Next Article in Journal
On the Computability of Primitive Recursive Functions by Feedforward Artificial Neural Networks
Previous Article in Journal
Multi-Objective Q-Learning-Based Brain Storm Optimization for Integrated Distributed Flow Shop and Distribution Scheduling Problems
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Exact Solution for the Production Planning Problem with Several Regimes Switching over an Infinite Horizon Time

by
Dragos-Patru Covei
Department of Applied Mathematics, The Bucharest University of Economic Studies, Piata Romana, 1st District, 010374 București, Romania
Mathematics 2023, 11(20), 4307; https://doi.org/10.3390/math11204307
Submission received: 13 September 2023 / Revised: 12 October 2023 / Accepted: 13 October 2023 / Published: 16 October 2023
(This article belongs to the Section Computational and Applied Mathematics)

Abstract

:
We consider a stochastic production planning problem with regime switching. There are k 1 regimes corresponding to different economic cycles. The problem is to minimize the production costs and analyze the problem by the value function approach. Our main contribution is to show that the optimal production is characterized by an exact solution of an elliptic system of partial differential equations. A verification result is given for the determined solution.
MSC:
35B08; 35B09; 35J67; 49L12; 49K15; 60G46

1. Introduction and Proposal of the Paper

We consider a factory producing N 1 types of economic goods that stores them in an inventory-designated place. The model is described mathematically in the next.
Let ( Ω , F , F , P ) be a complete filtered probability space, where P is the historical probability and
F = { F t t 0 , } ,
is generated by an R N -valued Brownian motion denoted by w = w 1 , , w N with respect to the probability P.
In the production planning problem, the regime switching is captured by a continuous-time homogeneous Markov chain ε t adapted to F that can take k different values, modeling k regimes, which should be noted by 1 , 2 , , k . The Markov chain’s rate matrix that denotes the strongly ireductible generator of ε , is denoted by G = [ ϑ i j ] k × k where
ϑ i i = a i i < 0 for all i , ϑ i j = a i j 0 for all i j ,
and the diagonal elements ϑ i i may be expressed as
ϑ i i = Σ j i ϑ i j .
In this case, if P t t = E [ ε ( t ) ] R , then
d P t ( t ) d t = G ε ( t ) .
Moreover, ε ( t ) s explicitly described by the integral form
ε ( t ) = ε ( 0 ) + 0 t G ε ( u ) d u + M ( t ) ,
where M ( t ) is a martingale with respect to F . Here and hereafter, we use the notation from other papers to keep the applicative character of the problem,
p t = p 1 ( t ) , , p N ( t ) ,
which represents the production rate at time t (control variable) adjusted for the demand rate.
These adjusted-for-demand inventory levels are modeled by the following system of stochastic differential equations
d y i ( t ) = p i d t + σ ε ( t ) d w i , y i ( 0 ) = y i 0 for i = 1 , , N ,
where y i ( t ) is an Itô process in R (i.e., the inventory level of good i, at times t ,  adjusted for demand), p i is the deterministic part, σ ε ( t ) is a random regime-dependent constant (non-zero) diffusion coefficient taking on the values σ 1 , σ 2 , …, σ k , and y i 0 is the initial condition (i.e., initial inventory level of goods i).
The stochasticity here is due to demand adjustment, which is random and dependent on the regime. This is the most commonly used process when the demand is more volatile in some periods (e.g., some states of the Markov chain) and less volatile in other periods.
The performance over time of a demand-adjusted production
p t = p 1 ( t ) , , p N ( t ) ,
is measured by means of its cost. At this point, we introduce the cost functional, which yields the cost
J p 1 , , p N : = E 0 ( | p ( t ) | 2 + | y ( t ) | 2 ) e α ε ( t ) t d t , y ( t ) = y 1 ( t ) , , y N ( t ) ,
which measures the quadratic loss.
We measure deviations from the demand, from what place the loss. Here, α ε ( t ) is a regime-dependent, taking on the values α 1 > 0 , α 2 > 0 , …, α k > 0 , constant psychological rate of time discount from what place the exponential discounting.
At the moment, we are ready to frame our objective, which is to minimize the cost functional, i.e.,
inf p 1 , , p N J p 1 , , p N ,
Subject to the Itô Equation (4), the cost functional involves adjusted-for-demand inventory levels y whose dynamic is given by (4), and it depends on the choice of the demand-adjusted production p. Minimizing the cost functional in (6) means selecting the demand-adjusted production p so that it minimizes J (of (5)). Notice that J involves both y and p.
This model problem was proposed by Bensoussan, Sethi, Vickson, and Derzko [1] in the context of no regime switching in the economy and for the case of a factory producing one type of economic goods. Later, many other authors were concerned with regime switching.
In production management, Cadenillas, Lakner, and Pinedo [2] adapted the model problem in [1] to study the optimal production stochastic control planning problem of a company within an economy characterized by two-state regime switching with limited/unlimited information. Later, Dong, Malikopoulos, Djouadi, and Kuruganti [3] applied in civil engineering the model described by [2] to the study of the optimal stochastic control problem for home energy systems with solar and energy storage devices when the demand is subject to Brownian motion; the two switching regimes are the peak and off peak energy demand.
A good deal of attention to this subject has been also devoted by Pirvu and Zhang [4], where the authors studied the effect of high versus low discount rates to a consumption-investment decision problem.
After that, there have been numerous applications of regime switching in many important problems in economics, operations research, actuarial science, finance, reinsurance, and other fields, for example, the portfolio optimization problem in a defaultable market with finitely-many economical regimes is considered by Capponi and Figueroa-López in [5]; the pricing of derivatives using a stochastic discount factor modeled as a regime-switching geometric Brownian motion is discussed by Elliott and Hamada in [6]; the production control in a manufacturing system with multiple machines, which are subject to breakdowns and repairs, is considered by Gharbi and Kenne in [7]; the problem of the pricing of European-style options with switches among a finite number of states is discussed by Yao, Zhang, and Zhou in [8]; and no later, Wang, Chang, and Fang [9] considered the optimal portfolio and consumption rule with a Cox–Ingersoll–Ross (CIR) model in a general utility framework.
There are of course other research studies that may also serve to better explain the importance of regime switching in the real world.
In a precursor to this article, Covei and Pirvu [10] formulate and analyze the production-planning problem in the continuous-time case, with no regime switching in the economy over an infinite time. In [11], the author improved the results of [10], in the sense that the value function in the production model is given in the closed form. Related works that deal with no regime switching in the economy are Sheng-Zhu-Wang [12] and Qin-Bai-Ralescu [13].
Recently, Canepa, Covei, and Pirvu [14] considered the production planning problem with regime switching in the economy over a finite horizon time. Here, the solution is obtained through numerical approaches. However, a closed-form expression for the corresponding case of regime switching on a particular state space consisting of two regimes over an infinite horizon time is available in the paper of [15]. So, at least one question suggested by the paper of [16] has some nice features: can we obtain a closed-form solution when the state space consists of several numbers of states? Our present paper fills the gap in the literature by proving a closed-form solution to the stochastic production planning problem with regime switching in the economy over an infinite horizon in a general state space.
To conclude this introduction, our paper is structured as follows. In Section 2, we give the relationship of our model with a system of partial differential equations (PDE) system. Section 3 presents a closed-form solution and the uniqueness of the solution for our production planning problem. A numerical approximation of the solution for the production planning problem is also given in Section 4. In Section 5, we present a verification result. We introduce in Section 6 the equilibrium production rates as the subgame perfect production rates. They are the output of an interpersonal game between the present self and future selves. The equilibrium production rates are time consistent, meaning there is no incentive to deviate from them. It turns out that in our setting the optimal production rates are among the equilibrium ones so they are time consistent. In Section 7, we give some applications. Finally, in Section 8, we discuss our strategy.
The technique presented in this paper makes a methodological contribution that is of independent interest in other considerable numbers of works on regime switching.
Having presented the model that we want to solve, now we provide our means to tackle it.

2. Reduction of the Model to a PDE System

Our approach is based on the value function and dynamic programming, which leads to the Hamilton–Jacobi–Bellman (HJB) system of equations.
To characterize the value function, we apply the probabilistic approach. We search for functions V x , 1 , …, V x , k such that the stochastic process S p ( t ) defined below
S p t = e α ε ( t ) t V y t , ε ( t ) 0 t [ | p ( s ) | 2 + | y ( s ) | 2 ] e α ε ( s ) s d s ,
is supermartingale for all
p t = p 1 ( t ) , , p N ( t ) ,
And martingale for the optimal control
p * t = p 1 * ( t ) , , p N * ( t ) .
As shown by [10], if this is achieved, with the following transversality condition
lim t E [ e α ε ( t ) t V y t , ε ( t ) ] = 0 ,
some estimates on the value function yield that
V ( x , i ) = inf J p 1 , , p N ,
where x = x 1 , , x N R N assumes values
( y 1 0 , , y N 0 ) .
Once such a function is found, it turns out that ( u 1 , , u k ) with
u 1 ( x ) = V x , 1 , , u k ( x ) = V x , k ,
is the value function. We search for u 1 , , u k , the functions in C 2 0 , , and the supermartingale/martingale requirement yields by using Itô’s Lemma for Markov-modulated diffusion, the HJB system of equations, which characterizes the value function
σ 1 2 2 Δ u 1 σ k 2 2 Δ u k + G a , α u 1 u k x 2 x 2 = inf p { p u 1 + p 2 } inf p { p u k + p 2 } ,
where
G a , α = a 11 + α 1 a 12 a 1 k a 21 a 22 + α 2 a 2 k a k 1 a k 2 a k k + α k .
For the transformation of the HJB system, it is essential to observe that
inf p { p u i + p 2 } = 1 4 u i 2 , i = 1 , 2 , , k .
Thus, the HJB system (10) can be written as a PDE system
σ 1 2 2 Δ u 1 + a 11 + α 1 u 1 i = 2 k a 1 i u i x 2 = 1 4 u 1 2 , σ k 2 2 Δ u k + a k k + α k u k i = 1 k 1 a k i u i x 2 = 1 4 u k 2 .
To perform the verification, i.e., show that the HJB system gives the solution to the optimization problem, one should write (12) with the following boundary condition
u 1 ( x ) , , u k ( x ) , as | x | .
The value function will give us in turn the candidate optimal control. The first-order optimality conditions on the left-hand side of (11) are sufficient for optimality since we deal with a quadratic (convex) function, and they produce the candidate optimal control as follows:
p i * ( t ) = p ¯ i ( y 1 t , , y N t , ε ( t ) ) , i = 1 , , N ,
and
p ¯ i ( x 1 , , x N , j ) = 1 2 u j x i x 1 , , x N , for i { 1 , , n } , j { 1 , , k } .
The production rate p ¯ i is allowed to be negative. A negative production rate would correspond to a write-off or disposal of inventory (for example, due to obsolescence or perishability).
Our next goal of this paper is to determine the candidate optimal control in closed form.

3. Closed-Form Solution for the PDE System

In spite of their clear simplicity, the PDE system (12) with boundary conditions (13) presents a host of mathematical difficulties arising from the presence of nonlinear gradient terms u 1 2 , …, u k 2 , see for details [17].
The following result will be proved and is the main original element of the article.
Theorem 1.
Assume that G a , α is a positive definite matrix with all elements of G a , α 1 positive. Then, the PDE system (12) with boundary condition (13) has a unique radially symmetric convex positive classical solution with quadratic growth.
Proof of Theorem 1.
In the following, we construct the function
u 1 , , u k C 2 0 , × × C 2 0 , ,
which satisfies (12) with boundary condition (13). One way of solving this partial differential equation is to show that there exists
u 1 x , , u k x = β 1 x 2 + η 1 , , β k x 2 + η k , with β 1 , , β k , η 1 , , η k 0 , ,
that solves (1).
The main task for the proof of existence of (15) is performed by proving that there exists
β 1 , , β k , η 1 , , η k 0 , ,
such that
2 β 1 N σ 1 2 2 + a 11 + α 1 β 1 x 2 + η 1 i = 2 k a 1 i β i x 2 + η i x 2 = 1 4 2 β 1 x 2 , 2 β k N σ k 2 2 + a k k + α k β k x 2 + η k i = 1 k 1 a k i β i x 2 + η i x 2 = 1 4 2 β k x 2 ,
or equivalently, after grouping the terms
x 2 i = 2 k a 1 i β i + ( a 11 + α 1 ) β 1 + β 1 2 1 β 1 N σ 1 2 i = 2 k a 1 i η i + ( a 11 + α 1 ) η 1 = 0 , x 2 i = 1 k 1 a k i β i + ( a k k + α k ) β k + β k 2 1 β k N σ k 2 i = 1 k 1 a k i η i + a k k + α k η k = 0 .
Now, we consider the system of equations
i = 2 k a 1 i β i + ( a 11 + α 1 ) β 1 + β 1 2 1 = 0 i = 1 k 1 a k i β i + ( a k k + α k ) β k + β k 2 1 = 0 β 1 N σ 1 2 i = 2 k a 1 i η i + ( a 11 + α 1 ) η 1 = 0 β k N σ k 2 i = 1 k 1 a k i η i + a k k + α k η k = 0 .
To solve (16), we can rearrange those equations 1, …, k such
a 11 + α 1 a 1 k a k 1 a k k + α k β 1 β k = 1 β 1 2 1 β k 2 .
The arguments in [18,19] say that System (17) has a unique positive solution. In fact, denoting by
h 1 β 1 , , β k = i = 2 k a 1 i β i + ( a 11 + α 1 ) β 1 + β 1 2 1 , h k β 1 , , β k = i = 1 k 1 a k i β i + ( a k k + α k ) β k + β k 2 1 ,
It happens that
h 1 β 1 β 1 , . . , β k = ( a 11 + α 1 ) + 2 β 1 > 0 , h 1 0 , β 2 , , β k < 0 , lim β 1 h 1 = , h k β k β 1 , , β k = ( a k k + α k ) + 2 β k > 0 , h k β 1 , , β k 1 , 0 < 0 , lim β k h k = ,
concluding the arguments in [18,19]. Next, letting
β 1 , , β k 0 , × × 0 ,
a unique solution of (17), we observe that the Equations k + 1 , …, 2 k of (16) can be written equivalently as
β 1 N σ 1 2 β k N σ k 2 = a 11 + α 1 a 1 k a k 1 a k k + α k η 1 η k ,
from where using the fact that G a , α 1 has all elements positive, we can see that there exist and are unique η 1 , …, η k 0 , that solve (16) and then
u 1 x , , u k x ,
solve (12). This finishes the proof of Theorem 1. □
Because our solution depends on solving a nonlinear algebraic system of equations, the exact solution of the PDE system cannot be determined using a computer software. In order to be implemented, the solution of the PDE system (12) in a software application in the next section, it is necessary to give the numerical approximation of solution to (16), and therefore, the arguments in [18,19] are used again.

4. Numerical Solution of an Algebraic Nonlinear System in Building the Solution for the PDE System

We intend to approximate β 1 , , β k , η 1 , , η k 0 , in (15) by the Newton–Raphson method. To do this, we denote h 1 β 1 , , β k , …, h k β 1 , , β k as in (18) and
J h 1 , , h k = a 11 + α 1 + 2 β 1 a 1 k a k 1 β 1 a k k + α k + 2 β k ,
The Jacobian matrix of (18). For n = 1 , 2 , we find the approximate of the unique parameters
β 1 , , β k 0 , × × 0 , ,
In the following way,
β 1 n + 1 β k n + 1 = β 1 n β k n a 11 + α 1 + 2 β 1 n a 1 k a k 1 a k k + α k + 2 β k n 1 h 1 β 1 n , , β k n h k β 1 n , , β k n ,
with β 1 0 , , β k 0 0 , . Clearly η 1 ,…, η k 0 , are easily determined from (19). Some other interesting numerical iterations can be applied in obtaining an optimal numerical solution of (15), which might be efficiently computed with reduced number of iterations and quick CPU time. For example, quasi-Newton variants: the AGD method (see [20]), the SM method (see [21]), or the accelerated double-step-size method (see [22]).
Now, we will move on to the verification result, which is also inspired from [15].

5. Verification

Next, we show that the control of (14) obtained in our reduction strategy is indeed optimal. We apply the supermartingale and martingale approaches.
Repeating the same argument in [14], as the first step, we can show that the stochastic process S p ( t ) defined below
S p t = e α ε ( t ) t V y t , ε ( t ) 0 t [ | p ( s ) | 2 + | y ( s ) | 2 ] e α ε ( s ) s d s ,
is supermartingale for all
p t = p 1 ( t ) , , p N ( t ) ,
And martingale for the optimal control
p * t = p 1 * ( t ) , , p N * ( t ) .
Owing to the well-known Itô Lemma for Markov-modulated diffusion (see [8] for more on this), we have
d S p s = e α ε ( s ) s [ σ ε s 2 2 Δ V y s , ε ( s ) y s 2 + p s V y s , ε ( s ) p s 2 ( α ε ( s ) + a ε ( s ) ε ( s ) ) V y s , ε s + i = 1 , i ε ( s ) k a ε ( s ) i V y s , i ] d s + d Z ( s ) ,
for some martingale Z ( s ) , and Z ( 0 ) = 0 . Therefore,
E S p t = S p 0 + E [ 0 t e α ε ( s ) s [ σ ε s 2 2 Δ V y s , ε ( s ) y s 2 + p s V y s , ε ( s ) ] d s ] + E [ 0 t e α ε ( s ) s [ p s 2 ( α ε ( s ) + a ε ( s ) ε ( s ) ) V y s , ε s ] d s ] + E [ 0 t e α ε ( s ) s [ i = 1 , i ε ( s ) k a ε ( s ) i V y s , i ] d s ] .
Then, the claim yields considering HJB Equations (10) and (12), which says that S p t is martingale for the optimal control and supermartingale otherwise. This last fact combined with the transversality condition yields the claim.
In the second step, let us establish the optimality of p 1 * , , p N * . Consider the quadratic estimate on the value function
V x , 1 = β 1 x 2 η 1 , , V x , k = β k x 2 η k ,
where β i and η i 0 , are the solutions of (16).
Let us provide a lower-bound estimate for α 1 , , α k so that the transversality condition (8) is met and
lim t E [ e α ϵ ( t ) t | y t | 2 ] = 0
holds true. The SDE system (4) in this case becomes
d y i t = β ϵ ( t ) y i t d t + σ ε ( t ) d W i ( t ) , i = 1 , N .
Using Itô’s Lemma, one obtains
d ( y i t ) 2 = 2 y i t d y i t + d y i t d y i t = [ 2 β ϵ ( t ) ( y i t ) 2 + σ ϵ ( t ) 2 ] d t + 2 y i t σ ϵ ( t ) d W i ( t ) .
We introduce
F i ( t ) = E [ ( y i t ) 2 ] .
By taking expectations in the above equation, we obtain
F i ( t ) = E 0 t [ 2 β ϵ ( s ) ( y i s ) 2 + σ ϵ ( s ) 2 ] d s + [ ( y i 0 ) 2 ] = E 0 t [ 2 β ϵ ( s ) ( y i s ) 2 + σ ϵ ( s ) 2 ] d s + y i 2 0 ) .
Let
D 2 = max { σ 1 2 , , σ k 2 } , D 3 = max ( [ ( y 1 0 ) 2 ] , , [ ( y k 0 ) 2 ] ) .
Then, in the light of the above equation, we obtain
F i ( t ) 0 t D 2 d s + D 3 .
Hence, we have that
F i ( t ) D 2 t + D 3 .
Therefore, one must choose α 1 , , α k 0 , for the transversality condition to hold true, and the proof is completed. Finally, a simple system of nonlinear Equations (16) remains to be solved.

6. The Equilibrium Production

For a production rate { p i ( t ) } t 0 and its corresponding inventory level { y i ( t ) } t 0 given by (4), we introduce equilibrium production as the subgame perfect production in the definition below (for more on this economic concept see [23]).
Definition 1.
Let F = ( F i , i = 1 , N ) : R × { 1 , 2 , k } R N be a vector map such that for any x > 0 and i { 1 , 2 , , k }
lim inf ϵ 0 J ( p ¯ i ) J ( p i ϵ ) ϵ 0 ,
where the subgame perfect production
p ¯ i ( s ) : = F i ( y ¯ i ( s ) , ϵ ( s ) ) .
Here, the process { y ¯ i ( s ) } s 0 is the inventory level process corresponding to { p ¯ i ( s ) } s 0 . The production rate { p ϵ i ( s ) } s 0 is defined by
p ϵ i ( s ) = p ¯ i ( s ) , s [ 0 , ] E ϵ , 0 p i ( s ) , s E ϵ , 0 ,
E ϵ , 0 = [ 0 , ϵ ] ; { p i ( s ) } s E ϵ , 0 is any production rate. If (21) holds true, then p ¯ i ( s ) ,   i = 1 N , is a subgame perfect production rate.
The equilibrium production is by design time consistent, meaning that they will be implemented at a future date even if the optimization criterion is updated. In some situations, the optimal production may be time inconsistent meaning that they will fail to be implemented in the future because they are not optimal anymore if the optimization criterion is updated; they will be implementable only in the presence of a commitment mechanism, that is why sometimes they are referred to as pre commitment production. Let us remark that, in our setting, the optimal production rate
p ¯ i , i { 1 , , N } ,
is a subgame perfect production with
F i ( x , j ) : = 1 2 u j x i ( x ) ,
Since
( p ¯ i , i = 1 N ) = arg min p 1 , , p N J p 1 , , p N
and thus (21) is automatically satisfied. Therefore, the equilibrium production is time consistent.

7. Applications

We offer some applications, which also are inspired by the paper of Ghosh, Arapostathis, and Marcus [16].
Application 1. Suppose there is one machine producing two products, and let ε t be the machine state that can take values in two regimes, 1 = good and 2 = bad, i.e., for every t 0 , , we have ε t { 1 , 2 } . We consider ε t a continuous-time Markov chain with generator
1 2 1 2 1 2 1 2 ,
And the inventory y i t , which is governed by the Itô system of stochastic differential Equations (4) with the diffusion σ 1 = σ 2 = 1 2 , and let α 1 = α 2 = 1 2 be the discount factor. Under these assumptions, the system (17) becomes
a 11 + α 1 a 11 a 22 a 22 + α 2 β 1 β 2 = 1 β 1 2 1 β 2 2 ,
or, with our data
β 1 2 + β 1 1 2 β 2 1 = 0 β 2 2 1 2 β 1 + β 2 1 = 0
which has a unique positive solution
β 1 = 1 4 17 1 , β 2 = 1 4 17 1 .
On the other hand, System (19) becomes
β 1 N σ 1 2 β 2 N σ 2 2 = a 11 + α 1 a 11 a 22 a 22 + α 2 η 1 η 2 ,
or, with our data
β 1 β 2 = 1 1 2 1 2 1 η 1 η 2 ,
which has a unique positive solution
η 1 = 4 3 β 1 + 2 3 β 2 = 1 2 17 1 , η 2 = 2 3 β 1 + 4 3 β 2 = 1 2 17 1 .
Then,
V x 1 , x 2 , 1 = V x 1 , x 2 , 2 = 1 4 17 1 x 1 2 + x 2 2 1 2 17 1
and furthermore, the production rate is
p ¯ i ( x 1 , x 2 , j ) = 1 2 17 1 x i , for i { 1 , 2 } , j { 1 , 2 } .
We also give the approximate of β 1 and β 2 , and η 1 and η 2 , by using the Newton–Raphson method. Denote
h 1 β 1 , β 2 = a 12 β 2 + ( a 11 + α 1 ) β 1 + β 1 2 1 h 2 β 1 , β 2 = a 21 β 1 + ( a 22 + α 2 ) β 2 + β 2 2 1
and
J h 1 , , h k = 2 β 1 + 1 1 2 1 2 2 β 2 + 1 .
We construct
β 1 n + 1 β 2 n + 1 = β 1 n β 2 n a 11 + α 1 + 2 β 1 n a 1 k a k 1 a k k + α k + 2 β k n 1 h 1 β 1 n , β 2 n h 2 β 1 n , β 2 n β 1 0 = β 2 0 = 0.1 .
Using the standard computation, approximations to four digits are
n = 1 β 1 1 = 1.4429 and β 2 1 = 1.4429 n = 2 β 1 2 = 0.9102 and β 2 2 = 0.9102 n = 3 β 1 3 = 0.7808 and β 2 3 = 0.7808 n = 4 β 1 4 = 0.7808 and β 2 4 = 0.7808
On the other hand,
β 1 = β 2 = 1 4 17 1 0.780 7 .
Clearly, the approximations for η 1 and η 2 are
η 1 = η 2 1 . 561 6 .
Application 2. Suppose there is one machine producing three products, and let ε t be the machine state that can take values in three regimes 1, 2, and 3, i.e., for every t 0 , , we have ε t { 1 , 2 , 3 } . We consider ε t a continuous-time Markov chain with generator
3 3 0 4 7 3 0 4 4 ,
and the inventory y i t , which is governed by (4) with σ 1 = σ 2 = σ 3 = 1 3 , and let α 1 = α 2 = α 3 = 1 be the discount factor. Under these assumptions, System (17) becomes
a 11 + 1 a 11 0 a 22 a 22 + a 11 + 1 a 11 0 a 22 a 22 + 1 β 1 β 2 β 3 = 1 β 1 2 1 β 2 2 1 β 3 2 ,
or, with our data
β 1 2 + 4 β 1 3 β 2 1 = 0 β 2 2 + 8 β 2 4 β 1 3 β 3 1 = 0 β 3 2 + 5 β 3 4 β 2 1 = 0
which has a unique positive solution
β 1 = β 2 = β 3 = 1 2 5 1 .
On the other hand, System (19) becomes
β 1 N σ 1 2 β 2 N σ 2 2 β 3 N σ 3 2 = a 11 + 1 a 11 0 a 22 a 22 + a 11 + 1 a 11 0 a 22 a 22 + 1 η 1 η 2 η 3 ,
or, with our data
β 1 β 2 β 3 = 3 + 1 3 0 4 4 + 3 + 1 3 0 4 4 + 1 η 1 η 2 η 3 ,
from where
η 1 η 2 η 3 = 7 13 15 52 9 52 5 13 5 13 3 13 4 13 4 13 5 13 β 1 β 2 β 3 ,
has a unique positive solution
η 1 = η 2 = η 3 = 1 2 5 1 2 .
Then,
V x 1 , x 2 , x 3 , 1 = V x 1 , x 2 , x 3 , 2 = V x 1 , x 2 , x 3 , 3 = 1 2 5 1 x 1 2 + x 2 2 + x 3 2 + 1
and furthermore, the production rate is
p ¯ i ( x 1 , x 2 , x 3 , j ) = 1 2 5 1 x i , for i { 1 , 2 , 3 } , j { 1 , 2 , 3 } .
We also point out that the numerical approximations for β 1 , β 2 , and β 3 , using the Newton–Raphson method described, are
n = 1 β 1 1 = 0.8418 β 2 1 = 1.017 β 3 1 = 1.2789 n = 2 β 1 2 = 0.6575 β 2 2 = 0.6761 β 3 2 = 0.7066 n = 3 β 1 3 = 0.6192 β 2 3 = 0.6196 β 3 3 = 0.6202 n = 4 β 1 4 = 0.618 β 2 4 = 0.618 β 3 4 = 0.618
when β 1 0 = 1 , β 2 0 = 2 , and β 3 0 = 3 . Clearly, 1 2 5 1 0.618 .

8. Final Remarks and Conclusions

When w i is correlated with correlation ρ , the HJB system (10) becomes
σ 1 2 2 Δ u 1 σ k 2 2 Δ u k + G a , α u 1 u k ρ 2 σ 1 2 i j 2 u 1 x i x j σ k 2 i j 2 u k x i x j x 2 x 2 = inf p { p u 1 + p 2 } inf p { p u k + p 2 } ,
which has the same solution as (10), due to the mixed derivative terms (see [17] for details).
In summary, we have reduced the stochastic production-planning problem with several regime switching in the economy to demonstrate that there is an exact solution for the PDE system that models the stochastic production problem.

Funding

This research received no external funding.

Data Availability Statement

No data were used.

Acknowledgments

The author would like to thank the referees for their valuable discussions.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Bensoussan, A.; Sethi, S.P.; Vickson, R.; Derzko, N. Stochastic production planning with production constraints. SIAM J. Control 1984, 22, 920–935. [Google Scholar] [CrossRef]
  2. Cadenillas, A.; Lakner, P.; Pinedo, M. Optimal production management when demand depends on the business cycle. Oper. Res. 2013, 61, 1046–1062. [Google Scholar] [CrossRef]
  3. Dong, J.; Malikopoulos, A.; Djouadi, S.M.; Kuruganti, T. Application of Optimal Production Control theory for Home Energy Management in a Micro Grid. In Proceedings of the 2016 American Control Conference (ACC), Boston, MA, USA, 6–8 July 2016; pp. 5014–5019. [Google Scholar]
  4. Pirvu, T.A.; Zhang, H. Investment-consumption with regime-switching discount rates. Math. Soc. Sci. 2014, 71, 142–150. [Google Scholar] [CrossRef]
  5. Capponi, A.; Figueroa-López, J.E. Dynamic Portfolio Optimization with a Defaultable Security and Regime-Switching. Math. Financ. 2012, 207–249. [Google Scholar] [CrossRef]
  6. Elliott, R.; Hamada, A.S. Option Pricing Using A Regime Switching Stochastic Discount Factor. Int. J. Theor. Appl. Financ. 2014, 17, 1–26. [Google Scholar] [CrossRef]
  7. Gharbi, A.; Kenne, J.P. Optimal production control problem in stochastic multiple-product multiple-machine manufacturing systems. IIE Trans. 2003, 35, 941–952. [Google Scholar] [CrossRef]
  8. Yao, D.D.; Zhang, Q.; Zhou, X.Y. A regime-switching model for european options. In Stochastic Processes, Optimization, and Control Theory: Applications in Financial Engineering, Queueing Networks, and Manufacturing Systems; Yan, H., Yin, G., Zhang, Q., Eds.; International Series in Operations Research & Management Science; Springer: New York, NY, USA, 2006; Chapter 14; Volume 94, pp. 281–300. [Google Scholar]
  9. Wang, C.F.; Chang, H.; Fang, Z.M. Optimal Portfolio and Consumption Rule with a CIR Model under HARA Utility. J. Oper. Res. Soc. China 2018, 6, 107–137. [Google Scholar] [CrossRef]
  10. Covei, D.-P.; Pirvu, T.A. An elliptic partial differential equation and its application. Appl. Math. Lett. 2020, 101, 1–7. [Google Scholar] [CrossRef]
  11. Covei, D.-P. An elliptic partial differential equation modeling the production planning problem. J. Appl. Anal. Comput. 2021, 11, 903–910. [Google Scholar]
  12. Sheng, L.; Zhu, Y.; Wang, K. Uncertain dynamical system-based decision making with application to production-inventory problems. Appl. Math. Model. 2018, 56, 275–288. [Google Scholar] [CrossRef]
  13. Qin, Z.; Bai, M.; Ralescu, D. A fuzzy control system with application to production planning problems. Inf. Sci. 2011, 181, 1018–1027. [Google Scholar] [CrossRef]
  14. Canepa, E.C.; Covei, D.-P.; Pirvu, T.A. Stochastic production planning with regime switching. J. Ind. Manag. 2023, 19, 1697–1713. [Google Scholar]
  15. Covei, D.-P.; Pirvu, T.A. An elliptic partial differential equations system and its applications. Carpathian J. Math. 2021, 37, 427–440. [Google Scholar] [CrossRef]
  16. Ghosh, M.K.; Arapostathis, A.; Marcus, S.I. Optimal Control of Switching Diffusions with Application to Flexible Manufacturing Systems. SIAM J. Control Optim. 1992, 31, 1183–1204. [Google Scholar] [CrossRef]
  17. Covei, D.-P. On a parabolic partial differential equation and system modeling a production planning problem. Electron. Arch. 2022, 30, 1340–1353. [Google Scholar] [CrossRef]
  18. Győri, I.; Hartung, F.; Mohamady, N.A. Existence and uniqueness of positive solutions of a system of nonlinear algebraic equations. Period. Math. Hung. 2017, 75, 114–127. [Google Scholar] [CrossRef]
  19. Győri, I.; Hartung, F.; Mohamady, N.A. Boundedness of positive solutions of a system of nonlinear delay differential equations. Discret. Contin. Dyn. Syst.—Ser. B 2018, 23, 809–836. [Google Scholar] [CrossRef]
  20. Andrei, N. An acceleration of gradient descent algorithm with backtracking for unconstrained optimization. Numer. Algorithms 2006, 42, 63–73. [Google Scholar] [CrossRef]
  21. Stanimirovic, P.S.; Miladinovic, M.B. Accelerated gradient descent methods with line search. Numer. Algorithms 2010, 54, 503–520. [Google Scholar] [CrossRef]
  22. Petrovic, M.J. An Accelerated Double Step Size model in unconstrained optimization. Appl. Math. Comput. 2015, 250, 309–319. [Google Scholar] [CrossRef]
  23. Ekeland, I.; Pirvu, T.A. Investment and consumption without commitment. Math. Financ. Econ. 2008, 2, 57–86. [Google Scholar] [CrossRef]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Covei, D.-P. Exact Solution for the Production Planning Problem with Several Regimes Switching over an Infinite Horizon Time. Mathematics 2023, 11, 4307. https://doi.org/10.3390/math11204307

AMA Style

Covei D-P. Exact Solution for the Production Planning Problem with Several Regimes Switching over an Infinite Horizon Time. Mathematics. 2023; 11(20):4307. https://doi.org/10.3390/math11204307

Chicago/Turabian Style

Covei, Dragos-Patru. 2023. "Exact Solution for the Production Planning Problem with Several Regimes Switching over an Infinite Horizon Time" Mathematics 11, no. 20: 4307. https://doi.org/10.3390/math11204307

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop