In ...Missing: incrementation | Show results with:incrementation
  74. [74]
    Comparison of Models for Describing; Corn Yield Response to ...
    Jan 1, 1990 · All models indicated similar maximum yields, but there were marked discrepancies among models when predicting economic optimum rates of ...Missing: logistic curve
  75. [75]
    [PDF] Superiority of S-Shaped (Sigmoidal) Yield Curves for Explaining ...
    Introductory crop production, general soils, and soil fertility texts often use examples of simple asymptotic (Mitscherlich) or quadratic-plus-plateau yield ...Missing: approximation | Show results with:approximation
  76. [76]