Fact-checked by Grok 2 weeks ago

Mean value theorem

The Mean Value Theorem (MVT) is a cornerstone theorem in that establishes a relationship between the average rate of change of a over an and its instantaneous rate of change at some point within that . Specifically, if a f is continuous on the closed [a, b] and differentiable on the open (a, b), then there exists at least one point c in (a, b) such that f'(c) = \frac{f(b) - f(a)}{b - a}. This geometric interpretation implies that the tangent line to the curve at c is parallel to the connecting the endpoints (a, f(a)) and (b, f(b)). The theorem's origins trace back to early mathematical developments, with an initial version appearing in the 14th or 15th century through the work of the Indian astronomer and mathematician Parameshvara from the Kerala school. It was further refined in the late 17th century by , whose theorem (a special case of the MVT when f(a) = f(b)) provided a proof for roots of . The modern formulation is attributed to in his 1797 treatise Théorie des fonctions analytiques, where he used it to justify Taylor expansions, though formalized it rigorously in 1823 using limits. Later refinements by Joseph Serret and Pierre Ossian Bonnet in 1868 removed the need for continuity of the derivative, aligning with the standard statement taught today. The MVT holds profound importance as a foundational tool in , serving as a bridge between global function behavior and local properties. It underpins proofs of key results, such as the , by linking integrals (average rates) to (instantaneous rates). For instance, it demonstrates that if f'(x) = 0 on an interval, then f is , and if two functions have the same , they differ by a . Applications extend to optimization, where it guarantees critical points for monotonicity , and to physics for modeling and in . Generalizations of the MVT include Cauchy's Mean Value Theorem, which extends to quotients of functions, stating that for continuous f and g (with g' \neq 0) on [a, b], there exists c in (a, b) such that \frac{f'(c)}{g'(c)} = \frac{f(b) - f(a)}{g(b) - g(a)}. This leads to for limits of indeterminate forms. In higher dimensions, vector versions apply to fields and , while integral forms relate to average values over intervals. Overall, the theorem's elegance and utility make it indispensable for understanding continuity, differentiability, and real-world modeling.

Core Theorem

Statement

The mean value theorem states that if a function f is continuous on the closed interval [a, b] and differentiable on the open interval (a, b), then there exists at least one point c \in (a, b) such that f'(c) = \frac{f(b) - f(a)}{b - a}. This formulation assumes the standard conditions of continuity at the endpoints and differentiability in the interior, ensuring the existence of such a c without specifying its uniqueness. Geometrically, the theorem implies that there is a point on the of f where the line is parallel to the connecting the endpoints (a, f(a)) and (b, f(b)), reflecting the average rate of change over the interval. For example, consider f(x) = x^2 on [0, 1]. Here, \frac{f(1) - f(0)}{1 - 0} = 1, and f'(x) = 2x, so c = \frac{1}{2} satisfies f'(c) = 1. This special case where f(a) = f(b) reduces to , asserting f'(c) = 0 for some c.

Proof

The proof of the Mean Value Theorem proceeds by constructing an and applying , a prerequisite result that guarantees the existence of a point where the vanishes for a satisfying certain conditions. Consider the g(x) = f(x) - f(a) - \frac{f(b) - f(a)}{b - a}(x - a) for x \in [a, b]. This represents the difference between f(x) and the () connecting the points (a, f(a)) and (b, f(b)). First, verify the boundary values: g(a) = f(a) - f(a) - \frac{f(b) - f(a)}{b - a}(a - a) = 0, and g(b) = f(b) - f(a) - \frac{f(b) - f(a)}{b - a}(b - a) = f(b) - f(a) - (f(b) - f(a)) = 0. Thus, g(a) = g(b) = 0./04:_Applications_of_Derivatives/4.04:_The_Mean_Value_Theorem) Next, confirm the hypotheses of Rolle's theorem hold for g. Since f is continuous on [a, b] and the secant line term \frac{f(b) - f(a)}{b - a}(x - a) is a linear (hence continuous) function on [a, b], their difference g is continuous on [a, b]. Similarly, f is differentiable on (a, b), and the secant term is differentiable everywhere, so g is differentiable on (a, b). With g(a) = g(b) = 0, Rolle's theorem applies, yielding some c \in (a, b) such that g'(c) = 0. Differentiating g gives g'(x) = f'(x) - \frac{f(b) - f(a)}{b - a}. Setting g'(c) = 0 therefore implies f'(c) - \frac{f(b) - f(a)}{b - a} = 0, or equivalently, f'(c) = \frac{f(b) - f(a)}{b - a}, which is the conclusion of the Mean Value Theorem.

Historical Context

Precursors

The foundations of the mean value theorem trace back to medieval mathematical developments, with an early version appearing in the work of the Indian mathematician and astronomer Parameshvara (c. 1370–1460) from the school of astronomy and mathematics. In his astronomical calculations, Parameshvara stated a mean value-type formula for the inverse tangent function, providing one of the first known instances of a result akin to the MVT. Earlier geometric investigations into tangents and areas laid groundwork for differential and integral concepts. In the 11th century, (Alhazen, c. 965–1040 ), in his , used summation techniques similar to Riemann sums to compute volumes of paraboloids and model light propagation, contributing to early ideas of integration. In the , as emerged, key figures refined methods for finding tangents and relating them to s. (1607–1665) developed a method for finding tangents by considering limiting slopes using s, applied to curves like y = x^n. (1630–1677), in his Lectiones geometricae (1670), geometrically connected tangents to areas under curves through increments, anticipating links between average and instantaneous changes. Michel Rolle (1652–1719) further advanced this in 1691 with , proving that if a is continuous on [a, b], differentiable on (a, b), and f(a) = f(b), then there exists c in (a, b) such that f'(c) = 0—a special case of the MVT. Leonhard Euler (1707–1783) extended these ideas in the early through his Institutiones calculi differentialis (1755), discussing averages in infinite series expansions to approximate functions, building toward more general theorems. These precursors culminated in 's formal statement of the theorem later in the century.

Lagrange's Contribution

formalized the mean value theorem in his 1797 publication Théorie des fonctions analytiques, specifically in Section 4 (articles 45–53), where he presented it as a foundational result for theory. This work, printed in by L’Imprimerie de la République, marked a significant advancement in by avoiding infinitesimals and emphasizing algebraic methods. Lagrange's primary motivation was to rigorously connect finite differences to derivatives within the framework of expansions, thereby establishing a solid basis for that aligned with his vision of pure analysis. He built upon earlier ideas from precursors such as Leonhard Euler, whom he explicitly credited for contributions to concepts in articles 13 and 14. By doing so, Lagrange provided a clean, general statement of the , describing it as “a which is new and remarkable for its simplicity and generality.” The key innovation in Lagrange's approach was his proof of the mean value theorem using an , which demonstrated the existence of a point where the equals the average rate of change. This method was instrumental in deriving the remainder term for , allowing for precise error estimates in series approximations and underscoring the theorem's utility in function analysis. Lagrange's formulation thus elevated the theorem from informal precursors to a cornerstone of modern .

Implications

Rolle's Theorem Connection

Rolle's theorem states that if a f is continuous on the closed [a, b], differentiable on the open (a, b), and f(a) = f(b), then there exists at least one point c in (a, b) such that f'(c) = 0. This result, named after the French mathematician Michel Rolle, was first proved in 1691 as part of his work on the separation of roots of polynomials by their derivatives. Rolle's theorem is limited to cases where the function values at the endpoints are equal, implying the existence of a within the . The mean value theorem generalizes by relaxing the condition f(a) = f(b) and instead guaranteeing a point where the equals the rate of change, or secant slope, \frac{f(b) - f(a)}{b - a}. This extension accounts for non-zero endpoint differences, broadening the theorem's applicability to scenarios involving linear approximations and error bounds in analysis. A standard proof of the mean value theorem relies on applied to an auxiliary function. Define g(x) = f(x) - f(a) - \frac{f(b) - f(a)}{b - a}(x - a); then g(a) = 0 and g(b) = 0, so by , there exists c \in (a, b) such that g'(c) = 0. Substituting yields f'(c) = \frac{f(b) - f(a)}{b - a}. This construction demonstrates how serves as a foundational special case, enabling the derivation of the more general mean value theorem. In optimization, underpins techniques for locating critical points where derivatives vanish, forming a basis for analyzing extrema in constrained intervals.

Applications in Analysis

The mean value theorem (MVT) provides a criterion for the monotonicity of differentiable functions. Specifically, if f is continuous on [a, b] and differentiable on (a, b) with f'(x) > 0 for all x \in (a, b), then f is strictly increasing on [a, b]. To prove this using the MVT, suppose for contradiction that there exist points x, y \in [a, b] with a \leq x < y \leq b such that f(x) \geq f(y). By the MVT applied to f on [x, y], there exists c \in (x, y) such that f'(c) = \frac{f(y) - f(x)}{y - x} \leq 0, since y - x > 0 and f(y) - f(x) \leq 0. This contradicts the assumption that f'(x) > 0 everywhere in (a, b). A similar argument shows that f'(x) < 0 implies f is strictly decreasing. The MVT also underpins L'Hôpital's rule, a key tool for evaluating limits of indeterminate forms. Consider functions f and g that are differentiable near a point a with f(a) = g(a) = 0 and g'(x) \neq 0 near a, where \lim_{x \to a} \frac{f(x)}{g(x)} is of the form \frac{0}{0}. By Cauchy's generalized mean value theorem (a direct extension of the MVT), for x near a, there exists c between a and x such that \frac{f(x) - f(a)}{g(x) - g(a)} = \frac{f'(c)}{g'(c)}, so \frac{f(x)}{g(x)} = \frac{f'(c)}{g'(c)}. Taking the limit as x \to a yields \lim_{x \to a} \frac{f(x)}{g(x)} = \lim_{x \to a} \frac{f'(x)}{g'(x)}, provided the latter limit exists. This connection extends to other indeterminate forms like \frac{\infty}{\infty} via appropriate transformations. In approximation theory, the MVT yields bounds on function growth controlled by the derivative. If f is continuous on [a, b] and differentiable on (a, b) with |f'(x)| \leq M for some constant M > 0 and all x \in (a, b), then |f(b) - f(a)| \leq M(b - a). This follows directly from the MVT: there exists c \in (a, b) such that f(b) - f(a) = f'(c)(b - a), so |f(b) - f(a)| = |f'(c)|(b - a) \leq M(b - a). Such estimates are fundamental for assessing error in linear approximations and understanding when M is finite.

Cauchy's Extension

Statement

Cauchy's mean value theorem is a of the mean value theorem to two functions. If functions f and g are on the closed interval [a, b], differentiable on the open interval (a, b), with g'(x) \neq 0 for all x \in (a, b) and g(b) \neq g(a), then there exists at least one point c \in (a, b) such that \frac{f'(c)}{g'(c)} = \frac{f(b) - f(a)}{g(b) - g(a)}. This assumes the standard conditions of and differentiability, ensuring the existence of such a c without specifying uniqueness. When g(x) = x, it reduces to the standard mean value theorem. Geometrically, considering the parametric curve (g(x), f(x)) for x \in [a, b], the theorem implies there is a point where the (g'(c), f'(c)) is to the vector connecting the endpoints (g(a), f(a)) and (g(b), f(b)). For example, consider f(x) = x^2 and g(x) = x on [1, 3]. Here, \frac{f(3) - f(1)}{g(3) - g(1)} = \frac{9 - 1}{3 - 1} = 4, and f'(x) = 2x, g'(x) = 1, so \frac{f'(c)}{g'(c)} = 2c = 4 gives c = 2.

Proof

The proof of Cauchy's mean value theorem proceeds by constructing an and applying . Assume g(b) \neq g(a); the case g(b) = g(a) follows similarly if f(b) = f(a), or requires separate handling. Consider the \phi(x) = f(x) - f(a) - \frac{f(b) - f(a)}{g(b) - g(a)} (g(x) - g(a)) for x \in [a, b]. This represents the difference between f(x) and the relative to g. Verify the boundary values: \phi(a) = f(a) - f(a) - \frac{f(b) - f(a)}{g(b) - g(a)}(g(a) - g(a)) = 0, and \phi(b) = f(b) - f(a) - \frac{f(b) - f(a)}{g(b) - g(a)}(g(b) - g(a)) = f(b) - f(a) - (f(b) - f(a)) = 0. Thus, \phi(a) = \phi(b) = 0. Since f and g are continuous on [a, b] and differentiable on (a, b), the linear term is also continuous and differentiable, so \phi is continuous on [a, b] and differentiable on (a, b). By , there exists c \in (a, b) such that \phi'(c) = 0. Differentiating gives \phi'(x) = f'(x) - \frac{f(b) - f(a)}{g(b) - g(a)} g'(x). Setting \phi'(c) = 0 implies f'(c) = \frac{f(b) - f(a)}{g(b) - g(a)} g'(c), or equivalently, \frac{f'(c)}{g'(c)} = \frac{f(b) - f(a)}{g(b) - g(a)}, assuming g'(c) \neq 0.

Multivariable Extensions

Several Variables

The mean value theorem extends to scalar-valued s of several variables, providing a relationship between the difference in function values at two points and the at an intermediate point along the connecting . Specifically, consider a f: \mathbb{R}^n \to \mathbb{R} that is continuous on the closed joining points a, b \in \mathbb{R}^n and differentiable on the open segment between them. Then there exists some point c on the open segment such that f(b) - f(a) = \nabla f(c) \cdot (b - a), where \nabla f(c) denotes the vector of f at c. This formulation highlights the vector form of the theorem, where the change in f is expressed as the dot product of the gradient at c with the displacement vector b - a. The gradient \nabla f(c) = \left( \frac{\partial f}{\partial x_1}(c), \dots, \frac{\partial f}{\partial x_n}(c) \right) captures the directional sensitivity of f in the direction of b - a. For the theorem to hold, stronger assumptions are typically required, such as f being continuously differentiable (i.e., C^1) along the , ensuring the partial derivatives exist and are continuous in a neighborhood of the segment. To illustrate, take f(x, y) = x^2 + y^2 on \mathbb{R}^2, which is continuously differentiable everywhere, and consider the segment from a = (0, 0) to b = (1, 1). Here, f(b) - f(a) = (1^2 + 1^2) - (0^2 + 0^2) = 2. The is \nabla f(x, y) = (2x, 2y), so the theorem guarantees some c = (t, t) with $0 < t < 1 such that $2 = \nabla f(c) \cdot (1, 1) = (2t, 2t) \cdot (1, 1) = 4t, yielding t = 1/2 and c = (1/2, 1/2). This example demonstrates how the theorem equates the net change to a directional derivative scaled by the segment length. This multivariable version aligns with the first-order term in the multivariable , approximating the function's linear behavior near a point.

Vector-Valued Functions

The mean value theorem extends naturally to vector-valued functions \mathbf{F}: [a, b] \to \mathbb{R}^m that are continuous on the closed interval [a, b] and differentiable on the open interval (a, b). Since \mathbf{F}(t) = (f_1(t), \dots, f_m(t)) where each f_i: [a, b] \to \mathbb{R} satisfies the hypotheses of the scalar mean value theorem, the result applies componentwise: for each i = 1, \dots, m, there exists c_i \in (a, b) such that f_i(b) - f_i(a) = f_i'(c_i) (b - a). Thus, \mathbf{F}(b) - \mathbf{F}(a) = (b - a) \begin{pmatrix} f_1'(c_1) \\ \vdots \\ f_m'(c_m) \end{pmatrix}, where the derivative vector \mathbf{F}'(t) = (f_1'(t), \dots, f_m'(t))^T is evaluated at possibly different points c_i. This componentwise form highlights that, unlike the scalar case, there is no guarantee of a single point c where the full vector equality \mathbf{F}(b) - \mathbf{F}(a) = \mathbf{F}'(c) (b - a) holds, as the average rate of change need not align exactly with any single instantaneous derivative vector. A unified perspective uses directional derivatives via projections. For any vector \mathbf{v} \in \mathbb{R}^m, the scalar function g(t) = \mathbf{v} \cdot \mathbf{F}(t) is differentiable on (a, b) and continuous on [a, b], so the mean value theorem yields a single c \in (a, b) such that \mathbf{v} \cdot (\mathbf{F}(b) - \mathbf{F}(a)) = \mathbf{v} \cdot \mathbf{F}'(c) (b - a). This holds for all \mathbf{v}, providing a matrix-like interpretation where the Jacobian (here, the row vector \mathbf{F}'(c)) acts linearly on the scalar difference b - a. From the directional form, the mean value inequality follows by normalizing: let \mathbf{u} = (\mathbf{F}(b) - \mathbf{F}(a)) / \|\mathbf{F}(b) - \mathbf{F}(a)\| (assuming \mathbf{F}(b) \neq \mathbf{F}(a)), then \|\mathbf{F}(b) - \mathbf{F}(a)\| = \mathbf{u} \cdot \mathbf{F}'(c) (b - a) \leq \|\mathbf{F}'(c)\| (b - a) \leq (b - a) \sup_{t \in [a, b]} \|\mathbf{F}'(t)\|, where the first inequality uses the and the second uses the definition of the supremum. This bounds the net change in terms of the maximum speed, analogous to arc length estimates in curve analysis. For vector-valued functions \mathbf{F}: U \to \mathbb{R}^m where U \subset \mathbb{R}^n is open and \mathbf{F} is differentiable, consider points x, y \in U such that the line segment joining them lies in U. Parametrize the segment as \gamma(t) = x + t(y - x) for t \in [0, 1]. By the chain rule, \mathbf{F}(y) - \mathbf{F}(x) = \int_0^1 \frac{d}{dt} \mathbf{F}(\gamma(t)) \, dt = \int_0^1 D\mathbf{F}(\gamma(t)) (y - x) \, dt, where D\mathbf{F} is the m \times n Jacobian matrix. This integral form reduces the multivariable case to a 1D integral along the path, from which componentwise or directional mean value estimates apply by integrating the scalar results for each entry or projection. In general, no single Jacobian satisfies \mathbf{F}(y) - \mathbf{F}(x) = D\mathbf{F}(c) (y - x) for some c on the segment, distinguishing it from the scalar multivariable case where the gradient provides an exact linear approximation in direction y - x. The corresponding inequality is \|\mathbf{F}(y) - \mathbf{F}(x)\| \leq \|y - x\| \sup \{\|D\mathbf{F}(z)\|_{\mathrm{op}} : z \text{ on segment}\}, with the operator norm \| \cdot \|_{\mathrm{op}} measuring the maximum stretch.

Applicability Limits

Required Conditions

The Mean Value Theorem (MVT) for a real-valued function f on an interval requires two core hypotheses: f must be continuous on the closed interval [a, b] and differentiable on the open interval (a, b). These conditions ensure the existence of at least one point c \in (a, b) where f'(c) = \frac{f(b) - f(a)}{b - a}. Continuity on [a, b] is essential to guarantee that f attains its endpoint values and that the secant line slope is well-defined, avoiding jumps or discontinuities that could prevent the function from satisfying the intermediate value property needed for the theorem's validity. Without this, a function might connect equal endpoint values without crossing zero in its difference quotient, as in the case of , a special instance of the MVT. Differentiability on (a, b) is equally critical, as the theorem equates the secant slope to a derivative value, which must exist at some interior point. This condition fails for functions that are continuous everywhere but differentiable nowhere, such as the Weierstrass function f(x) = \sum_{n=0}^{\infty} a^n \cos(b^n \pi x) where $0 < a < 1 and ab > 1 + \frac{3\pi}{2}; in such cases, no c exists where f'(c) is defined, so the MVT conclusion cannot hold on any interval. For Cauchy's extension of the MVT, which involves two functions f and g, the conditions mirror the standard MVT—both must be continuous on [a, b] and differentiable on (a, b)—with the additional requirement that g'(x) \neq 0 for all x \in (a, b) to ensure the ratio \frac{f'(c)}{g'(c)} is defined and avoids singularities in the generalized slope equality \frac{f(b) - f(a)}{g(b) - g(a)} = \frac{f'(c)}{g'(c)}. In multivariable extensions for a scalar f: \mathbb{R}^n \to \mathbb{R}, the hypotheses adapt to the joining points a, b \in \mathbb{R}^n: f must be on the closed segment and differentiable on its relative interior, with the segment's straight-line regularity ensuring the path is smooth enough for the conclusion f(b) - f(a) = \nabla f(c) \cdot (b - a) to hold for some c on the open segment. For vector-valued functions or more general paths, similar and differentiability conditions apply, but the path must exhibit sufficient regularity, such as being C^1, to parameterize it appropriately and extend the theorem componentwise or along the .

Failure Cases

The Mean Value Theorem requires a function to be continuous on the closed interval [a, b] and differentiable on the open interval (a, b); violations of these conditions lead to failures where no point c \in (a, b) exists such that f'(c) = \frac{f(b) - f(a)}{b - a}. A classic failure due to discontinuity occurs with step functions. Consider f(x) = \begin{cases} x + 1 & \text{if } x < 1 \\ x - 1 & \text{if } x \geq 1 \end{cases} on [0, 2]. This function has a jump discontinuity at x = 1, so it is not continuous on [0, 2]. The secant slope is \frac{f(2) - f(0)}{2 - 0} = \frac{(2-1) - (0+1)}{2} = 0, but where defined, f'(x) = 1 everywhere in (0, 2), so no c satisfies f'(c) = 0. Non-differentiability also causes failure, even if continuity holds. For f(x) = |x| on [-1, 1], the function is continuous on [-1, 1] but not differentiable at x = 0 due to the corner. The secant slope is \frac{f(1) - f(-1)}{1 - (-1)} = \frac{1 - 1}{2} = 0, but f'(x) = -1 for x < 0 and f'(x) = 1 for x > 0, so no c \in (-1, 1) has f'(c) = 0. In the multivariable setting for scalar-valued functions f: \mathbb{R}^n \to \mathbb{R}, the theorem states that if f is continuous on the joining x and y and differentiable on its relative interior, then there exists z on the segment such that f(y) - f(x) = \nabla f(z) \cdot (y - x). Failure occurs similarly if the restriction to the segment violates continuity or differentiability. For instance, the f(\mathbf{r}) = 0 if the first coordinate is negative and $1 otherwise, restricted to the segment from (-1, 0, \dots, 0) to (1, 0, \dots, 0), yields a step discontinuity at the , with f(1,0,\dots,0) - f(-1,0,\dots,0) = 1 but \nabla f = 0 where defined, analogous to the single-variable step case. Darboux's theorem establishes that derivatives, where they exist on an interval, satisfy the intermediate value property: if f' exists on (a, b) and f'(p) < k < f'(q) for p, q \in (a, b), then there exists r \in (p, q) with f'(r) = k. However, the Mean Value Theorem demands differentiability throughout (a, b) to guarantee a derivative matching the secant slope, which Darboux's property alone cannot ensure without full differentiability, as seen in the absolute value example.

Integral Variants

First Theorem for Integrals

The first mean value theorem for integrals states that if a f is continuous on the closed [a, b], then there exists some point c \in [a, b] such that \int_a^b f(x) \, dx = f(c) (b - a). This formulation expresses that the definite of f over [a, b] equals the value of the function at some point c multiplied by the length of the interval, effectively linking the integral to an "" function value. The proof relies on the continuity of f on the compact [a, b], which guarantees that f attains its minimum value m and maximum value M somewhere in the by the . Dividing the integral by the length yields the average value \frac{1}{b-a} \int_a^b f(x) \, dx, which must satisfy m \leq \frac{1}{b-a} \int_a^b f(x) \, dx \leq M. Since f is continuous, the ensures there exists c \in [a, b] where f(c) equals this average value. Unlike the classical mean value theorem for derivatives, which requires differentiability, this integral variant depends solely on continuity and does not involve derivatives or secant lines. For a constant function f(x) = k on [a, b], the theorem holds trivially for any c \in [a, b], as the integral is k(b - a) and f(c) = k. For a linear function such as f(x) = x on [0, 1], the integral is \frac{1}{2}, and c = \frac{1}{2} (the midpoint) satisfies f(c) \cdot 1 = \frac{1}{2}.

Second Theorem for Integrals

The second mean value theorem for integrals extends the basic mean value property to weighted integrals where one provides a monotonic weighting factor. Specifically, if f is continuous on the closed [a, b] and g is monotonic and non-negative on [a, b], then there exists some \xi \in [a, b] such that \int_a^b f(x) g(x) \, dx = f(\xi) \int_a^b g(x) \, dx. This form assumes g is non-increasing (by or sign adjustment if increasing). A more general variant, applicable when g is monotonic but not necessarily vanishing at an , states that there exist \xi, \eta \in [a, b] such that \int_a^b f(x) g(x) \, dx = g(a) \int_a^\xi f(x) \, dx + g(b) \int_\eta^b f(x) \, dx. The theorem, often referred to as Bonnet's theorem, was established by Pierre Ossian Bonnet in 1849 as part of his work on properties of definite integrals. To prove Bonnet's form, assume g is continuous, non-negative, and non-increasing on [a, b] with g(b) = 0 (a common case; the general case follows by subtracting g(b)). Define F(x) = \int_a^x f(t) \, dt. Integration by parts yields \int_a^b f(x) g(x) \, dx = \int_a^b F'(x) g(x) \, dx = F(b) g(b) - F(a) g(a) - \int_a^b F(x) g'(x) \, dx = - \int_a^b F(x) g'(x) \, dx, since F(a) = 0 and g(b) = 0. Here, -g' is non-negative (as g is non-increasing), so the first mean value theorem for integrals applies to F(x) (-g'(x)), giving some \xi \in [a, b] such that \int_a^b F(x) (-g'(x)) \, dx = F(\xi) \int_a^b (-g'(x)) \, dx = F(\xi) (g(a) - g(b)) = F(\xi) g(a). Thus, \int_a^b f(x) g(x) \, dx = g(a) F(\xi) = g(a) \int_a^\xi f(t) \, dt. Since f is continuous, by the intermediate value theorem, there exists \eta \in [a, \xi] such that f(\eta) (\xi - a) = \int_a^\xi f(t) \, dt, but adjusting for the weighting leads to the equivalent form with a single \xi. For the non-vanishing case, the proof adapts by applying the theorem to g(x) - g(b). This theorem differs from the first mean value theorem for integrals, which applies to unweighted averages (\int_a^b f(x) \, dx = f(\xi) (b - a)) by incorporating the monotonic weighting function g(x), allowing analysis of integrals where the weight varies systematically across the interval. Bonnet's result builds briefly on earlier ideas akin to Cauchy's mean value theorem but focuses on integral forms with monotonic integrators.

Broader Generalizations

Linear Algebra

In the context of linear algebra, the mean value theorem extends naturally to linear operators on normed vector spaces. For a bounded linear operator T: V \to W between normed spaces, the difference T(x) - T(y) equals T(x - y) exactly, since the Fréchet derivative of the map x \mapsto T(x) is the constant operator T itself. This yields an exact mean value property, analogous to the scalar case where the derivative is constant. Taking norms, \|T(x) - T(y)\| \leq \|T\| \cdot \|x - y\|, where \|T\| denotes the operator norm \sup_{\|v\| \leq 1} \|T(v)\|. This inequality mirrors the mean value inequality for differentiable functions and establishes Lipschitz continuity of T with constant \|T\|. In finite-dimensional spaces, such as \mathbb{R}^n with the Euclidean , the mean value theorem for s reduces to its componentwise form. A T: \mathbb{R}^n \to \mathbb{R}^m represented by a A satisfies T(x) - T(y) = A(x - y), and applying the scalar mean value theorem to each component of T along the from y to x confirms the exact equality without needing an intermediate point c. The bound then follows from the induced by the , ensuring the holds uniformly. This highlights how the theorem's persists in coordinate representations. A related generalization appears in , where serves as a mean value inequality for . For a convex function f: \mathbb{R}^n \to \mathbb{R} and weights \lambda_i \geq 0 summing to 1, f\left( \sum \lambda_i x_i \right) \leq \sum \lambda_i f(x_i), interpreting the left side as f at the (a "mean") and the right as the of the function values. This can be derived using the mean value theorem on the supporting hyperplane or by integrating the subgradient along line segments, as convexity implies the derivative (subgradient) is non-decreasing. thus extends the mean value concept to inequalities for nonlinear convex maps, foundational in optimization and matrix theory. In matrix perturbation theory, mean value principles underpin bounds on eigenvalue variations. For Hermitian matrices A and perturbation E, states |\lambda_k(A + E) - \lambda_k(A)| \leq \|E\|_2 for the k-th largest eigenvalue, providing an bound analogous to the mean value inequality. More precisely, the first-order perturbation to an eigenvalue \lambda of A under small E is the \langle u, E u \rangle, where u is the corresponding eigenvector—a "mean value" of E over the eigenspace. This exact linear term in the expansion reflects the constant derivative property for the eigenvalue map in the Hermitian setting.

Probability Theory

In , serves as a key analog of the mean value theorem for functions applied to . For a f: \mathbb{R} \to \mathbb{R} and a X with finite \mathbb{E}[X], the inequality asserts that \mathbb{E}[f(X)] \geq f(\mathbb{E}[X]). This result follows from the geometric property of functions, where the graph lies above any tangent line, a consequence of the mean value theorem applied to the or test for convexity. The inequality quantifies how variability in X increases the expected value of f(X) relative to the function at the mean, with equality holding if X is constant or f is affine. In the context of stochastic processes, Doob's martingale framework extends the mean value principle via conditional expectations. A stochastic process (X_n)_{n \geq 0} adapted to a filtration (\mathcal{F}_n)_{n \geq 0} is a martingale if \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] = X_n almost surely for each n, implying that the conditional expected future value equals the current value given available information. This property, introduced by Doob, mirrors the mean value theorem by preserving the "average" across time steps under uncertainty, enabling convergence theorems and optional stopping results essential for analyzing random walks, branching processes, and diffusion approximations. For submartingales and supermartingales, analogous inequalities hold, bounding deviations from the mean. A discrete analog of the mean value theorem arises in probability when approximating expectations through finite sums, bridging discrete and continuous models. For a function f on a partition of an interval, the difference f(b) - f(a) equals f'(c)(b - a) for some c \in (a, b), and in probabilistic settings, Riemann sums \sum f(x_i) \Delta x_i approximate \mathbb{E}[f(X)] = \int f(x) dP(x) with error controlled by the MVT, particularly for uniform partitions or Monte Carlo integration. This connection facilitates the transition from discrete sums of independent random variables to integral expectations in limit theorems. The mean value theorem underpins proofs of the law of large numbers in settings with bounded differences, via concentration inequalities like McDiarmid's. For a function g of independent random variables X_1, \dots, X_n where changing one X_i alters g by at most c_i, McDiarmid's inequality bounds the deviation \mathbb{P}(|g(X_1, \dots, X_n) - \mathbb{E}| \geq t) \leq 2 \exp(-2t^2 / \sum c_i^2), ensuring the sample average converges to the true mean with probability approaching 1 as n \to \infty. This bounded differences approach, relying on martingale increments from the MVT, proves uniform laws of large numbers for empirical measures and is pivotal in high-dimensional probability.

Complex Analysis

In complex analysis, the mean value property characterizes holomorphic functions through their behavior on disks. Specifically, if f is holomorphic on an open disk centered at a point a with radius r, then the value of the function at the center equals the average of its values over the circumference of the circle of radius r: f(a) = \frac{1}{2\pi} \int_0^{2\pi} f(a + r e^{i\theta}) \, d\theta. This integral represents the arithmetic mean of f on the boundary circle |z - a| = r. The proof of this property follows directly from , which states that for holomorphic f inside and on a simple closed C, f(a) = \frac{1}{2\pi i} \oint_C \frac{f(z)}{z - a} \, dz, where a is inside C. Parametrizing the circle C as z = a + r e^{i\theta} for \theta from 0 to $2\pi, with dz = i r e^{i\theta} \, d\theta, substitution yields f(a) = \frac{1}{2\pi i} \int_0^{2\pi} \frac{f(a + r e^{i\theta})}{r e^{i\theta}} i r e^{i\theta} \, d\theta = \frac{1}{2\pi} \int_0^{2\pi} f(a + r e^{i\theta}) \, d\theta, confirming the mean value expression. Holomorphic functions, being complex differentiable at every point in their domain, satisfy this property due to the integral representation enabled by their analyticity. Unlike the mean value theorem in , which involves averages over line segments or intervals, the complex analogue applies to integrals over closed curves such as circles, highlighting the conformal invariance and global nature of holomorphic functions. This property has significant implications, notably in deriving the . For a f on a bounded domain, if |f| attains its maximum at an interior point, then f is constant; otherwise, the maximum occurs on the boundary. The proof proceeds by contradiction: supposing a strict interior maximum at a, the mean value property implies |f(a)| \leq \frac{1}{2\pi} \int_0^{2\pi} |f(a + r e^{i\theta})| \, d\theta \leq |f(a)|, with equality only if |f| is constant on , propagating constancy throughout the domain by .