% Upper-case A B C D E F G H I J K L M N O P Q R S T U V W X Y Z % Lower-case a b c d e f g h i j k l m n o p q r s t u v w x y z % Digits 0 1 2 3 4 5 6 7 8 9 % Exclamation ! Double quote " Hash (number) # % Dollar $ Percent % Ampersand & % Acute accent ' Left paren ( Right paren ) % Asterisk * Plus + Comma , % Minus - Point . Solidus / % Colon : Semicolon ; Less than < % Equals = Greater than > Question mark ? % At @ Left bracket [ Backslash \ % Right bracket ] Circumflex ^ Underscore _ % Grave accent ` Left brace { Vertical bar | % Right brace } Tilde ~ % ---------------------------------------------------------------------| % --------------------------- 72 characters ---------------------------| % ---------------------------------------------------------------------| % % Optimal Foraging Theory Revisited: Appendix. Lagrange Multiplier % Method % (note: this material not included in final version of document) % % (c) Copyright 2007 by Theodore P. Pavlic % \chapter{Lagrange Multiplier Method} \label{app:lagrange} In this \appname{}, we discuss a well-known method for optimizing a real functional of multiple real variables subject to equality and inequality constraints. This is a summary of the discussion by \citet{Bertsekas95} of \emph{Lagrange multiplier theory}. First, take $n \in \N$, set $\set{D} \subseteq \R^n$, and function % \begin{equation*} f: \set{D} \mapsto \R \end{equation*} % This will be the function we wish to \emph{minimize}. The solution to this problem will be a vector from $\set{D}$ that we will call the \emph{local minimum of $f$ subject to constraints}. Now, take $m \in \W$ and a family $(h_i)_{i=1}^m$ where % \begin{equation*} h_i: \set{D} \mapsto \R \end{equation*} % is a function called an \emph{equality constraint} for all $i \in \{1,2,\dots,m\}$. The local minimum of $f$ will be a vector $\v{x} \in \set{D}$ such that $h_i(\v{x})=0$ for all $i \in \{1,2,\dots,m\}$. Finally, take $r \in \W$ and a family $(g_j)_{j=1}^r$ where % \begin{equation*} g_j: \set{D} \mapsto \R \end{equation*} % is a function called an \emph{inequality constraint} for all $j \in \{1,2,\dots,r\}$. The local minimum of $f$ will be a vector $\v{x} \in \set{D}$ such that $g_i(\v{x}) \leq 0$ for all $i \in \{1,2,\dots,r\}$. In summary, the problem is to % \begin{align*} \text{minimize} &\enskip f(\v{x})\\ \text{subject to} &\enskip h_i(\v{x}) = 0 \text{ for all } i \in \{1,2,\dots,m\}\\ &\enskip g_j(\v{x}) \leq 0 \text{ for all } j \in \{1,2,\dots,r\} \end{align*} % To change this to a maximization problem, define a function $f^\star: \set{D} \mapsto \R$ with $f^\star(\v{x})=-f(\v{x})$ for all $\v{x} \in \set{D}$ and minimize $f^\star$ instead of $f$; the \emph{local maximum} for $f$ will be a local minimum of $f^\star$. Thus, we will only handle the minimization problem here. For the rest of this \appname{}, assume the above problem. Note that it is assumed that $f$, $g$, and $h$ are all continuously differentiable functions. At some points, it may also be assumed that they are twice continuously differentiable functions. Also note that while we assume this type of differentiability holds for all points in $\R^n$, it is only necessary that it holds in some open set around the minimum of $f$; we make the stronger assumption for notational simplicity. \section{Terminology} A number of terms and constructs need to be defined before giving a solution to the minimization problem. % \begin{description} \item\emph{Feasible Set:} For this problem, the \emph{feasible set} of points is denoted $\set{F}$ and defined by % \begin{equation*} \set{F} \triangleq \{ \v{x} \in \set{D} : h_i(\v{x}) = 0, g_j(\v{x})\leq 0, i \in \{1,2,\dots,m\}, j \in \{1,2,\dots,r\} \} \end{equation*} % That is, the set $\set{F}$ represents all possible candidates for a minimum of $f$. Any point $\v{x} \in \set{F}$ is called a \emph{feasible point} or a \emph{feasible vector}. For all intents and purposes, the function $f$ need only be defined on the set $\set{F}$. That is, $\set{F}$ can be viewed as the domain of interest for function $f$. Thus, minimizing $f$ subject to $g$ and $h$ is equivalent to minimizing the restriction of $f$ to $\set{F}$ (\ie, $f|_\set{F}$). Therefore, consider the following statements. % \begin{itemize} \item If $\set{F}$ is a compact set, by the extreme value theorem, $f|_\set{F}$ achieves its minimum bound. \item If $\set{F}$ is a convex set and $f$ is a convex function \emph{over $\set{F}$} then $f|_\set{F}$ is a convex function. \item Assume that all of the following are the case. % \begin{enumerate}[(i)] \item $\set{D}$ is a convex set. \item $f$ and $g$ are convex functions over $\set{D}$. \item $h$ is \emph{affine} (\ie, there exists some $\v{a} \in \R^n$ and $b \in \R$ such that $h(\v{x}) = \v{a}^\T \v{x} + b$ for all $\v{x} \in \set{D}$). \end{enumerate} % In this case, $\set{F}$ is a convex set. \end{itemize} % \item\emph{Local Minimum:} To say that $\v{x}^* \in \set{F}$ is a \emph{local minimum} of $f$ subject to $g$ and $h$ (\ie, $f$ over $\set{F}$) means that $\varepsilon \in \R_{>0}$, % \begin{equation} f(\v{x}^*) \leq f(\v{y}) \text{ for all } \v{y} \in \set{F} \setdiff \{\v{x}^*\} \text{ with } \|\v{x}^*-\v{y}\|_2 < \varepsilon \label{eq:lagrange_minimum} \end{equation} % A point $\v{x}^* \in \set{F}$ is a \emph{strict local minimum} if \longref{eq:lagrange_minimum} is true when the $\leq$ relation replaced by $<$. Note that we will often omit reference to $g$, $h$, and $\set{F}$ when discussing minima of $f$; however, it is implied that $f$ is not being minimized over $\set{F}$ rather than its entire domain. % \item\emph{Global Minimum:} To say that $\v{x}^* \in \set{F}$ is a \emph{global minimum} of $f$ (subject to $g$ and $h$) means that % \begin{equation} f(\v{x}^*) \leq f(\v{y}) \text{ for all } \v{y} \in \set{F} \setdiff \{\v{x}^*\} \label{eq:lagrange_global_minimum} \end{equation} % Clearly, all global minima are local minima. A point $\v{x}^* \in \set{F}$ is \emph{the} \emph{strict global minimum} if \longref{eq:lagrange_global_minimum} is true when the $\leq$ relation replaced by $<$. Note the following statements. % \begin{itemize} \item By the extreme value theorem, if $\set{F}$ is a compact set then a feasible point exists that is a global minimum of $f$. \item By the discussion in \longref{app:math_euclidean_convexity}, if $\set{F}$ is a convex set and $f$ is a convex function over $\set{F}$ then every local minimum of $f$ is a global minimum of $f$. \item By the discussion in \longref{app:math_euclidean_convexity}, if $\set{F}$ is a convex set and $f$ is a \emph{strict} convex function over $\set{F}$ then there exists at most one global minimum of $f$. \item If $\set{F}$ is a convex compact set and $f$ is a strictly convex function over $\set{F}$ then a feasible point exists that is the single global minimum of $f$. \end{itemize} % \item\emph{Active Inequality Constraints:} For any point $\v{x} \in \set{F}$, define the \emph{active inequality constraints} $\set{A}(\v{x})$ by % \begin{equation*} \set{A}(\v{x}) \triangleq \{ j \in \{1,2,\dots,r\} : g_j(\v{x}) = 0 \} \end{equation*} % In other words, for any $\v{x} \in \set{F}$, $\set{A}(\v{x})$ represents a set of indices for inequality constraints in which the point $\v{x}$ is nearly infeasible. Note that if $\set{A}(\v{x}) = \emptyset$ for all $\v{x} \in \R^n$ then this minimization problem should not be impacted by the inequality constraints. Thus, this set really represents the inequality constraints that alter the solution from the one that would exist without inequality constraints. \item\emph{Equality Constraint Gradients:} For each $\v{x} \in \set{F}$ and $i \in \{1,2,\dots,m\}$, the gradient $\nabla_\v{x} h_i(\v{x})$ is called an \emph{equality constraint gradient} at $\v{x}$. \item\emph{Inequality Constraint Gradients:} For each $\v{x} \in \set{F}$ and $j \in \set{A}(\v{x})$, the gradient $\nabla_\v{x} g_j(\v{x})$ is called an \emph{(active) inequality constraint gradient} at $\v{x}$. Notice the use of the set of active inequality constraints (\ie, $\set{A}(\v{x})$ for each $\v{x} \in \set{F}$). \item\emph{Linear Independence:} Take $q \in \N$ and vector $\v{a} \in \R^q$ and family $(\v{x}^i)_{i=1}^q$ where $\v{x}^i \in \R^n$ for all $i \in \{1,2,\dots,q\}$. Recall that $\v{a}=0$ is equivalent to the statement that $a_1=a_2=\cdots=a_q=0$. The vectors in $(\v{x}^i)_{i=1}^q$ are said to be \emph{linearly independent} if $\v{a}=0$ is the \emph{only} vector for which % \begin{equation*} a_1 \v{x}^1 + a_2 \v{x}^2 + a_3 \v{x}^3 + \cdots + a_n \v{x}^n = 0 \end{equation*} % \item\emph{Regular Vectors:} A point $\v{x} \in \set{F}$ is called a \emph{regular point} or a \emph{regular vector} if the single family consisting of all of the equality constraint gradients and all \emph{active} inequality constraint gradients are linearly independent at $\v{x}$. In other words, all active constraints (both inequality and equality) need to be linearly independent at the point. \item\emph{(First-Order) Feasible Variations:} For any point $\v{x}^* \in \set{D}$, the \emph{(first-order) feasible variations} at $\v{x}^*$ are denoted $\set{V}(\v{x}^*)$ and defined by % \begin{align*} \set{V}(\v{x}^*) \triangleq \{ \v{\Delta} \in \R^n : (\nabla_\v{x} h_i(\v{x}^*))^\T \v{\Delta} = 0, i \in \{1,2,\dots,m\} \}\\ \mathrel{\phantom{\triangleq}} {}\cap \{ \v{\Delta} \in \R^n : (\nabla_\v{x} g_j(\v{x}^*))^\T \v{\Delta} = 0, j \in \set{A}(\v{x}^*) \} \end{align*} % For any point $\v{x}^* \in \set{D}$, a direction $\v{\Delta} \in \set{V}(\v{x}^*)$ is roughly such that there exists some small $\delta \in \R_{>0}$ with $\v{x}^* + \delta \v{\Delta} \in \set{F}$. Thus, the first-order feasible variations are roughly the directions in Euclidean space along which feasible vectors can move and still stay feasible. \item\emph{The Lagrangian:} The \emph{Lagrangian function} $L: \set{D} \times \R^m \times \R^r \mapsto \R$ is defined by % \begin{equation*} L(\v{x},\v{\lambda},\v{\mu}) \triangleq f(\v{x}) + \sum\limits_{i=1}^m \lambda_i h_i(\v{x}) + \sum\limits_{j=1}^r \mu_j g_j(\v{x}) \end{equation*} % This function \emph{adjoins} the constraints to the function to minimize so that minimization of $L$ results in the minimization of $f$ with respect to its constraints. The coefficients represented by the elements of $\v{\lambda}$ and $\v{\mu}$ are known as \emph{Lagrange multipliers}, and so $\v{\lambda}$ and $\v{\mu}$ are known as \emph{Lagrange multiplier vectors}. \end{description} \section{Necessary Conditions} \label{app:math_lagrange_necessary} The following are typically known as \emph{\acro{KKT}{Karush-Kuhn-Tucker} conditions} or \emph{Kuhn-Tucker conditions}. They represent the conditions that are necessary for a (regular) point to be a minimum. \paragraph{First-Order Necessary Conditions:} Assume that $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to the constraints by $f$ and $g$ and assume that $\v{x}^*$ is regular. As \citet{Bertsekas95} notes, this regularity condition is not necessary in a class of problems with \emph{linear constraints}; however, we leave out that interesting case for brevity. Given these assumptions, there must exist unique Lagrange multiplier vectors $\v{\lambda}^* \in \R^m$ $\v{\mu}^* \in \R^r$ defined by % \begin{equation*} \v{\lambda}^* \triangleq \begin{bmatrix} \lambda^*_1 \\ \lambda^*_2 \\ \vdots \\ \lambda^*_m \end{bmatrix} \quad \text{ and } \quad \v{\mu}^* \triangleq \begin{bmatrix} \mu^*_1 \\ \mu^*_2 \\ \vdots \\ \mu^*_r \end{bmatrix} \end{equation*} % such that % \begin{equation} \nabla_\v{x} L( \v{x}^*, \v{\lambda}^*, \v{\mu}^* ) = 0 \label{eq:lagrange_if} \end{equation} % where % \begin{equation*} \mu^*_j \geq 0 \text{ for all } j \in \set{A}(\v{x}^*) \end{equation*} % and % \begin{equation*} \mu^*_j = 0 \text{ for all } j \notin \set{A}(\v{x}^*) \end{equation*} % This is known as the \emph{first-order necessary condition} for any point to be a local minimum of $f$ subject to the constraints of $f$ and $g$. That is, if $\v{x}^*$ is a point such that \longref{eq:lagrange_if} is not satisfied then $\v{x}^*$ cannot be a local minimum of $f$ subject to $g$ and $h$. Note that not all points that satisfy \longref{eq:lagrange_if} will be local minima of $f$ subject to $g$ and $h$; however, if it is also the case that $\set{F}$ is a convex set and $f$ is a convex function over $\set{F}$ then \longref{eq:lagrange_if} is also a sufficient condition for a minimum. That is, for the convex case, any point $\v{x}^*$ that satisfies \longref{eq:lagrange_if} will be a local minimum of $f$ subject to $g$ and $h$. We discuss this convex case further in \longref{app:math_lagrange_convex}. \paragraph{Second-Order Necessary Conditions:} Assume that the point $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to the constraints by $f$ and $g$ and assume that $\v{x}^*$ is regular. Take $\v{\lambda}^* \in \R^m$ and $\v{\mu}^* \in \R^r$ to be the Lagrange multiplier vectors where the first-order necessary conditions hold for $\v{x}^*$. If, in addition, $f$, $g$, and $h$ are twice continuously differentiable then % \begin{equation*} \v{\Delta}^\T \nabla^2_{\v{x}\v{x}} L(\v{x}^*,\v{\lambda}^*,\v{\mu}^*) \v{\Delta} \geq 0 \text{ for all } \v{\Delta} \in \set{V}( \v{x}^* ) \end{equation*} % These are known as a \emph{second-order necessary condition} for any point to be a local minimum of $f$ subject to the constraints of $f$ and $g$. That is, any point $\v{x}^*$ that does not satisfy this condition cannot be a local minimum of $f$ subject to $g$ and $h$. \paragraph{Local Minimum Candidates:} Take set $\set{C} \subseteq \set{F}$ defined by % \begin{equation*} \set{C} \triangleq \{ \v{x}^* \in \set{F} : \v{x}^* \text{ satisfies \longref{eq:lagrange_if} for some } \v{\lambda}^* \in \R^m,\v{\mu}^* \in \R^r \} \end{equation*} % The set $\set{C}$ is a set of candidates for a local minimum of $f$ subject to $g$ and $h$. That is, it is impossible for any point that is not an element of $\set{C}$ to be a local minimum of $f$ subject to $g$ and $h$. Of course, when $f$, $g$, and $h$ are twice continuously differentiable, the size of this candidate set can be further reduced by the second-order necessary condition. If a local minimum exists, it must be an element of $\set{C}$. Now assume that a global minimum is \emph{known} to exist (\eg, $\set{F}$ is a compact set). In this case, it must be an element of $\set{C}$. That is, if a global minimum is known to exist, the point $\v{x}^*$ such that % \begin{equation*} f(\v{x}^*) = \min\{ f(\v{x}) : \v{x} \in \set{C} \} \end{equation*} % will be a global minimum of $f$ subject to $g$ and $h$. Therefore, if $\set{C}$ is a finite set and a global minimum is known to exist, a simple strategy for finding a global minimum is to evaluate $f$ at every point of $\set{C}$ and take the point that results in the least $f$. \section{Sufficiency Conditions} \longrefs{app:math_lagrange_sufficient} and \shortref{app:math_lagrange_sufficient_alt} give different sets of conditions that are sufficient to call a point a strict local minimum. These conditions are similar to the \ac{KKT} conditions for a minimum given in \longref{app:math_lagrange_necessary}; however, there are some important differences. In \longref{app:math_lagrange_convex}, we discuss how convexity allows further conclusions to be made about points which satisfy these sufficiency conditions. \subsection{Standard Sufficiency Conditions} \label{app:math_lagrange_sufficient} Assume that $f$, $g$, and $h$ are all twice continuously differentiable. Take $\v{x}^* \in \R^n$, $\v{\lambda}^* \in \R^m$, and $\v{\mu}^* \in \R^r$ such that % \begin{enumerate}[(i)] \item $\nabla_\v{x} L( \v{x}^*, \v{\lambda}^*, \v{\mu}^* ) = 0$ \item $h_i(\v{x}^*) = 0$ for all $i \in \{1,2,\dots,m\}$ \item $g_j(\v{x}^*) \leq 0$ for all $j \in \{1,2,\dots,r\}$ \item $\mu_j^* > 0$ for all $j \in \set{A}(\v{x}^*)$ \label{item:lagrange_sufficient_strict_mu} \item $\mu_j^* = 0$ for all $j \notin \set{A}(\v{x}^*)$ \item $\v{\Delta}^\T \nabla^2_{\v{x}\v{x}} L(\v{x}^*,\v{\lambda}^*,\v{\mu}^*) \v{\Delta} > 0$ for all $\v{\Delta} \in \set{V}(\v{x}^*) \setdiff \{0\}$ \label{item:lagrange_sufficient_hessian} \end{enumerate} % In this case, $\v{x}^*$ must be a strict local minimum of $f$ with respect to constraints from $g$ and $h$. Notice that properties (\shortref{item:lagrange_sufficient_strict_mu}) and (\shortref{item:lagrange_sufficient_hessian}) use a \emph{strict} inequality. Also notice that property (\shortref{item:lagrange_sufficient_hessian}) excludes $0$ from the set of feasible variations. These are the classical \emph{second-order sufficiency conditions} for a minimum of $f$ subject to the constraints of $f$ and $g$. \subsection{Alternate Sufficiency Conditions} \label{app:math_lagrange_sufficient_alt} Assume that $f$, $g$, and $h$ are all twice continuously differentiable. Take $\v{x}^* \in \R^n$, $\v{\lambda}^* \in \R^m$, and $\v{\mu}^* \in \R^r$ such that % \begin{enumerate}[(i)] \item $\nabla_\v{x} L( \v{x}^*, \v{\lambda}^*, \v{\mu}^* ) = 0$ \item $h_i(\v{x}^*) = 0$ for all $i \in \{1,2,\dots,m\}$ \item $g_j(\v{x}^*) \leq 0$ for all $j \in \{1,2,\dots,r\}$ \item $\mu_j^* \geq 0$ for all $j \in \set{A}(\v{x}^*)$ \label{item:lagrange_sufficient_alt_mu} \item $\mu_j^* = 0$ for all $j \notin \set{A}(\v{x}^*)$ \item $\v{\Delta}^\T \nabla^2_{\v{x}\v{x}} L(\v{x}^*,\v{\lambda}^*,\v{\mu}^*) \v{\Delta} > 0$ for all $\v{\Delta} \in \set{V}_2(\v{x}^*)$ where % \begin{align*} \set{V}_2(\v{x}^*) &\triangleq \{ \v{\Delta} \in \R^n : (\nabla_\v{x} h_i(\v{x}^*))^\T \v{\Delta} = 0, i \in \{1,2,\dots,m\} \}\\ &\mathrel{\phantom{\triangleq}} {}\cap \{ \v{\Delta} \in \R^n : (\nabla_\v{x} g_j(\v{x}^*))^\T \v{\Delta} = 0, \mu_j^* > 0, j \in \set{A}(\v{x}^*) \}\\ &\mathrel{\phantom{\triangleq}} {}\cap \{ \v{\Delta} \in \R^n : (\nabla_\v{x} g_j(\v{x}^*))^\T \v{\Delta} \leq 0, \mu_j^* = 0, j \in \set{A}(\v{x}^*) \}\\ &\mathrel{\phantom{\triangleq}} {}\cap \{ \v{\Delta} \in \R^n : \v{\Delta} \neq 0 \} \end{align*} \label{item:lagrange_sufficient_alt_hessian} \end{enumerate} % In this case, $\v{x}^*$ must be a strict local minimum of $f$ with respect to constraints from $g$ and $h$. These are alternate second-order sufficiency conditions for a minimum of $f$ subject to constraints from $g$ and $h$. Notice that property (\shortref{item:lagrange_sufficient_alt_mu}) does \emph{not} require a strict inequality; because of this, property (\shortref{item:lagrange_sufficient_alt_hessian}) is a stronger condition than the analogous condition from \longref{app:math_lagrange_sufficient}. \section{Lagrange Multipliers and Convexity} \label{app:math_lagrange_convex} Recall the discussion of convexity in Euclidean spaces from \longref{app:math_euclidean_convexity}. Under convexity conditions, finding local and global minima of $f$ subject to $g$ and $h$ is greatly simplified. \subsection{Convex Sets and Convex Functions} Recall the definition of a convex set and a convex function introduced in \longref{app:math_euclidean_convexity}. % \begin{description} \item\emph{Convex Set:} Assume that for all $\v{x},\v{y} \in \set{F}$ with $\v{x} \neq \v{y}$, % \begin{equation*} t \v{x} + (1-t) \v{y} \in \set{F} \end{equation*} % for all $t \in (0,1)$. Thus, $\set{F}$ is a \emph{convex set}. It can be shown that $\set{F}$ can be represented as the Cartesian product of $n$ intervals of $\R$. \item\emph{Convex Function over a Convex Set:} Assume that $\set{F}$ is a convex set. To say that function $f$ is \emph{convex over $\set{F}$} means that for all $\v{x},\v{y} \in \set{F}$ with $\v{x} \neq \v{y}$, % \begin{equation*} f( t \v{x} + (1-t) \v{y} ) \leq t f(\v{x}) + (1-t) \v{y} \end{equation*} % for all $t \in (0,1)$. \item\emph{Strictly Convex Function over a Convex Set:} Assume that $\set{F}$ is a convex set. To say that function $f$ is \emph{strictly convex over $\set{F}$} means that for all $\v{x},\v{y} \in \set{F}$ with $\v{x} \neq \v{y}$, % \begin{equation*} f( t \v{x} + (1-t) \v{y} ) < t f(\v{x}) + (1-t) \v{y} \end{equation*} % for all $t \in (0,1)$. \end{description} % Now assume that $f$ is twice continuously differentiable and $\set{F}$ is a convex set. If it is the case that % \begin{equation*} \v{\Delta}^\T \nabla^2_{\v{x}\v{x}} f(\v{x}) \v{\Delta} \geq 0 \text{ for all } \v{\Delta} \in \R^n \text{ and } \v{x} \in \set{F} \end{equation*} % then $f$ is convex over convex set $\set{F}$. Additionally, if it is the case that % \begin{equation*} \v{\Delta}^\T \nabla^2_{\v{x}\v{x}} f(\v{x}) \v{\Delta} > 0 \text{ for all } \v{\Delta} \in \R^n \setdiff \{0\} \text{ and } \v{x} \in \set{F} \end{equation*} % then $f$ is strictly convex over convex set $\set{F}$. These are sufficiency conditions for convex functions over convex sets. \subsection{Convex Feasible Set} Assume that $\set{F}$ is a convex set (\eg, the sufficient conditions for convexity given in the definition of $\set{F}$ above). Assume that $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to $g$ and $h$. It must be the case that % \begin{equation} ( \nabla_{\v{x}} f(\v{x}^*) )^\T (\v{x} - \v{x}^*) \geq 0 \label{eq:lagrange_convex_if} \end{equation} % for all $\v{x} \in \set{F}$. That is, this is a necessary condition for any minimum of $f$ subject to $g$ and $h$. If there is a point $\v{x}^* \in \set{F}$ such that \longref{eq:lagrange_convex_if} does not hold for all $\v{x} \in \set{F}$ then $\v{x}^*$ cannot be a minimum of $f$ subject to $g$ and $h$. In fact, for the case where $\set{F}$ is convex, \longref{eq:lagrange_convex_if} is equivalent to the condition in \longref{eq:lagrange_if}. \subsection{Convex Function over Convex Set} Assume that $\set{F}$ is a convex set. Also assume that $f$ is convex over $\set{F}$. Consider the following consequences of the convexity of $f$ over convex set $\set{F}$. % \begin{description} \item\emph{Necessary Conditions Become Sufficient:} In this case, the necessary conditions in \longrefs{eq:lagrange_if} and \shortref{eq:lagrange_convex_if} also become sufficient conditions. In other words, % \begin{itemize} \item A point $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to $g$ and $h$ if and only if \longref{eq:lagrange_if} holds. \item A point $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to $g$ and $h$ if and only if \longref{eq:lagrange_convex_if} holds. \end{itemize} % That is, for this case of a convex function over a convex feasible set, the following are equivalent % \begin{enumerate}[(i)] \item A point $\v{x}^* \in \set{F}$ is a local minimum of $f$ subject to $g$ and $h$. \item A point $\v{x}^* \in \set{F}$ satisfies \longref{eq:lagrange_if}. \item A point $\v{x}^* \in \set{F}$ satisfies \longref{eq:lagrange_convex_if}. \end{enumerate} % \item\emph{Local Minima Become Global Minima:} Now assume that $\v{x}^*$ is known to be a local minimum of $f$ subject to $g$ and $h$ (\eg, any of the sufficiency conditions given above are met). Because $f$ is convex over the convex set $\set{F}$, the following are true. % \begin{itemize} \item The point $\v{x}^*$ is a global minimum of $f$ subject to $g$ and $h$. \item If $\v{x}^*$ is a strict local minimum of $f$ subject to $g$ and $h$ then $\v{x}^*$ is a strict global minimum of $f$ subject to $g$ and $h$. In fact, it can be shown that $\v{x}^*$ is the unique (strict) global minimum of $f$ subject to $g$ and $h$. \item If $f$ is strictly convex over convex set $\set{F}$ then $\v{x}^*$ must be the unique (strict) global minimum of $f$ subject to $g$ and $h$. \end{itemize} % \end{description}