Misplaced Pages

Hamilton–Jacobi–Bellman equation

Article snapshot taken from Wikipedia with creative commons attribution-sharealike license. Give it a read and then ask your questions in the chat. We can research this topic together.

The Hamilton-Jacobi-Bellman ( HJB ) equation is a nonlinear partial differential equation that provides necessary and sufficient conditions for optimality of a control with respect to a loss function . Its solution is the value function of the optimal control problem which, once known, can be used to obtain the optimal control by taking the maximizer (or minimizer) of the Hamiltonian involved in the HJB equation.

#984015

71-518: The equation is a result of the theory of dynamic programming which was pioneered in the 1950s by Richard Bellman and coworkers. The connection to the Hamilton–Jacobi equation from classical physics was first drawn by Rudolf Kálmán . In discrete-time problems, the analogous difference equation is usually referred to as the Bellman equation . While classical variational problems , such as

142-498: A − c T − j ≥ 0 {\displaystyle Ak^{a}-c_{T-j}\geq 0} . Working backwards, it can be shown that the value function at time t = T − j {\displaystyle t=T-j} is where each v T − j {\displaystyle v_{T-j}} is a constant, and the optimal amount to consume at time t = T − j {\displaystyle t=T-j}

213-399: A − c t {\displaystyle k_{t+1}=Ak_{t}^{a}-c_{t}} , where A is a positive constant and 0 < a < 1 {\displaystyle 0<a<1} . Assume capital cannot be negative. Then the consumer's decision problem can be written as follows: Written this way, the problem looks complicated, because it involves solving for all

284-1579: A cost-to-go function J ∗ {\displaystyle J^{\ast }} . The latter obeys the fundamental equation of dynamic programming: a partial differential equation known as the Hamilton–Jacobi–Bellman equation , in which J x ∗ = ∂ J ∗ ∂ x = [ ∂ J ∗ ∂ x 1         ∂ J ∗ ∂ x 2         …         ∂ J ∗ ∂ x n ] T {\displaystyle J_{x}^{\ast }={\frac {\partial J^{\ast }}{\partial \mathbf {x} }}=\left[{\frac {\partial J^{\ast }}{\partial x_{1}}}~~~~{\frac {\partial J^{\ast }}{\partial x_{2}}}~~~~\dots ~~~~{\frac {\partial J^{\ast }}{\partial x_{n}}}\right]^{\mathsf {T}}} and J t ∗ = ∂ J ∗ ∂ t {\displaystyle J_{t}^{\ast }={\frac {\partial J^{\ast }}{\partial t}}} . One finds that minimizing u {\displaystyle \mathbf {u} } in terms of t {\displaystyle t} , x {\displaystyle \mathbf {x} } , and

355-414: A recursive relationship called the Bellman equation . For i  = 2, ...,  n , V i −1 at any state y is calculated from V i by maximizing a simple function (usually the sum) of the gain from a decision at time i  − 1 and the function V i at the new state of the system if this decision is made. Since V i has already been calculated for

426-411: A complicated problem by breaking it down into simpler sub-problems in a recursive manner. While some decision problems cannot be taken apart this way, decisions that span several points in time do often break apart recursively. Likewise, in computer science, if a problem can be solved optimally by breaking it into sub-problems and then recursively finding the optimal solutions to the sub-problems, then it

497-551: A continuous time interval t 0 ≤ t ≤ t 1 {\displaystyle t_{0}\leq t\leq t_{1}} that minimizes a cost function The solution to this problem is an optimal control law or policy u ∗ = h ( x ( t ) , t ) {\displaystyle \mathbf {u} ^{\ast }=h(\mathbf {x} (t),t)} , which produces an optimal trajectory x ∗ {\displaystyle \mathbf {x} ^{\ast }} and

568-538: A control problem by applying Bellman's principle of optimality and then working out backwards in time an optimizing strategy can be generalized to stochastic control problems. Consider similar as above now with ( X t ) t ∈ [ 0 , T ] {\displaystyle (X_{t})_{t\in [0,T]}\,\!} the stochastic process to optimize and ( u t ) t ∈ [ 0 , T ] {\displaystyle (u_{t})_{t\in [0,T]}\,\!}

639-530: A decision by breaking it down into a sequence of decision steps over time. This is done by defining a sequence of value functions V 1 , V 2 , ..., V n taking y as an argument representing the state of the system at times i from 1 to n . The definition of V n ( y ) is the value obtained in state y at the last time n . The values V i at earlier times i  =  n  −1,  n  − 2, ..., 2, 1 can be found by working backwards, using

710-671: A given n {\displaystyle n} . For example, when n = 4 , five possible solutions are There are at least three possible approaches: brute force , backtracking , and dynamic programming. Brute force consists of checking all assignments of zeros and ones and counting those that have balanced rows and columns ( n / 2 zeros and n / 2 ones). As there are 2 n 2 {\displaystyle 2^{n^{2}}} possible assignments and ( n n / 2 ) n {\displaystyle {\tbinom {n}{n/2}}^{n}} sensible assignments, this strategy

781-556: A graph G=(V,E) , the shortest path p from a vertex u to a vertex v exhibits optimal substructure: take any intermediate vertex w on this shortest path p . If p is truly the shortest path, then it can be split into sub-paths p 1 from u to w and p 2 from w to v such that these, in turn, are indeed the shortest paths between the corresponding vertices (by the simple cut-and-paste argument described in Introduction to Algorithms ). Hence, one can easily formulate

SECTION 10

#1732801570985

852-550: A quadratic form for the value function, we obtain the usual Riccati equation for the Hessian of the value function as is usual for Linear-quadratic-Gaussian control . Dynamic programming Dynamic programming is both a mathematical optimization method and an algorithmic paradigm . The method was developed by Richard Bellman in the 1950s and has found applications in numerous fields, from aerospace engineering to economics . In both contexts it refers to simplifying

923-400: A sequence of smaller decisions. To do so, we define a sequence of value functions V t ( k ) {\displaystyle V_{t}(k)} , for t = 0 , 1 , 2 , … , T , T + 1 {\displaystyle t=0,1,2,\ldots ,T,T+1} which represent the value of having any amount of capital k at each time t . There

994-523: A solution V {\displaystyle V\,\!} of the latter does not necessarily solve the primal problem, it is a candidate only and a further verifying argument is required. This technique is widely used in Financial Mathematics to determine optimal investment strategies in the market (see for example Merton's portfolio problem ). As an example, we can look at a system with linear stochastic dynamics and quadratic cost. If

1065-618: A typical optimal control problem is to subject to with initial state variable x ( t 0 ) = x 0 {\displaystyle x(t_{0})=x_{0}} . The objective function J ( t 0 , x 0 ; u ) {\displaystyle J(t_{0},x_{0};u)} is to be maximized over all admissible controls u ∈ U [ t 0 , t 1 ] {\displaystyle u\in U[t_{0},t_{1}]} , where u {\displaystyle u}

1136-460: Is subject to the terminal condition As before, the unknown scalar function V ( x , t ) {\displaystyle V(x,t)} in the above partial differential equation is the Bellman value function , which represents the cost incurred from starting in state x {\displaystyle x} at time t {\displaystyle t} and controlling

1207-459: Is which can be simplified to We see that it is optimal to consume a larger fraction of current wealth as one gets older, finally consuming all remaining wealth in period T , the last period of life. There are two key attributes that a problem must have in order for dynamic programming to be applicable: optimal substructure and overlapping sub-problems . If a problem can be solved by combining optimal solutions to non-overlapping sub-problems,

1278-447: Is (by assumption) no utility from having capital after death, V T + 1 ( k ) = 0 {\displaystyle V_{T+1}(k)=0} . The value of any quantity of capital at any previous time can be calculated by backward induction using the Bellman equation . In this problem, for each t = 0 , 1 , 2 , … , T {\displaystyle t=0,1,2,\ldots ,T} ,

1349-1132: Is a Lebesgue measurable function from [ t 0 , t 1 ] {\displaystyle [t_{0},t_{1}]} to some prescribed arbitrary set in R m {\displaystyle \mathbb {R} ^{m}} . The value function is then defined as V ( t , x ( t ) ) = max u ∈ U ∫ t t 1 I ( τ , x ( τ ) , u ( τ ) ) d τ + ϕ ( x ( t 1 ) ) {\displaystyle V(t,x(t))=\max _{u\in U}\int _{t}^{t_{1}}I(\tau ,x(\tau ),u(\tau ))\,\mathrm {d} \tau +\phi (x(t_{1}))} with V ( t 1 , x ( t 1 ) ) = ϕ ( x ( t 1 ) ) {\displaystyle V(t_{1},x(t_{1}))=\phi (x(t_{1}))} , where ϕ ( x ( t 1 ) ) {\displaystyle \phi (x(t_{1}))}

1420-604: Is already known, so using the Bellman equation once we can calculate V T ( k ) {\displaystyle V_{T}(k)} , and so on until we get to V 0 ( k ) {\displaystyle V_{0}(k)} , which is the value of the initial decision problem for the whole lifetime. In other words, once we know V T − j + 1 ( k ) {\displaystyle V_{T-j+1}(k)} , we can calculate V T − j ( k ) {\displaystyle V_{T-j}(k)} , which

1491-894: Is applied maps vectors of n pairs of integers to the number of admissible boards (solutions). There is one pair for each column, and its two components indicate respectively the number of zeros and ones that have yet to be placed in that column. We seek the value of f ( ( n / 2 , n / 2 ) , ( n / 2 , n / 2 ) , … ( n / 2 , n / 2 ) ) {\displaystyle f((n/2,n/2),(n/2,n/2),\ldots (n/2,n/2))} ( n {\displaystyle n} arguments or one vector of n {\displaystyle n} elements). The process of subproblem creation involves iterating over every one of ( n n / 2 ) {\displaystyle {\tbinom {n}{n/2}}} possible assignments for

SECTION 20

#1732801570985

1562-403: Is called memoization ; this is the top-down approach, since we first break the problem into subproblems and then calculate and store values. In the bottom-up approach, we calculate the smaller values of fib first, then build larger values from them. This method also uses O( n ) time since it contains a loop that repeats n − 1 times, but it only takes constant (O(1)) space, in contrast to

1633-586: Is capital, and f {\displaystyle f} is a production function satisfying the Inada conditions . An initial capital stock k 0 > 0 {\displaystyle k_{0}>0} is assumed. Let c t {\displaystyle c_{t}} be consumption in period t , and assume consumption yields utility u ( c t ) = ln ⁡ ( c t ) {\displaystyle u(c_{t})=\ln(c_{t})} as long as

1704-424: Is either zero or one, depending on whether the vector is a permutation of n / 2 ( 0 , 1 ) {\displaystyle (0,1)} and n / 2 ( 1 , 0 ) {\displaystyle (1,0)} pairs or not. Value function The value function of an optimization problem gives the value attained by the objective function at a solution, while only depending on

1775-409: Is given, and he only needs to choose current consumption c t {\displaystyle c_{t}} and saving k t + 1 {\displaystyle k_{t+1}} . To actually solve this problem, we work backwards. For simplicity, the current level of capital is denoted as k . V T + 1 ( k ) {\displaystyle V_{T+1}(k)}

1846-495: Is known as the Bellman equation , which can be solved for an exact solution of the discrete approximation of the optimization equation. In economics, the objective is generally to maximize (rather than minimize) some dynamic social welfare function . In Ramsey's problem, this function relates amounts of consumption to levels of utility . Loosely speaking, the planner faces the trade-off between contemporaneous consumption and future consumption (via investment in capital stock that

1917-412: Is not guaranteed in most situations. Instead, the notion of a viscosity solution is required, in which conventional derivatives are replaced by (set-valued) subderivatives . Consider the following problem in deterministic optimal control over the time period [ 0 , T ] {\displaystyle [0,T]} : where C [ ⋅ ] {\displaystyle C[\cdot ]}

1988-523: Is not practical except maybe up to n = 6 {\displaystyle n=6} . Backtracking for this problem consists of choosing some order of the matrix elements and recursively placing ones or zeros, while checking that in every row and column the number of elements that have not been assigned plus the number of ones or zeros are both at least n / 2 . While more sophisticated than brute force, this approach will visit every solution once, making it impractical for n larger than six, since

2059-498: Is referred to as call-by-need ). Some languages make it possible portably (e.g. Scheme , Common Lisp , Perl or D ). Some languages have automatic memoization built in, such as tabled Prolog and J , which supports memoization with the M. adverb. In any case, this is only possible for a referentially transparent function. Memoization is also encountered as an easily accessible design pattern within term-rewrite based languages such as Wolfram Language . Dynamic programming

2130-437: Is said to have optimal substructure . If sub-problems can be nested recursively inside larger problems, so that dynamic programming methods are applicable, then there is a relation between the value of the larger problem and the values of the sub-problems. In the optimization literature this relationship is called the Bellman equation . In terms of mathematical optimization, dynamic programming usually refers to simplifying

2201-510: Is the "scrap value". If the optimal pair of control and state trajectories is ( x ∗ , u ∗ ) {\displaystyle (x^{\ast },u^{\ast })} , then V ( t 0 , x 0 ) = J ( t 0 , x 0 ; u ∗ ) {\displaystyle V(t_{0},x_{0})=J(t_{0},x_{0};u^{\ast })} . The function h {\displaystyle h} that gives

Hamilton–Jacobi–Bellman equation - Misplaced Pages Continue

2272-508: Is the control vector that we are trying to find. Thus, V ( x , t ) {\displaystyle V(x,t)} is the value function . The system must also be subject to where F [ ⋅ ] {\displaystyle F[\cdot ]} gives the vector determining physical evolution of the state vector over time. For this simple system, the Hamilton–Jacobi–Bellman partial differential equation

2343-436: Is the maximum of ln ⁡ ( c T − j ) + b V T − j + 1 ( A k a − c T − j ) {\displaystyle \ln(c_{T-j})+bV_{T-j+1}(Ak^{a}-c_{T-j})} , where c T − j {\displaystyle c_{T-j}} is the choice variable and A k

2414-511: Is the scalar cost rate function and D [ ⋅ ] {\displaystyle D[\cdot ]} is a function that gives the bequest value at the final state, x ( t ) {\displaystyle x(t)} is the system state vector, x ( 0 ) {\displaystyle x(0)} is assumed given, and u ( t ) {\displaystyle u(t)} for 0 ≤ t ≤ T {\displaystyle 0\leq t\leq T}

2485-398: Is used in production), known as intertemporal choice . Future consumption is discounted at a constant rate β ∈ ( 0 , 1 ) {\displaystyle \beta \in (0,1)} . A discrete approximation to the transition equation of capital is given by where c {\displaystyle c} is consumption, k {\displaystyle k}

2556-574: Is widely used in bioinformatics for tasks such as sequence alignment , protein folding , RNA structure prediction and protein-DNA binding. The first dynamic programming algorithms for protein-DNA binding were developed in the 1970s independently by Charles DeLisi in the US and by Georgii Gurskii and Alexander Zasedatelev in the Soviet Union . Recently these algorithms have become very popular in bioinformatics and computational biology , particularly in

2627-558: The Hamiltonian , H ( t , x , u , λ ) = I ( t , x , u ) + λ ( t ) f ( t , x , u ) {\displaystyle H\left(t,x,u,\lambda \right)=I(t,x,u)+\lambda (t)f(t,x,u)} , as with ∂ V ( t , x ) / ∂ x = λ ( t ) {\displaystyle \partial V(t,x)/\partial x=\lambda (t)} playing

2698-519: The Taylor expansion of the first term on the right-hand side is where o ( d t ) {\displaystyle {\mathcal {o}}(dt)} denotes the terms in the Taylor expansion of higher order than one in little- o notation . Then if we subtract V ( x ( t ) , t ) {\displaystyle V(x(t),t)} from both sides, divide by dt , and take

2769-478: The brachistochrone problem , can be solved using the Hamilton–Jacobi–Bellman equation, the method can be applied to a broader spectrum of problems. Further it can be generalized to stochastic systems, in which case the HJB equation is a second-order elliptic partial differential equation . A major drawback, however, is that the HJB equation admits classical solutions only for a sufficiently smooth value function, which

2840-465: The parameters of the problem. In a controlled dynamical system , the value function represents the optimal payoff of the system over the interval [t, t 1 ] when started at the time- t state variable x(t)=x . If the objective function represents some cost that is to be minimized, the value function can be interpreted as the cost to finish the optimal program, and is thus referred to as "cost-to-go function." In an economic context, where

2911-485: The Bellman equation is This problem is much simpler than the one we wrote down before, because it involves only two decision variables, c t {\displaystyle c_{t}} and k t + 1 {\displaystyle k_{t+1}} . Intuitively, instead of choosing his whole lifetime plan at birth, the consumer can take things one step at a time. At time t , his current capital k t {\displaystyle k_{t}}

Hamilton–Jacobi–Bellman equation - Misplaced Pages Continue

2982-500: The Fibonacci sequence: F i = F i −1 + F i −2 , with base case F 1  =  F 2  = 1. Then F 43 =  F 42  +  F 41 , and F 42 =  F 41  +  F 40 . Now F 41 is being solved in the recursive sub-trees of both F 43 as well as F 42 . Even though the total number of sub-problems is actually small (only 43 of them), we end up solving

3053-729: The HJB equation is a necessary and sufficient condition for an optimum when the terminal state is unconstrained. If we can solve for V {\displaystyle V} then we can find from it a control u {\displaystyle u} that achieves the minimum cost. In general case, the HJB equation does not have a classical (smooth) solution. Several notions of generalized solutions have been developed to cover such situations, including viscosity solution ( Pierre-Louis Lions and Michael Crandall ), minimax solution ( Andrei Izmailovich Subbotin  [ ru ] ), and others. Approximate dynamic programming has been introduced by D. P. Bertsekas and J. N. Tsitsiklis with

3124-400: The HJB equation with respect to x {\displaystyle x} , which after replacing the appropriate terms recovers the costate equation where λ ˙ ( t ) {\displaystyle {\dot {\lambda }}(t)} is Newton notation for the derivative with respect to time. The value function is the unique viscosity solution to

3195-438: The choice variables c 0 , c 1 , c 2 , … , c T {\displaystyle c_{0},c_{1},c_{2},\ldots ,c_{T}} . (The capital k 0 {\displaystyle k_{0}} is not a choice variable—the consumer's initial capital is taken as given.) The dynamic programming approach to solve this problem involves breaking it apart into

3266-577: The consumer lives. Assume the consumer is impatient, so that he discounts future utility by a factor b each period, where 0 < b < 1 {\displaystyle 0<b<1} . Let k t {\displaystyle k_{t}} be capital in period t . Assume initial capital is a given amount k 0 > 0 {\displaystyle k_{0}>0} , and suppose that this period's capital and consumption determine next period's capital as k t + 1 = A k t

3337-400: The current state x ( t ) {\displaystyle x(t)} as "new" initial condition must be optimal for the remaining problem. If the value function happens to be continuously differentiable , this gives rise to an important partial differential equation known as Hamilton–Jacobi–Bellman equation , where the maximand on the right-hand side can also be re-written as

3408-726: The exact optimization relationship. Alternatively, the continuous process can be approximated by a discrete system, which leads to a following recurrence relation analog to the Hamilton–Jacobi–Bellman equation: at the k {\displaystyle k} -th stage of n {\displaystyle n} equally spaced discrete time intervals, and where f ^ {\displaystyle {\hat {f}}} and g ^ {\displaystyle {\hat {\mathbf {g} }}} denote discrete approximations to f {\displaystyle f} and g {\displaystyle \mathbf {g} } . This functional equation

3479-423: The limit as dt approaches zero, we obtain the HJB equation defined above. The HJB equation is usually solved backwards in time , starting from t = T {\displaystyle t=T} and ending at t = 0 {\displaystyle t=0} . When solved over the whole of state space and V ( x ) {\displaystyle V(x)} is continuously differentiable,

3550-716: The mathematical definition: Notice that if we call, say, fib(5) , we produce a call tree that calls the function on the same value many different times: In particular, fib(2) was calculated three times from scratch. In larger examples, many more values of fib , or subproblems , are recalculated, leading to an exponential time algorithm. Now, suppose we have a simple map object, m , which maps each value of fib that has already been calculated to its result, and we modify our function to use it and update it. The resulting function requires only O ( n ) time instead of exponential time (but requires O ( n ) space): This technique of saving values that have already been calculated

3621-467: The minimal path from P {\displaystyle P} to R {\displaystyle R} . is a paraphrasing of Bellman's famous Principle of Optimality in the context of the shortest path problem . Using dynamic programming in the calculation of the n th member of the Fibonacci sequence improves its performance greatly. Here is a naïve implementation, based directly on

SECTION 50

#1732801570985

3692-480: The needed states, the above operation yields V i −1 for those states. Finally, V 1 at the initial state of the system is the value of the optimal solution. The optimal values of the decision variables can be recovered, one by one, by tracking back the calculations already performed. In control theory , a typical problem is to find an admissible control u ∗ {\displaystyle \mathbf {u} ^{\ast }} which causes

3763-713: The number of solutions is already 116,963,796,250 for n  = 8, as we shall see. Dynamic programming makes it possible to count the number of solutions without visiting them all. Imagine backtracking values for the first row – what information would we require about the remaining rows, in order to be able to accurately count the solutions obtained for each first row value? We consider k × n boards, where 1 ≤ k ≤ n , whose k {\displaystyle k} rows contain n / 2 {\displaystyle n/2} zeros and n / 2 {\displaystyle n/2} ones. The function f to which memoization

3834-523: The objective function usually represents utility , the value function is conceptually equivalent to the indirect utility function . In a problem of optimal control , the value function is defined as the supremum of the objective function taken over the set of admissible controls. Given ( t 0 , x 0 ) ∈ [ 0 , t 1 ] × R d {\displaystyle (t_{0},x_{0})\in [0,t_{1}]\times \mathbb {R} ^{d}} ,

3905-500: The optimal control u ∗ {\displaystyle u^{\ast }} based on the current state x {\displaystyle x} is called a feedback control policy, or simply a policy function. Bellman's principle of optimality roughly states that any optimal policy at time t {\displaystyle t} , t 0 ≤ t ≤ t 1 {\displaystyle t_{0}\leq t\leq t_{1}} taking

3976-406: The path of minimum total length between two given nodes P {\displaystyle P} and Q {\displaystyle Q} . We use the fact that, if R {\displaystyle R} is a node on the minimal path from P {\displaystyle P} to Q {\displaystyle Q} , knowledge of the latter implies the knowledge of

4047-580: The role of the costate variables . Given this definition, we further have d λ ( t ) / d t = ∂ 2 V ( t , x ) / ∂ x ∂ t + ∂ 2 V ( t , x ) / ∂ x 2 ⋅ f ( x ) {\displaystyle \mathrm {d} \lambda (t)/\mathrm {d} t=\partial ^{2}V(t,x)/\partial x\partial t+\partial ^{2}V(t,x)/\partial x^{2}\cdot f(x)} , and after differentiating both sides of

4118-405: The same problems over and over if we adopt a naive recursive solution such as this. Dynamic programming takes account of this fact and solves each sub-problem only once. This can be achieved in either of two ways: Some programming languages can automatically memoize the result of a function call with a particular set of arguments, in order to speed up call-by-name evaluation (this mechanism

4189-543: The solution for finding shortest paths in a recursive manner, which is what the Bellman–Ford algorithm or the Floyd–Warshall algorithm does. Overlapping sub-problems means that the space of sub-problems must be small, that is, any recursive algorithm solving the problem should solve the same sub-problems over and over, rather than generating new sub-problems. For example, consider the recursive formulation for generating

4260-429: The steering. By first using Bellman and then expanding V ( X t , t ) {\displaystyle V(X_{t},t)} with Itô's rule , one finds the stochastic HJB equation where A {\displaystyle {\mathcal {A}}} represents the stochastic differentiation operator , and subject to the terminal condition Note that the randomness has disappeared. In this case

4331-401: The strategy is called " divide and conquer " instead. This is why merge sort and quick sort are not classified as dynamic programming problems. Optimal substructure means that the solution to a given optimization problem can be obtained by the combination of optimal solutions to its sub-problems. Such optimal substructures are usually described by means of recursion . For example, given

SECTION 60

#1732801570985

4402-479: The studies of nucleosome positioning and transcription factor binding. From a dynamic programming point of view, Dijkstra's algorithm for the shortest path problem is a successive approximation scheme that solves the dynamic programming functional equation for the shortest path problem by the Reaching method. In fact, Dijkstra's explanation of the logic behind the algorithm, namely Problem 2. Find

4473-405: The system x ˙ ( t ) = g ( x ( t ) , u ( t ) , t ) {\displaystyle {\dot {\mathbf {x} }}(t)=\mathbf {g} \left(\mathbf {x} (t),\mathbf {u} (t),t\right)} to follow an admissible trajectory x ∗ {\displaystyle \mathbf {x} ^{\ast }} on

4544-412: The system dynamics is given by and the cost accumulates at rate C ( x t , u t ) = r ( t ) u t 2 / 2 + q ( t ) x t 2 / 2 {\displaystyle C(x_{t},u_{t})=r(t)u_{t}^{2}/2+q(t)x_{t}^{2}/2} , the HJB equation is given by with optimal action given by Assuming

4615-433: The system optimally from then until time T {\displaystyle T} . Intuitively, the HJB equation can be derived as follows. If V ( x ( t ) , t ) {\displaystyle V(x(t),t)} is the optimal cost-to-go function (also called the 'value function'), then by Richard Bellman's principle of optimality , going from time t to t  +  dt , we have Note that

4686-399: The top row of the k × n board and recursively compute the number of solutions to the remaining ( k − 1) × n board, adding the numbers of solutions for every admissible assignment of the top row and returning the sum, which is being memoized. The base case is the trivial subproblem, which occurs for a 1 × n board. The number of solutions for this board

4757-401: The top row of the board, and going through every column, subtracting one from the appropriate element of the pair for that column, depending on whether the assignment for the top row contained a zero or a one at that position. If any one of the results is negative, then the assignment is invalid and does not contribute to the set of solutions (recursion stops). Otherwise, we have an assignment for

4828-524: The top-down approach which requires O( n ) space to store the map. In both examples, we only calculate fib(2) one time, and then use it to calculate both fib(4) and fib(3) , instead of computing it every time either of them is evaluated. Consider the problem of assigning values, either zero or one, to the positions of an n × n matrix, with n even, so that each row and each column contains exactly n / 2 zeros and n / 2 ones. We ask how many different assignments there are for

4899-581: The unknown function J x ∗ {\displaystyle J_{x}^{\ast }} and then substitutes the result into the Hamilton–Jacobi–Bellman equation to get the partial differential equation to be solved with boundary condition J ( t 1 ) = b ( x ( t 1 ) , t 1 ) {\displaystyle J\left(t_{1}\right)=b\left(\mathbf {x} (t_{1}),t_{1}\right)} . In practice, this generally requires numerical techniques for some discrete approximation to

4970-497: The use of artificial neural networks ( multilayer perceptrons ) for approximating the Bellman function in general. This is an effective mitigation strategy for reducing the impact of dimensionality by replacing the memorization of the complete function mapping for the whole space domain with the memorization of the sole neural network parameters. In particular, for continuous-time systems, an approximate dynamic programming approach that combines both policy iterations with neural networks

5041-417: Was introduced. In discrete-time, an approach to solve the HJB equation combining value iterations and neural networks was introduced. Alternatively, it has been shown that sum-of-squares optimization can yield an approximate polynomial solution to the Hamilton–Jacobi–Bellman equation arbitrarily well with respect to the L 1 {\displaystyle L^{1}} norm. The idea of solving

#984015