Envelope theorem
   HOME

TheInfoList



OR:

In mathematics and
economics Economics () is the social science that studies the production, distribution, and consumption of goods and services. Economics focuses on the behaviour and interactions of economic agents and how economies work. Microeconomics analyzes ...
, the envelope theorem is a major result about the differentiability properties of the
value function The value function of an optimization problem gives the value attained by the objective function at a solution, while only depending on the parameters of the problem. In a controlled dynamical system, the value function represents the optimal payof ...
of a parameterized optimization problem. As we change parameters of the objective, the envelope theorem shows that, in a certain sense, changes in the optimizer of the objective do not contribute to the change in the objective function. The envelope theorem is an important tool for
comparative statics In economics, comparative statics is the comparison of two different economic outcomes, before and after a change in some underlying exogenous parameter. As a type of ''static analysis'' it compares two different equilibrium states, after the ...
of
optimization Mathematical optimization (alternatively spelled ''optimisation'') or mathematical programming is the selection of a best element, with regard to some criterion, from some set of available alternatives. It is generally divided into two subfi ...
models. The term envelope derives from describing the graph of the value function as the "upper envelope" of the graphs of the parameterized family of functions \left\ _ that are optimized.


Statement

Let f(x,\alpha) and g_(x,\alpha), j = 1,2, \ldots, m be real-valued continuously
differentiable function In mathematics, a differentiable function of one real variable is a function whose derivative exists at each point in its domain. In other words, the graph of a differentiable function has a non-vertical tangent line at each interior point in it ...
s on \mathbb^, where x \in \mathbb^ are choice variables and \alpha \in \mathbb^ are parameters, and consider the problem of choosing x, for a given \alpha, so as to: : \max_ f(x, \alpha) subject to g_(x,\alpha) \geq 0, j = 1,2, \ldots, m and x \geq 0. The Lagrangian expression of this problem is given by :\mathcal (x, \lambda, \alpha) = f(x, \alpha) + \lambda \cdot g(x, \alpha) where \lambda \in \mathbb^ are the
Lagrange multiplier In mathematical optimization, the method of Lagrange multipliers is a strategy for finding the local maxima and minima of a function subject to equality constraints (i.e., subject to the condition that one or more equations have to be satisfied e ...
s. Now let x^(\alpha) and \lambda^(\alpha) together be the solution that maximizes the objective function ''f'' subject to the constraints (and hence are saddle points of the Lagrangian), :\mathcal^ (\alpha) \equiv f(x^(\alpha), \alpha) + \lambda^(\alpha) \cdot g(x^(\alpha), \alpha), and define the
value function The value function of an optimization problem gives the value attained by the objective function at a solution, while only depending on the parameters of the problem. In a controlled dynamical system, the value function represents the optimal payof ...
:V(\alpha) \equiv f(x^(\alpha), \alpha). Then we have the following theorem. Theorem: ''Assume that V and \mathcal are continuously differentiable. Then'' : \frac = \frac = \frac, k = 1, 2, \ldots, l ''where \partial \mathcal / \partial \alpha_ = \partial f / \partial \alpha_ + \lambda \cdot \partial g / \partial \alpha_''.


For arbitrary choice sets

Let X denote the choice set and let the relevant parameter be t\in \lbrack 0,1]. Letting f:X\times \lbrack 0,1]\rightarrow R denote the parameterized objective function, the value function V and the optimal choice correspondence (set-valued function) X^ are given by: "Envelope theorems" describe sufficient conditions for the value function V to be differentiable in the parameter t and describe its derivative as where f_ denotes the partial derivative of f with respect to t. Namely, the derivative of the value function with respect to the parameter equals the partial derivative of the objective function with respect to t holding the maximizer fixed at its optimal level. Traditional envelope theorem derivations use the first-order condition for (), which requires that the choice set X have the convex and topological structure, and the objective function f be differentiable in the variable x. (The argument is that changes in the maximizer have only a "second-order effect" at the optimum and so can be ignored.) However, in many applications such as the analysis of incentive constraints in contract theory and game theory, nonconvex production problems, and "monotone" or "robust" comparative statics, the choice sets and objective functions generally lack the topological and convexity properties required by the traditional envelope theorems.
Paul Milgrom Paul Robert Milgrom (born April 20, 1948) is an American economist. He is the Shirley and Leonard Ely Professor of Humanities and Sciences at Stanford University, the Stanford University School of Humanities and Sciences, a position he has held ...
and Segal (2002) observe that the traditional envelope formula holds for optimization problems with arbitrary choice sets at any differentiability point of the value function, provided that the objective function is differentiable in the parameter: Theorem 1: Let t\in \left( 0,1\right) and x\in X^\left(t\right) . If both V^\left( t\right) and f_\left(x,t\right) exist, the envelope formula () holds. Proof: Equation () implies that for x\in X^\left( t\right) , : \max_\left f\left( x,s\right) -V\left( s\right)\right=f\left( x,t\right) -V\left( t\right) =0. Under the assumptions, the objective function of the displayed maximization problem is differentiable at s=t, and the first-order condition for this maximization is exactly equation (). Q.E.D. While differentiability of the value function in general requires strong assumptions, in many applications weaker conditions such as
absolute continuity In calculus, absolute continuity is a smoothness property of functions that is stronger than continuity and uniform continuity. The notion of absolute continuity allows one to obtain generalizations of the relationship between the two central ope ...
, differentiability almost everywhere, or left- and right-differentiability, suffice. In particular, Milgrom and Segal's (2002) Theorem 2 offers a sufficient condition for V to be absolutely continuous, which means that it is differentiable almost everywhere and can be represented as an integral of its derivative: Theorem 2: Suppose that f(x,\cdot ) is absolutely continuous for all x\in X. Suppose also that there exists an integrable function b: ,1/math> \rightarrow \mathbb_ such that , f_(x,t), \leq b(t) for all x\in X and almost all t\in \lbrack 0,1]. Then V is absolutely continuous. Suppose, in addition, that f(x,\cdot ) is differentiable for all x\in X, and that X^(t)\neq \varnothing almost everywhere on ,1/math>. Then for any selection x^(t)\in X^(t), Proof: Using ()(1), observe that for any t^,t^\in \lbrack 0,1] with t^, : , V(t^)-V(t^), \leq \sup_, f(x,t^)-f(x,t^), =\sup_\left\vert \int_^f_(x,t)dt\right\vert \leq \int_^\sup_, f_(x,t), dt\leq \int_^b(t)dt. This implies that V is absolutely continuous. Therefore, V is differentiable almost everywhere, and using () yields (). Q.E.D. This result dispels the common misconception that nice behavior of the value function requires correspondingly nice behavior of the maximizer. Theorem 2 ensures the
absolute continuity In calculus, absolute continuity is a smoothness property of functions that is stronger than continuity and uniform continuity. The notion of absolute continuity allows one to obtain generalizations of the relationship between the two central ope ...
of the value function even though the maximizer may be discontinuous. In a similar vein, Milgrom and Segal's (2002) Theorem 3 implies that the value function must be differentiable at t=t_ and hence satisfy the envelope formula () when the family \left\ _ is equi-differentiable at t_\in \left( 0,1\right) and f_\left(X^\left( t\right) ,t_\right) is single-valued and continuous at t=t_, even if the maximizer is not differentiable at t_ (e.g., if X is described by a set of inequality constraints and the set of binding constraints changes at t_).


Applications


Applications to producer theory

Theorem 1 implies Hotelling's lemma at any differentiability point of the profit function, and Theorem 2 implies the producer surplus formula. Formally, let \pi \left( p\right) denote the profit function of a price-taking firm with production set X\subseteq \mathbb^ facing prices p\in \mathbb^, and let x^\left( p\right) denote the firm's supply function, i.e., : \pi (p)=\max_p\cdot x=p\cdot x^\left( p\right) \text Let t=p_ (the price of good i) and fix the other goods' prices at p_\in \mathbb^. Applying Theorem 1 to f(x,t)=tx_+p_\cdot x_ yields \frac=x_^(p) (the firm's optimal supply of good i). Applying Theorem 2 (whose assumptions are verified when p_ is restricted to a bounded interval) yields :\pi (t,p_)-\pi (0,p_)=\int_^x_^(s,p_)ds, i.e. the producer surplus \pi (t,p_)-\pi (0,p_) can be obtained by integrating under the firm's supply curve for good i.


Applications to mechanism design and auction theory

Consider an agent whose utility function f(x,t) over outcomes x\in \bar depends on his type t\in \lbrack 0,1]. Let X\subseteq \bar represent the "menu" of possible outcomes the agent could obtain in the mechanism by sending different messages. The agent's equilibrium utility V(t) in the mechanism is then given by (1), and the set X^(t) of the mechanism's equilibrium outcomes is given by (2). Any selection x^(t)\in X^(t) is a choice rule implemented by the mechanism. Suppose that the agent's utility function f(x,t) is differentiable and absolutely continuous in t for all x\in Y, and that \sup_, f_(x,t), is integrable on ,1/math>. Then Theorem 2 implies that the agent's equilibrium utility V in any mechanism implementing a given choice rule x^ must satisfy the integral condition (4). The integral condition (4) is a key step in the analysis of mechanism design problems with continuous type spaces. In particular, in Myerson's (1981) analysis of single-item auctions, the outcome from the viewpoint of one bidder can be described as x=\left( y,z\right) , where y is the bidder's probability of receiving the object and z is his expected payment, and the bidder's expected utility takes the form f\left( \left( y,z\right) ,t\right) =ty-z. In this case, letting \underline denote the bidder's lowest possible type, the integral condition (4) for the bidder's equilibrium expected utility V takes the form : V(t)-V(\underline)=\int_^y^(s)ds. (This equation can be interpreted as the producer surplus formula for the firm whose production technology for converting numeraire z into probability y of winning the object is defined by the auction and which resells the object at a fixed price t). This condition in turn yields Myerson's (1981) celebrated Revenue equivalence, revenue equivalence theorem: the expected revenue generated in an auction in which bidders have independent private values is fully determined by the bidders' probabilities y^\left( t\right) of getting the object for all types t as well as by the expected payoffs V(\underline) of the bidders' lowest types. Finally, this condition is a key step in Myerson's (1981) of optimal auctions. For other applications of the envelope theorem to mechanism design see Mirrlees (1971), Holmstrom (1979), Laffont and Maskin (1980), Riley and Samuelson (1981), Fudenberg and Tirole (1991), and Williams (1999). While these authors derived and exploited the envelope theorem by restricting attention to (piecewise) continuously differentiable choice rules or even narrower classes, it may sometimes be optimal to implement a choice rule that is not piecewise continuously differentiable. (One example is the class of trading problems with linear utility described in chapter 6.5 of Myerson (1991).) Note that the integral condition (3) still holds in this setting and implies such important results as Holmstrom's lemma (Holmstrom, 1979), Myerson's lemma (Myerson, 1981), the revenue equivalence theorem (for auctions), the Green–Laffont–Holmstrom theorem (Green and Laffont, 1979; Holmstrom, 1979), the Myerson–Satterthwaite inefficiency theorem (Myerson and Satterthwaite, 1983), the Jehiel–Moldovanu impossibility theorems (Jehiel and Moldovanu, 2001), the McAfee–McMillan weak-cartels theorem (McAfee and McMillan, 1992), and Weber's martingale theorem (Weber, 1983), etc. The details of these applications are provided in Chapter 3 of Milgrom (2004), who offers an elegant and unifying framework in auction and mechanism design analysis mainly based on the envelope theorem and other familiar techniques and concepts in demand theory.


Applications to multidimensional parameter spaces

For a multidimensional parameter space T\subseteq \mathbb^, Theorem 1 can be applied to partial and directional derivatives of the value function. If both the objective function f and the value function V are (totally) differentiable in t, Theorem 1 implies the envelope formula for their gradients: \nabla V\left( t\right) =\nabla _f\left( x,t\right) for each x\in X^\left( t\right) . While total differentiability of the value function may not be easy to ensure, Theorem 2 can be still applied along any smooth path connecting two parameter values t_ and t. Namely, suppose that functions f(x,\cdot ) are differentiable for all x\in X with , \nabla _f(x,t), \leq B for all x\in X, t\in T. A smooth path from t_ to t is described by a differentiable mapping \gamma :\left 0,1\right\rightarrow T with a bounded derivative, such that \gamma \left( 0\right) =t_ and \gamma \left( 1\right) =t. Theorem 2 implies that for any such smooth path, the change of the value function can be expressed as the path integral of the partial gradient \nabla _f(x^(t),t) of the objective function along the path: : V(t)-V(t_)=\int_\nabla _f(x^(s),s)\cdot ds. In particular, for t=t_, this establishes that cyclic path integrals along any smooth path \gamma must be zero: :\int \nabla _f(x^(s),s)\cdot ds=0. This "integrability condition" plays an important role in mechanism design with multidimensional types, constraining what kind of choice rules x^ can be sustained by mechanism-induced menus X\subseteq \bar. In application to producer theory, with x\in X\subseteq \mathbb^ being the firm's production vector and t\in \mathbb^ being the price vector, f\left( x,t\right) =t\cdot x, and the integrability condition says that any rationalizable supply function x^ must satisfy : \int x^(s)\cdot ds=0. When x^ is continuously differentiable, this integrability condition is equivalent to the symmetry of the
substitution matrix In bioinformatics and evolutionary biology, a substitution matrix describes the frequency at which a character in a nucleotide sequence or a protein sequence changes to other character states over evolutionary time. The information is often in ...
\left(\partial x_^\left( t\right) /\partial t_\right) _^. (In
consumer theory The theory of consumer choice is the branch of microeconomics that relates preferences to consumption expenditures and to consumer demand curves. It analyzes how consumers maximize the desirability of their consumption as measured by their pref ...
, the same argument applied to the expenditure minimization problem yields symmetry of the Slutsky matrix.)


Applications to parameterized constraints

Suppose now that the feasible set X\left( t\right) depends on the parameter, i.e., : V(t) =\sup_f(x,t) : X^(t) =\\text where X\left( t\right) =\left\ for some g:X\times \left 0,1\right\rightarrow \mathbb^. Suppose that X is a convex set, f and g are concave in x, and there exists \hat\in X such that g\left( \hat,t\right) >0 for all t\in \left 0,1\right. Under these assumptions, it is well known that the above constrained optimization program can be represented as a saddle-point problem for the Lagrangian L\left( x,\lambda,t\right) =f(x,t)+\lambda\cdot g\left( x,t\right) , where \lambda \in \mathbb_^ is the vector of
Lagrange multiplier In mathematical optimization, the method of Lagrange multipliers is a strategy for finding the local maxima and minima of a function subject to equality constraints (i.e., subject to the condition that one or more equations have to be satisfied e ...
s chosen by the adversary to minimize the Lagrangian. This allows the application of Milgrom and Segal's (2002, Theorem 4) envelope theorem for saddle-point problems, under the additional assumptions that X is a compact set in a normed linear space, f and g are continuous in x, and f_ and g_ are continuous in \left( x,t\right) . In particular, letting \left( x^(t),\lambda^\left( t\right) \right) denote the Lagrangian's saddle point for parameter value t, the theorem implies that V is absolutely continuous and satisfies : V(t)=V(0)+\int_^L_(x^(s),\lambda^\left( s\right) ,s)ds. For the special case in which f\left( x,t\right) is independent of t, K=1, and g\left( x,t\right) =h\left( x\right) +t, the formula implies that V^(t)=L_(x^(t),\lambda^\left( t\right) ,t)=\lambda^\left( t\right) for a.e. t. That is, the Lagrange multiplier \lambda^\left( t\right) on the constraint is its "
shadow price A shadow price is the monetary value assigned to an abstract or intangible commodity which is not traded in the marketplace. This often takes the form of an externality. Shadow prices are also known as the recalculation of known market prices in o ...
" in the optimization program.


Other applications

Milgrom and Segal (2002) demonstrate that the generalized version of the envelope theorems can also be applied to convex programming, continuous optimization problems, saddle-point problems, and optimal stopping problems.


See also

* Maximum theorem * Danskin's theorem * Hotelling's lemma *
Le Chatelier's principle Le Chatelier's principle (pronounced or ), also called Chatelier's principle (or the Equilibrium Law), is a principle of chemistry used to predict the effect of a change in conditions on chemical equilibria. The principle is named after French c ...
*
Roy's identity Roy's identity (named after French economist René Roy) is a major result in microeconomics having applications in consumer choice and the theory of the firm. The lemma relates the ordinary (Marshallian) demand function to the derivatives of the i ...
*
Value function The value function of an optimization problem gives the value attained by the objective function at a solution, while only depending on the parameters of the problem. In a controlled dynamical system, the value function represents the optimal payof ...


References

{{Reflist, 30em Production economics Calculus of variations Economics theorems Theorems in analysis