Research article

An iterative technique for solving path planning in identified environments by using a skewed block accelerated algorithm

  • Received: 26 August 2022 Revised: 18 October 2022 Accepted: 26 October 2022 Published: 22 December 2022
  • MSC : 35A35, 65F10, 68T40

  • Currently, designing path-planning concepts for autonomous robot systems remains a topic of high interest. This work applies computational analysis through a numerical approach to deal with the path-planning problem with obstacle avoidance over a robot simulation. Based on the potential field produced by Laplace's equation, the formation of a potential function throughout the simulation configuration regions is obtained. This potential field is typically employed as a guide in the global approach of robot path-planning. An extended variant of the over-relaxation technique, namely the skewed block two-parameter over relaxation (SBTOR), otherwise known as the explicit decoupled group two-parameter over relaxation method, is presented to obtain the potential field that will be used for solving the path-planning problem. Experimental results with a robot simulator are presented to demonstrate the performance of the proposed approach on computing the harmonic potential for solving the path-planning problem. In addition to successfully validating pathways generated from various locations, it is also demonstrated that SBTOR outperforms existing over-relaxation algorithms in terms of the number of iterations, as well as the execution time.

    Citation: A'qilah Ahmad Dahalan, Azali Saudi. An iterative technique for solving path planning in identified environments by using a skewed block accelerated algorithm[J]. AIMS Mathematics, 2023, 8(3): 5725-5744. doi: 10.3934/math.2023288

    Related Papers:

    [1] A'qilah Ahmad Dahalan, Azali Saudi . Pathfinding algorithm based on rotated block AOR technique in structured environment. AIMS Mathematics, 2022, 7(7): 11529-11550. doi: 10.3934/math.2022643
    [2] Shuting Tang, Xiuqin Deng, Rui Zhan . The general tensor regular splitting iterative method for multilinear PageRank problem. AIMS Mathematics, 2024, 9(1): 1443-1471. doi: 10.3934/math.2024071
    [3] Huiling Wang, Nian-Ci Wu, Yufeng Nie . Two accelerated gradient-based iteration methods for solving the Sylvester matrix equation AX + XB = C. AIMS Mathematics, 2024, 9(12): 34734-34752. doi: 10.3934/math.20241654
    [4] Yajun Xie, Changfeng Ma . The hybird methods of projection-splitting for solving tensor split feasibility problem. AIMS Mathematics, 2023, 8(9): 20597-20611. doi: 10.3934/math.20231050
    [5] Yang Cao, Quan Shi, Sen-Lai Zhu . A relaxed generalized Newton iteration method for generalized absolute value equations. AIMS Mathematics, 2021, 6(2): 1258-1275. doi: 10.3934/math.2021078
    [6] Yajun Xie, Minhua Yin, Changfeng Ma . Novel accelerated methods of tensor splitting iteration for solving multi-systems. AIMS Mathematics, 2020, 5(3): 2801-2812. doi: 10.3934/math.2020180
    [7] Jin-Song Xiong . Generalized accelerated AOR splitting iterative method for generalized saddle point problems. AIMS Mathematics, 2022, 7(5): 7625-7641. doi: 10.3934/math.2022428
    [8] Zui-Cha Deng, Fan-Li Liu, Liu Yang . Numerical simulations for initial value inversion problem in a two-dimensional degenerate parabolic equation. AIMS Mathematics, 2021, 6(4): 3080-3104. doi: 10.3934/math.2021187
    [9] Wen-Ning Sun, Mei Qin . On maximum residual block Kaczmarz method for solving large consistent linear systems. AIMS Mathematics, 2024, 9(12): 33843-33860. doi: 10.3934/math.20241614
    [10] Lina Zhang, Xiaoting Rui, Jianshu Zhang, Guoping Wang, Junjie Gu, Xizhe Zhang . A framework for establishing constraint Jacobian matrices of planar rigid-flexible-multibody systems. AIMS Mathematics, 2023, 8(9): 21501-21530. doi: 10.3934/math.20231096
  • Currently, designing path-planning concepts for autonomous robot systems remains a topic of high interest. This work applies computational analysis through a numerical approach to deal with the path-planning problem with obstacle avoidance over a robot simulation. Based on the potential field produced by Laplace's equation, the formation of a potential function throughout the simulation configuration regions is obtained. This potential field is typically employed as a guide in the global approach of robot path-planning. An extended variant of the over-relaxation technique, namely the skewed block two-parameter over relaxation (SBTOR), otherwise known as the explicit decoupled group two-parameter over relaxation method, is presented to obtain the potential field that will be used for solving the path-planning problem. Experimental results with a robot simulator are presented to demonstrate the performance of the proposed approach on computing the harmonic potential for solving the path-planning problem. In addition to successfully validating pathways generated from various locations, it is also demonstrated that SBTOR outperforms existing over-relaxation algorithms in terms of the number of iterations, as well as the execution time.



    Path planning is a computational problem to find a sequence of valid configurations that move the object of interest from an arbitrary source to a specific destination. The term path planning has been used widely, mostly in the robotics area. Good path planning for a mobile robot means it should be executed while avoiding walls and dodging any obstacles in between. Autonomous robot path planning is very impactful, as it is extremely beneficial not only for transportation, but also for households, pharmaceuticals, industries, manufacturing, aerospace and others. In general, the ability of a robot to navigate autonomously is a prerequisite and foundation for the development of intelligent robots. To ensure the successful completion of the many tasks that mobile robots are designed to accomplish, path-planning algorithms must be both effective and practical.

    On the whole, navigation problems are divided into four main categories, i.e., localization, path planning, motion control and cognitive mapping [1]. In essence, the localization of a robot is telling where the robot is currently located, whereas path planning reveals in which direction the robot is meant to travel and what is the best route to arrive at the destination. Robot motion control gives instructions on the manner in which the robot is able to move. Cognitive mapping concerns the previous, current and next positions of the robot, as well as to what extent the robot should remember. Hence, from all of these category's functions, it is not wrong to say that path planning is a crucial issue to be solved in achieving intelligent robots. With the aid of a path-planning algorithm, the robot should be able to select and recognize the ideal route inside the given configuration region.

    It is common knowledge that there are two types of navigation, i.e., global and local approaches. This work considers global navigation, where prior knowledge of the configuration region is given, which is also known as an off-line mode for path planning. The robot's ability to represents the real world and to execute the algorithm are the two key elements of global path planning. These two elements are interconnected and significantly influence one another while determining the optimal route and minimum duration for the robot to travel across the known environment [2]. The main objective of this study is to utilize the harmonic potential in simulating point-robot path-planning in the identified environment following the resemblance of heat transfer. The model of the aforementioned heat transfer problem was designed by using Laplace's equation [3]. Harmonic functions are literally the solutions to Laplace's equation, which are also called Laplacian potentials. These Laplacian potentials portray the analogy of temperature values from heat transfer in the configuration region that will be used to simulate the path generation. Heat transfer's capacity to surpass the local minima enigma is one of its most crucial characteristics, making it very encouraging for robot navigation control. The following sections provide further explanations. Numerous approaches can be used to achieve harmonic functions, but the most prevalent approaches rely on numerical techniques due to the availability of rapid processing machines, as well as on their elegance and competence in solving the problem [4,5,6].

    Furthermore, the iterative approach could be implemented to produce the Laplacian potential by computationally solving Laplace's equation and discretizing the environment through the use of mesh grid points [7,8,9,10]. The cost of computation for iterative techniques is substantially higher than analytical techniques since the discretization process produces a sparse linear system. However, the discretization resolution determines the computational complexity. Using a finite difference approximation, the Laplace equation can be translated to a linear system. When this linear system is documented in matrix notation, the majority of the elements are zero, and the matrix size of the end product is usually large and sparse. Therefore, the iterative technique is employed to solve the linear system to prevent the higher requirement for memory storage of those large and sparse matrices. Several experiments were executed in this work to measure the competency of the skewed block accelerated iterative approach in terms of the computation of the Laplacian potential when generating robot-point paths. Fundamentally, a complete process of the path-planning construction phase in this study consists of the following steps:

    Begin

    Step 1: Mapping of the robot's configuration region.

    Step 2: Formulation and modeling of the finite-difference approximation of the proposed iterative schemes.

    Step 3: Algorithms of the proposed iterative schemes.

    Step 4: Numerical simulations.

    Step 5: Evaluation and analysis.

    End.

    Consider Laplace's partial differential equation in two dimensions:

    2Ux2+2Uy2=0, (1)

    where U(x,y) is some unknown function of two variables. A harmonic function in the region Ωn should be able to satisfy the generalization of Eq (1), as below.

    2U=ni=12Ux2i=0, (2)

    in which 2 is the n-dimensional Laplace operator (or Laplacian as it is often called), the xi is the i-th coordinate in a Cartesian system and n is the dimension of the region. This work only focuses on two-dimensional problems with xy notation represented as ij. The region boundary of Ω comprises internal and external walls, boundaries of obstaclesand the goal position. According to [7], the existence of the spontaneous formation of local minima in the solution region cannot arise because harmonic functions obey the min-max principle. Thus, the only critical points that can happen are saddle points. The escape from such a critical point would then be found by searching the neighborhood. Any path deviation or disturbance from the critical point will eventually cause a smooth path throughout the region. For this reason, harmonic potential provides a highly beneficial option in navigation, as it offers a complete path-planning algorithm. The equation of Laplace can be solved with a range of numerical techniques, i.e., Jacobi, Gauss-Seidel and successive over-relaxation (SOR) are the standard approaches [11,12]. The development of the SOR iterative method has sparked interest in examining and resolving several issues. Since its establishment, the generalization of over-relaxation iterative approaches has been the subject of extensive research, and it has had promising outcomes. Therefore, this work will solve Eq (2) by exploiting the combination of skewed block accelerated iterative approach with the accelerated over-relaxation (AOR) and two-parameter over-relaxation (TOR) methods, for rapid computation.

    In the configuration region for this model, the robot is described as a node point and the designated area is defined in the mesh grid pattern. As stated before, the off-line mode approach is used to compute the Laplacian operator of the point-robot designated area by utilizing the analogy of heat transfer. All wall boundaries and obstacles (with the highest potential value) stand for heat sources, while the goal position (with the lowest potential value) represents the heat sink. This heat transfer activity is modeled from Laplace's equation and then numerically solved to gain the heat distribution, which represents the harmonic potential for each nodal point in the mesh grid. By making use of the heat distribution property which flows from higher to lower temperatures, a gradient search can be used to generate the path from any starting position with a high potential value to the goal which has the lowest potential value. The path-planning algorithm utilizes the gradient descent search (GDS) to determine a feasible route for the robot to travel around the environment safely from the start point to reach the goal position. From the current point, the GDS method examines the potential values of its eight neighboring points on the finite-difference grid and simply picks the node with the lowest Laplacian potential value. This procedure is repeated until it reaches the target point [7,13,14,15].

    This work aims to imitate the aforesaid paradigm for path planning, describing the solution of Laplace's Eq (2) over the analogy of temperature (for the potential) and heat flow (for the pathway). The experiment was carried out in a two-dimensional domain with walls and various forms of obstacles. A new technique called the skewed block two-parameter over-relaxation (SBTOR) iterative approach is proposed to solve Eq (2) and obtain the potential values for each node. For the purpose of performance comparison, the existing block successive over-relaxation (BSOR), block accelerated over-relaxation (BAOR), block two-parameter over-relaxation (BTOR), skewed block successive over-relaxation (SBSOR), and skewed block accelerated over-relaxation (SBAOR) techniques are also investigated.

    The pioneering work of the skewed block technique (originally known as explicit decoupled group, EDG) arose from Abdullah's [16] focus on solving two-dimensional Poisson equations. The EDG iterative scheme essentially engaged the half-sweep technique in a block frame on the rotated mesh grid. By virtue of the proficiency of this technique, it has been widely used to solve numerous differential equations [17,18,19,20,21,22,23,24,25]. In the robotics literature, the standard Gauss-Seidel [18,19] and SOR [3,20] schemes have been employed to solve Eq (2). This work, however, generalizes a reliable numerical solver, when addressing the solutions of Laplace's equation, by utilizing a skewed block technique.

    Fundamentally, merely half of the node points in the mesh region were computed while using a skewed block approach. In contrast, the explicit group (EG) scheme is basically a full-sweep (FS) technique in a block frame, evaluated every node point in the mesh region. Figure 1 provides an example of each EG and EDG scheme in a 10×10 mesh region. To obtain the Laplacian potentials in the configuration region, only the black node points (see Figure 1) will be computed throughout the iteration process, at least until the convergence condition is satisfied. The convergence criteria in this work are denoted as ε, and the stopping condition is subject to u(k+1)u(k)ε. The rest of the residual nodes, or the white node points (see Figure 1(b)), will be calculated independently using a direct technique [16,26,27]. Moreover, the block iterative techniques involve four Laplacian potentials simultaneously per calculation, thus indirectly speeding up the computation. As shown in Figure 1, a set of one-node point and two-node points can be earmarked to measure groups of nodal points that are adjacent to the boundary.

    Figure 1.  Interior mesh nodes for the (a) block (EG) and (b) skewed block (EDG) techniques.

    From Figure 1(a), each and every formulation that practices the EG iterative approaches shall compute a set of four nodes in a group all at once throughout the iteration process (excluding the unique set adjacent to the boundary). Meanwhile, the EDG is essentially derived from a skewed 5-point finite-difference approximation (5-FDA) [16]. As illustrated in Figure 1(b), the configuration domain for the EDG technique is distributed with two different types of nodal points, i.e., the black node () and the white node (). By pairing the identical node, the solutions of each group could be executed independently for each pair. On account of this independency, roughly half of the iterations across the solution domain are to be executed using either type of node, reducing performance time as well as computation complexity. Additionally, to compute the unique sets adjacent to the boundary, the direct method [27] is computed by Eq (1). One of EDG's main advantages is that these approaches diminish the computational complexity by measuring only half of all node points. The FS and half-sweep (HS) iterative approaches' computational stencils are shown in Figure 2, where h is the distance between node points for each direction. These molecules can be displayed in matrix form for FS and HS cases, respectively, as

    2U=1h2[010141010]
    Figure 2.  5-point stencil for (a) FS/standard and (b) HS/skewed FDA.

    and

    2U=12h2[101040101].

    Laplace's equation is a special case of Poisson's equation 2U=f, in which the function f is equal to zero. Hence, a two-dimensional Laplace equation (2D-Le) could be written using Eqs (1) and (2) altogether.

    The system must be discretized using the finite-difference approach in order to solve the 2D-Le before it can be computed effectively using the numerical technique. The simplest second derivative of 5-FDA (also known as an FS iteration) is

    2U(x,y)1h2[u(x+h,y)+u(xh,y)+u(x,y+h)+u(x,yh)4u(x,y)],

    in which U is a function to satisfy Laplace's equation, while u is the potential node at the point (x,y). In addition, the xy plane of the configuration region is rotated by 45 clockwise to provide the approximation that is based on the cross-orientation operator [28,29]. As a result, a skewed 5-point approximation, commonly referred to as the HS iteration, is produced and denoted by

    2U(x,y)12h2[u(xh,yh)+u(x+h,yh)+u(xh,y+h)+u(x+h,y+h)4u(x,y)].

    Using the notation Uij to represent the solution of the potential node for 2D-Le through the mesh grid point (xi,yj), the Laplace equation for the FS case is discretized using a standard 5-point stencil, as follows:

    Ui1,j+Ui+1,j+Ui,j1+Ui,j+14Ui,j=0, (3)

    whereas the discretization of a rotated case via the 5-point stencil is

    Ui1,j1+Ui1,j+1+Ui+1,j1+Ui+1,j+14Ui,j=0. (4)

    A linear system is then generated by applying Eqs (3) and (4) to the Laplace's problem subject to the 2D-Le. Matrix notation can be used to build this linear system, which results in large and sparse matrices, and is represented as

    Au=b, (5)

    in which A represents known coefficients that are expressed in matrix form and b is a known vector, while u is the unknown vector. The expansion of each coefficient for both the FS and HS cases is discussed in depth in [25]. Since the linear system of Eq (5) creates large and sparse matrices, iterative approaches can be used to solve this problem using conventional point or block techniques [16]. In order to elucidate the block technique, the fundamental concept behind the point technique must first be described. Thus, from the second-order central finite difference of Eqs (3) and (4), the point Gauss-Seidel iterative schemes for FS and HS can be reformulated and designated respectively as below.

    U(k+1)i,j=14[U(k+1)i1,j+U(k)i+1,j+U(k+1)i,j1+U(k)i,j+1], (6)
    U(k+1)i,j=14[U(k+1)i1,j1+U(k+1)i+1,j1+U(k)i1,j+1+U(k)i+1,j+1]. (7)

    By integrating a weighted parameter [30], the point SOR iterative technique for FS and HS cases is given respectively as

    U(k+1)i,j=ω4[U(k+1)i1,j+U(k)i+1,j+U(k+1)i,j1+U(k)i,j+1]+(1ω)U(k)i,j, (8)
    U(k+1)i,j=ω4[U(k+1)i1,j1+U(k+1)i+1,j1+U(k)i1,j+1+U(k)i+1,j+1]+(1ω)U(k)i,j. (9)

    Normally, the optimal weighted parameter ω is determined in the range of 1ω<2 [31]. If the parameter value is equal to 1, the approach is essentially simplified to the conventional Gauss-Seidel method. The SOR and Gauss-Seidel methods are actually rather comparable, with the exception that SOR uses a scaling factor to reduce the approximation error.

    The AOR contributes a supplementary relaxation parameter that, in this work is denoted as r, along with the weighted parameter ω from the SOR technique. Both parameters are used to generate iterative schemes that are able to accelerate the convergence rates. AOR is a simple yet powerful scheme, due to the fact that two parameters are involved, for the larger linear system first introduced by Hadjidimos [32]. The point AOR schemes for standard and skewed approaches are shown respectively as

    U(k+1)i,j=r4[U(k+1)i1,jU(k)i1,j+U(k+1)i,j1U(k)i,j1]+ω4[U(k)i1,j+U(k)i+1,j+U(k)i,j1+U(k)i,j+1]+(1ω)U(k)i,j, (10)
    U(k+1)i,j=r4[U(k+1)i1,j1U(k)i1,j1+U(k+1)i+1,j1U(k)i+1,j1]+ω4[U(k)i1,j1+U(k)i+1,j1+U(k)i1,j+1+U(k)i+1,j+1]+(1ω)U(k)i,j. (11)

    According to Hadjidimos, the r value is typically selected to be close to the corresponding ω value. Additionally, he claimed that the uncertain optimal values of these two parameters did not impose any limitations on the ability in to obtain the minimum iterations number.

    The TOR iterative method is basically a generalization of the AOR method with another additional parameter, r. The TOR scheme reduces to the AOR iterative scheme provided the value of r is equal to r [33]. As stated in [33], the additional parameter r could provide faster convergence if an appropriate value is chosen. The point TOR iterative schemes for FS and HS are stated respectively as

    U(k+1)i,j=r4U(k+1)i,j1+r4U(k+1)i1,j+ω4(U(k)i,j+1+U(k)i+1,j)+(ωr4)U(k)i,j1+(ωr4)U(k)i1,j+(1ω)U(k)i,j, (12)
    U(k+1)i,j=r4U(k+1)i+1,j1+r4U(k+1)i1,j2+ω4(U(k)i1,j+1+U(k)i+1,j+1)+(ωr4)U(k)i+1,j1+(ωr4)U(k)i1,j1+(1ω)U(k)i,j. (13)

    Similar to the AOR approach, all of the parameters are in the range of [1,2) and they are chosen to be near to the value of the corresponding SOR parameters [32,34]. In accordance with the 2D-Le, all of the proposed iterative schemes simply substitute each node's value with the average of its four neighbors' values. In this work, the node values that represent the walls, obstacles and target position remain constant.

    As discussed earlier, the block iterative technique obtains four Laplacian potentials per computation. As illustrated in Figure 1(a), all formulations using EG iterative methods compute a group of four nodes at once during the iteration process. Basically, the EG iteration is based on a group of a small number of points, and it is derived using the standard 5- FDA. By analyzing Eq (3) and the point SOR iterative scheme in Eq (8), the block SOR scheme [16,28,31] can be written as

    [4110140110410114][Ui,jUi+1,jUi,j+1Ui+1,j+1]=[S1S2S3S4] (14)

    for

    S1=Ui1,j+Ui,j1,S2=Ui+2,j+Ui+1,j1,S3=Ui1,j+1+Ui,j+2,S4=Ui+2,j+1+Ui+1,j+2.

    For this scheme, it is also possible to generate a linear system of the same form as Eq (5). The scheme can later be translated to the coefficient A after determining the inverse of its matrix, as represented below.

    [Ui,jUi+1,jUi,j+1Ui+1,j+1]=124[6S1+Sa6S2+Sb6S3+Sb6S4+Sa], (15)

    with

    Sa=2(S2+S3)+S1+S4,Sb=2(S1+S4)+S2+S3.

    Based on the point SOR concept using 5-FDA, the block SOR (BSOR) iterative scheme for Eq (15) is now denoted as

    [Ui,jUi+1,jUi,j+1Ui+1,j+1](k+1)=ω24[6S1+Sa6S2+Sb6S3+Sb6S4+Sa]+(1ω)[Ui,jUi+1,jUi,j+1Ui+1,j+1](k). (16)

    Besides, by examining Eq (3) and the point AOR approximation in Eq (10), the formulation of the block AOR method is stated as [35]

    [4100140000410014][Ui,jUi+1,j+1Ui+1,jUi,j+1]=[S1S2S3S4], (17)

    with

    S1=r(U(k+1)i1,jU(k)i1,j+U(k+1)i,j1U(k)i,j1)+ω(U(k)i1,j+U(k)i,j1),S2=r(U(k+1)i+1,j1U(k)i+1,j1)+ω(U(k)i+1,j1+U(k)i+2,j),S3=r(U(k+1)i1,j+1U(k)i1,j+1)+ω(U(k)i1,j+1+U(k)i,j+2),S4=ω(U(k)i+2,j+1+U(k)i+1,j+2).

    Again, Eq (17) may also be transformed into a linear system in the form of Eq (5) and translated as in Eq (15). Hence, the block AOR (BAOR) iterative scheme can also be expressed as Eq (16).

    The general expression of the block TOR (BTOR) iterative scheme, considering Eq (3) and the point TOR iterative method in Eq (12), can be written as

    [Ui,jUi+1,jUi,j+1Ui+1,j+1](k+1)=124[6S1+Sa6S2+Sb6S3+Sb6S4+Sa]+(1ω)[Ui,jUi+1,jUi,j+1Ui+1,j+1](k), (18)

    where

    S1=r(U(k+1)i1,jU(k)i1,j)+r(U(k+1)i,j1U(k)i,j1)+ω(U(k)i1,j+U(k)i,j1),S2=r(U(k+1)i+1,j1U(k)i+1,j1)+ω(U(k)i+1,j1+U(k)i+2,j),S3=r(U(k+1)i1,j+1U(k)i1,j+1)+ω(U(k)i1,j+1+U(k)i,j+2),S4=ω(U(k)i+2,j+1+U(k)i+1,j+2),Sa=2(S2+S3)+S1+S4,Sb=2(S1+S4)+S2+S3.

    As noted in the previous section, the EDG is derived from skewed 5-FDA [16]. The solution domain for the EDG technique is shown in Figure 1(b). From the solution domain, the block EDGSOR formulation may be expressed in matrix form, as Eq (14), with

    S1=Ui1,j1+Ui1,j+1+Ui+1,j1,S2=Ui,j+2+Ui+2,j+2+Ui+2,j, (19)

    and

    S3=Ui,j1+Ui+2,j1+Ui+2,j+1,S4=Ui1,j+Ui1,j+2+Ui+1,j+2. (20)

    Both Eqs (19) and (20) are deduced from the system of linear equations in Eq (14). These equations can be solved independently, as shown below.

    [Ui,jUi+1,j+1]=115[4114][S1S2], (21)
    [Ui+1,jUi,j+1]=115[4114][S3S4]; (22)

    hence, it can be defined as follows (respectively):

    [Ui,jUi+1,j+1](k+1)=ω15[4S1+S2S1+4S2]+(1ω)[Ui,jUi+1,j+1](k), (23)
    [Ui+1,jUi,j+1](k+1)=ω15[4S3+S4S3+4S4]+(1ω)[Ui+1,jUi,j+1](k). (24)

    The algorithm for SBSOR scheme can be enforced by using either Eq (23) or (24). Both equations lead to an equivalent solution.

    Similarly, adopting the skewed AOR formula from Eqs (7) and (11) in the block scheme yields a new formula known as SBAOR [36]. As a result, the SBAOR iteration scheme is described as

    [Ui,jUi+1,j+1](k+1)=r15[4S2S2]+ω15[4S1+S3S1+4S3]+(1ω)[Ui,jUi+1,j+1](k), (25)

    where

    S1=U(k)i1,j1+U(k)i1,j+1+U(k)i+1,j1,S2=U(k+1)i1,j1+U(k+1)i1,j+1+U(k+1)i+1,j1S1,S3=U(k)i,j+2+U(k)i+2,j+2+U(k)i+2,j.

    The same goes for the rotated block TOR, which applies the skewed TOR formula from Eq (7), as well as Eq (13), to the block scheme, and it provides the SBTOR iterative scheme as follows:

    [Ui,jUi+1,j+1](k+1)=r15[4S3S3]+r15[4S4S4]+ω15[4S1+4S2+S5S1+S2+4S5]+(1ω)[Ui,jUi+1,j+1](k), (26)

    in which

    S1=U(k)i1,j1+U(k)i+1,j1,S2=U(k)i1,j+1,S3=U(k+1)i1,j1+U(k+1)i+1,j1S1,S4=U(k+1)i1,j+1S2,S5=U(k)i,j+2+U(k)i+2,j+2+U(k)i+2,j.

    As stated before, there is no generic formula for defining the optimal value for all weighted parameters, which results in a minimum number of iterations. Consistent with [32], the values of r and r are commonly appointed as close to the corresponding ω value. Note that not every parameter value leads to convergence. The optimal values for all three relaxation parameters for each FS and HS case are different. Due to each relaxation value being predetermined before execution, the complexity related to determining the relaxation values for the overall computation is unaffected. Table 1 provides a list of the optimal relaxation values used throughout the experiments. Sensitivity analysis has been used to specify each relaxation parameter's value. The values are the same for every N-size grid, although they differ from other grid sizes and are not significant enough to be revealed; they vary by roughly 0.0001.

    Table 1.  Grid search of relaxation parameter values.
    Methods ω r r
    BSOR 1.82 - -
    BAOR 1.83 1.82 -
    BTOR 1.83 1.86 1.89
    SBSOR 1.81 - -
    SBAOR 1.82 1.84 -
    SBTOR 1.82 1.87 1.88

     | Show Table
    DownLoad: CSV

    Algorithm 1 describes the implementation of the SBTOR scheme in accordance with Eq (26) to solve the two-dimensional Laplace problem.

    Algorithm 1: SBTOR scheme
    i. Setup the configuration region with start and goal positions.
    ii. Initialize the starting point U,ε1015,iteration0.
    iii. Set the variables
    S1U(k)i1,j1+U(k)i+1,j1,S2U(k)i1,j+1,S3U(k+1)i1,j1+U(k+1)i+1,j1S1,S4U(k+1)i1,j+1S2,S5U(k)i,j+2+U(k)i+2,j+2+U(k)i+2,j.
    iv. For each non-occupied black nodes (), calculate
    U(k+1)i,jr154S3+r154S4+ω15[4S1+4S2+S5]+(1ω)U(k)i,j,U(k+1)i+1,j+1r15S3+r15S4+ω15[S1+S2+4S5]+(1ω)U(k)i+1,j+1.
    v. Compute the unique sets adjacent to the boundary via the direct method using Eq (13).
    Then, evaluate the remaining white nodes () by using
    U(k+1)i,j14[U(k+1)i1,j+U(k)i+1,j+U(k+1)i,j1+U(k)i,j+1].
    vi. Encounter the convergence test ε1015, go to (vii). Else, back to (iii).
    vii. Run GDS for path construction.

     | Show Table
    DownLoad: CSV

    Following the aim of this study, an experiment has been conducted by using a robot 2D simulator [37] built by the authors to analyze the competency of the proposed algorithm, i.e., SBTOR scheme, as a tool to solve the path-planning problem. In the block mesh grid domain (see Figure 1) with the execution of the skewed block iterative method, approximately half of the group of node points are computed throughout the iteration process. This will logically result in a substantial reduction of the computational complexity by nearly 50% of the entire operation. The simulations are carried out on four different configuration spaces with five different grid sizes in order to assess the performance of the proposed algorithms. The aforementioned performances were analyzed with three criteria: (i) number of iterations, (ii) CPU time in seconds and (iii) success of path planning. The amount of CPU time and iteration count required by every algorithm is recorded once the tolerance rate for the iteration process is reached. The GDS approach then makes use of the computed Laplacian potentials to guarantee that a path from the starting point to the goal position is successfully constructed.

    In the designated areas, several obstacles with different sizes and shapes are positioned. The Dirichlet boundary condition was utilized in the initial configuration, in which the walls and obstacles are set at the high potential value (the highest temperature). The start position had no initial value specified, whereas the goal location was set at the low potential value (viewed as the lowest temperature). The remainder of mesh grid points were initialized to zero. A machine running a 2.50-GHz processor with 8 GB of RAM has been used to carry out the simulation experiments. Numerical computation was performed iteratively until the stopping condition was met. Provided the Laplacian potential values give no further changes and the difference of current iterations (k) with the next iterations (k+1) is extremely small, e.g., 1.015, the loop should be stopped. This accuracy level was necessary to prevent any point on the surface which has critical point slopes that result in an abrupt absence of descending gradient patterns.

    Each iterative scheme compared in the experiments is shown in Table 2, along with the iteration counts (k) and the execution time (t). Both the block (EG) and skewed block (EDG) schemes evidently show that the TOR iterative approach surpassed its generalization techniques. The computational complexity of the skewed block iterative approach will decrease by approximately half of that relative to the HS process. Since only half of the total nodes in the mesh grid are examined, the skewed schemes surpass the standard block schemes that utilize the FS process. The size of the mesh grid and pattern of the region plays a significant role in producing the outputs of k and t. Obviously, the larger grid size will require more iterations and a longer time to be executed. The improvement ratio between each over-relaxation technique for the smallest grid size (300) to the largest (1500) varies. In terms of the iteration count, on average, the BTOR efficiently decreased by 3-6% compared to BAOR, and 12-20% next to BSOR, whereas SBTOR reduced it by more than 3-8% compared to SBAOR, and roughly 19-32% of SBSOR. However, in terms of execution time, BTOR is between 3-7% lower relative to BAOR, and 14-18% in comparison with BSOR, while SBTOR effectively depreciated approximately 1-4% from SBAOR, and 15-23% in comparison with SBSOR. This study still contributes good results since previous studies [3,13,38] only did the experiments on extremely small grid meshes, i.e., 50 by 50 and 70 by 70 grid. Figures 3 and 4 show the results of Table 2 graphically for the number of iterations and CPU time, respectively. From the graphs, again, the skewed block TOR undoubtedly offers the optimum performance in contrast to its predecessor.

    Table 2.  Number of iterations (k) and execution time (t) in seconds for the proposed iterative schemes. N × N is the size of the grid mesh, e.g., N = 300.
    Techniq-ues N × N
    300 600 900 1200 1500
    k t k t k t k t k t
    BSOR 1258 6.88 5899 163.72 12844 871.66 22227 2694.80 34055 6286.69
    BAOR 1042 6.05 4994 137.87 10928 751.78 19107 2442.66 29306 5551.02
    BTOR 997 5.05 4812 133.00 10581 720.29 18549 2394.62 28445 5404.33
    SBSOR 953 4.34 4495 110.88 9916 603.41 17653 1947.12 27037 3878.32
    SBAOR 766 3.67 3730 95.08 8200 498.19 14338 1599.91 22141 3336.68
    SBTOR 743 3.69 3649 91.03 8038 490.77 14077 1626.65 21676 3303.54
    BSOR 1729 7.67 6782 199.59 14874 1009.48 26007 2827.46 39968 6925.80
    BAOR 1610 8.25 6368 185.36 13953 926.49 24429 3003.98 32926 5909.85
    BTOR 1489 7.64 5957 169.39 13062 867.20 22905 2787.69 31552 5700.19
    SBSOR 1324 4.70 5599 142.64 13252 787.53 22898 2386.46 42248 6170.35
    SBAOR 1188 4.17 4767 122.85 10490 624.48 18371 1930.06 24644 3900.10
    SBTOR 1131 4.63 4561 119.23 10032 603.66 17542 1881.31 24001 3815.88
    BSOR 2666 13.24 11076 315.87 24519 1602.81 42897 5591.93 65977 12331.17
    BAOR 2480 13.83 10389 301.27 22995 1633.35 40322 5261.60 62423 11975.63
    BTOR 2371 11.66 9977 296.46 22111 1883.36 38917 5094.93 59912 10921.11
    SBSOR 2035 8.00 10243 261.74 24864 1529.74 45435 5017.92 70627 10771.29
    SBAOR 1835 7.67 7741 203.61 17131 1072.61 30027 3437.95 46055 7612.81
    SBTOR 1784 6.69 7557 202.95 16707 1251.76 29270 3398.37 45091 7337.51
    BSOR 1629 7.80 6487 187.33 14194 990.20 24913 2979.14 38195 6919.25
    BAOR 1392 7.56 5648 167.65 12367 891.51 21724 2609.11 33518 6139.29
    BTOR 1328 7.10 5428 163.21 11907 850.26 20963 2573.12 34842 6494.66
    SBSOR 1238 4.33 5958 152.73 13501 828.24 23920 2527.82 38175 5948.26
    SBAOR 1116 4.53 4223 109.96 9369 615.38 16313 1725.41 25023 3991.87
    SBTOR 1125 5.02 4106 107.90 9149 605.69 15908 1774.41 26360 4380.35

     | Show Table
    DownLoad: CSV
    Figure 3.  Performance graph in relation to the number of iterations.
    Figure 4.  Performance graph in relation to the execution time.

    The ideology of path-planning flow within this experiment starts with establishing the initial start position and the target location. Next, we identify the optimal parameter according to the corresponding iterative schemes. Once the harmonic potential is generated from the selected algorithm, a complete smooth path is developed through the use of the GDS technique. This impression could describe the iterative scheme that tracks the descending gradient from its starting point to the next consecutive points with lower potentials from previous points, up to the target location (with the lowest potential value). These successful trails can be observed as demonstrated in Figure 5. In this simulation, the starting point is indicated by a green square point, while a red circle point is allocated for the target. As shown in Figure 5, each start location from every region has efficiently accomplished the route by arriving at the target position along while escaping any walls and obstacles in between (if any). The path trajectories can be really fast because they only involve the gradient evaluation of the precomputed Laplacian potential [13]. All four configuration regions in this experiment are relatively simple relative to those in [38].

    Figure 5.  Path construction in an identified stationary environment.

    There are various techniques to solve navigation problems. According to Patle et al. [1], these techniques can be divided into two main groups, i.e., classical approaches and reactive approaches. Due to the lack of artificial intelligence tools at the time, the classical approach was initially quite common as a tool to address robot navigational issues. When a task is performed utilizing a classical approach, it can be noticed that either a result will be acquired, or it will be confirmed that there is no result. This method is less suitable for real-time implementation due to its significant disadvantages, including its high processing cost and inability to respond to the uncertainty existing in the environment. Reactive techniques have surpassed traditional ways in recent years as the preferred method for mobile robot navigation. They are extremely adept at managing the environment's uncertainty. Meanwhile, Sanchez-Ibanez et al. [39] classified path planning into four categories with two major subcategories (see Figure 6). There are characteristics shared by two adjacent subcategories from distinct categories. The schematic also shows how some subcategories lean more toward global planning or local planning than others. From the reviewed papers [1,39], it is safe to say the proposed algorithm in this study has its own edge and drawback. Algorithm 1 has clearly implemented the improved version of potential field approaches. In essence, the goal and obstacles operate as charged surfaces, and the overall potential creates the imaginary forces on the robot. The robot is drawn toward the goal and kept clear of the obstacles by this imaginary force [8]. Later, the robot will travel along the negative gradient to avoid obstacles and arrive at its destination point. To prevent a local minimum problem, this work utilizes the harmonic function [7]. Moreover, Algorithm 1 exhibits superior computing execution when the skewed accelerated relaxation technique is implemented, as it performs much faster to obtain the solution of Laplace's equation to solve the path-planning problem.

    Figure 6.  Classification of existing path-planning techniques [39].

    The iterative approaches investigated in this work are from over-relaxation families including the implementation of the block and skewed block schemes to enhance the execution performance as well as reduce the execution time. The introduction of accelerated weighted parameters to respective skewed nodes further improves the overall performance of the newly proposed SBTOR iterative schemes, yielding promising results. The experiments evidently demonstrate that the solution to the robot path-planning problem is feasibly solved by using numerical approaches owing to advanced algorithms and the accessibility of fast machines these days. From the table of results, the SBTOR scheme was found to significantly outperform its predecessor's techniques in terms of both the number of iterations and the time taken. The computational performance is unaffected by increasing the number of obstacles; in fact, it will complete more quickly because the calculation disregards the regions engaged by the obstacles. It can be said that the larger the area of obstacles taken, the less the required calculations and storage, or, in other words, the obstacles reduce computational complexity. Once again, the authors would like to highlight that the SBTOR outperformed the SBAOR (approximately by 3-8%) and SBSOR (by about 19-32%) in terms of the iteration count, while the SBTOR saved roughly 1-4% over SBAOR and 15-23% over SBSOR in terms of the processing time. The application of the skewed block TOR scheme families in robot pathfinding and in Algorithm 1 is another aspect of this study's originality or novelty. The authors intend to investigate the red and black strategy [23,24,40,41,42] by applying the proposed approaches in future work. It is believed that the aforementioned strategy will enhance and refine the overall computation.

    This research was financially supported by Universiti Pertahanan Nasional Malaysia. The authors also acknowledge support from Science Foundation Ireland (SFI) under Grant Number SFI/16/RC/3918 (Confirm), and the Marie Skłodowska-Curie grant agreement No. 847577, co-funded by the European Regional Development Fund.

    The researchers declare that there is no conflict of interest regarding the publication of this study.



    [1] B. Patle, B. Ganesh, A. Pandey, D. Parhi, A. Jagadeesh, A review: On path planning strategies for navigation of mobile robot, Def. Technol., 15 (2019), 582–606. https://doi.org/10.1016/j.dt.2019.04.011 doi: 10.1016/j.dt.2019.04.011
    [2] N. Buniyamin, N. Sariff, W. Wan Ngah, Z. Mohamad, Robot global path planning overview and a variation of ant colony system algorithm, International Journal of Mathematics and Computers in Simulation, 5 (2011), 9–16.
    [3] S. Sasaki, A practical computational technique for mobile robot navigation, Proceedings of the IEEE International Conference on Control Applications, 1998, 1323–1327. https://doi.org/10.1109/CCA.1998.721675 doi: 10.1109/CCA.1998.721675
    [4] H. Chou, P. Kuo, J. Liu, Numerical streamline path planning based on log-space harmonic potential function: a simulation study, Proceedings of IEEE International Conference on Real-time Computing and Robotics, 2017,535–542. https://doi.org/10.1109/RCAR.2017.8311918 doi: 10.1109/RCAR.2017.8311918
    [5] Y. An, S. Wang, C. Xu, L. Xie, 3D path planning of quadrotor aerial robots using numerical optimization, Proceedings of the 32nd Chinese Control Conference, 2013, 2305–2310.
    [6] M. Silva, W. Silva, R. Romero, Performance analysis of path planning techniques based on potential fields, Proceedings of Latin American Robotics Symposium and Intelligent Robotics Meeting, 2010,115–119. https://doi.org/10.1109/LARS.2010.33 doi: 10.1109/LARS.2010.33
    [7] C. Connolly, R. Gruppen, The applications of harmonic functions to robotics, Journal of Robotic Systems, 10 (1993), 931–946. https://doi.org/10.1002/rob.4620100704 doi: 10.1002/rob.4620100704
    [8] O. Khatib, Real-time obstacle avoidance for manipulators and mobile robots, Proceedings of IEEE International Conference on Robotics and Automation, 1985,500–505. https://doi.org/10.1109/ROBOT.1985.1087247 doi: 10.1109/ROBOT.1985.1087247
    [9] H. Ghassemi, S. Panahi, A. Kohansal, Solving the Laplace's equation by the FDM and BEM using mixed boundary conditions, American Journal of Applied Mathematics and Statistics, 4 (2016), 37–42. https://doi.org/10.12691/ajams-4-2-2 doi: 10.12691/ajams-4-2-2
    [10] A. Dahalan, A. Saudi, Pathfinding algorithm based on rotated block AOR technique in structured environment, AIMS Mathematics, 7 (2022), 11529–11550. https://doi.org/10.3934/math.2022643 doi: 10.3934/math.2022643
    [11] K. Al-Khaled, Numerical solutions of the Laplace's equation, Appl. Math. Comput., 170 (2005), 1271–1283. https://doi.org/10.1016/j.amc.2005.01.018 doi: 10.1016/j.amc.2005.01.018
    [12] K. Shivaram, H. Jyothi, Finite element approach for numerical integration over family of eight node linear quadrilateral element for solving Laplace equation, Mater. Today: Proc., 46 (2021), 4336–4340. https://doi.org/10.1016/j.matpr.2021.03.437 doi: 10.1016/j.matpr.2021.03.437
    [13] C. Connolly, J. Burns, R. Weiss, Path planning using Laplace's equation, Proceedings of IEEE International Conference on Robotics and Automation, 1990, 2102–2106. https://doi.org/10.1109/ROBOT.1990.126315 doi: 10.1109/ROBOT.1990.126315
    [14] J. Barraquand, B. Langlois, J. Latombe, Numerical potential field techniques for robot path planning, IEEE T. Syst. Man. Cy., 22 (1992), 224–241. https://doi.org/10.1109/21.148426 doi: 10.1109/21.148426
    [15] M. Karnik, B. Dasgupta, V. Eswaran, A comparative study of Dirichlet and Neumann conditions for path planning through harmonic functions, In: Lecture notes in computer science, Berlin: Springer, 2002,442–451. https://doi.org/10.1007/3-540-46080-2_46
    [16] A. Abdullah, The four point explicit decoupled group (EDG) method: a fast Poison solver, Int. J. Comp. Math., 38 (1991), 61–70. https://doi.org/10.1080/00207169108803958 doi: 10.1080/00207169108803958
    [17] J. Sulaiman, M. Hassan, M. Othman, The half-sweep iterative alternating decomposition explicit (HSIADE) method for diffusion equation, In: Computational and information science, Berlin: Springer, 2004, 57–63. https://doi.org/10.1007/978-3-540-30497-5_10
    [18] A. Saudi, J. Sulaiman, Block iterative method for robot path planning, Proceedings of the 2nd Seminar on Engineering and Technology, 2009, 1–4.
    [19] A. Saudi, J. Sulaiman, Half-Sweep Gauss-Seidel (HSGS) iterative method for robot path planning, Proceedings of the 3rd International Conference on Informatics and Technology, 2009, 34–39.
    [20] A. Saudi, J. Sulaiman, Robot path planning using Laplacian Behaviour-Based Control (LBBC) via half-sweep SOR, Proceedings of the International Conference on Technological Advances in Electrical, Electronics and Computer Engineering, 2013,424–429. https://doi.org/10.1109/TAEECE.2013.6557312 doi: 10.1109/TAEECE.2013.6557312
    [21] S. Matsui, H. Nagahara, R. Taniguchi, Half-sweep imaging for depth from defocus, Image Vision Comput., 32 (2014), 954–964. https://doi.org/10.1016/j.imavis.2014.09.001 doi: 10.1016/j.imavis.2014.09.001
    [22] J. Chew, J. Sulaiman, A. Sunarto, Solving one-dimensional porous medium equation using unconditionally stable half-sweep finite difference and SOR method, Mathematics and Statistics, 9 (2021), 166–171. https://doi.org/10.13189/ms.2021.090211 doi: 10.13189/ms.2021.090211
    [23] F. Musli, J. Sulaiman, A. Saudi, Numerical simulations of agent navigation via half-sweep modified two-parameter over-relaxation (HSMTOR), Journal of Physics Conference Series, 1988 (2021), 012035. https://doi.org/10.1088/1742-6596/1988/1/012035 doi: 10.1088/1742-6596/1988/1/012035
    [24] A. Saudi, A. Dahalan, An efficient red-black skewed modified accelerated arithmetic mean iterative method for solving two-dimensional Poisson equation, Symmetry, 14 (2022), 993. https://doi.org/10.3390/sym14050993 doi: 10.3390/sym14050993
    [25] A. Dahalan, A. Saudi, Pathfinding algorithm based on rotated block AOR technique in structured environment, AIMS Mathematics, 7 (2022), 11529–11550. https://doi.org/10.3934/math.2022643 doi: 10.3934/math.2022643
    [26] M. Othman, A. Abdullah, An efficient four points modified explicit group Poisson solver, Int. J. Comput. Math., 76 (2000), 203–217. https://doi.org/10.1080/00207160008805020 doi: 10.1080/00207160008805020
    [27] D. Evans, Group explicit iterative methods for solving large linear systems, Int. J. Comput. Math., 17 (1985), 81–108. https://doi.org/10.1080/00207168508803452 doi: 10.1080/00207168508803452
    [28] G. Dahlquist, Å. Björck, Numerical Methods, New Jersey: Prentice Hall, 1974.
    [29] W. Yousif, D. Evans, Explicit group over-relaxation methods for solving elliptic partial differential equations, Math. Comput. Simulat., 28 (1986), 453–466. https://doi.org/10.1016/0378-4754(86)90040-6 doi: 10.1016/0378-4754(86)90040-6
    [30] D. Young, Iterative methods for solving partial difference equations of elliptic type, T. Am. Math. Soc., 76 (1954), 92–111. https://doi.org/10.2307/1990745 doi: 10.2307/1990745
    [31] D. Young, Iterative methods for solving partial difference equations of elliptic type, Ph. D. Thesis, Harvard University, 1950.
    [32] A. Hadjidimos, Accelerated overrelaxation method, Math. Comput., 32 (1978), 149–157. https://doi.org/10.2307/20006264 doi: 10.2307/20006264
    [33] J. Kuang, J. Ji, A survey of AOR and TOR methods, J. Comput. Appl. Math., 24 (1988), 3–12. https://doi.org/10.1016/0377-0427(88)90340-8 doi: 10.1016/0377-0427(88)90340-8
    [34] N. Ali, F. Pin, Modified explicit group AOR methods in the solution of elliptic equations, Applied Mathematical Sciences, 6 (2012), 2465–2480.
    [35] M. Martins, W. Yousif, D. Evans, Explicit group AOR method for solving elliptic partial differential equations, Neural, Parallel and Scientific Computations, 10 (2002), 411–422.
    [36] N. Ali, L. Chong, Group accelerated OverRelaxation methods on rotated grid, Appl. Math. Comput., 191 (2007), 533–542. https://doi.org/10.1016/j.amc.2007.02.131 doi: 10.1016/j.amc.2007.02.131
    [37] A. Saudi, Robot path planning using family of SOR iterative methods with Laplacian behavior-based control, Ph. D. Thesis, University Malaysia Sabah, 2015.
    [38] S. Chen, Collision-free path planning, Ph. D. Thesis, Iowa State University, 1997. https://doi.org/10.31274/rtd-180813-10480
    [39] J. Sanchez-Ibanez, C. Perez-del-Pulgar, A. Garcia-Cerezo, A path planning for autonomous mobile robots: a review, Sensors, 21 (2021), 7898. https://doi.org/10.3390/s21237898 doi: 10.3390/s21237898
    [40] T. Iwashita, M. Shimasaki, Block red-black ordering: a new ordering strategy for parallelization of ICCG method, Int. J. Parallel Prog., 31 (2003), 55–75. https://doi.org/10.1023/A:1021738303840 doi: 10.1023/A:1021738303840
    [41] N. Saad, A. Sunarto, A. Saudi, Accelerated red-black strategy for image composition using Laplacian operator, IJCDS, 10 (2021), 1085–1095. https://doi.org/10.12785/ijcds/100198 doi: 10.12785/ijcds/100198
    [42] A. Dahalan, A. Saudi, J. Sulaiman, Pathfinding for mobile robot navigation by exerting the quarter-sweep modified accelerated overrelaxation (QSMAOR) iterative approach via the Laplacian operator, Mod. Simul. Eng., 2022 (2022), 9388146. https://doi.org/10.1155/2022/9388146 doi: 10.1155/2022/9388146
  • This article has been cited by:

    1. A’Qilah Ahmad Dahalan, Azali Saudi, Jumat Sulaiman, Enhancing Autonomous Guided Vehicles with Red-Black TOR Iterative Method, 2023, 11, 2227-7390, 4393, 10.3390/math11204393
  • Reader Comments
  • © 2023 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(1743) PDF downloads(100) Cited by(1)

Figures and Tables

Figures(6)  /  Tables(2)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog