3.2. Manipulator Path Planning
The input to the path planning procedure is a model of the processed surface in the form of a triangular mesh obtained at the scanning stage and the output is a program in AS language.
When the working tool of a robotic manipulator moves along the surface being processed, the intersection point of the working tool axis with the surface will move along a curve on the surface, which we will call the trace of the working tool on the surface. Of course, for a given surface, the manipulator path, i.e., a given curve in space along which the center of the working tool moves with the spatial orientation of the working tool specified at each point of this path (the direction of the axis of the working tool) uniquely determines the trace of the path on the surface. We assume that the orientation of the working tool is directed normal to the surface at the point of intersection of the axis of the working tool and the surface, and that there is a certain rule for calculating the distance on the axis of the working tool from the point of intersection of the axis with the surface to the center of the working tool (in the simplest case, this distance is given fixed distance from the nozzle to the surface being processed), then the trace of the working tool uniquely determines the trajectory of the working tool. Thus, the problem of path planning is reduced to the problem of determining a set of curves (in the general case) on the surface corresponding to the trace of the working tool on the surface being processed.
Thus, the path planning procedure for the manipulator can be divided into three stages:
1) constructing a trace of the manipulator path on the surface;
2) constructing a set of points lying on the path of the manipulator's working tool (thus, the trajectory of the working tool is specified by a broken line);
3) transformation of the geometrically specified path of the manipulator's working tool into the manipulator program in AS language.
The main research is focused on the first stage since the second and third stages are quite simple. The description of some subtle issues related to the representation of the trajectory of a working tool in the form of a sequence of geometric primitives (circular arcs and straight segments) was discussed earlier in the paper [
23].
The proposed method for constructing the manipulator's trajectory trace involves a three-step sequence:
1) defining a starting curve on the surface;
2) calculating a distance-to-line function defined on the surface. For any point (p) on the surface represents the geodesic distance from p to ;
3) constructing isolines , on the surface for a given parameter d, where k is an integer index. These isolines form a set of curves on the surface. Each curve is considered a trace of the manipulator's working tool trajectory on the surface.
Further, we will focus on the task of constructing the geodesic distance field, defined by the distance function
. This particular task belongs to a broad class of problems involving the computation of geodesic paths and distances on a polyhedral surface. It is a well-developed area of applied mathematics that continues to attract researchers' attention (particularly due to its practical significance). The problems of computing geodesic distances are divided into two classes: computing geodesic distances from a single point (Single Source Geodesic Distance SSGD) or from multiple points, which can be geometric objects such as curves or flat figures on the surface (Multiple Source Geodesic Distance MSGD). Many methods for computing geodesic distances reduce the problem to solving partial differential equations on a smooth manifold. When working with surfaces represented by meshes, the equation is discretized and the finite element method (FEM) or mesh method is used to solve it. These methods are applied to both SSGD and MSGD problems, and often the same method can be applied to both classes of problems by differently setting the boundary conditions for one differential equation. Methods that reduce the problem to solving differential equations (PDE-based methods) are divided into two groups: Wavefront-based methods and Diffusion-based methods [
34].
The most straightforward approach to computing the geodesic distance function is solving the eikonal equation (3) for given boundary conditions:
Most methods in the first group implement this approach. The eikonal equation (3) can serve as a formal definition of the geodesic distance function and has a clear physical meaning (if we consider the special case of the eikonal equation (3) as an equation specifically for the distance function). However, it is a nonlinear hyperbolic equation, which is extremely difficult to solve directly in most cases.
Methods in the second group operate on parabolic-type equations, offering significant computational advantages and enabling the creation of efficient software implementations. The problem with these methods is that they calculate, at best, only an approximation of the geodesic distance function. A breakthrough in the field of geodesic distance function computation methods based on numerical solutions of differential equations was the emergence of the heat flow method [
35]. With sufficiently high computational efficiency, this method allows (subject to certain requirements for the quality of mesh triangulation) the calculation of the distance function with reasonable accuracy.
When attempting to apply geodesic distance function computation methods based on solving differential equations to our task, a specific problem arises. This lies in defining the boundary conditions on the boundary of the domain. In general, at least for methods like heat flow [
35], this problem is solvable but requires the use of certain artificial techniques. In this paper, we propose a new method for calculating the geodesic distance function on a surface, specifically designed to compute the distance from a curve defined in the plane, free from such problems. The method is based on the energy minimization procedure of a vector field proposed in the paper [
36]. The application of discrete exterior calculus, suggested by the authors of the paper [
36], for the computational implementation of their method, offers, in our opinion, a significant advantage of our proposed method over traditional PDE-based distance function computation methods in terms of implementation efficiency. Before proceeding to the algorithm description, we will discuss the main mathematical ideas and constructions underlying the proposed method.
Most methods for finding the geodesic distance function involve solving partial differential equations on smooth manifolds. We propose a method for calculating the geodesic distance function based on computing a vector field on the surface whose integral lines coincide with the isolines of the function This interpretation of the problem has a clear motivation. We can interpret the geometric procedure of constructing the isoline = l as a physical process. Let us imagine we have two small pucks A and B, connected by an inextensible string of length l. We move puck A along the starting curve γ, while moving puck B across the surface such that two conditions are met: firstly, the string must always be taught and lie on the surface. If we associate the pucks with two points A and B on the surface, we can say that the string lies on the segment of the geodesic line AB and the geodesic distance between points A and B is maintained during the motion. Secondly, the direction of the geodesic line AB at point A is orthogonal to the tangent of γ at point A. The trajectory of point B on the surface will be a curve, which we will denote as pl. In general, the curve pl may have self-intersections and even intersect with the curve γ may have self-intersections and even intersect with the curve pl may not be smooth everywhere. However, we can assume that for a segment of the curve γ a parameter value lmax > 0 is defined, such that if l ∈ [0, lmax], then pl will represent a smooth curve without self-intersections, on which a smooth one-dimensional tangent vector field is defined. By varying the value of the parameter l, we obtain a smooth tangent vector field U, defined in a region of the surface adjacent to the starting curve γ.
Figure 7 schematically illustrates the kinematic procedure for constructing the isolines of the vector field function
and the corresponding tangent vector field
U.
For the simplest case – when the starting curve
is given in the plane, we can easily obtain an analytical expression for the corresponding field
. Let us assume that puck
moves along the curve
with a constant unit speed, i.e.
. The system of two pucks connected by a string can be considered as a rigid body. If we denote
as the angular velocity vector of this rigid body and denote
, where
and
are the radius vectors of points A and B, respectively, then for any time moment (t) the equation (4) holds:
Since
, we obtain equation (5):
It is well known that the magnitude and direction of the vector
have a simple geometric interpretation, namely (6):
where
is the curvature of
at the point
, and
R(
s) is the radius of the circle tangent to the curve
at the point
. The vector
is orthogonal to the tangent vector
. It can be said that the vector
points towards the center of the tangent circle
. The point
O will also be the instantaneous center of rotation of our "rigid body" and will lie on the same line with the points
and
. The point
O will also be the instantaneous center of rotation of our "rigid body". Let us define a scalar function
on the curve
such that
, and the sign of
is determined by the relative position of the center of the tangent circle
O and the point
With this definition of "signed curvature"
from (7) follows the equation (8) for
:
The corresponding vector field
for the coordinates
will be analytically given by formula (9):
3.2.1. Isometric Flows of Tangent Vector Fields
In fact, the kinematic procedure described above for constructing the tangent field on the surface defines a function . The concept of a flow of a vector field will play a key role in the analysis of the vector fields . Formally, the flow of a tangent vector field on a smooth manifold is defined as a one-parameter group of mappings of the manifold onto itself.
Definition 1. Let
is a tangent vector field defined on a smooth manifold
M. The vector field defines a parameterized set of mapping
. For a given value of the parameter
, the action of the mapping
on a point
is defined as
, where
is the solution of the differential equation (10) with the initial condition
In courses on the theory of ODEs (ordinary differential equations) it is proven that, under certain conditions imposed on the smoothness of the field (which we will assume to be fulfilled), the solution to equation (10) exists and is unique. The curves γ are called integral curves of the vector field . Just as linear transformations of the plane, the mappings form a group with respect to the composition operation. In particular, it is easy to show that and . The one-parameter group of mappings is called the flow of the tangent vector field on the manifold .
This definition will be required later for constructing analytical procedures. Still, to show the main ideas on which the proposed method is based, it is more convenient to use physical interpretations of the concept of the flow of a vector field. First, if we use the hydrodynamic interpretation of the concept of a flow of a vector field, we will immediately see the following fact. The flow of the vector field has a property invariant to the choice of the starting curve and the distribution – if we place an elastic filament in the flow so that at any point, the line of the filament is orthogonal to the streamlines of the field, the length of the filament will not change with time as it moves in the flow. Moreover, in some cases, the flow of the tangent vector field can be isometric. Let us imagine that we place a small piece of a thin film made of elastic material on a surface where a tangent vector field at the point p of the surface over which point q is located at that moment in time. In general, the piece of film will move in such a flow, rotating and deforming as it moves. But if the vector field U on the surface is such that no matter where on the surface the piece of film begins its movement, it will then move along the surface without deformation, the flow of such a field will be isometric. Let us give a simple example. Let the surface be a segment of a sphere of radius , centered at point . As the starting curve we choose the arc of the circle formed by the intersection of the surface , passing through the point with the sphere. Let us call the line passing through the center of the sphere and perpendicular to the surface . If we construct the level line of the function of the geodesic distance , moving the puck along the arc at a unit speed, then this is equivalent to the fact that we would rotate the sphere around the axis with an angular velocity , while keeping the pucks and so that they remain motionless, and the filament stretched. It is clear that the corresponding vector field on the segment of the sphere coincides with the field of velocities of the points of the sphere arising when the sphere rotates around the axis with an angular velocity . Obviously, the flow of such a vector field will be isometric. Tangent vector fields whose flows are isometric are called Killing vector fields. The given example is a classic example of a Killing vector field on the surface of a sphere.
The converse statement can be readily proven. Let be a Killing vector field on a smooth two-dimensional surface. Denote as an arbitrary integral line of this field. Then any integral line of the field will be an isoline of the geodesic distance function . For a special case, characterized by the fulfillment of two conditions formulated below, this statement allows us to analytically formulate the problem for a tangential vector field such that if is its solution, then the integral lines of the field will be isolines of the geodesic distance function .
- 1)
On the surface we are working with, Killing vector fields exist. It should be noted that Killing vector fields do not exist on all surfaces.
- 2)
As the initial curve we choose an integral line of some Killing field on the surface.
Subsequently, we will extend this analytical problem to the general case of a smooth two-dimensional surface by introducing the concept of an approximate Killing vector field. For this, we will need some definitions from the field of Riemannian geometry.
3.2.2. Two-Dimensional Surfaces Embedded in Three-Dimensional Space as Riemannian Manifolds. Analytical Definition of Killing Fields
Henceforth, we will consider a smooth two-dimensional surface embedded in three-dimensional Euclidean space as a smooth Riemannian manifold with the metric induced by the Euclidean metric in . This last statement implies the following definition of the metric on the manifold . Let be an arbitrary point on the surface, be a chart of the manifold , such that and we choose some local coordinate system in the region .
If we choose a Cartesian coordinate system in the three-dimensional space into which our surface is embedded, with origin at an arbitrary point
, then we define a local parameterization of the surface in the neighborhood of the point
of the form
, where
is the radius vector of the point
in this coordinate system. In this case, the vectors
, defined by the equalities
and
are invariant under shifts of the point
and form a basis in the surface tangent to the surface at the point
. The Gram matrix of the basis
will be the matrix of the metric tensor
at the point
in the local coordinates
(11):
Note: If we denote as the values at the point of the coefficients of the first fundamental form on this surface, calculated in local coordinates , then the equalities , , will hold, and we can say that the induced metric is given by the first fundamental form on the surface. At any point the metric defines a scalar product in the tangent space . When a metric is given on a smooth manifold, the concept of an isometric flow of a vector field and, along with it the concept of a Killing vector field can be defined as follows:
Definition 2. A tangent vector field
on a smooth Riemannian manifold
M will be a Killing vector field if and only if for any point
and any two tangent vectors
the relation (12) holds:
where
is the flow of the vector field
for
for some small
, and
d is the differential of the mapping.
From Definition 2 it is easy to obtain equation (13), which is often considered as the definition of a Killing vector field
on a smooth Riemannian manifold:
where
is the Lie derivative.
3.2.3. The Laplace-de Rham Operator. The Relation Between Killing and Harmonic Fields on 2-Dimensional Surfaces
Using the intuitive physical interpretation of the concept of an isometric flow of a tangent vector field, it is easy to formulate two necessary conditions that a Killing vector field on a 2-dimensional surface embedded in three-dimensional space must satisfy:
- 1)
the geodesic curvature of any integral line of the field must be constant;
- 2)
on any curve , which is an integral line of the condition must hold.
Therefore, the vector field
, constructed using the above-described kinematic procedure, can only be a Killing vector field if it satisfies condition (14) and the geodesic curvature of
is constant:
Let us now try to reverse our perspective on condition (14). We will consider it as a condition imposed on the vector field in an analytically formulated problem. Note that condition (14) uniquely defines a Killing vector field on the surface (if there exists a Killing vector field on the surface such that γ is its integral line). This fact is intuitively obvious if we use the physical interpretation of the concept of an isometric flow of a vector field. If we consider the surface as a smooth Riemannian manifold, then a stronger statement holds: A Killing field is determined uniquely by a vector at some point and its gradient (i.e. all covariant derivatives of the field at the point). A number of important insights for constructing an analytical procedure for constructing vector fields associated with the geodesic distance function are provided by considering vector fields
on the surface. Earlier we saw that on the surface, the field
U is given by the distribution of curvature
along the initial curve
. Consider the simple but important case where the curvature of
is constant. On the surface there are two types of curves with constant curvature – straight lines and arcs of circles. Two types of initial curves – straight lines and arcs of circles correspond to two types of vector fields on the surface. Straight lines correspond to constant vector fields
, analytically defined by formula (15):
where
is the unit vector (the direction vector of the straight line
). If
is an arc of a circle with center at point
, then the field
will coincide with the velocity field of points of the rotating disk
. Let's choose the center of the Cartesian coordinate system at point
and imagine that the axis of the disk rotating with angular velocity
, where
is the constant curvature of the arc of the circle
is projected onto this point. In such a coordinate system, the components
,
of the field
, are given by formulas (16):
It is evident, that if the flow of the field
on the surface is isometric, then the field must be divergence-free, i.e.
. As for the curl of the field, as it is easy to check
, and for vector fields
on the surface whose flows are isometric, we can formulate the requirement of the constant curl of the field
∀
p , where
is a constant vector perpendicular to the plane under consideration. It follows that any Killing vector field on the plane will belong to the kernel of the Hodge operator (sometimes also called the Laplace de Rham operator), defined on
by equation (17):
As can be easily seen, if
is a harmonic field on the surface, i.e., conditions (18) hold for
then
belongs to the kernel of the Laplace-de Rham operator, i.e., equality (18) follows from the fulfillment of conditions (19):
Let us note that the fields
are both harmonic and Killing. The fields
, obviously, are not harmonic, but as will be shown below, they are closely related to harmonic fields. Let us remind that an arbitrary harmonic vector field is uniquely defined by its restriction on the arbitrary segment of its flow line. Let
be an integral curve of a harmonic vector field
on the surface, and let
be the natural parameter of the curve
, so
. If the function
is given on an arbitrary interval
, then this function can be used to reconstruct the vector field
over the entire plane. In other words, the harmonic vector field is fully determined by its restriction to a segment of the streamline. Accordingly, if the starting curve
is a segment of the integral line of some harmonic vector field
, then the vector field
is uniquely determined by condition (14). Let the condition
be given on the arc
of a circle of radius
centered at the point
for the harmonic vector field
. In a Cartesian coordinate system, centered at the point
, the components
and
of the harmonic vector field
are determined by equations (20):
The integral lines of this field will be arcs of circles centered at the point
. Thus, if
is a Killing vector field, and
is a harmonic vector field on the surface, given by the condition
, where
is a curve of constant curvature, then the integral lines of the fields
and
coincide. Let's transfer these results to the general case of curved two-dimensional surfaces. The Laplace-de Rham operator on Riemannian manifolds is defined by formula (21) [
37]:
where
is the Ricchi operator, and
is the rough Laplacian (or connection Laplacian), defined by the formula. On a Riemannian manifold
the Ricci operator
is defined as (22), where
is the Ricci tensor. The space of smooth tangent vector fields on
we will denote as
.
It is known that if
is a Killing vector field on a Riemannian smooth manifold
and
is a harmonic vector field on
, then
[
38].
Let us pose the following Problem 1: find a tangent vector field on a two-dimensional Riemannian manifold such that and , where is a smooth curve on .
As follows from the above, either this problem will have no solution, or the set of solutions
will consist of vector fields of the form (23):
where
is a Killing vector field defined by the condition
,
is a harmonic vector field defined by the condition
, and
is a parameter. The integral lines of the fields
and
coincide and represent the level curves of the geodesic distance function
. Due to this, for any value of the parameter
the integral lines of the vector field
will also be level lines of the function
, and we can consider the set
as a set of vector fields whose integral lines are level lines of the function
. Let us consider now the case of an arbitrary starting curve
on the surface, without assuming that the curvature of
is constant. We will assume that
is parameterized by the natural parameter
. Assume that on some interval
the condition (24) holds:
where
is a given constant, and
, where
is a small parameter. Then, in the surface region
the field
will be approximately Killing. Analytically, this means that for any
the magnitude of
is small. It should be noted that the known problem is approximating a curve with two ends on the plane of a piecewise smooth curve consisting of a finite number of smooth segments, each of which is either a straight-line segment or a circular arc. This issue is often formulated as the problem of the decomposition of a planar curve into arcs and line segments. Problems of such decomposition, in a certain sense, optimal decomposition, arise in digital image processing, approximation of curves by splines of certain types, and in robotics and several methods of such decomposition are currently known [
39,
40,
41]. If the seed curve on the plane allows approximation by a piecewise smooth curve, which is a sequence of straight-line segments and circular arc segments, then the vector field U corresponding to this starting curve will be "approximate Killing vector field" in the above sense. Assuming that we can draw an analogy between the above-considered case of a curve on a plane and the general case of a two-dimensional surface embedded in a three-dimensional Euclidean space, we consider the segmentation of a two-dimensional surface, in which the surface is divided into "geometrically homogeneous regions", that within each such region of homogeneity, the values of Gaussian and average curvature of the surface are approximately constant. Segmentation of this kind has been described previously in articles [
39,
40,
41]. If a geodesic line is chosen as the starting curve on the surface, then within each region of homogeneity, the vector field U determined by the starting curve will be the approximate Killing vector field, in the sense of the relative smallness of the value |◻U|. As the results of numerical experiments with both scanning data and synthetic data show, in the vast majority of cases when applying the segmentation method described above, most of the surface area is covered by such regions of homogeneity. However, the theoretical question is: what conditions are imposed on the surface the vector fields that correspond to the starting geodesic curve can be considered "approximate Killing vector fields" remains open, and the above reasoning is based on empirically obtained data in favor of the validity of the hypothesis that such vector fields are approximately Killing vector fields.
3.2.4. The Energy Functional of the Vector Field
As it is known, such constructions of classical differential analysis of vector fields as the gradient of a scalar function, divergence, and curl of a vector field, have their analogues and generalizations in the calculus of smooth differential forms on smooth manifolds, sometimes called exterior calculus. In many cases, operating with dual forms of the first order (covector fields) to vector fields is much easier than directly with vector fields. In our case, the application of the language of k-forms will have, as will be seen from the following, great advantages. Below, we present basic information about the relationship between vector fields and covector fields and introduce the notation to be used in the further exposition.
The metric defines a canonical isomorphism on the manifold (which is usually called the musical isomorphism) between the tangent bundle
and the cotangent bundle
, given by two mappings (25):
For any point
and any vector
the covector
is defined so that the condition (26) is satisfied:
Correspondingly, at an arbitrary point
the inverse transformation
for an arbitrary covector
is defined so that the condition (27) is satisfied:
We will denote the spaces of smooth k-forms on the manifold as . In this case, as is known, is the space of smooth scalar functions on (the space of smooth mappings ), and is the space of smooth covector fields on (i.e., each element will be an element of and vice versa, each element of will be an element of ).
The de-Rham Laplacian operator
, which we considered above, is the operator dual to the Hodge Laplacian, acting in the space
dual to
(28):
The Hodge-Laplacian operator
is defined as (29):
where
is the exterior derivative operator, and
is the so-called codifferential. In those cases where the operator
acts on
-1 forms of the space
, the formula (30) holds:
Thus, the codifferential can be considered as a generalization of the divergence operator. The exterior derivative operator
can be considered as a generalization of the curl of a vector field, since the formula (31) holds:
where
is the Hodge star operator. A harmonic form is a form
, satisfying the conditions (32):
Of course, with such a definition, the harmonic form will be the form dual to the harmonic vector field. The space of harmonic forms on a manifold gives the kernel of the Hodge–Laplace operator (for any harmonic form ).
The authors of the paper [
35] defined the energy functional of the vector field as a mapping
, the action of which on the
k-1 - form (covector field)
is determined by the formula (33):
It is evident that if
is a harmonic form, then
. The authors of the paper [
35] arrived at the construction of the functional (33), by considering the bilinear form on covector fields
defined as
. If
is an arbitrary tangent vector field on
, then the magnitude
can be considered as a quantitative criterion for the dissimilarity of the vector field
from the harmonic field, interpreting the term
is the integral divergence of the field
, and the term
as the integral curl of the field
. The task of minimizing the functional
under given constraints on the vector field
can be interpreted as the problem of finding a vector field that, while satisfying the imposed conditions, is as close as possible to the harmonic field. Further, we assume that the vector field
is subject to only one constraint of the form
, where
is a geodesic line on the surface. If, for a given surface and a given initial curve
there exists a set of solutions to task 1
, then by virtue of (28)
. This fact and the considerations presented in the previous section lead to Proposition 1:
Proposition 1. If the covector field minimizes the functional , defined by formula (31), under the condition , then the integral lines of the vector field will coincide with the isolines of the function .
Since we do not have a proof of Proposition 1 for the general case of arbitrary surfaces in three-dimensional space, Proposition 1 in the form stated above should be considered a hypothesis. The proofs of this proposition for some specific cases that we have obtained are not rigorous, but the intuitively convincing considerations that led to the formulation of this hypothesis and the results of numerical experiments make the assumption of the validity of this hypothesis highly probable in our opinion. Until a proof (or disproof) of the validity of this hypothesis is obtained, we consider the numerical method for calculating the geodesic distance function based on this hypothesis as a heuristic method that provides good practical results and is convenient to use.
3.2.5. Remarks on the Numerical Implementation of the Method
The proposed method for constructing the geodesic distance function is based on a computational point of view on the numerical methods for minimizing the energy functional of a vector field proposed by Fisher et al. [
36]. The authors of the article [
36] proposed a new approach to the problem of designing tangent vector fields with user-defined constraints based on the application of Discrete Exterior calculus methods. Previously proposed approaches to the vector field design problem by various authors were based on the general idea of interpolating vectors specified in individual mesh nodes representing the surface. All these approaches rely on explicit coordinate frames and vectors represented through coefficients in these frames, be either 2D or 3D. The parallel transport of tangent vectors between these coordinate frames generally makes vector field optimization a nonlinear problem. In contrast, in article [
36], the vector field design problem was formulated as a linear problem by using an intrinsic, coordinate-free approach based on discrete differential forms and the associated Discrete Exterior Calculus. The use of the Discrete Exterior Calculus technique, in which covector fields are represented by scalar quantities on the edges of the mesh and operators acting on differential forms by matrices, allows us to construct highly efficient algorithms, in the software implementation of which it is convenient to use advanced libraries of numerical methods of linear algebra. A detailed presentation of numerical methods for minimizing the energy functional of a vector field is given in the article [
36]; here, in the following presentation, the focus is on aspects specific to the problem of constructing the geodesic distance function.
The algorithm for constructing the distance function consists of a sequence of three procedures:
A) constructing a geodesic line on the surface, which we will refer to as the initial curve ;
B) calculating the vector field , which minimizes the energy functional of the vector field (33) under the condition (14);
C) reconstructing the distance function from the vector field , whose integral lines approximately represent the isolines of the function .
We work with a discrete representation of the surface using a triangular mesh. The triangular mesh is programmatically represented as a linear list of vertices (each vertex corresponds to a data structure representing a tuple of three elements—the coordinates of the vertex in space) and a so-called triangle list. Each triangle corresponds to one of the faces of the polyhedron that describes the surface. Each triangle is represented as a tuple of three elements—the indices of the triangle's vertices in the vertex list. Naturally, the vertex and triangle list indirectly define the set of edges of the polyhedron representing the surface. We will denote the vertices as
, where
is the index of the vertex in the list. The edge connecting the vertices
and
will be denoted as
, where the order of the indices indicates the orientation of the edge. The triangles (mesh faces) will be denoted as
, where we will assume that the order of the vertex indices determines the orientation of the triangle. The covector
, dual to a given tangent vector field
on the surface, is represented by its values on the edges
, the values
theoretically represent the integral of
along the edge
. Step
B) is implemented as a program module that takes as input a triangular mesh and a list
of tuples
, specifying the predetermined values
. The list
provides a discrete representation of the boundary condition (1). In the described program implementation of the method for constructing the geodesic distance function on a surface, the user selects two points
and
on the boundary of the region
. Then, a discrete representation of the geodesic line connecting points
and
is constructed on the mesh. The representation of the geodesic line as a path through the faces of the surface is then converted into the list
and fed into the program module that implements step
. The code for this module implements numerical methods for minimizing the energy functional of the vector field, as detailed in the article [
36]. To reconstruct the tangent vector field on the surface from the calculated discrete covector field, the Whitney finite elements method is applied. The module provides an interface that allows obtaining the coordinates of the tangent vector to the surface at a given point, based on the barycentric coordinates of the point on a specified face of the mesh. The first stage of the procedure for reconstructing the geodesic distance function
step
) is the construction of the tangent vector field
, which is an approximation of the gradient field
. The construction of the vector field
is carried out in three steps:
- 1)
in the center of each face (triangle of the mesh), the vector is calculated;
- 2)
each vector is rotated by an angle of in the surface of the face. The choice of turning direction (clockwise or counterclockwise) is indifferent. We will interpret this step as the computation of the discrete field , where is a linear rotation operator by ;
- 3)
for each face, a vector is calculated.
Thus, we have a standard problem in numerical methods – reconstructing a scalar function defined on a surface from the gradient field of the function. We use a standard approach. The problem is formulated as finding the minimum of the functional
with respect to
which is an approximation of the function
, that we are seeking.
As is known, the Euler-Lagrange equation for this extremal problem has the form (35) [
42]:
The previously computed discrete tangential vector defines a Dirichlet boundary condition on the boundary . However, as practice has shown, it is better to solve the problem with mixed boundary conditions separately for the two regions into which the initial curve divides the original region . In this case, for both regions, the initial curve γ will be part of the boundary where the Cauchy boundary condition is satisfied.
To test the software implementation of the described method for constructing the geodesic distance function from a curve, we used both synthetic data (triangular meshes calculated for analytically defined surfaces) and data obtained from scanning.
Software implementation of the third stage (the translation of the sequence of geometric primitive into the sequence of AS language commands) is straightforward, and will not be discussed here; it was previously described in articles [
23,
24,
25].
3.3. Experimental Application of the Developed Control Algorithms
The developed method of automatic manipulator program generation was tested and the robotic MPS of a wear-resistant Cr-based coating was performed on the worn parts of the jaw's crushing plate for crushing mineral raw materials. A robotic scanning of the crushing plate was carried out, and then a robotic MPS of a protective coating on the worn-out sections of the plate was performed. A visual representation of the sequence of the process of robotic MPS of the powder coating on a crushing plate, including preliminary 3D scanning, is given in
Figure 8.
Specifically, for scanning the crushing plate (
Figure 8 a), a scanning step of 30 mm was used, the speed of movement of the end effector was 100 mms
-1, the distance between the scanning plane and the plane of the crushing plate was on average 70 mm, the distance sensor was polled 30 times per second , while the total scanning time was 2 minutes for a scanning working area of 600 cm
2. The crushing plate model was visualized (
Figure 8b). Although this is not the purpose or even a mandatory step of scanning, this was done solely for clarity of comparison with the real object. The scanning accuracy was good, new 3D model reconstruction algorithms made it possible to obtain a model with the same distances determined with an accuracy of 0.5 mm, namely 100 mm between the centers of the ribbed projections and a projection height of 30 mm. During the study, different scanning speeds were tested: from 10 cms-
1 to 1 ms
-1, depending on the complexity of the object geometry; however, we came to the conclusion that it is reasonable to reduce the speed of the distance sensor on the working segments of the trajectory to ensure a set of more points for statistical processing.
The MPS was carried out by moving the robot arm with the microplasmatron installed on it in accordance with the obtained 3D model of the crushing plate. The spraying distance was maintained equal to 100 mm (
Table 3). As previously noted in
Section 2.2, the spraying parameter was chosen based on the requirements for the structure of the coating.
As shown in
Figure 8 a, the substrate surface is first scanned using a distance sensor attached to the robotic arm. A robotic arm with a sensor moves in a plane along a U-shaped trajectory with a constant modulus speed.
Figure 8a shows how, by measuring the Z coordinates from the surface of the plate, 3D scanning is carried out, based on the results of which a 3D model of the substrate is reconstructed (
Figure 8 b). In practice, the 3D model is not visualized since it is a set of point clouds, that is, three-dimensional coordinates that remain in the robot's memory.
Then the distance sensor on the robotic arm is replaced with a microplasmatron (
Figure 8 c) and the robot's movement is generated along a 3D trajectory (that is, along three-dimensional coordinates, but with the addition of the spraying distance H to Z-coordinate). As a result, the microplasmatron moves at a given spraying distance H from the surface of the plate with strict adherence to the linear speed of movement V, while maintaining the perpendicularity of the plasma jet to the substrate. When the robot reaches the edge of the plate, it moves horizontally by a given distance - the spraying step τ, and turns, moving at the same linear speed in the opposite direction, ensuring that the spraying tracks overlap by approximately one-third. As a result, a coating of uniform thickness is sprayed (
Figure 8 d).
Ideally, a robot with a microplasmatron could itself move to the area of the worn surface and perform spraying, but this task has not yet been solved, since additional analysis of the image of the plate surface was required to determine the boundaries of the wear area. The method implemented here did not use a video camera, did not perform image analysis, and used only distance sensor data. Therefore, in practice, the manipulator was manually brought to the boundaries of the worn area, where the coordinates were measured. Then a code for moving the robot was generated, stopping the MPS process when the worn area was completely coated. This was a pilot experiment, during which robotic MPS of a protective coating was performed on worn areas of the surface of a crusher plate. In contrast, the movement of a robotic arm with a microplasmatron was automatically generated along the reconstructed 3D model of the product with precision maintaining a given linear speed and distance to the surface.
Production tests of the crushing plate were carried out under conditions of grinding polymetallic ores of various hardness with continuous operation of the crusher for at least 30 minutes with a loading from 70 % to 100% of the height of the crushing chamber. The movable plate of the jaw crusher, coated after its main resource had been depleted, was additionally operated for 6 months, being subjected to the optimal load for 5-6 hours per shift. Thus, the plate's service life was extended with the same crusher performance. A production test certificate was received (Ust-Kamenogorsk, Kazakhstan), and it was concluded that the service life of the movable plate of the jaw crusher restored by robotic plasma spraying of the Cr - based protective coating was increased by 15% compared to the plates not subjected to restoration.
Thus, the practical application of the new approach to control the robot manipulator performing microplasma spraying of a protective coating on the surface of the jaw crusher plate was successful. The novelty of the proposed approach lies in the use of the technique of constructing the geodesic distance function from the starting curve, which provides two advantages in comparison with other known methods. Firstly, the proposed method is insensitive to the choice of the starting geodesic curve, whereas in existing seed curve methods, the starting curve is often chosen as the geodesic max height. Thus, the new method offers the prospect of fully automatic trajectory generation. Secondly, this technique provides better accuracy, which allows to operate with trajectories with a small step between the working segments of the trajectory, which improves compliance with technological parameters and, accordingly, the quality of the coating.