You are on page 1of 25

Computational and Applied Mathematics (2022) 41:124

https://doi.org/10.1007/s40314-022-01831-4

Radial basis function-based vector field algorithm for wildfire


boundary tracking with UAVs

Licheng Feng1 · Jay Katupitiya1

Received: 10 October 2021 / Revised: 8 January 2022 / Accepted: 11 March 2022 /


Published online: 1 April 2022
© The Author(s) 2022

Abstract
This paper tackles the problem of dynamic wildfire boundary tracking with UAVs. Wildfire
boundary is treated as the zero-level set curve of an implicit function and is approximated
with radial basis functions. Its propagation is modeled with the Hamilton–Jacobi equation
with an arbitrary initial boundary as the input. To navigate UAVs to the wildfire boundary,
an analytical velocity vector field, whose integral curves converge to the wildfire boundary,
is constructed on the basis of the typical radial basis function thin-plate spline. Computer
simulations with a single UAV and multiple UAVs have been conducted for the evaluation
of the proposed solution, and numerical results show that the proposed algorithm can ensure
the successful tracking of an arbitrarily shaped wildfire boundary.

Keywords Wildfire boundary tracking · Level set curve · Hamilton–Jacobi equation ·


Radial basis functions · Vector field

Mathematics Subject Classification 93C85 · 93C95

1 Introduction

Wildfire, also known as forest fire or bushfire, is an area of combustible vegetation that has
caught fire and requires significant amounts of resources for its suppression. Wildfires are
becoming more and more prevalent throughout the world and cause considerable ecological
and economical loss to the wildland environment (Ambroz et al. 2019). In the study of
combating wildfires, one main concern is the live wildfire boundary monitoring, but it is
risky and costly to adopt manned aerial wildfire monitoring, especially when dealing with
uncontrolled fires. Given these situations, it is essential to adopt automated approaches to

Communicated by Antonio José Silva Neto.

B Licheng Feng
licheng.feng@student.unsw.edu.au
1 School of Mechanical and Manufacturing Engineering, University of New South Wales, High St
Kensington, Sydney, NSW 2052, Australia

123
124 Page 2 of 25 L. Feng, J. Katupitiya

monitor wildfire frontiers, which can help improve significantly the wildfire suppression
efficiency, risk management, and policy implementation.
Unmanned aerial vehicles (UAVs), due to the ease of deployment, low maintenance cost,
high mobility, and flexibility, have gained considerable attention in numerous military and
civilian applications, such as ground traffic surveillance (Kanistras et al. 2013), geological
surveying (Adams and Friedland 2011), reconnaissance, and tracking (Obermeyer et al.
2010), to name a few. In the context of forestry, autonomous UAVs have been proven to be
an effective tool for studying environmental features and surveying hazardous phenomena-
prone places (Tang and Shao 2015). A survey in Yuan et al. (2015) analyzes the sensing
hardware and algorithms for wildfire detection. The work in Twidwell et al. (2016) presents
a comprehensive analysis of the usage of UAVs in fire management operations. Survey in
Bailon-Ruiz and Lacroix (2020) reviews the state of the art on wildfire remote sensing with
UAVs from the perspective of autonomy.
In recent years, UAV-based monitoring has become a prominent study area and plays
a more integral role in providing situational awareness (SA) (Stanton et al. 2001), which
refers to perception of elements of interest in the environment within a volume of time and
space, and the comprehension of their influences on the evolution of the event state over
time. Although a great number of monitoring algorithms (Bertozzi et al. 2005; Susca et al.
2008; Cassandras et al. 2011; Lan and Schwager 2013; Smith et al. 2011; Smith and Rus
2010) have emerged, fire frontier monitoring is a largely under investigated domain. The
focus of this paper is to develop a strategy for a UAV to track the boundary of an ongoing
wildfire event, with the aim of providing wildfire fighters with the up-to-date information of
active wildfire frontiers. The major components of this paper include using an easy algorithm
to reconstruct the wildfire boundary based on initial information from sensing instruments,
such as airborne sensors, and to efficiently solve the wildfire propagation equation. Based
on the reconstructed wildfire boundary, develop a velocity vector field for UAV guidance
control. In line with the main research points, the main novelties of this paper include: (1)
The investigation into applying radial basis functions to construct a level set function from
sensor data to describe the wildfire boundary, (2) the application of radial basis functions to
estimate the propagated wildfire boundary modeled by Hamilton–Jacobi equation, and (3)
the design of an analytical velocity vector field from radial basis functions to guide the UAV
to track the wildfire boundary.
To the best knowledge of the authors, there are no other publications presenting such
strategies to track dynamic wildfire boundaries. The remainder of the paper is organized as
follows. Section 2 surveys the previous literature and establishes the current state of similar
research. Section 3 describes the problem statement, assumptions, and models used in this
work. Section 4 describes the strategies for wildfire boundary construction and evolution
and presents a guidance methodology. Section 5 is devoted to the presentation of simulation
results collected to assess the effectiveness and efficiency of the proposed algorithm. Section
6 closes the paper with a discussion of the results and future work.

2 Related work

In this section, we review current state of similar research. The dynamic nature of wild-
fire propagation poses challenges to keep a continuous and exhaustive surveillance of the
advancing fire boundary. To effectively track the dynamic wildfire boundary, it is imperative
to understand the mechanism of wildfire growth, and accurately predict its propagation. In

123
Radial basis function-based vector... Page 3 of 25 124

the existing literature, a great deal of methods have been developed to estimate the propaga-
tion of wildfire boundaries, such as Huygen’s principle (Richards 1990), Lagrangian (direct)
approach (Balažovjech and Mikula 2011; Dziuk 1999; Hou et al. 1994), and the Eulerian
(level set) approach (Ambroz et al. 2019; Osher and Fedkiw 2006; Sethian 1999). However,
when it comes to the issue of tracking wildfire boundary, the propagation of wildfire bound-
ary is seldom considered. Only the work Kumar et al. (2011) encloses the points on the fire
front with an elliptical envelop model and deployed UAVs to persistently covering the fire
front under the UAV footprints. This paper concludes that the fire fronts follow the Huygen’s
principle to propagate in an elliptical fashion when assuming a uniform fuel distribution,
uniform landscape and weather, and constant wind direction. However, the initial shapes of
wildfire boundaries are not constricted to an elliptical envelop. Therefore, the model pre-
sented in Kumar et al. (2011) is not suitable for an initial fire front with an arbitrary shape.
The level set approach, devised by Osher and Sethian (1988), is highly robust and accurate
to solve the problem of curve evolution under complex conditions. If we can approximate
the initial wildfire boundary with a highly accurate level set function, the level set algorithm
is a good choice for computing the advancing wildfire boundary.
The trajectory control problem, defined as making a vehicle follow a pre-determined
course in space, is essential for safe and efficient operation of UAVs. The problem of track-
ing wildfire boundary requires that the UAV should reach the wildfire frontier and then follow
the wildfire boundary as accurately as possible. In the past decades, a number of tracking
algorithms (Bertozzi et al. 2005; Susca et al. 2008; Cassandras et al. 2011; Lan and Schwager
2013; Smith et al. 2011; Smith and Rus 2010) have emerged, whereas the wildfire boundary
tracking is a largely unexplored problem. In the literature, boundary tracking is also named
as level set tracking (Matveev et al. 2012), curve tracking (Malisoff et al. 2017), and iso-
line tracking (Dong et al. 2020). In these studies, the environment is always mapped with
a scalar concentration function to represent any spatially varying variable, such as density
of combustible material, temperature, etc. UAVs are required to track the constant level sets
(boundary) determining the extent of the spatial field. These problems have been widely stud-
ied in the application of environmental exploration, for instance, tracking an oil or chemical
spill in the ocean (Li et al. 2014a; Fahad et al. 2015; Jiang and Li 2018; Wang et al. 2019) and
tracking ash cloud (Dong and You 2021). Roughly speaking, existing control algorithms for
boundary tracking can be categorized as the gradient-based approach and the gradient-free
approach, depending on whether the gradient information is available or not.
If the target boundary can be described as a geometric curve rather than a function of
time, its tracking can be handled by path following algorithms. There are already many
existing path following approaches, and some of them can be applied to deal with track-
ing dynamic boundaries. Among different path-following algorithms, vector field method
is shown to achieve high path-following accuracy while requiring low control efforts (Sujit
et al. 2014). As a typical gradient-based method, the gradient vector field algorithm decouples
the guidance law into convergence and circulation two terms and weighs them by scalars to
influence field strength and direction (Wilhelm and Clem 2019). The resulted vector field
has the feature that its integral curve approaches the target path asymptotically, and it can
ensure convergence of any trajectory that does not reach the “critical” locations where the
vector filed is degenerate (Goncalves et al. 2010; Kapitanyuk et al. 2017). The study of vector
field is extensive. The earliest application of vector field approach to the problem of isoline
tracking can be found in work (Marthaler and Bertozzi 2003; Triandaf and Schwartz 2005),
where UAVs are controlled to track a certain level of an explicitly formulated isoline. Work
(Wilhelm and Clem 2019; Lim et al. 2014; Nelson et al. 2007) developed plane vector field
path-following control laws for straight-line paths and circular arcs and orbits. To navigate a

123
124 Page 4 of 25 L. Feng, J. Katupitiya

single mobile robot along a general n-dimensional smooth curve, the work Goncalves et al.
(2009) provided a universal algorithm to compute an artificial vector field and extended it in
Goncalves et al. (2010), where the implicit functions that define the target curve were used
as the convergence term in the designed guidance law. Inspired by the idea of constructing a
general vector field, work (Kapitanyuk et al. 2017; De Marina et al. 2017) designed vector
field-based controllers to drive the UAV to track closed smooth curves described by explicit
equations. In the application of vector field algorithms, the mathematical expression of the
desired curve is essential for the construction of the vector field controller. In aforementioned
studies, they all assumed that the explicit representation of the target curve is available by
default. In gradient-based algorithms, the knowledge of the spatial field gradient is highly
desirable. If the gradient information is not explicitly available, some work estimates gradi-
ent and their spatial divergence for designing guidance laws. Work (Li et al. 2014a; Fahad
et al. 2015; Jiang and Li 2018; Wang et al. 2019) studied the problem of tracking a dynamic
ocean plume whose dispersion is modeled by an advection–diffusion equation. As the tracked
dynamic concentration-level curves have a real physical meaning, robots are assumed to be
equipped with onboard sensors like fluorometer to obtain the measurement of the chemi-
cal concentration. By data processing, the estimated gradient of the concentration and the
corresponding divergence of the concentration gradient are available. Thus, a control law
only depending on concentration measurements was successfully constructed for tracking
the chemicals poured in a marine environment.
If it is impossible to obtain the gradient information directly or by estimation, some
studies turn to gradient-free algorithms. “Bang-Bang” approach is one of the most common
and straightforward gradient-free methods and has been adopted to solve some boundary
tracking problems (Jin and Bertozzi 2007; Nelson et al. 2007), however, the Bang-Bang type
control switches between alternate steering angles if current measurement is above or below a
predefined threshold, and results in unavoidable zigzagging behavior in the tracking process.
Some work employed the sliding mode control technique to the tracking problem. In Matveev
et al. (2012), the UAV equipped with a onboard sensor was limited to measure the field value
of the environmental level set at its current position and was driven to track the static spatial
level set with a sliding mode controller. A suboptimal sliding mode algorithm, designed with
the density measurement of the contaminant from an onboard sensor of the autonomous
vehicle, was proposed in Menon et al. (2014) to steer the single autonomous vehicle along
the smooth contour of the static spatial contaminant phenomena. The work (Dong et al.
2020; Dong and You 2021) also assumed that the concentration measurement at the current
location is in a point-wise fashion (Matveev et al. 2012) with onboard sensors, and presented
a gradient-free controller in a PI-like form for a Dubins vehicle to track a desired isoline of an
unknown scalar field using only the concentration feedback. It validated the effectiveness of
the controller via simulations using a fixed-wing UAV on the real dataset of the environmental
pollution level. For above-mentioned level curve tracking problems solved with gradient-free
algorithms, the concentration value of the level curve has a physical meaning and must be
available at the UAV’s current position with an onboard sensor.
Regardless of gradient-based or gradient-free controllers, most above-mentioned studies
focus on tracking static level curves with a given explicit expression, rather than time-varying
curves. As for variable curve tracking based on an advection–diffusion pollution dispersion
model, it should be noticed that the concentration is meaningful and the UAV must measure
the chemical concentration value at its current position for implementing the guidance law.
In this paper, the wildfire boundary is modeled with level set function and its propagation
is described by the Hamilton–Jacobi equation; nevertheless, the “concentration” value at the
UAV’s position is unavailable as we do not have sensors that can measure a meaningful

123
Radial basis function-based vector... Page 5 of 25 124

“concentration”. The only UAV data available are its current position and velocity measured
by its onboard sensors. As for the tracked wildfire boundary, only the positions of way-
points on the wildfire boundary are available. To apply the idea of vector field algorithm
for wildfire boundary tracking, it is necessary to deal with the scattered waypoints on the
wildfire boundary with an effective and efficient approximation algorithm, thus obtaining
an explicit level set function describing the wildfire boundary. What is required then is the
construction of the vector field using the approximated level set function. In this paper, we
consider tracking a time-varying wildfire boundary whose dependence on time cannot be
neglected. Different from existing studies which approximate the wildfire boundary with an
elliptical envelope, we apply a meshless numerical interpolation approach to reconstruct the
wildfire boundary, regardless of the shape of the boundary. Also, the introduced meshless
algorithm is incorporated with the Hamilton–Jacobi equation to exploit the propagation of
wildfire boundary. In addition, to keep the integral of the velocity vector field converges to
the desired curve, it will be derived from the approximated level set function by obeying the
principle of gradient-based methods.

3 Problem formulation

Generally speaking, the main problem studied in this paper is to establish a control law for the
UAV, such that it converges to, and follows the boundary of wildfires, in a specified direction
such as clockwise or counterclockwise. In this section, first, the kinematic model of the UAV
used in this paper is presented, and then, the details of the wildfire boundary reconstruction
and propagation are provided. The following is a formal statement of the control problem.

3.1 UAV model

In the existing literature, it is assumed that the dynamic controller has a quick and precise
response, thus considering a simpler kinematic model and dealing with the “outer” kinematic
control. The kinematic model of the UAV used in this paper is described by the Dubins car
model in 2D plane
ẋr = νr cosθr
ẏr = νr sinθr
θ̇r = ωr , (1)
where x r = [xr , yr ]T and θr represent the global Cartesian coordinates and the heading
angle of the UAV, and vr and ωr are the corresponding translational and rotational velocities,
respectively.
The position x r and the orientation θr in dynamics (1) cannot be simultaneously stabilized
by a continuous static (time-invariant feedback) (Brockett 1983). To avoid this problem, for
a positive constant l0 , a new variable x is defined
x = [x, y]T = [xr + l0 cosθr , yr + l0 sinθr ]T , (2)
as the output of the system and the control objective is set based on the position of that output
(Sahin et al. 2007). Using feedback linearization, the kinematic model (1) can be transformed
into the single-integrator model (Li et al. 2014b)
ẋ = u, (3)

123
124 Page 6 of 25 L. Feng, J. Katupitiya

where the new control variable u = [u 1 , u 2 ]T is defined as


  
cosθr −l0 sinθr νr
u= (4)
sinθr l0 cosθr ωr .
 
cosθr −l0 sinθr
Note that the determinant of matrix is always l0 no matter the value
sinθr l0 cosθr
of
 the heading angle θr ; therefore, the matrix’s inverse always exists and it is given by
cosθr sinθr
− sin θr cosθr . After coordinate transformation, designed control input u can be directly
l0 − l0
applied to the integrator model (Sahin et al. 2007; Jiang et al. 2020). In the following content,
we only consider the single-integrator model (3).

3.2 Radial basis function

Wildfire boundary can be regarded as a series of scattered points (t). With the feature
of level set algorithm, fire boundary is implicitly defined as a zero level set of a smooth
time-dependent function (x, t) : R2 → R, namely
(t) = {x ∈ R2 |(x, t) = 0}. (5)
The initialization of  usually uses signed distance function, given by

⎨ − d(t) (x) x inside the fire boundary

(x, 0) = 0 x on the fire boundary , (6)


d(t) (x) x outside the fire boundary
where d(t) (x) is the distance between x and its nearest point on the wildfire boundary (Zhai
et al. 2020).
Radial basis functions (RBFs) are a set of real-valued axial symmetric functions, namely
the value only relies on the distance to the origin. To build a geometric model from multivariate
scattered data, the radial basis function approach is a popular and effective tool due to its
ease of implementation with a simple form and good approximation behavior. There are
many types of popular radial-based functions and thin-plate splines are used in this paper to
construct the level set function for wildfire boundary.
The thin-plate spline can be written as
gi (x) = x − x i 2 log( x − x i  +1), (7)
where the x i = [xi , yi ∈
]T R2
are referred to the centers of radial basis function,  · 
represents the Euclidean norm operator. The points on wildfire boundary can be approximated
by a N number of thin-plate splines, centered at N fixed centers, which can be written as
N
(x, t) = λi (t)gi (x) + p(x, t), (8)
i=1

where coefficients λi (t) are real numbers, p(x, t) is a first-order polynomial, determined by
the selected RBF (Koushki et al. 2020) and changed with time to account for the linear and
constant portion of (x, t) and to ensure positive definiteness of the solution (Morse et al.
2005).
The polynomial p(x, t) is not required for positive definite RBFs; however, it is necessary
for semi-positive RBFs to guarantee singularity (Buhmann 2003). The adopted thin-plate

123
Radial basis function-based vector... Page 7 of 25 124

spline RBF is semi-positive; therefore, we take the polynomial part as p(x, t) = c1 (t) +
c2 (t)x + c3 (t)y for the 2D problem (Xie and Mirmehdi 2007; Wei et al. 2018). To ensure a
unique solution of RBF interpolation of the level set function, the expansion coefficients in
Eq. (8) must be subjected to the following orthogonality conditions (Morse et al. 2005):
N N N
λi (t) = λi (t)xi = λi (t)yi = 0. (9)
i=1 i=1 i=1

The function (8) combined with the constraints in (9) can be rewritten in a matrix form
(Xie and Mirmehdi 2007; Wei et al. 2018)

Gα = f , (10)
⎡ ⎤ ⎡ ⎤
  g1 (x 1 ) . . . g N (x 1 ) 1 x1 y1
⎢ .. ⎥, P = ⎢ .. .. .. ⎥, α =
, A = ⎣ ... ..
A P
where G = . . ⎦ ⎣. . . ⎦
P T 03×3
g1 (x N ) . . . g N (x N ) 1 x N yN
 T  T
λ1 . . . λ N c1 c2 c3 and f = (x 1 , t) . . . (x N , t) 0 0 0 .

3.3 Wildfire model

Existing wildfire propagation models can be summarized into physical models, empirical
models, and quasi-empirical models (Sullivan 2009a, b; Sullivan and Andrew 2009). Wildfires
involve complex combustion chemistry, heat transfer, and fluid dynamics, making many
data is unmeasurable for physical models. Empirical models derive statistical correlations
between observations and the wildfire, while they are restricted by the lack of physical basis
in different environments. Quasi-empirical models use some forms of physical framework
for statistical modeling and they often include simplified equations with less number of
parameters (Karouni et al. 2014). Miyanishi (2001) proposed a semi-empirical fire spread
model focusing on front velocities at the head, the rear, the flanks, and elsewhere between
them (Fig. 1). The fire spread velocities primarily depend on the wind velocity.
In this paper, homogeneous topography, vegetation, and meteorology are assumed in the
propagation of wildfire boundary. Without loss of generality, a simplified Fedell’s model
(Mallet et al. 2009) which is easier to tune, either via direct trials or with systemic methods
for parameter estimation, is adopted in this paper. The rate of spread (RoS) at a point x on
the fire boundary is given by
⎧ 
⎪ π
⎨ 0 + a νw cosnu ϕ if |ϕ| ≤
U (x, t) = 2
⎪ π , (11)
⎩ 0 [α + (1 − α)|sinϕ|] if |ϕ| >
2
where νw is the wind velocity, ϕ is the angle (measured counterclockwise) between the wind
direction and the outward normal to the wildfire boundary at point x, 0 and a are determined
by fuel, and α ∈ [0, 1] is the ratio between RoS at the flanks (ϕ = π2 ) and that at the rear
(ϕ = π). In this paper, we assume that the parameters of RoS are known.
In this work, level set algorithm is implemented to describe the propagation of wildfire
boundary. Although (x, t) has no physical meaning, the evolution of its zero level set reveals
the propagation of wildfire boundary. Assuming that the wildfire boundary only propagates
in its normal directions with RoS U (x, t); therefore, the evolution of function (x, t) is

123
124 Page 8 of 25 L. Feng, J. Katupitiya

Fig. 1 A schematic of wildfire propagation during the time interval t under a wind of magnitude vw in the
direction ϕ = 0

governed by the Hamilton–Jacobi equation (Alessandri et al. 2021)


∂(x, t)
+ U (x, t)  ∇(x, t) = 0, (12)
∂t
where ∇(x) = [ ∂(x) ∂(x) T
∂ x , ∂ y ] is the gradient of the level set function. Although Eq. (12)
can be solved with the conventional finite difference methods proposed in Peng et al. (1999),
Mallet et al. (2009), the parameterized level set function based on radial basis functions Wei
et al. (2018) is adopted here to obtain smooth wildfire boundaries. Therefore, the original
time-dependent Hamilton–Jacobi partial differential equation (12) is discretized into a system
of coupled ODEs governing the motion of the dynamics interfaces, which can be regarded
as a collocation formulation of the general method of lines (Wei et al. 2018)

G + B(α, t) = 0, (13)
dt
 T
where B(α, t) = U (x 1 , t)|∇ g(x 1 )α| . . . U (x 1 , t)|∇ g(x N )α| 0 0 0 and g(x) =
[g1 (x) . . . g N (x) 1 x y]T . The system (13) is coupled nonlinear ODEs which can be
solved by Euler methods. Thus, it is more computationally efficient to evolve boundaries by
updating a set of parameterized coefficients than finite difference methods.

3.4 Control objective

We assume that the velocity of UAV is much faster than that of an advancing wildfire boundary
and the information of wildfire boundary is available to each UAV. All deployed UAVs can
evaluate their positions and speeds and have information on nearby UAVs. According to the
definition of the level set function, it can be regarded as the tracking error. We define the notion
of error distance as |(x, t)|. The goal is to design a controller for the UAV to eliminate the
tracking error limt→∞ |(x, t)| → 0, bringing thus the UAV to the wildfire boundary. Upon
reaching the boundary of wildfire, the UAV should patrol along the boundary. The control
objective is formally stated as follows.

123
Radial basis function-based vector... Page 9 of 25 124

Problem: For the dynamic wildfire boundary, whose propagation is modeled by the
Hamilton–Jacobi equation (12), design a control law to drive the UAV, which is subject to the
kinematic constraint (3), to track the propagated wildfire boundary {x ∈ R2 |(x, t) = 0}
with a desired velocity νd , eliminating the tracking error limt→∞ |(x, t)| → 0.

4 Trajectory tracking control

In this section, we implement the principle of vector field technique to design a guidance law
for the single-integrator modeled UAV to track the wildfire boundary.
The accurate expression of the wildfire boundary is vital for designing the vector field-
based controller. Thus, we first use the meshless thin-plate radial basis function to approximate
the wildfire boundary with an explicit function before vector field construction. To avoid the
meaningless trivial solution of Eq. (10), additional points outside and inside the closed wildfire
boundary are appended. A common practice to extend points (t) is generating off-surface
points in their normalized external and internal normal directions (Carr et al. 2001), which
are defined as

x i+ = x i + δni , x i− = x i − δni , i = 1, . . . , n, (14)

where x i+ and x i− are off-surface points in external and internal normal directions, respec-
tively. ni is the normalized normal vector at point x i on the wildfire boundary. δ is a small
step size and its value can be selected follow the rule in Zhu and Wathen (2015), Cuomo
et al. (2017) .To make the reconstructed boundary is relatively insensitively to the projection
distance mentioned in Eq. (6), care must be taken when projecting off-surface points x i+ and
x i− along the normals to ensure that they do not intersect other parts of the surface. Thus, the
closest points to these new constructed points are the corresponding base points generated
them (Carr et al. 2001). The set of points

˜ =  ∪  + ∪  − , (15)

with  + = {x i+ |x i ∈ (t)},  − = {x i− |x i ∈ (t)} is the extended dataset for reconstructing


wildfire boundary. The value of extended dataset is given according to Eq. (6).
Applying the new dataset as the input to Eq. (10), the resulting zero level set of its solution
is the approximation of the wildfire boundary. After obtaining the approximation level set
function of wildfire boundary, the Hamilton–Jacobi equation will be solved with a RBF-
based parameterized level set method (Wei et al. 2018) to estimate the propagated wildfire
boundary.
According to the problem definition, the studied tracking problem can be decomposed into
two subproblems: (i) convergence to the wildfire boundary and (ii) patrol along the wildfire
boundary. Each subproblem will be solved separately by designing corresponding vector
fields. Then, the composition of individual vector field for two subproblems will ensure both
convergence and patrol. The total vector field can be represented in the form as

u = V con + V pat , (16)

where V con produces vectors that converge to the path and V pat produces vectors that patrol
along the path.
The wildfire boundary is the zero level set, and the value (x, t) can be regarded as a signed
distance function from the multirotor to the desired path, or the tracking error (Kapitanyuk
et al. 2017; Kapitanyuk and Chepinsky 2013). Inspired by this property of level set function,

123
124 Page 10 of 25 L. Feng, J. Katupitiya

we define a Lyapunov function V p as


1
Vp = (x, t)2 , (17)
2
which achieves a minimum at the boundary of the wildfire along the gradient descent direction
−∇V p . Therefore, the vector field for convergence is modeled as
∇(x, t)
V con = −kn · (x, t) , (18)
 ∇(x, t) 
where kn is a scalar weight, the normal vector ∇(x, t) is given by
N  
∂  x − xi 
= c2 + λi (x − xi ) 2log( x − x i  +1) +
∂x  x − x i  +1
i=1
N  . (19)
∂  x − xi 
= c3 + λi (y − yi ) 2log( x − x i  +1) +
∂y  x − x i  +1
i=1

When the UAV reaches the boundary of wildfire, the convergent item becomes zero and
the UAV is required to patrol along the wildfire boundary. The direction of the UAV’s velocity
should be aligned with the orientation of the vector field. Therefore, an additional behavior
constraint related to the velocity vector ẋ and the tangent vector (E·∇(x,t))
T
∇(x,t) can be described
as
(E · ∇(x, t))T
ẋ = kt , (20)
 ∇(x, t) 
 
0 1
where E = is an orthogonal rotation matrix and determines the moving direction
−1 0
of the UAV along the wildfire boundary, and kt is a scalar weight. According to Eq. (20), the
vector field for patrol is defined as
E · ∇(x, t)
V pat = kt · . (21)
 ∇(x, t) 
Combining the two terms defined in (18) and (21), the vector field-based controller is repre-
sented as
∇(x, t) E · ∇(x, t)
u = −kn · (x, t) + kt · . (22)
 ∇(x, t)   ∇(x, t) 
To ensure that the velocity of the UAV is the desired value νd , Eq. (22) is multiplied with
a coefficient √ 2 νd 2 2 , and thus, the controller is
kn (x,t) +kt

kn νd · (x, t) ∇(x, t) kt νd E · ∇(x, t)


u = − · + · . (23)
kn2 (x, t)2 + kt2  ∇(x, t)  kn2 (x, t)2 + kt2  ∇(x, t) 

In the constructed vector field (23), the values of kn and kt do not influence the velocity
magnitude of the UAV, as the magnitude of the constructed vector field is vd . However, they
influence the time taken by the UAV to converge to the tracked trajectory. The increase of
kn or the decrease of kt will raise the portion of convergence controller term, thus increasing
the UAV’s rate of convergence to the target curve.

123
Radial basis function-based vector... Page 11 of 25 124

For the case of multiple UAVs, the potential conflicts among UAVs are avoided by intro-
ducing the repelling field. We define a function Vr as
m x−x oi 
Vr = kr e − lr , (24)
i=1

where x oi is the position of the ith obstacle, and kr and lr are positive coefficients. The
repulsive vector can be defined as
m
kr − x−x oi  x − x oi
V rep = −∇Vr = e lr · . (25)
lr  x − x oi 
i=1

Therefore, the final controller for the UAV is the combination of Eqs. (23) and (25)
kn νd · (x, t) ∇(x, t) kt νd E · ∇(x, t)
u = − · + ·
kn2 (x, t)2 + kt2  ∇(x, t)  kn2 (x, t)2 + kt2  ∇(x, t) 
m
kr − x−x oi  x − x oi
+ e lr · . (26)
lr  x − x oi 
i=1

The scheme of the wildfire boundary tracking mission can be explained in these steps:
• The initial map of the environment with wildfire locations is provided by sensing instru-
ments
• Generate extended data in the internal and external normal directions of sampled way-
points on the wildfire boundary
• Apply thin-plate spline RBFs to reconstruct a level set function to approximate the wildfire
boundary with the extended dataset
• Based on available information about wind, topography, and vegetation, the above level
set function is then used to estimate the propagation of wildfire boundary by solving the
Hamilton–Jacobi equation with RBFs
• If there are new sensing data, reconstruct a new level set function with the updated data,
otherwise using the existing level set function to generate a velocity vector field for
designing guidance law
• Update the trajectory for the UAV with the generated guidance law.
The constructed velocity vector field is analytical and tolerant to sensed changes, as the
characteristics of radial basis functions ensure that any pop-up changes on the wildfire bound-
ary can be modeled and updated in the approximated wildfire boundary, thus refreshing the
constructed vector field. It means that once new sensing data of wildfire boundary are avail-
able from other instruments, the wildfire boundary will be renewed with the approximation
of radial basis functions. Thus, the wildfire boundary will expand with the renewed boundary.
Correspondingly, the reconstructed velocity field is refreshed.

5 Simulation results

In this section, numerical simulations produced in MATLAB are used to evaluate the perfor-
mance of the proposed controller. UAVs are first employed to track a simulated advancing
wildfire boundary, and then follows the tracking of a real dynamic wildfire boundary.

123
124 Page 12 of 25 L. Feng, J. Katupitiya

30

t = 20s

25 t = 25s

t = 30s

t = 35s
20
t = 40s
t = 5s
t = 0s t = 10s
y(m)

15 t = 15s

10

0
0 5 10 15 20 25 30
x(m)
Fig. 2 Propagated wildfire boundary by solving Hamilton–Jacobi equation

5.1 Simulated wildfire boundary application

The initial position of an arbitrarily shaped wildfire boundary is described in a polar coordinate
whose origin is [xo , yo ]T

r = ro + ks (sin6β + sin3β), (27)

where ks is the parameter that shapes the curve, ro is a constant and β ∈ [0, 2π] is the
angular coordinate. We consider the positions of waypoints on wildfire boundary in x and
y directions; thus, the polar coordinate will be converted to Cartesian coordinate. In this
simulation, parameters for initial boundary are defined as xo = 15, yo = 15, ro = 5, and
ks = 0.8. The parameters for RoS model (11) are set to be the same as corresponding
coefficients in Mallet et al. (2009), namely 0 = 0.2, a = 0.5, n u = 3, and α = 0.5. The
magnitude of wind velocity is set as νw = 0.1m/s, which is smaller than the velocity of the
UAV. With the RoS model, the Hamilton–Jacobi equation (12) is numerically solved with the
radial basis function-based approach (Wei et al. 2018) to generate the propagation of wildfire
boundary. An example of the propagated wildfire with initial boundary described in (27) is
visualized in Fig. 2.
At each time the wildfire boundary is updated, radial basis functions are applied to recon-
struct the boundary curve to update the velocity vector field. First, an extended dataset should
be constructed in both internal and external normal directions with a small step δ = 0.01
of sampled waypoints on wildfire boundary. Then follows the approximation with radial
basis functions and the construction of velocity vector field. The extended dataset is shown
in Fig. 3, which indicates the sampled waypoints on the initial wildfire boundary and their

123
Radial basis function-based vector... Page 13 of 25 124

22
sampled point
external normal
internal normal
20

18

16
y(m)

14

12

10

8
8 10 12 14 16 18 20 22
x(m)
Fig. 3 Sampled waypoints on initial wildfire boundary with internal (blue) and external (red) waypoint normals
indicated

corresponding external and internal normal vectors. Applying RBF algorithm to the new
extended dataset, we obtain an approximated surface and its intersection with the zero level
set, as shown in Fig. 4. On the basis of the approximated level set function , a vector field for
velocity (Fig. 5) is generated. Obviously, the constructed vector field can ensure convergence
to the desired boundary.

5.1.1 Single UAV scenario

In this section, we present the performance of the proposed algorithm with a single UAV and
also compare it with the tracking strategy presented in Kumar et al. (2011). The work Kumar
et al. (2011) formulated the problem of tracking the fire front as an optimization of a utility
function, and designed a controller for a double integrator system on the basis of the gradient
descent information of the utility function. In this paper, the square of the approximated level
set function is regarded as the utility function. Its gradient descent term and a damping force
relating to UAV’s current velocity are combined as the input control.
Simulation with a single UAV is set as follows. A single UAV with initial condition
x = [25m, 9m]T is used to track the dynamic wildfire boundary. The desired velocity is
νd = 10m/s and control parameters are set as kn = 15, kt = 10. Simulation time and time
step are t f = 30s and t = 0.01s. The wildfire boundary is updated every 1s. Corresponding
simulation results generated from the application of the presented controller in this paper and
the controller in Kumar et al. (2011) are shown in Figs. 6 and 7. As shown in Fig. 6a, although
the wildfire boundary propagates with time, the single UAV’s trajectory is adaptively adjusted

123
124 Page 14 of 25 L. Feng, J. Katupitiya

Fig. 4 Approximated surface whose intersection with  = 0, i.e., the approximated surface’s zero level set is
the desired wildfire boundary

22

20

18

16
y(m)

14

12

10

8 10 12 14 16 18 20 22
x(m)
Fig. 5 Vector field constructed with RBF approximation method

123
Radial basis function-based vector... Page 15 of 25 124

to patrol along the dynamic wildfire boundary with the proposed guidance law. While in
Fig. 6b, the controller in Kumar et al. (2011) guides the UAV to track the wildfire boundary
only in the descending direction of the designed utility function. The change of tracking error
at UAV’s position is shown in Fig. 7. Regardless of the algorithm used [i.e., the proposed
algorithm or the comparison algorithm (Kumar et al. 2011)], the computed tracking error ||
has fluctuations at the critical time that the wildfire boundary is updated. The reason for these
fluctuations is that when the UAV is commanded the updated wildfire boundary, it receives a
command that is is a sudden change from the current boundary to the new boundary. Under
the guidance of the proposed controller, it takes some time for the UAV to move across from
the current tracking boundary to the updated boundary. However, as shown in Fig. 7, we can
see that until such time the updated wildfire boundary is received, the proposed algorithm
performs really well, rapidly decreasing the tracking error towards zero. By comparison, the
proposed algorithm can ensure that the UAV can converge to and patrol along the wildfire
boundary, while the controller (Kumar et al. 2011) only ensures the convergence to the
boundary and the trajectory tracking in the outdoor expanding direction. In addition, if the
fire boundary is not dynamic, the control law (Kumar et al. 2011) would have the issue of
local minimum, where the gradient of utility function is zero. This is because the dynamic
nature of the propagated wildfire boundary ensures that the local minimum changes, and that
the UAV is not stuck in the local minimum for long.

5.1.2 Multi-UAV scenario

For multi-UAV scenario, we also present a comparison of the proposed guidance law with the
method in Kumar et al. (2011). Both algorithms use the same controller for collision avoid-
ance. Three UAVs initially located at [26 m, 8.5 m]T , [25.5 m, 11 m]T and [24.5 m, 7.5 m]T ,
are employed to track the dynamic wildfire boundary. Control parameters for collision avoid-
ance are kr = 10, lr = 2, other simulation parameters are same as the single case. Simulation
results with three UAVs are displayed in Figs. 8, 9, and 10. Figure 8 shows the dynamic track-
ing trajectories to the expanding wildfire boundary with three UAVs. The presented algorithm
ensures that all UAVs converge to the boundary and then patrol along, as it shown in Fig. 8a.
In Fig. 8b, the controller in Kumar et al. (2011) also ensures that all UAVs converge to the
boundary, but all UAVs move in the expanding direction of the boundary after reaching the
wildfire boundary. Although the three UAVs are deployed at very close positions at the start
time, both controllers guarantee that the overall trend of distance between any two UAVs
increases along with time, as shown in Fig. 9, indicating that no occurrence of collisions
among UAVs. The tracking errors resulted from the application of the presented algorithm
and the algorithm in Kumar et al. (2011) are shown in Fig. 10. The tracking error || of
each UAV displayed in Fig. 10a indicates that although it has a sharp increase at the critical
time that the wildfire boundary is updated, it reduces dramatically and converges to zero
along with time. In comparison, the tracking error of each UAV resulted from the controller
in Kumar et al. (2011) (Fig. 10b) decreases slower to zero at the beginning and it also has
fluctuations arising from updating of wildfire boundary.

5.2 Actual wildfire boundary application

We also apply the proposed algorithm to track actual wildfire boundaries 1 extracted using
the image processing technique. An example of extracted wildfire boundaries at different
time instants is shown in Fig. 11. The initial position of the single UAV is [45m, 18m]T ;

123
124 Page 16 of 25 L. Feng, J. Katupitiya

UAV trajectory
Initial position
Final position
Fire boundary
25

20
y(m)

15

10

5
5 10 15 20 25
x(m)
(a) Proposed controller

UAV trajectory
Initial position
Final position
Fire boundary
25

20
y(m)

15

10

5 10 15 20 25
x(m)
(b) Controller in [22]
Fig. 6 Tracking trajectory of a single UAV resulted from the proposed controller (a) and the controller in
Kumar et al. (2011) (b); filled square represents the initial position of UAV, filled star represents the final
positions of UAV, red line is the trajectory of UAV, and black dot-dash line is wildfire boundary

123
Radial basis function-based vector... Page 17 of 25 124

4.5

3.5

2.5
| |

1.5

0.5

0
0 5 10 15 20 25 30
t(s)
(a) Proposed controller
4.5

3.5

2.5
| |

1.5

0.5

0
0 5 10 15 20 25 30
t(s)
(b) Controller in [22]
Fig. 7 The time profile of wildfire boundary tracking error of a single UAV with the proposed controller (a)
and the controller in Kumar et al. (2011) (b), the time at which the impulse happens is the critical time for
updating wildfire boundary

other parameters are same as previous. The resulted tracking trajectory and tracking error are
displayed in Figs. 12 and 13, respectively. From Fig. 12, we can find that the single UAV can
adapt its trajectory to closely follow along the wildfire boundary. Although the tracking error
(Fig. 13) has fluctuations when the UAV is commanded to track a renewed wildfire boundary
from its current position, this error can be reduced rapidly when the UAV moves across the
current boundary to converge to and patrol along the updated wildfire boundary under the
guidance of the proposed controller.

123
124 Page 18 of 25 L. Feng, J. Katupitiya

28 UAV1 trajectory
UAV2 trajectory
26 UAV3 trajectory
Fire boundary
24

22

20
y(m)

18

16

14

12

10

6 8 10 12 14 16 18 20 22 24 26
x(m)
(a) Proposed controller

28 UAV1 trajectory
UAV2 trajectory
26 UAV3 trajectory
Fire boundary
24

22

20
y(m)

18

16

14

12

10

6 8 10 12 14 16 18 20 22 24 26
x(m)
(b) Controller in [22]
Fig. 8 Tracking trajectories of three UAVs resulted from the proposed controller (a) and the controller in
Kumar et al. (2011) (b); filled squares represent the initial positions of UAVs, filled stars represent the final
positions of UAVs, red dot-dash line is the trajectory of UAV 1, blue dash line is the trajectory of UAV 2, and
pink line is the trajectory of UAV 3

123
Radial basis function-based vector... Page 19 of 25 124

20
d 12
18 d 13
d 23
16

14

12
d(m)

10

0
0 5 10 15 20 25 30
t(s)
(a) Proposed controller
15
d 12
d 13
d 23

10
d(m)

0
0 5 10 15 20 25 30
t(s)
(b) Controller in [22]
Fig. 9 The time profiles of distance between any two UAVs with the proposed controller (a) and the controller
in Kumar et al. (2011) (b); red dot-dash line is the distance between UAV 1 and UAV 2, blue dash line is the
distance between UAV 1 and UAV 3, and pink line is the distance between UAV 2 and UAV 3

To summarize, the proposed algorithm works well no matter deploying a single UAV
or multiple UAVs to track the dynamic wildfire boundary. Different from the the controller
which only ensures the track of the dynamic wildfire boundary in its expanding direction, the
presented controller in this paper ensures both convergence to and patrolling along the wildfire
boundary with a constant velocity. In addition, the presented guidance law is robust to deal
with abrupt changes in wildfire boundary, as the introduction of RBF-based approximation
algorithm ensures that any change in the wildfire boundary can be modeled and reflected in
the constructed vector field. Thus, the abrupt position differences between the UAV and the
expanding wildfire boundary can rapidly reduce and converge to zero along with time.

123
124 Page 20 of 25 L. Feng, J. Katupitiya

5
UAV1
4.5
UAV2
4 UAV3

3.5

3
| |

2.5

1.5

0.5

0
0 5 10 15 20 25 30
t(s)
(a) Proposed controller
5
UAV1
4.5 UAV2
UAV3
4

3.5

3
| |

2.5

1.5

0.5

0
0 5 10 15 20 25 30
t(s)
(b) Controller in [22]
Fig. 10 The time profiles of wildfire boundary tracking errors of three UAVs with the proposed controller (a)
and the controller in Kumar et al. (2011) (b); the time at which the impulse happens is the critical time for
updating wildfire boundary, red dot-dash line represents UAV 1, blue dash line represents UAV 2, and the pink
line represents UAV 3

6 Conclusion

In this paper, we presented a guidance strategy for UAVs to track an arbitrarily shaped wild-
fire boundary using the level set approach. Radial basis functions are used to approximate the
wildfire boundary and construct the vector field for tracking dynamic wildfire fronts. Numer-
ical simulations with single and multiple UAVs demonstrate that the proposed guidance law
is robust and ensures successful tracking of an expanding wildfire boundary, regardless of
the shape of the boundary. In this paper, only the kinematic model of the UAV is considered,
and the vector field-based control law for more realistically modeled UAV will be studied.

123
Radial basis function-based vector... Page 21 of 25 124

(a) t = 0s (b) t = 5s

(c) t = 10s (d) t = 15s


Fig. 11 Extracted real wildfire boundaries at different time instants

UAV trajectory
Initial position
Final position
35 Fire boundary

30
y(m)

25

20

26 28 30 32 34 36 38 40 42 44 46
x(m)
Fig. 12 Tracking trajectory of a real dynamic wildfire boundary

123
124 Page 22 of 25 L. Feng, J. Katupitiya

4.5

3.5

2.5
| |

1.5

0.5

0
0 2 4 6 8 10 12 14 16
t(s)
Fig. 13 The time profile of wildfire tracking error of a single UAV, the time at which the impulse happens is
the critical time for updating wildfire boundary

In addition, we assumed that the wildfire boundary is closed and the open boundaries were
not considered. However, the open boundaries can be artificially closed and the same method
can be applied to track those open boundary segments.

Acknowledgements Open Access funding enabled and organized by CAUL and its Member Institutions. This
research did not receive any specific funding.

Funding Open Access funding enabled and organized by CAUL and its Member Institutions.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which
permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give
appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence,
and indicate if changes were made. The images or other third party material in this article are included in the
article’s Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is
not included in the article’s Creative Commons licence and your intended use is not permitted by statutory
regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder.
To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

References
Adams SM, Friedland CJ (2011) A survey of unmanned aerial vehicle (UAV) usage for imagery collection
in disaster research and management. In: 9th international workshop on remote sensing for disaster
response, vol 8
Alessandri A, Bagnerini P, Gaggero M, Mantelli L (2021) Parameter estimation of fire propagation models
using level set methods. Appl Math Model 92:731–747
Ambroz M, Balažovjech M, Medl’a M, Mikula K (2019) Numerical modeling of wildland surface fire propa-
gation by evolving surface curves. Adv Comput Math 45(2):1067–1103

123
Radial basis function-based vector... Page 23 of 25 124

Bailon-Ruiz R, Lacroix S (2020) Wildfire remote sensing with UAVs: a review from the autonomy point of
view. In: 2020 International conference on unmanned aircraft systems (ICUAS). IEEE, pp 412–420
Balažovjech M, Mikula K (2011) A higher order scheme for a tangentially stabilized plane curve shortening
flow with a driving force. SIAM J Sci Comput 33(5):2277–2294
Bertozzi AL, Kemp M, Marthaler D (2005) Determining environmental boundaries: asynchronous communi-
cation and physical scales. Cooperative control. Springer, Berlin, pp 25–42
Brockett RW et al (1983) Asymptotic stability and feedback stabilization. Differ Geom Control Theory
27(1):181–191
Buhmann MD (2003) Radial basis functions: theory and implementations, vol 12. Cambridge University Press,
Cambridge
Carr JC, Beatson RK, Cherrie JB, Mitchell TJ, Fright WR, McCallum BC, Evans TR (2001) Reconstruction
and representation of 3D objects with radial basis functions. In: Proceedings of the 28th annual conference
on computer graphics and interactive techniques, pp 67–76
Cassandras CG, Ding XC, Lin X (2011) An optimal control approach for the persistent monitoring problem.
In: 2011 50th IEEE conference on decision and control and European control conference. IEEE, pp
2907–2912
Cuomo S, Galletti A, Giunta G, Marcellino L (2017) Reconstruction of implicit curves and surfaces via RBF
interpolation. Appl Numer Math 116:157–171
De Marina HG, Kapitanyuk YA, Bronz M, Hattenberger G, Cao M (2017) Guidance algorithm for smooth
trajectory tracking of a fixed wing UAV flying in wind flows. In: 2017 IEEE international conference on
robotics and automation (ICRA). IEEE, pp 5740–5745
Dong F, You K (2021) The isoline tracking in unknown scalar fields with concentration feedback. Automatica
133:109779
Dong F, You K, Wang J (2020) Coordinate-free isoline tracking in unknown 2-D scalar fields. In: 2020
IEEE/RSJ international conference on intelligent robots and systems (IROS). IEEE, pp 2496–2501
Dziuk G (1999) Discrete anisotropic curve shortening flow. SIAM J Numer Anal 36(6):1808–1830
Fahad M, Saul N, Guo Y, Bingham B (2015) Robotic simulation of dynamic plume tracking by unmanned
surface vessels. In: 2015 IEEE international conference on robotics and automation (ICRA). IEEE, pp
2654–2659
Goncalves VM, Pimenta LC, Maia CA, Pereira GA (2009) Artificial vector fields for robot convergence and
circulation of time-varying curves in n-dimensional spaces. In: American control conference. IEEE, pp
2012–2017
Goncalves VM, Pimenta LC, Maia CA, Dutra BC, Pereira GA (2010) Vector fields for robot navigation along
time-varying curves in n-dimensions. IEEE Trans Rob 26(4):647–659
Hou TY, Lowengrub JS, Shelley MJ (1994) Removing the stiffness from interfacial flows with surface tension.
J Comput Phys 114(2):312–338
Jiang X, Li S (2018) Plume front tracking in unknown environments by estimation and control. IEEE Trans
Ind Inf 15(2):911–921
Jiang C, Chen Z, Guo Y (2020) Multi-robot formation control: a comparison between model-based and
learning-based methods. J Control Decis 7(1):90–108
Jin Z, Bertozzi AL (2007) Environmental boundary tracking and estimation using multiple autonomous vehi-
cles. In: 2007 46th IEEE conference on decision and control. IEEE, pp 4918–4923
Kanistras K, Martins G, Rutherford MJ, Valavanis KP (2013) A survey of unmanned aerial vehicles (UAVs)
for traffic monitoring. In: 2013 international conference on unmanned aircraft systems (ICUAS). IEEE,
pp 221–234
Kapitanyuk YA, Chepinsky S (2013) Control of mobile robot following a piecewise-smooth path. Gyrosc
Navig 4(4):198–203
Kapitanyuk YA, Proskurnikov AV, Cao M (2017) A guiding vector-field algorithm for path-following control
of nonholonomic mobile robots. IEEE Trans Control Syst Technol 26(4):1372–1385
Karouni A, Daya B, Bahlak S, Chauvet P (2014) A simplified mathematical model for fire spread predictions in
wildland fires combining between the models of Anderson and Rothermel. Int J Model Optim 4(3):197–
200
Koushki M, Jabbari E, Ahmadinia M (2020) Evaluating RBF methods for solving PDEs using Padua points
distribution. Alex Eng J 59(5):2999–3018
Kumar M, Cohen K, HomChaudhuri B (2011) Cooperative control of multiple uninhabited aerial vehicles for
monitoring and fighting wildfires. J Aerosp Comput Inf Commun 8(1):1–16
Lan X, Schwager M (2013) Planning periodic persistent monitoring trajectories for sensing robots in gaussian
random fields. In: 2013 IEEE international conference on robotics and automation. IEEE, pp 2415–2420
Li S, Guo Y, Bingham B (2014a) Multi-robot cooperative control for monitoring and tracking dynamic plumes.
In: 2014 IEEE international conference on robotics and automation (ICRA). IEEE, pp 67–73

123
124 Page 24 of 25 L. Feng, J. Katupitiya

Li S, Kong R, Guo Y (2014b) Cooperative distributed source seeking by multiple robots: algorithms and
experiments. IEEE/ASME Trans Mechatron 19(6):1810–1820
Lim S, Jung W, Bang H (2014) Vector field guidance for path following and arrival angle control. In: 2014
international conference on unmanned aircraft systems (ICUAS). IEEE, pp 329–338
Malisoff M, Sizemore R, Zhang F (2017) Adaptive planar curve tracking control and robustness analysis under
state constraints and unknown curvature. Automatica 75:133–143
Mallet V, Keyes DE, Fendell F (2009) Modeling wildland fire propagation with level set methods. Comput
Math Appl 57(7):1089–1101
Marthaler D, Bertozzi AL (2003) Collective motion algorithms for determining environmental boundaries. In:
SIAM conference on applications of dynamical systems, Citeseer
Matveev AS, Teimoori H, Savkin AV (2012) Method for tracking of environmental level sets by a unicycle-like
vehicle. Automatica 48(9):2252–2261
Menon PP, Edwards C, Shtessel YB, Ghose D, Haywood J (2014) Boundary tracking using a suboptimal
sliding mode algorithm. In: 53rd IEEE conference on decision and control. IEEE, pp 5518–5523
Miyanishi K (2001) Forest fires: behavior and ecological effects. Elsevier, Amsterdam
Morse BS, Yoo TS, Rheingans P, Chen DT, Subramanian KR (2005) Interpolating implicit surfaces from
scattered surface data using compactly supported radial basis functions. In: ACM SIGGRAPH 2005
courses, pp 78–es
Nelson DR, Barber DB, McLain TW, Beard RW (2007) Vector field path following for miniature air vehicles.
IEEE Trans Rob 23(3):519–529
Obermeyer K, Oberlin P, Darbha S (2010) Sampling-based roadmap methods for a visual reconnaissance UAV.
In: AIAA guidance, navigation, and control conference, p 7568
Osher S, Fedkiw R (2006) Level set methods and dynamic implicit surfaces, vol 153. Springer Science and
Business Media, Berlin
Osher S, Sethian JA (1988) Fronts propagating with curvature-dependent speed: algorithms based on
Hamilton–Jacobi formulations. J Comput Phys 79(1):12–49
Peng D, Merriman B, Osher S, Zhao H, Kang M (1999) A PDE-based fast local level set method. J Comput
Phys 155(2):410–438
Richards GD (1990) An elliptical growth model of forest fire fronts and its numerical solution. Int J Numer
Methods Eng 30(6):1163–1179
Sahin E, Spears WM, Winfield AF (2007) Swarm robotics: second SAB 2006 international workshop, Rome,
Italy, September 30–October 1, 2006 Revised Selected Papers, vol 4433. Springer
Sethian JA (1999) Level set methods and fast marching methods: evolving interfaces in computational
geometry, fluid mechanics, computer vision, and materials science, vol 3. Cambridge University Press,
Cambridge
Smith SL, Rus D (2010) Multi-robot monitoring in dynamic environments with guaranteed currency of obser-
vations. In: 49th IEEE conference on decision and control (CDC). IEEE, pp 514–521
Smith RN, Schwager M, Smith SL, Jones BH, Rus D, Sukhatme GS (2011) Persistent ocean monitoring with
underwater gliders: adapting sampling resolution. J Field Robot 28(5):714–741
Stanton NA, Chambers PR, Piggott J (2001) Situational awareness and safety. Saf Sci 39(3):189–204
Sujit P, Saripalli S, Sousa JB (2014) Unmanned aerial vehicle path following: a survey and analysis of algo-
rithms for fixed-wing unmanned aerial vehicless. IEEE Control Syst Mag 34(1):42–59
Sullivan AL (2009) Wildland surface fire spread modelling, 1990–2007. 1: physical and quasi-physical models.
Int J Wildland Fire 18(4):349–368
Sullivan AL (2009b) Wildland surface fire spread modelling, 1990-2007. 2: Empirical and quasi-empirical
models. Int J Wildland Fire 18(4):369–386
Sullivan, Andrew L (2009a) Wildland surface fire spread modelling, 1990–2007. 3: simulation and mathemat-
ical analogue models. Int J Wildland Fire 18(4):387–403
Susca S, Bullo F, Martinez S (2008) Monitoring environmental boundaries with a robotic sensor network.
IEEE Trans Control Syst Technol 16(2):288–296
Tang L, Shao G (2015) Drone remote sensing for forestry research and practices. J For Res 26(4):791–797
Triandaf I, Schwartz IB (2005) A collective motion algorithm for tracking time-dependent boundaries. Math
Comput Simul 70(4):187–202
Twidwell D, Allen CR, Detweiler C, Higgins J, Laney C, Elbaum S (2016) Smokey comes of age: unmanned
aerial systems for fire management. Front Ecol Environ 14(6):333–339
Wang J-W, Guo Y, Fahad M, Bingham B (2019) Dynamic plume tracking by cooperative robots. IEEE/ASME
Trans Mechatron 24(2):609–620
Wei P, Li Z, Li X, Wang MY (2018) An 88-line MATLAB code for the parameterized level set method based
topology optimization using radial basis functions. Struct Multidiscip Optim 58(2):831–849

123
Radial basis function-based vector... Page 25 of 25 124

Wilhelm JP, Clem G (2019) Vector field UAV guidance for path following and obstacle avoidance with minimal
deviation. J Guid Control Dyn 42(8):1848–1856
Xie X, Mirmehdi M (2007) Implicit active model using radial basis function interpolated level sets. In: BMVC.
Citeseer, pp 1–10
Yuan C, Zhang Y, Liu Z (2015) A survey on technologies for automatic forest fire monitoring, detection, and
fighting using unmanned aerial vehicles and remote sensing techniques. Can J For Res 45(7):783–792
Zhai C, Zhang S, Cao Z, Wang X (2020) Learning-based prediction of wildfire spread with real-time rate of
spread measurement. Combust Flame 215:333–341
Zhu S, Wathen AJ (2015) Convexity and solvability for compactly supported radial basis functions with
different shapes. J Sci Comput 63(3):862–884

Publisher’s Note Springer Nature remains neutral with regard to jurisdictional claims in published maps and
institutional affiliations.

123

You might also like