Fact-checked by Grok 2 weeks ago

Concave function

In , a concave function is a f: S \to \mathbb{R}, defined on a S in a , that satisfies the f(tx + (1-t)y) \geq t f(x) + (1-t) f(y) for all points x, y \in S and all scalars t \in [0,1]. This condition ensures that the of f lies above or on any chord (straight ) connecting two points on the graph, providing a geometric interpretation of the function's in the sense of supporting lines from below. Concave functions exhibit several important properties that distinguish them from other function classes. A function f is concave if and only if its negative -f is , establishing a direct duality between the two concepts. For twice continuously differentiable functions, concavity holds if and only if the D^2 f(x) is negative semidefinite at every point x in the domain, meaning all eigenvalues of the Hessian are less than or equal to zero. Additionally, the superlevel sets \{ x \in S \mid f(x) \geq \alpha \} of a concave function are sets, which facilitates analysis in geometric and optimization contexts. Concave functions play a central role in optimization theory, where maximizing a objective function over a feasible set guarantees that any local maximum is a global maximum, simplifying the solution of such problems via conditions. In , they model key behaviors such as diminishing marginal in consumer preference functions, where the marginal rate of substitution decreases as increases, and diminishing marginal returns in production functions. These applications extend to broader fields like and , where concavity ensures well-behaved algorithms for problems involving and .

Definition and Interpretation

Formal Definition

A f: X \to \mathbb{R}, where X is a subset of \mathbb{R}^n, is if for all x, y \in X and \lambda \in [0,1], f(\lambda x + (1-\lambda)y) \geq \lambda f(x) + (1-\lambda) f(y). This inequality expresses the idea that the value at any of points lies above the corresponding of values, ensuring the of the bends "upward" relative to the connecting any two points in the . An equivalent formulation is that f is concave if and only if -f is convex. This duality highlights the symmetry between concavity and convexity in optimization theory. The domain X must be convex to guarantee that convex combinations remain within X, and the function is typically real-valued over vector spaces like \mathbb{R}^n. Affine functions, of the form h(x) = a^\top x + b for a \in \mathbb{R}^n and b \in \mathbb{R}, satisfy both the concave and convex inequalities with equality and thus represent the boundary case between the two classes.

Geometric Interpretation

A concave function exhibits a distinctive geometric shape in its over a . Unlike the of a , which lies below any connecting two points on it, the of a concave function lies above such chords, meaning the segment connecting any two points on the falls entirely below or on the itself. This visually distinguishes concavity as a form of "upward bowing" or dished , where the function values between points exceed the straight-line connection, evoking an inverted bowl or saucer-like appearance in plots. The hypograph of a concave function f, defined as the set \{(x, t) \mid x \in \dom f, \, t \leq f(x)\} consisting of all points below or on the graph, forms a in the product space. This convexity of the hypograph provides a foundational geometric equivalent to the function's concavity, mirroring how the epigraph characterizes convexity for the opposite case. Geometrically, the condition illustrates this through : for any distinct points x and y in the domain and \theta \in (0,1), the value f(\theta x + (1-\theta) y) lies at or above the interpolated value \theta f(x) + (1-\theta) f(y), ensuring the graph arches above the . This intuitive bending reinforces the function's tendency toward higher intermediate values relative to straight-line approximations, underscoring its role in capturing peaking or diminishing behaviors in graphical representations.

Properties and Characterizations

Univariate Case

In the univariate case, a f: \mathbb{R} \to \mathbb{R} is if it satisfies the general definition of , meaning the lies above any connecting two points in its domain. For differentiable , a first-order condition characterizes : f is if and only if its f' is nonincreasing on the domain, which implies that for all x, y in the domain with x < y, f'(y) \leq f'(x). Equivalently, the lies below its tangent lines, satisfying f(y) \leq f(x) + f'(x)(y - x) for all x, y in the domain. For twice continuously differentiable functions f: \mathbb{R} \to \mathbb{R}, the second derivative test provides a simple criterion: f is concave if and only if f''(x) \leq 0 for all x in the domain. This condition ensures the graph curves downward or is linear, reflecting the nonincreasing nature of f'. A key inequality for concave functions is , which states that for weights \lambda_i \geq 0 with \sum \lambda_i = 1 and points x_i in the domain, f\left( \sum \lambda_i x_i \right) \geq \sum \lambda_i f(x_i). This follows directly from the definition and extends to expectations: if X is a random variable with existing moments, then f(\mathbb{E}[X]) \geq \mathbb{E}[f(X)]. Concave functions on \mathbb{R} exhibit specific monotonicity behavior: they are either nonincreasing (f'(x) \leq 0 for all x), nondecreasing (f'(x) \geq 0 for all x), or achieve a global maximum at some point c, being nondecreasing on (-\infty, c] and nonincreasing on [c, \infty). Concavity is preserved under certain compositions: if g is concave and nondecreasing, and f is concave, then the composition g \circ f is concave. Additionally, composition with affine functions, such as f(Ax + b) where A is a scalar and b \in \mathbb{R}, maintains concavity.

Multivariate Case

In the multivariate setting, a twice continuously differentiable function f: \mathbb{R}^n \to \mathbb{R} defined on a convex domain is concave if and only if its Hessian matrix \nabla^2 f(x) is negative semidefinite at every point x in the domain. This condition means that for all vectors v \in \mathbb{R}^n and all x in the domain, v^T \nabla^2 f(x) v \leq 0. The negative semidefiniteness of the Hessian ensures that the function does not curve upward in any direction, generalizing the univariate second-derivative test to higher dimensions. Jensen's inequality extends to multivariate concave functions as follows: for a concave function f and a probability measure \mu on a convex set in \mathbb{R}^n, the integral satisfies \int f \, d\mu \geq f\left( \int x \, d\mu \right). This formulation captures the expectation-based version where, for a random vector X with distribution \mu, f(\mathbb{E}[X]) \geq \mathbb{E}[f(X)], provided X lies in the domain almost surely. The inequality reflects the function's tendency to lie above its chords, even under general probability measures beyond finite convex combinations. Concave functions are quasi-concave, meaning their superlevel sets \{x \mid f(x) \geq \alpha\} are convex for every \alpha \in \mathbb{R}. However, the converse does not hold; quasi-concave functions may fail to satisfy the full concavity inequality, as their upper-level sets are convex without requiring the function to lie above all secant lines. This distinction arises because quasi-concavity only demands convexity of superlevel sets, allowing for functions that are "bowl-shaped" in a weaker sense than full concavity. Certain operations preserve concavity for multivariate functions. Specifically, non-negative linear combinations of concave functions—such as \sum \alpha_i f_i where \alpha_i \geq 0 and \sum \alpha_i = 1—remain concave, as they correspond to convex combinations of the original functions. The pointwise infimum of a family of concave functions is concave, since the hypograph of the infimum is the intersection of the hypographs of the individual functions, which are convex sets. In contrast, the pointwise supremum of a family of concave functions is not necessarily concave, as the hypograph of the supremum is the union of the individual hypographs, which is not necessarily convex. For differentiability in the multivariate case, concave functions defined on open convex domains in \mathbb{R}^n are analyzed using Gâteaux and Fréchet derivatives. A concave function on such a domain is Gâteaux differentiable at an interior point x if the directional derivative exists in every direction h, given by \lim_{t \to 0} \frac{f(x + t h) - f(x)}{t}. If Gâteaux differentiability holds at x, then the function is necessarily Fréchet differentiable there, meaning the derivative is linear and continuous in the direction, satisfying f(x + h) = f(x) + \nabla f(x)^T h + o(\|h\|) as \|h\| \to 0. This equivalence stems from the sublinear growth bound imposed by concavity on convex domains.

Relation to Convexity

Equivalence with Convex Functions

A function f defined on a convex set is concave if and only if its negation -f is convex. This equivalence follows directly from the definitions: the concavity condition f(\lambda x + (1-\lambda)y) \geq \lambda f(x) + (1-\lambda) f(y) for \lambda \in [0,1] and x, y in the domain negates to -f(\lambda x + (1-\lambda)y) \leq \lambda (-f(x)) + (1-\lambda) (-f(y)), which is precisely the convexity inequality for -f. Conversely, if -f is convex, negating the convexity inequality recovers the concavity of f. This relationship has significant implications in optimization, where maximizing a concave objective function over a convex set is equivalent to minimizing the convex function -f over the same set, allowing the application of convex optimization techniques to concave maximization problems. In convex analysis, concave functions play a key role in duality theory, particularly in , where the dual problem involves maximizing a concave perturbation of the primal objective to derive strong duality bounds and optimality conditions. Borderline cases highlight the symmetry: affine functions, which satisfy equality in the convexity inequality, are both convex and concave, as are their negations. A function is strictly concave if and only if its negation is strictly convex, meaning the inequality is strict for \lambda \in (0,1) and x \neq y. The terminology for concave functions evolved alongside convex analysis in the early 20th century, with Johan Jensen's 1906 work on and establishing foundational concepts that directly influenced the definition of concavity as the negation of convexity.

Strict and Strong Concavity

A strictly concave function satisfies a strengthened version of the concavity inequality, where the function value at any convex combination of distinct points exceeds the corresponding convex combination of the function values. Formally, a function f defined on a convex set is strictly concave if, for all x, y in the domain with x \neq y and all \lambda \in (0,1), f(\lambda x + (1-\lambda) y) > \lambda f(x) + (1-\lambda) f(y). This strict inequality ensures that the of f lies strictly above the connecting any two points on the graph, precluding flat segments where equality would hold. Strong imposes an even stronger condition by quantifying the deviation above the with a term. A f is strongly concave with m > 0 if, for all x, y in the and all \lambda \in (0,1), f(\lambda x + (1-\lambda) y) \geq \lambda f(x) + (1-\lambda) f(y) + \frac{m}{2} \lambda (1-\lambda) \|x - y\|^2. This additional positive term guarantees growth in the amount by which the exceeds the , providing a uniform measure of "" that distinguishes it from mere strict . For twice-differentiable functions, strict concavity is characterized by the Hessian matrix being negative definite everywhere in the domain, meaning \nabla^2 f(x) \prec 0 for all x, which ensures the local second-order approximation is strictly concave. Strongly concave functions, being a subclass of strictly concave ones, inherit this property but with the stronger Hessian condition \nabla^2 f(x) \preceq -m I for some m > 0, implying unique global maximizers over convex sets when they exist. In contrast to non-strict concavity, strict concavity eliminates any linear segments in the function, while strong concavity's quadratic term ensures a bounded rate of deviation that aids in convergence analyses. The negation of a strictly (respectively, strongly) concave function is strictly (strongly) convex.

Examples

Basic Examples

Constant functions, such as f(x) = c where c is a , are trivially concave, as they satisfy the concavity with equality for all points in their domain. These functions have a zero , confirming their concavity everywhere. Linear functions of the form f(x) = ax + b, where a and b are real , are concave (and also ) over the entire real line, since affine functions preserve convex combinations exactly. Their is zero, aligning with the for concavity in the univariate case. A classic example of a strictly concave quadratic function is f(x) = -x^2, whose graph forms a downward-opening parabola. The second derivative f''(x) = -2 < 0 for all x demonstrates its strict concavity across the real numbers. The natural logarithmic function f(x) = \log x (using the natural base) is concave on the positive real numbers x > 0. This follows from its second derivative f''(x) = -1/x^2 < 0, which is negative everywhere in the domain. Piecewise linear functions can also be concave if their slopes are non-increasing across segments. For instance, the function defined as f(x) = x for $0 \leq x \leq 0.5 and f(x) = 1 - x for $0.5 \leq x \leq 1 (often called an inverted tent function) has slopes 1 and then -1, satisfying the condition for concavity on [0,1]. Such functions, with weakly decreasing slopes, are concave by the characterization for univariate cases.

Advanced Examples

One prominent example of a multivariate concave function arises in information theory with the Shannon entropy, defined for a probability vector p = (p_1, \dots, p_n) in the simplex as H(p) = -\sum_{i=1}^n p_i \log p_i. This function is concave on the probability simplex, as established by the concavity of the negative entropy term and Jensen's inequality applied to the convex function x \log x. In production theory, the Cobb-Douglas function f(x, y) = x^a y^{1-a} for x, y \geq 0 and a \in (0,1) is concave on the nonnegative orthant, since its Hessian matrix is negative semidefinite, reflecting constant returns to scale and diminishing marginal returns. Log-linear utility functions, such as u(x) = \log\left( \sum_{i=1}^n x_i \right) for x \in \mathbb{R}^n_{++}, are concave because the summation is affine and the logarithm is a concave nondecreasing function, preserving concavity under composition. The extended-real-valued indicator function of a convex set C \subseteq \mathbb{R}^n, defined as I_C(x) = 0 if x \in C and -\infty otherwise, is concave, as its hypograph coincides with the convex set C \times \mathbb{R}. Composition rules further illustrate advanced concavity: if g is concave and nondecreasing on [0, \infty) and f is concave with nonnegative values, then g \circ f is concave; for instance, \sqrt{f(x)} inherits concavity when f is nonnegative and concave.

Applications

Optimization Theory

In optimization theory, concave functions play a pivotal role in maximization problems due to their desirable properties that ensure global optimality. For an unconstrained concave function f defined on a convex domain, any local maximum is also a global maximum, as the function's graph lies below any tangent line, preventing higher values elsewhere. This property simplifies the search for optima, as local optimization techniques suffice to identify the global solution. When maximizing a concave objective function subject to convex constraints, the problem can be reformulated as a convex minimization problem by negating the objective, allowing the application of standard convex optimization techniques. Specifically, maximizing concave f(x) over a convex set is equivalent to minimizing the convex function -f(x) over the same set. The Karush-Kuhn-Tucker (KKT) conditions, which include stationarity, primal and dual feasibility, and complementary slackness, are then necessary and sufficient for global optimality under a constraint qualification such as . Thus, any point satisfying the KKT conditions for this reformulated problem yields the global maximum of the original concave maximization. For solving these problems algorithmically, gradient ascent is commonly used for smooth concave functions, iteratively updating the solution in the direction of the positive gradient to approach the maximum. For nonsmooth concave functions, subgradient methods extend this approach by selecting a subgradient from the subdifferential at each step and performing ascent updates, converging to an optimal solution under appropriate step-size rules. If the function is strictly concave, the maximizer is unique, ensuring a single global optimum.

Economics and Utility Theory

In economics, concave utility functions are fundamental to modeling consumer behavior, as their concavity reflects the principle of diminishing marginal utility, where additional units of a good provide progressively smaller increments in satisfaction. This property ensures that the second derivative of the utility function with respect to consumption is negative, leading to -averse preferences under uncertainty. Specifically, for a twice-differentiable utility function u(x), concavity implies u''(x) < 0, which aligns with the Arrow-Pratt measure of absolute aversion defined as r_A(x) = -\frac{u''(x)}{u'(x)} > 0, quantifying how much an individual dislikes . This framework, originating from expected utility theory, explains why -averse agents prefer certain outcomes over gambles with the same , as demonstrated by applied to concave utilities. Concave production functions similarly capture decreasing in economic modeling, where output increases at a diminishing rate with additional inputs, promoting realistic depictions of resource constraints and efficiency. A prominent example is the Constant (CES) production function, given by Y = A \left[ \sum_i \alpha_i K_i^\rho + (1 - \sum_i \alpha_i) L^\rho \right]^{1/\rho} with \rho \leq 1, which exhibits under these parameters and allows for flexible between factors like (K) and labor (L) while maintaining decreasing marginal productivity. This form has been widely adopted in growth models to analyze technological progress and factor shares, as it balances substitutability with concavity to avoid unbounded outputs. In consumer theory, concave utility functions interacting with linear budget constraints yield convex indifference curves, ensuring unique tangency solutions for optimal consumption bundles and supporting the existence of demand functions. The upper contour sets of such utilities are convex, implying that indifference curves bow inward toward the origin, which facilitates the graphical representation of trade-offs and substitution effects under price changes. This convexity is a direct consequence of the quasi-concavity of the utility (strengthened to strict concavity for smoother behavior), enabling the second-order conditions for maximization to hold in standard models. Welfare economics employs social functions to aggregate individual utilities in a manner that values , with the utilitarian —defined as W = \sum_i u_i(x_i)—being if each u_i is , thereby promoting Pareto-efficient allocations that weigh total without excessive tolerance. This approach underpins theorems like the First Theorem, where competitive equilibria maximize such functions under ideal conditions, and informs policy evaluations by penalizing distributions with high variance in utilities. An empirical application appears in portfolio theory, where the logarithmic utility function u(w) = \log(w) models risk-averse investors seeking to maximize long-term growth, as its relative risk aversion of unity leads to myopic allocation strategies independent of horizon length. In continuous-time settings, this yields the optimal fraction invested in the market portfolio as \pi = \frac{\mu - r}{\sigma^2}, balancing \mu against \sigma^2, and has been validated in models for its simplicity and alignment with observed diversification behavior.

References

  1. [1]
    [PDF] 21. Concave and Quasiconcave Functions
    Dec 6, 2022 · A function f is concave if for all x,y ∈ S and 0 ≤ t ≤ 1, f tx + (1 − t)y ≥ tf(x) + (1 − t)f(y). Every chord of the graph lies on or below the ...
  2. [2]
    [PDF] Concave functions in economics 1. Preliminaries 1 2. Concave ...
    With concave functions, solving maximization problems is so much easier. If you can find a vector satisfying the first order conditions for a maximum, then you ...Missing: optimization | Show results with:optimization
  3. [3]
    [PDF] Convex Optimization
    This book is about convex optimization, a special class of mathematical optimiza- tion problems, which includes least-squares and linear programming ...
  4. [4]
    [PDF] Convex functions
    f is concave if −f is convex. ▷ f is strictly convex if dom f is convex and for x, y ∈ dom f, x ≠ y, 0 <𝜃< 1, f (𝜃x + (1 − 𝜃)y) < 𝜃f (x)+(1 − 𝜃)f (y).<|control11|><|separator|>
  5. [5]
    [PDF] 1 Theory of convex functions - Princeton University
    Let's first recall the definition of a convex function. Definition 1. A function f : Rn → R is convex if its domain is a convex set and for all x, y.
  6. [6]
    None
    Below is a merged summary of the sections on concave functions, geometric interpretation, chords, and hypograph from "Convex Optimization" by Boyd & Vandenberghe, based on the provided summaries from various pages of the book (https://web.stanford.edu/~boyd/cvxbook/bv_cvxbook.pdf). To retain all information in a dense and organized manner, I will use a table in CSV format for detailed content, followed by a concise narrative summary. The table captures page-specific details, definitions, properties, and additional context, while the narrative provides an overarching synthesis.
  7. [7]
    [PDF] Introduction to Real Analysis Liviu I. Nicolaescu University of Notre ...
    May 2, 2025 · ... concave function defined on an interval. I. Then for any n P N, any ... graph lies above any chord and below any tangent. Denote by pk ...
  8. [8]
    [PDF] Geometry of Convex Functions - Stanford CCRMA
    Figure 85: When a real function f is differentiable at each point in its open domain, there is an intuitive geometric interpretation of function convexity in ...
  9. [9]
    [PDF] Quasi-concave functions and concave functions. - Faculty
    Quasi-concave functions and concave functions. ▷ If f is concave, then it is quasi-concave, so you might start by checking for concavity. ▷ If f is a ...
  10. [10]
    [PDF] Topic 18: Differentiability
    3 Lemma Let f be a concave function on the open convex set C ⊂ Rm. If f has a Gâteaux derivative at x, then it is a Fréchet derivative. Proof: Let v 7→ f.
  11. [11]
    [PDF] Negative Semidefiniteness, and Concave and Quasiconcave ...
    Oct 25, 2012 · If D2 f (x) is ND, then the function is strictly concave. Proof. We first show that concavity implies Hessian matrix is NSD. Suppose f is ...
  12. [12]
    [PDF] Convexity - CMU Math
    Constant functions f(x) = c are both convex and concave. Powers of x: f(x) = xr with r ≥ 1 are convex on the interval 0 <x< ∞, and with 0 < r ≤ 1 are concave ...
  13. [13]
    [PDF] Concave and Convex Functions1
    Given the graph of a function, the hypograph of f, written hypf, is the set of points that lies on or below the graph of f, while the epigraph of f, written ...
  14. [14]
    [PDF] 5.2 The Natural Logarithmic Function Definition: ln(x)
    5.2 The Natural Logarithmic Function. Reminders: 1. Sign up through WebAssign ... (Concave up/down) d2 dx2 ln(x) = 1/x2 < 0 for x > 0: always concave ...
  15. [15]
    [PDF] EE364a Homework 3 solutions
    Each activity generates revenue, which is a piecewise-linear concave function of the activity level: rj(xj) = ( pjxj. 0 ≤ xj ≤ qj pjqj + pdisc j. (xj − qj) ...
  16. [16]
    [PDF] Entropy, Relative Entropy and Mutual Information
    One of the consequences of the concavity of entropy is that mixing two gases of equal entropy results in a gas with higher entropy. Theorem. 2.7.4: Let (X, Y) ...
  17. [17]
    [PDF] Proving that a Cobb-Douglas function is concave if the sum ... - Faculty
    A function F is quasi-concave if h(x) = g(F(x)) is a concave function for some strictly increasing function g from < to <. You should be able to prove this.
  18. [18]
    [PDF] Optimization for ML
    Gradient descent is a first-order iterative optimization algorithm for finding the minimum/maximum of a function. For a optimization problem with a concave ...
  19. [19]
    [PDF] Simple Routines for Optimization
    Feb 12, 2004 · Figure 8 shows a piecewise-linear concave function. Figure 9 illustrates the subdifferential for a concave function. 4.4 Computing Subgradients.
  20. [20]
    Risk Aversion in the Small and in the Large - jstor
    This paper concerns utility functions for money. A measure of risk aversion in the small, the risk premium or insurance premium for an arbitrary risk, and a ...
  21. [21]
    Microeconomic Theory - Andreu Mas-Colell; Michael D. Whinston
    $$158.99A text that provides balanced and in-depth analysis of the essentials of microeconomics. Masterfully combining the results of years of teaching microeconomics ...
  22. [22]
    Capital-Labor Substitution and Economic Efficiency - jstor
    K. J. Arrow, H. B. Chenery, B. S. Minhas, and R. M. Solow. IN many branches ... CES production function. 16 The sectors covered are both consumer goods ...