ArticlePDF Available

New results about the geometric invariants in KCC-theory

Authors:

Abstract

The KCC-theory (Kosambi, [12], CARTAN, [7], and CHERN, [8]) of a system of second order ordinary differential equations (SODE) uses five geometric in-variants that determine, up to a change of coordinates, the solutions of the system. Geometrically speaking to a SODE corresponds a vector field, called a semispray (or alternatively , a second order vector field) that lives on the tangent bundle of a manifold. For a semispray S it is well known that it determines a nonlinear connection N and a Finsler connection D, called the Berwald connection, both of them living on the tangent space of the given manifold. We prove that all five invariants of the system can be expressed in this geometric framework. Using the dynamical covariant derivative and the covariant derivative induced by the Berwald connection we determine two invariant equations for the variational equations of a SODE and for the symmetries of the associated semispray. The KCC-theory has significant applications in biology, [2], [14].
ANALELE S¸TIINT¸ IFICE ALE UNIVERSIT
˘
AT¸II ”AL.I.CUZA” IAS¸I
Tomul XLVII, s.I a, Matematic˘a, 2001, f.2.
NEW RESULTS ABOUT THE GEOMETRIC INVARIANTS
IN KCC-THEORY
BY
P.L. ANTONELLI
and I. BUC
˘
ATARU
Abstract. The KCC-theory (Kosambi, [12], CARTAN, [7], and CHERN, [8]) of
a system of second order ordinary differential equations (SODE) uses five geometric in-
variants that determine, up to a change of coordinates, the solutions of the system.
Geometrically speaking to a SODE corresponds a vector field, called a semispray (or al-
ternatively, a second order vector field) that lives on the tangent bundle of a manifold. For
a semispray S it is well known that it determines a nonlinear connection N and a Finsler
connection D, called the Berwald connection, both of them living on the tangent space
of the given manifold. We prove that all five invariants of the system can be expresse d
in this geometric framework. Using the dynamical covariant derivative and the covariant
derivative induced by the Berwald connection we determine two invariant equations for
the variational equations of a SODE and f or the symmetries of the associated semispray.
The KCC-theory has significant applications in biology, [2], [14].
2000 Mathematics Subject Classification: 53C60, 58A20, 58A30, 37N25.
Key words and phrases: SODE, Jacobi field, Berwald connection.
Introduction. Second order systems of ODES are now recognized as
being important in Volterra-Hamilton theory, [4] and in Analytical tropho-
dynamics, [2], [14] where intrinsic properties like curvature determine the
stability of production processes.
From a geometric point of view, a system of second order ODE is equiv-
alent to a vector field S, called a semispray that lies on the tangent bundle
Partially supported by NSERC-7667
PIMS Postdoctoral Fellow
406 P.L. ANTONELLI and I. BUC
˘
ATARU 2
T M of a manifold M . As it is well known, a s em ispray yields two other
important geometrical objects, living also on the tangent bundle: a non-
linear connection, which is a distribution N supplementary to the vertical
distribution V , and a linear connection D adapted to V and N. This linear
connection was introduced in 1947 by Berwald in ([6]) and therefore is
called the Berwald connection. A remarkable direction for the geometry
of a SODE appeared in the thirties. The core of this new method defines
the so-called ”KCC-theory”(Kosambi, [12], Cartan, [7], and Chern, [8])
which establishes at five the number of geometric invariants that determine,
up to a change of coordinates, the solutions of the given system of second
order. Our main result is an interpretation of these five KCC-invariants in
terms of N and D, the curvature of N and the torsion and curvature of D.
For a semispray S on the tangent bundle T M of a manifold M we con-
sider the induced nonlinear connection N. The vertical component of S,
with respect to this nonlinear connection N , gives the first invariant of S,
called the deviation tensor of the semispray. The curvature of the nonlinear
connection, R
i
jk
is the third invariant. In section two we give a global ex-
pression of the induced Berwald connection D. It is proved that this linear
connection on the tangent bundle is a Finsler connection. With respect
to an adapted basis to the nonlinear connection, the Berwald connection
D has only one nonzero component of torsion R
i
jk
and two nonzero com-
ponents of curvature R
i
jkl
and D
i
jkl
. These are the third, the fourth, and
the fifth invariants of the semispray S, respectively. The second invariant
is expressed using the third invariant and the h-covariant derivative of the
first invariant. The Berwald connection appears also in [5], and [16] in local
coordinates as a Finsler connection on the tangent bundle, and in [9] in the
pullback bundle of the tangent bundle by its natural projection.
Besides the dynamical covariant derivative induced by a nonlinear con-
nection, or a semispray, ([16], [9]), we introduce a covariant derivative de-
termined by the Berwald connection. Only in the particular case when the
semispray is homogeneous, these two covariant derivatives coincide. The
path equations and the variational equations (Jacobi equations) are written
using both covariant derivatives. The relationship between Jacobi vector
fields for a system of SODE and the dynamical symmetries of the associated
semispray is given.
3 GEOMETRIC INVARIANTS IN KCC- THEORY 407
1. Dynamical covariant derivative induced by a semispray.
Let M be a real, smooth, n-dimensional manifold, and (T M, π, M ) be
its tangent bundle. For a local chart (U, φ = (x
i
)) on M, we denote by
(π
1
(U), Φ = (x
i
, y
i
)) the induced local chart on T M . The kernel of the
linear map induced by the natural submersion π : T M M, determines
the vertical distribution V : u T M 7→ V
u
= Kerπ
,u
T
u
T M. This is
an n-dimensional integrable distribution. If {
x
i
|
u
,
y
i
|
u
} is the natural
basis of the tangent space T
u
T M, then {
y
i
|
u
} is a basis for V
u
, u T M .
Consider J =
y
i
dx
i
, the almost tangent structure (J is also called the
vertical endomorphism of T M), and Γ = y
i
y
i
the Liouville vector field. A
vector field S on T M is called a semispray (or a second order vector field) if
JS = Γ. The local expression of a semispray is S = y
i
x
i
2G
i
(x, y)
y
i
.
The functions G
i
(x, y) are called the local coefficients of the semispray and
these are defined on domains of local chart.
An n-dimensional distribution N : u T M 7→ N
u
T
u
T M that is sup-
plementary to the vertical distribution V is called a nonlinear connection.
For every u T M we have the direct sum
(1.1) T
u
T M = N
u
V
u
.
An adapted basis to the previous direct sum is {
δ
δx
i
=
x
i
N
j
i
(x, y)
y
j
,
y
i
}. We call this basis the Berwald basis of the nonlinear connection N.
The functions N
i
j
(x, y) are defined on domains of loc al chart, and these
are called the local coefficients of the nonlinear connection N. It is well
known that every semispray S with local coefficients G
i
, induces a nonlinear
connection N with local coefficients N
i
j
=
G
i
y
j
, [11]. Next, we shall work
with this nonlinear connection. Denote by h and v the horizontal and
the vertical projectors, that correspond to the decomposition (1.1). In the
Berwald basis we have:
(1.2) h =
δ
δx
i
dx
i
, v =
y
i
δy
i
,
408 P.L. ANTONELLI and I. BUC
˘
ATARU 4
where (dx
i
, δy
i
= dy
i
+ N
i
j
(x, y)dx
j
) is the dual basis of the Berwald basis.
The vertical projector v is given by
(1.3) v(X) =
1
2
(X + [S, JX] + J[X, S]), X χ(T M).
A tensor field of (r, s)-type on T M is called a Finsler tensor field, [15], (or
a d-tensor field in [16]) if under a change of induced coordinates on T M ,
its components transform like the components of a (r, s)-type tensor field
on the base manifold M.
The local expression of a semispray in the Berwald basis is
(1.4) S = y
i
δ
δx
i
E
i
(x, y)
y
i
,
where
(1.5) E
i
(x, y) = 2G
i
(x, y) N
i
j
(x, y)y
j
is a (1,0)-type tensor Finsler field. This is called the first invariant of the
semispray in ([8], [7]), or the deviation tensor in [11].
For a vector field X = X
i
(x)
x
i
on the base manifold M, let us consider
(1.6) X
c
= X
i
(x)
x
i
+
X
i
x
j
(x)y
j
y
i
, X
h
= X
i
δ
δx
i
, and X
v
= X
i
(x)
y
i
,
the complete, the horizontal, and the vertical lift, respectively. It is very
easy to check that these lifts of a vector field X χ(M ) are related by
(1.6)
0
X
c
= 2X
h
+ [S, X
v
] = 2X
h
+ L
S
X
v
,
where L
S
is the Lie derivative with respect to S. The dynamical covariant
derivative of a Finsler vector field X
i
(x, y) is defined by, [9]:
(1.7) X
i
= S(X
i
) + N
i
j
X
j
=
X
i
x
j
y
j
2
X
i
y
j
G
j
+
G
i
y
j
X
j
.
Remark 1.1.
1
o
For a Finsler vector field X
i
(x, y), its dynamical covariant derivative
X
i
satisfies
(1.7)
0
[S, X
i
y
i
] = X
i
δ
δx
i
+ X
i
y
i
.
5 GEOMETRIC INVARIANTS IN KCC- THEORY 409
As a consequence we have that v[S, X
i
y
i
] = X
i
y
i
, and then X
i
is
also a Finsler vector field.
2
o
We have the properties: (X
i
+ Y
i
) = X
i
+ Y
i
, and (fX
i
) =
S(f )X
i
+ fX
i
.
3
o
If X = X
i
(x)
x
i
is a vector field on the base manifold M , then X
c
=
X
i
(x)
δ
δx
i
+ X
i
y
i
, so v(X
c
) = X
i
y
i
.
4
o
The dynamical covariant derivative of the Liouville vector field (y
i
) and
the first invariant are related by y
i
= −E
i
.
A curve c : t I 7→ c(t) = (x
i
(t)) M is called a path of the semispray
S if its lift to T M : ˜c : t I 7→ ˜c(t) = (x
i
(t),
dx
i
dt
(t)) T M is an integral
curve of S. In local coordinates the curve c(t) = (x
i
(t)) is a path of S if
and only if:
(1.8)
d
2
x
i
dt
2
+ 2G
i
(x,
dx
dt
) = 0.
An equivalent invariant form of the system (1.8) is given by:
(1.8)
0
(
dx
i
dt
) = −E
i
(x,
dx
dt
).
Remark 1.2. The semispray S is called a spray if S is homogeneous of
degree two with respect to y, and this is equivalent to E
i
= 0. In this case
the paths of the semispray S are horizontal curves of the induced nonlinear
connection, because S is a horizontal vector field.
2. The Berwald connection induced by a semispray. Consider
S a semispray, and N the induced nonlinear connection with the Berwald
basis {
δ
δx
i
,
y
i
}. Consider the tensor field θ =
δ
δx
i
δy
i
. We can see that
the restriction of θ to the vertical distribution is an isomorphism between
this vertical distribution and the horizontal distribution.
Definition 2.1. A linear connection D (Koszul connection) on T M
is called a Finsler connection if D prese rves by parallelism the horizontal
410 P.L. ANTONELLI and I. BUC
˘
ATARU 6
distribution N and the almost tangent structure J is absolutely parallel
with respect to D.
So, a linear connection D on T M is a Finsler connection if and only
if Dh = 0 and DJ = 0. For a Finsler connection D it is very easy to
check that D preserves also by parallelism the vertical distribution V , that
is Dv = 0. Moreover a linear connection D on T M is a Finsler connection
if and only if Dv = 0, and = 0.
With respect to the Berwald basis, a Finsler connection D has the local
expression, [16]:
(2.1)
D
δ
δx
i
δ
δx
j
= F
k
ji
δ
δx
k
, D
δ
δx
i
y
j
= F
k
ji
y
k
,
D
y
i
δ
δx
j
= C
k
ji
δ
δx
k
, D
y
i
y
j
= C
k
ji
y
k
.
Next, a Finsler connection will be indicated also by the set D=(N
i
j
, F
k
ij
, C
k
ij
).
Under a change of induced c oordinates on T M, the coefficients F
k
ij
transform
like the coefficients of a linear connection on the base manifold M. The
coefficients C
k
ij
are the components of a (1,2)-type Finsler tensor field.
For a Finsler vector field X
i
(x, y), we define the covariant derivative
induced by a Finsler connection D = (N
i
j
, F
k
ij
, C
k
ij
) as:
(2.2) DX
i
= S(X
i
) + F
i
jk
X
j
y
k
C
i
jk
X
j
E
k
,
or in the equivalent form
(2.2)
0
(DX
i
)
y
i
= D
S
(X
i
y
i
).
Denote by X
i
|k
, and X
i
|
k
the horizontal, and the vertical covariant deriva-
tives of X
i
, respectively. These are given by:
(2.3) X
i
|k
=
δX
i
δx
k
+ F
i
jk
X
j
, X
i
|
k
=
X
i
y
k
+ C
i
jk
X
j
.
Then, the Finsler covariant derivative takes the form:
(2.2)
00
DX
i
= X
i
|k
y
k
X
i
|
k
E
k
.
7 GEOMETRIC INVARIANTS IN KCC- THEORY 411
More generally, if T
i
1
···i
r
j
1
···j
s
are the components of a (r, s)-type Finsler tensor
field, the Finsler covariant derivative of this DT
i
1
···i
r
j
1
···j
s
is also a Finsler tensor
field of (r, s)-type, and it is defined by:
(2.2)
000
DT
i
1
···i
r
j
1
···j
s
= T
i
1
···i
r
j
1
···j
s
|k
y
k
T
i
1
···i
r
j
1
···j
s
|
k
E
k
,
where T
i
1
···i
r
j
1
···j
s
|k
and T
i
1
···i
r
j
1
···j
s
|
k
are the h and v-covariant derivatives, respec-
tively ([16], p.40). We can remark here that this covariant derivative satisfies
the Leibniz rule, that is for example D(T
i
j
S
j
k
) = D(T
i
j
)S
j
k
+ T
i
j
D(S
j
k
).
Theorem 2.1. The map D : χ(T M ) × χ(T M ) χ(T M), given by
(2.4) D
X
Y = v[hX, vY ] + h[vX, hY ] + J[vX, θY ] + θ[hX, JY ]
is a Finsler connection on T M . We call this the Berwald connection of the
semispray S.
Proof. As all the operators involved in (2.4) are additive we have
that D is additive too, with respect to b oth arguments. To prove that
D
fX
Y = fD
X
Y , f F(T M ) we have to use that vh = hv = Jv = θh = 0.
The other properties like D
X
fY = X(f)Y + f D
X
Y , Dh = 0, and D J = 0
follow easy using also Jθ = v, θJ = h, v
2
= v, and h
2
= h.
With resp ec t to the Berwald basis, the Berwald connection D has the
local coefficients, ([5], [16]):
(2.5) F
i
jk
=
N
i
j
y
k
=
2
G
i
y
j
y
k
, C
i
jk
= 0.
For the Berwald connection D one considers typically the torsion T (X, Y ) =
D
X
Y D
Y
X [X, Y ]. With respect to the Berwald basis, there is only one
nonzero component of the torsion, ([5], [16]):
(2.6) vT (
δ
δx
i
,
δ
δx
j
) =: R
k
ij
y
k
= (
δN
k
j
δx
i
δN
k
i
δx
j
)
y
k
, (v)h torsion.
The Finsler tensor field R
k
ij
is also called the curvature tensor of the non-
linear connection N , because:
(2.7) [
δ
δx
i
,
δ
δx
j
] = R
k
ij
y
k
.
412 P.L. ANTONELLI and I. BUC
˘
ATARU 8
Let us consider now R(X, Y )Z = D
X
D
Y
Z D
Y
D
X
Z D
[X,Y ]
Z, the cur-
vature of the Berwald connection. With respect to the Berwald basis, there
are only two nonzero components of the curvature ([5], [16]):
(2.8)
R
i
jkl
=
δF
i
jk
δx
l
δF
i
jl
δx
k
+ F
m
jk
F
i
ml
F
m
jl
F
i
mk
;
(the Riemann Christoffel curvature tensor)
D
i
jkl
=
F
i
jk
y
l
=
3
G
i
y
j
y
k
y
l
(the Douglas tensor of the semispray).
Remark 2.1. If S is a spray, a consequence of the homogeneity is that
the Finsler tensor fields R
i
jk
and R
i
ljk
are related by, ([6], [16]):
(2.9) R
i
jk
= R
i
ljk
y
l
.
In the general case, when S is not homogeneous, we have the following
result:
Proposition 2.1. Let D be the Berwald connection associated to a
semispray S, R
i
jk
the (v)h-component of the torsion, and R
i
ljk
the Riemann-
Christoffel curvature tensor. Then these are related by:
(2.10) R
i
ljk
y
l
R
i
kj
= y
i
|j|k
y
i
|k|j
,
or in the equivalent form
(2.10)
0
R
i
ljk
y
l
R
i
kj
= E
i
|
k|j
E
i
|
j|k
.
Proof. Using the Ricc i identities for the Berwald connection D ([17],
p.78) and the Liouville ve ctor field, we get (2.10). For (2.10)
0
it is enough
to prove that:
(2.11) y
i
|j
= −E
i
|
j
.
This (1,1)-type Finsler tensor field is called the deflection tensor of the
connection. According to (2.3) we have y
i
|j
=
δy
i
δx
j
+ F
i
kj
y
k
= F
i
kj
y
k
N
i
j
. As
9 GEOMETRIC INVARIANTS IN KCC- THEORY 413
E
i
= 2G
i
G
i
y
j
y
j
, the E
i
|
j
=
E
i
y
j
= 2
G
i
y
j
G
i
y
j
2
G
i
y
j
y
k
y
k
= N
i
j
F
i
kj
y
k
=
y
i
|j
.
Proposition 2.2. The dynamical covariant derivative (1.7) induced by
a semispray S, and the Berwald covariant derivative are related by:
(2.12) X
i
= DX
i
y
i
|j
X
j
= DX
i
+ E
i
|
j
X
j
.
Proof. From (2.2), the Berwald covariant derivative has the form:
DX
i
= S(X
i
) + F
i
jk
y
j
X
k
, so DX
i
= S(X
i
) + N
i
k
X
k
+ (F
i
jk
y
j
N
i
k
)X
k
=
X
i
+ y
i
|k
y
k
, and the proof is completed.
Remark 2.2. Using the Berwald covariant derivative, we can write the
path equation (1.8) in an equivalent form:
(2.13) D(
dx
i
dt
) = (E
i
+ E
i
|
j
dx
j
dt
).
3. Symmetries and Jacobi equations for a semispray. Consider
a semispray S. A path of the semispray is a solution of the system (1.8), or
of the equivalent forms (1.8)
0
, or (2.13). Consider c(t) = (x
i
(t)) a trajectory
of (1.8), and let vary it into nearby ones according to:
(3.1) ˜x
i
(t) = x
i
(t) + εξ
i
(t),
where ε denotes a scalar parameter with small value |ε|, and ξ
i
(t) are com-
ponents of a contravariant vector field along c(t). Substitution of (3.1) into
(1.8), asking for ˜c(t) = (˜x
i
(t)) to be also a trajectory of (1.8), and letting
ε 0 yields to the so-called variational equations:
(3.2)
d
2
ξ
i
dt
2
+ 2
G
i
x
j
ξ
j
+ 2
G
i
y
j
j
dt
= 0.
Theorem 3.1. For the variational equations (3.2) we have the equiva-
lent invariant form (Jacobi equations):
(3.3)
2
ξ
i
+ (R
i
jk
dx
k
dt
+ E
i
|j
)ξ
j
= 0.
414 P.L. ANTONELLI and I. BUC
˘
ATARU 10
A vector field (ξ
i
(t)) along a path c of the semispray S is called a Jacobi
vector field if it satisfies (3.3).
Proof. Denote by:
(3.4) B
i
j
= 2
G
i
x
j
S(
G
i
y
j
)
G
i
y
r
G
r
y
j
.
It can be proved that B
i
j
is a (1,1)-type Finsler tensor field. It has been
introduced in [6], for the homogeneous case. This tensor field is called the
second invariant of the given SODE in [12], [7], and [8], or the Jacobi endo-
morphism in [9]. It is easy to check that the equations (3.2) are equivalent
to:
(3.5)
2
ξ
i
+ B
i
j
ξ
j
= 0.
All we have to prove now is the following expression of the second invariant:
(3.6) B
i
j
= R
i
jk
y
k
+ E
i
|j
.
Let X
i
(x, y) be an arbitrary Finsler vector field, and consider the vector
field
(3.7)
˜
X = X
i
x
i
+ S(X
i
)
y
i
on T M . We have then:
(3.8) [S,
˜
X] = (
2
X
i
+ B
i
j
X
j
)
y
j
.
If we consider the expression of S and
˜
X in the Berwald basis S = y
i
δ
δx
i
E
i
y
i
and
˜
X = X
i
δ
δx
i
+ X
i
y
i
, respectively, then the bracket [S ,
˜
X] can
be expressed as follows:
(3.9) [S,
˜
X] = {∇
2
X
i
+ (R
i
jk
y
k
+ E
i
|j
)X
j
}
y
i
.
If we compare (3.8) and (3.9) and we take into account that X
i
(x, y) is an
arbitrary Finsler vector field, then the second invariant B
i
j
can be expressed
as in (3.6).
11 GEOMETRIC INVARIANTS IN KCC- THEORY 415
Definition 3.1. 1
o
A Lie symmetry of the semispray S is a vector field
X on the base manifold M such that [S, X
c
] = 0, where X
c
is the complete
lift of X.
2
o
A dynamical symmetry of the semispray S is a vector field
˜
X on T M
such that [S,
˜
X] = 0.
If X χ(M ) is a Lie symmetry of S then X
c
is a dynamical symmetry of
S. As for X χ(M) we have that X
c
= 2X
h
+ [S, X
v
], then X is a Lie
symmetry of S if and only if
(3.10) 2[S, X
h
] + [S, [S, X
v
]] = 2L
S
X
h
+ L
S
L
S
X
v
= 0.
Theorem 3.2.
1
o
A vector field
˜
X = X
i
(x, y)
δ
δx
i
+ Y
i
(x, y)
y
i
is a dynamical symmetry of
S if and only if
(3.11) Y
i
= X
i
, and
2
X
i
+ B
i
j
X
j
= 0.
2
o
If
˜
X = X
i
(x, y)
δ
δx
i
+ Y
i
(x, y)
y
i
is a dynamical symmetry of the semis-
pray S and c(t) = (x
i
(t)) is a path of S, then the restriction of X
i
(x, y)
along ˜c(t) = (x
i
(t),
dx
i
dt
(t)) is a Jacobi vector field for S.
Proof. 1
o
If we express the Lie bracket [S,
˜
X] using the Berwald basis,
we have
(3.11)
0
[S,
˜
X] = (X
i
Y
i
)
δ
δx
i
+ (Y
i
+ B
i
j
X
j
)
y
j
.
So,
˜
X is a dynamical symmetry of S if and only if (3.11) is true.
2
o
If X
i
(x, y) are the horizontal components of a dynamical symmetry
˜
X,
then
2
X
i
+ (R
i
jk
y
k
+ E
i
|j
)X
j
= 0. The restriction of this along the curve
˜c give us the equations (3.3), and then X
i
is a Jacobi vector field along c.
The Jacobi equations (3.3) are the invariant form of the variational
equations (3.2) using the dynamical covariant derivative. Also, in (3.11) we
found the invariant equations of dynamical symmetries (or Lie symmetries)
in terms of dynamical covariant derivative. Next we shall rewrite these
equations, using the Berwald covariant derivative.
As X
i
= DX
i
+E
i
|
j
X
j
, we have that
2
X
i
= (DX
i
)+(E
i
|
j
X
j
) =
D
2
X
i
+ 2E
i
|
j
DX
j
+ [D(E
i
|
j
) + E
i
|
k
E
k
|
j
]X
j
.
416 P.L. ANTONELLI and I. BUC
˘
ATARU 12
Here we have used that D(E
i
|
j
X
j
) = D(E
i
|
j
)X
j
+ E
i
|
j
DX
j
. Denote by
(3.12) R
i
j
= D(E
i
|
j
) + E
i
|
k
E
k
|
j
+ R
i
jk
y
k
+ E
i
|j
.
We have that R
i
j
is a (1, 1)-type Finsler tensor field. Thus we have the
following theorem:
Theorem 3.3.
1
o
The Jacobi equations of the system (1.8) have the invariant form, using
the Berwald covariant derivative:
(3.13) D
2
ξ
i
+ 2E
i
|
j
Dξ
j
+ R
i
j
ξ
j
= 0.
2
o
A Finsler vector field X
i
(x, y) is a dynamical symmetry of the semispray
S if and only if:
(3.14) D
2
X
i
+ 2E
i
|
j
DX
j
+ R
i
j
X
j
= 0.
For the homogeneous case, that is E
i
= 0, we have that the dynamical
covariant derivative and the Berwald covariant derivative coincide, and the
(1, 1)-type Finsler tensor fields B
i
j
and R
i
j
are equal.
4. The geometric invariants of a semispray S. In KCC-theory
of a semispray ([12], [7], [8]), there are five geometric invariants. The first
KCC-invariant is E
i
, defined by (1.5). The second KCC-invariant is B
i
j
,
defined by (3.4). The third, fourth, and fifth invariants are:
(4.1)
B
i
jk
:=
1
3
(
B
i
j
y
k
B
i
k
y
j
),
B
i
lkj
:=
B
i
jk
y
l
,
D
i
jkl
:=
F
i
jk
y
l
=
3
G
i
y
j
y
k
y
l
.
The Tensor D
i
jkl
is called the Douglas tensor, and we already saw in (2.8)
that it is one of the nonzero components of the curvature of the Berwald
connection.
13 GEOMETRIC INVARIANTS IN KCC- THEORY 417
Theorem 4.1.
1
o
The curvature R
i
jk
of t he nonlinear connection N (or the (v)h-torsion of
the Berwald connection D) is the third invariant of the semispray S.
2
o
The Riemann-Christoffel curvature tensor R
i
jkl
of the Berwald connection
D is the fourth invariant of the semispray S.
Proof. We have to prove that R
i
jk
= B
i
jk
and R
i
jkl
= B
i
jkl
. First we
prove that R
i
jk
and R
i
jkl
satisfy (4.1)
2
, that is R
i
lkj
=
R
i
jk
y
l
. From (2.6) we
have R
k
ij
=
δN
k
j
δx
i
δN
k
i
δx
j
, so
R
k
ij
y
l
=
y
l
(
δN
k
j
δx
i
)
y
l
(
δN
k
i
δx
j
). As [
δ
δx
j
,
y
i
] =
F
k
ji
y
k
, we have that
R
k
ij
y
l
=
δ
δx
i
(
N
k
j
y
l
)F
p
li
N
k
j
y
p
δ
δx
j
(
N
k
i
y
l
)+F
p
lj
N
k
i
y
p
=
δF
k
jl
δx
i
δF
k
il
δx
j
+ F
p
jl
F
k
pi
F
p
il
F
k
jp
= R
k
lji
. According to (3.6) we have for the
second invariant B
i
j
, the expression B
i
j
= R
i
jk
y
k
+ E
i
|j
. So,
B
i
j
y
k
= R
i
jk
+
R
i
klj
y
l
+E
i
|j
|
k
. Then,
B
i
j
y
k
B
i
k
y
j
= 2R
i
jk
+R
i
klj
y
l
+R
i
jkl
y
l
+E
i
|j
|
k
E
i
|k
|
j
. Using
the Ricci identities for the B erwald connection D, ([17], p.78) we have that
E
i
|j
|
k
E
i
|
k|j
= D
i
ljk
E
l
, and E
i
|k
|
j
E
i
|
j|k
= D
i
lkj
E
l
. As the Douglas tensor
is symmetric we have that: E
i
|j
|
k
E
i
|
k|j
= E
i
|k
|
j
E
i
|
j|k
. Consequently, we
have, E
i
|j
|
k
E
i
|k
|
j
= E
i
|
j|k
E
i
|
k|j
= R
i
lkj
y
l
R
i
kj
. Finally, we have that:
B
i
j
y
k
B
i
k
y
j
= 3R
i
jk
+ (R
i
klj
+ R
i
jkl
+ R
i
ljk
)y
l
. Using the Bianchi identity
for the Berwald connection D, we have that R
i
klj
+ R
i
jkl
+ R
i
ljk
= 0, so that
R
i
jk
=
1
3
(
B
i
j
y
k
B
i
k
y
j
), and the theorem is proved.
5. A particular case. In this section we use the theory developed in
the previous sections in order to study the system:
(5.1)
d
2
x
i
dt
2
+ γ
i
jk
(x)
dx
j
dt
dx
k
dt
+ γ
i
j
(x)
dx
j
dt
+ γ
i
(x) = 0,
where γ
i
jk
(x) are the lo c al coeffic ients of a linear symmetric connection on
the base manifold M , γ
i
j
(x) are the components of a (1,1)-type tensor field
418 P.L. ANTONELLI and I. BUC
˘
ATARU 14
and γ
i
(x) are the components of a vector field. The system (5.1) occurs in
modeling some processes in biology ([3]).
On a local chart on T M we define the functions 2G
i
(x, y) = γ
i
jk
(x)y
j
y
k
+
γ
i
j
(x)y
j
+ γ
i
(x). I t is very easy to check that G
i
are the local components
of a semispray S = y
i
x
i
2G
i
y
i
on T M. Let N be the nonlinear
connection induced by S. Then the local coefficients of N are N
i
j
(x, y) =
γ
i
jk
(x)y
k
+
1
2
γ
i
j
(x). The first invariant of the system (5.1) is given by:
(5.2) E
i
(x, y) =
1
2
γ
i
j
(x)y
j
+ γ
i
(x).
The Berwald connection D, that corresponds to the semispray S has the
local coefficients F
i
jk
= γ
i
jk
, C
i
jk
= 0. So, an immediate consequence is that
the fifth invariant of the system (5.1), the Douglas tensor D
i
jkl
vanishes and
the fourth invariant R
i
jkl
is the curvature tensor of the linear connection
γ
i
jk
(x). Now let us express the second and the third invariants of the system.
First we have to remark that the h and v-covariant derivatives of the
first invariant E
i
with respect to the Berwald D connection are given by:
(5.3)
E
i
|
j
=
1
2
γ
i
j
,
E
i
|j
=
1
2
γ
i
k|j
y
k
+ γ
i
|j
1
4
γ
i
k
γ
k
j
.
As we saw in the previous section, the third invariant is R
i
jk
, the curvature
of the nonlinear connection N . As R
k
ij
=
δN
k
j
δx
i
δN
k
i
δx
j
, by a straightforward
calculation we get that
(5.4) R
i
jk
(x, y) = R
i
lkj
(x)y
l
+
1
2
(γ
i
k|j
(x) γ
i
j|k
(x)).
The second invariant B
i
j
can be expressed, using (3.6), as follows:
(5.5) B
i
j
= R
i
lkj
y
l
y
k
+ γ
i
k|j
y
k
+ γ
i
|j
1
2
γ
i
j|k
y
k
1
4
γ
i
k
γ
k
j
.
All these invariants can be obtained also, using Schouten’s film space idea,
([18], [4]), and Kron’s work ([13]).
15 GEOMETRIC INVARIANTS IN KCC- THEORY 419
Acknowledgements. This paper has be en written during the second
author’s PDF at the Department of Mathematics, University of Alberta,
Edmonton, Canada.
REFERENCES
1. Anastasiei, m. and Bucataru, i. Jacobi fields in ge neralized Lagrange spaces,
Rev. Roumaine Math. P ures Appl., 42(1997), no. 9-10, 689-695.
2. Antonelli, p.l., Auger, p. and Bradbury, r.h. Corals and Starfish Waves on
the Great Barrier Reef: Analytical Trophodynamics and 2-Patch Aggregation Meth-
ods, Mathl. Comput. Modeling, 27, 4(1998), 121-135.
3. Antonelli, p.l. and Bradbury, r.h. Voltera- Hamilton Models in the Ecology
and Evolution of Colonial Organisms, World Scientific, 1996.
4. Antonelli, p.l. and Buc
˘
ataru, i. Voltera- Hamilton production models with
discounting: general theory and worked examples, Nonlinear Analysis: Real World
Applications, 2(2001), 337- 356.
5. Antonelli, p.l., Ingarden, r.s. and Matsumoto, m. The Theory of Sprays and
Finsler Spaces with Applications in Physics and Biology, Kluwer Academic Press,
Dordrecht-Boston-London, 1993.
6. Berwald, l. On systems of second order ODE’s whose integral curves are topo-
logically equivalent to the system of straight lines, Annals of Mathematics 48, no.22
(1947), 193-215.
7. Cartan, e. Observations sur le emoir pr´ec ´edent, Math. Zeitschrift 37(1933),
619-622.
8. Chern, s.s. Sur la eometrie d’un syst`eme d’equations differentialles du second
ordre, Bull. Sci. Math. 63 (1939), 206-212.
9. Crampin, m., martinez, e. and sarlet, w. Linear connections for systems
of second-order ordinary differential equations, Ann. Inst. Henri Poincare 65, no.2
(1996), 223-249.
10. Douglas, j. The general geometry of paths, Ann of Math., 29(1928), 143-169.
11. Grifone, j. Structure presque-tangente et connexions I, II, Ann. Inst. Henri
Poincare 22, no.1 (1972), 287-334, 22, no.3 (1972), 291-338.
12. Kosambi, d.d. Parallelism and path-space, Math.Zeitschrift 37 (1933), 608-618.
420 P.L. ANTONELLI and I. BUC
˘
ATARU 16
13. Kron, g. A physical interpretation of the Riemann- Christoffel curvature (the dis-
tribution of dumping and synchronizing torques in oscillatory transmission system),
Tensor, N.S. 4(1955), 150-172.
14. Lackey, b. A model of trophodynamics, Nonlinear Analysis, 35, 1(1999), 37-57.
15. Matsumoto, m. Foundations of Finsler geometry and special Finsler spaces, Kai-
seisha Press, 1986.
16. Miron, r. and anastasiei, m. The Geometry of Lagrange Spaces: Theory and
Applications, Kluwer Academic Press, Dordrecht-Boston-London, 1994.
17. Miron, r. and an astasiei, m. Vector bundles and Lagrange spaces with applica-
tions to relativity, Geometry Balkan Press, 1997.
18. Schouten, j.a. Tensor Analysis for Physicists, Clarendon Press, Oxford, 1951.
Received: 20.IV.2001 Department of Mathematical Sciences
University of Alberta
Edmonton, Alberta
CANADA, T6G 2G1
pa2@gpu.srv.ualberta.ca
Faculty of Mathematics
”Al.I.Cuza University
Ia¸si, 6600
ROMANIA
bucataru@mail.uaic.ro
... Of these five geometrical invariants, the most relevant, from a mathematical and physical point of view, is the second invariant P i j , which is named the deviation curvature tensor. From the general perspective of the scientific, engineering and mathematical applications, its importance is given by its main property of determining the so-called Jacobi stability of the considered system of second order nonlinear differential equation [8][9][10][11][12]. Various engineering, physical, chemical, biochemical, or medical systems have been thoroughly investigated with the help of the KCC stability theory [13][14][15][16][17][18][19][20][21][22][23][24][25][26][27]. ...
... , n give the components of the external force F, which cannot be derived from a potential. By assuming that the Lagrangian L is regular, by using some simple calculations, we obtain the fundamental results that the Euler-Lagrange equations Equation (12) are equivalent mathematically to a complicated system of second-order ordinary, strongly nonlinear differential equations, given by [8,24] ...
... Now, we consider the open subset Ω ⊆ R n × R n × R 1 on which the system of differential equation (13) is defined, together with the coordinate transformations, defined by Equation (11), and assumed to be non-singular. On this mathematical structure, we can define the KCC-covariant derivative of an arbitrary vector field ξ i (x) by means of the definition [9][10][11][12], ...
Article
Full-text available
We analyze the stability of the geodesic curves in the geometry of the Gibbons–Maeda–Garfinkle–Horowitz–Strominger black hole, describing the space time of a charged black hole in the low energy limit of the string theory. The stability analysis is performed by using both the linear (Lyapunov) stability method, as well as the notion of Jacobi stability, based on the Kosambi–Cartan–Chern theory. Brief reviews of the two stability methods are also presented. After obtaining the geodesic equations in spherical symmetry, we reformulate them as a two-dimensional dynamic system. The Jacobi stability analysis of the geodesic equations is performed by considering the important geometric invariants that can be used for the description of this system (the nonlinear and the Berwald connections), as well as the deviation curvature tensor, respectively. The characteristic values of the deviation curvature tensor are specifically calculated, as given by the second derivative of effective potential of the geodesic motion. The Lyapunov stability analysis leads to the same results. Hence, we can conclude that, in the particular case of the geodesic motion on circular orbits in the Gibbons–Maeda–Garfinkle–Horowitz–Strominger, the Lyapunov and the Jacobi stability analysis gives equivalent results.
... (11), and assumed to be non-singular. On this mathematical structure we can define the KCC-covariant derivative of an arbitrary vector field ξ i (x) by means of the definition [9][10][11][12], ...
... (18) into Eqs. (13), and by considering the limit η → 0, we arrive to the deviation, or Jacobi, equations, representing the central mathematical result of the KCC theory, and which are given by [9][10][11][12] ...
... In Eq. (21) we have defined the important tensor G i jl , given by [8][9][10][11][12]35] ...
Preprint
Full-text available
We analyze the stability of the geodesic curves in the geometry of the Gibbons-Maeda-Garfinkle-Horowitz-Strominger black hole, describing the space time of a charged black hole in the low energy limit of the string theory. The stability analysis is performed by using both the linear (Lyapunov) stability method, as well as the notion of Jacobi stability, based on the Kosambi-Cartan-Chern theory. Brief reviews of the two stability methods are also presented. After obtaining the geodesic equations in spherical symmetry, we reformulate them as a two-dimensional dynamic system. The Jacobi stability analysis of the geodesic equations is performed by considering the important geometric invariants that can be used for the description of this system (the nonlinear and the Berwald connections), as well as the deviation curvature tensor, respectively. The characteristic values of the deviation curvature tensor are specifically calculated, as given by the second derivative of effective potential of the geodesic motion. The Lyapunov stability analysis leads to the same results. Hence, we can conclude that in the particular case of the geodesic motion on circular orbits in the Gibbons-Maeda-Garfinkle-Horowitz-Strominger, the Lyapunov and the Jacobi stability analysis gives equivalent results.
... This approach of Jacobi stability appears as a prolongation of the geometric stability approach of the geodesic flow, from a Riemann or Finsler manifold to a differentiable manifold without any metric [24][25][26][27][28][29]. More precisely, the concept of Jacobi stability plays the role of proof of the resilience of a dynamical system defined by a system of second-order differential equations (semi-spray or SODE), where this resilience reflects the adaptability and the preservation of the system's basic behavior to changes in internal parameters and to influences from outside circumstances. ...
... Once we have given a semi-spray, we can associate a nonlinear connection on the differential manifold, and conversely, by giving a nonlinear connection, we can associate a semi-spray. In conclusion, starting with any SODE (or semi-spray), we can construct a geometry on the base manifold by using the geometric associated objects [27,[41][42][43]. Moreover, all these geometrical objects are invariant with respect to any local coordinates change, i.e., they are tensors that can fulfilled the symmetry conditions or skew-symmetry conditions, or not, according to the type of the second-order differential equations system. ...
Article
Full-text available
In this paper will be considered a three-dimensional autonomous quadratic polynomial system of first-order differential equations with three real parameters, the so-called T-system. This system is symmetric relative to the Oz-axis and represents a special type of the generalized Lorenz system. The approach of this work will consist of the study of the nonlinear dynamics of this system through the Kosambi–Cartan–Chern (KCC) geometric theory. More exactly, we will focus on the associated system of second-order differential equations (SODE) from the point of view of Jacobi stability by determining the five invariants of the KCC theory. These invariants determine the internal geometrical characteristics of the system, and particularly, the deviation curvature tensor is decisive for Jacobi stability. Furthermore, we will look for necessary and sufficient conditions that the system parameters must satisfy in order to have Jacobi stability for every equilibrium point.
... The Jacobi stability is a natural generalization of the geometric approaches related to the stability of the geodesic flow, from a Riemannian manifold of a Finslerian manifold to a manifold with no metric [10][11][12][13][14][15]. More exactly, the Jacobi stability is an indicator of the vigour of a dynamical system given by a system of second-order differential equations (SODE or semi-spray), where this vigour represents the adaptation and the conservation of the basis behaviour to both the changes of the internal parameters of the system and the influences from the external environment. ...
... Starting with a semi-spray, we can define a nonlinear connection on the manifold, and, conversely, by using a nonlinear connection, we can define a semi-spray. Therefore, to any semi-spray (or SODE) we can associate a geometry on the manifold by the corresponding geometric objects [13,[30][31][32]. Conversely, these geometric objects are invariant relative to local coordinate changes, which means that they are tensors that can satisfy the conditions of symmetry or skew-symmetry or neither, depending on the form of the system of second-order differential equations (SODE). ...
Article
Full-text available
In the present work, two SIR patterns with demography will be considered: the classical pattern and a modified pattern with a linear coefficient of the infection transmission. By reformulating of each first-order differential systems as a system with two second-order differential equations, we will examine the nonlinear dynamics of the system from the Jacobi stability perspective through the Kosambi–Cartan–Chern (KCC) geometric theory. The intrinsic geometric properties of the systems will be studied by determining the associated geometric objects, i.e., the zero-connection curvature tensor, the nonlinear connection, the Berwald connection, and the five KCC invariants: the external force εi—the first invariant; the deviation curvature tensor Pji—the second invariant; the torsion tensor Pjki—the third invariant; the Riemann–Christoffel curvature tensor Pjkli—the fourth invariant; the Douglas tensor Djkli—the fifth invariant. In order to obtain necessary and sufficient conditions for the Jacobi stability near each equilibrium point, the deviation curvature tensor will be determined at each equilibrium point. Furthermore, we will compare the Jacobi stability with the classical linear stability, inclusive by diagrams related to the values of parameters of the system.
... The classical stability (linear or Lyapunov stability) of the Rosenzweig-MacArthur predator-prey system was completely studied in [11,12]. In this work, we will study another type of stability for this predator-prey system for the first time, namely the Jacobi stability.The Jacobi stability is a natural extension of the geometric stability of the geodesic flow, from a manifold with a Riemann metric or a Finsler metric to a manifold with no metric [14][15][16][17][18][19]. Practically, the Jacobi stability indicates the robustness of a dynamical system defined by a system of second-order differential equations (or SODE), where this robustness measures the lack of sensitivity and adaptation to both the modifications of the internal parameters of the system and the change in the external environment. ...
... Using a semispray, we can define a nonlinear connection on the manifold, and conversely, using a nonlinear connection, we can define a semispray. Consequently, any SODE can define a geometry on the manifold with the associated geometric objects, and the other way around [17,[31][32][33]. ...
Article
Full-text available
In this paper, we consider an autonomous two-dimensional ODE Kolmogorov-type system with three parameters, which is a particular system of the general predator–prey systems with a Holling type II. By reformulating this system as a set of two second-order differential equations, we investigate the nonlinear dynamics of the system from the Jacobi stability point of view using the Kosambi–Cartan–Chern (KCC) geometric theory. We then determine the nonlinear connection, the Berwald connection, and the five KCC invariants which express the intrinsic geometric properties of the system, including the deviation curvature tensor. Furthermore, we obtain the necessary and sufficient conditions for the parameters of the system in order to have the Jacobi stability near the equilibrium points, and we point these out on a few illustrative examples.
... [Antonelli et al., 1993;Yamasaki & Yajima, 2017;Gupta & Yadav, 2019;Huang et al., 2019;Chen et al., 2020]). KCC theory has been well studied in the field of differential geometry, especially Finsler geometry [Antonelli & Bucataru, 2001;Balan & Neagu, 2010;Neagu, 2013], and it has been applied to various fields such as biology [Antonelli et al., 2002] and physics [Boehmer & Harko, 2010;Harko & Sabȃu, 2008;Harko et al., 2015;Dȃnilȃ et al., 2016;Gupta & Yadav, 2017b], among others [Liu et al., 2021b]. For instance, Yamasaki and Yajima [2020] performed KCC analysis of bifurcation phenomena including catastrophe. ...
Article
Full-text available
This paper analyzes the properties of the nonequilibrium singular point in one-dimensional elementary catastrophe. For this analysis, the Kosambi–Cartan–Chern (KCC) theory is applied to characterize the dynamical system based on differential geometrical quantities. When both the nonlinear connection and deviation curvature are zero, that is, when the geometric stability of the KCC theory is neutral, two bifurcation curves are obtained: one is the known curve with an equilibrium singular point, and the other is a new curve with a nonequilibrium singular point. The two singular points are distinguished based on the vanishing condition of the Berwald connection. Applied to the ecosystem described by the Hill function, the absolute value of the cuspidal curvature of the nonequilibrium singular point is larger than that of the equilibrium singular point. The ecological interpretation of this result is that the range of bistability of the ecosystem in the nonequilibrium state is greater than that in the equilibrium state. The type of singular points in equilibrium and nonequilibrium bifurcation curves are not necessarily the same. For instance, there is a combination in which even if the former has one cusp, the latter may show various types, depending on the parametric space. These results demonstrate that there are cases where simply shifting the system from the equilibrium to nonequilibrium state expands the range of bistability and changes the type of singularity. Although singularity analysis is often performed near the equilibrium point, nonequilibrium analysis, i.e. analysis based on the KCC theory, provides a useful perspective for analyzing singularity theory according to the bifurcation phenomenon.
... The Kosambi-Cartan-Chern geometrical approach (KCC-theory) was developed in details in numerous mathematical books and papers [3][4][5]. KCC-theory allowed to describe the evolution of a dynamical system in a configuration space of the Lagrange type. At that, the dynamical system is governed by the system of second-order differentials equations. ...
Article
Full-text available
The geometrical Kosambi–Cartan–Chern approach has been applied to study the systems of differential equations which arise in quantum-mechanical problems of a particle on the background of non-Euclidean geometry. We calculate the geometrical invariants for the radial system of differential equations arising for electromagnetic and spinor fields on the background of the Schwarzschild spacetime. Because the second invariant is associated with the Jacobi field for geodesics deviation, we analyze its behavior in the vicinity of physically meaningful singular points r = M, ∞. We demonstrate that near the Schwarzschild horizon r = M the Jacobi instability exists and geodesics diverge for both considered problems.
Article
Full-text available
In this study, we discuss Jacobi stability in equilibrium and nonequilibrium regions for a first-order one-dimensional system using deviation curvatures. The deviation curvature is calculated using the Kosambi-Cartan-Chern theory, which is applied to second-order differential equations. The deviation curvatures of the first-order one-dimensional differential equations are calculated using two methods as follows. Method 1 is only differentiating both sides of the equation. Additionally, Method 2 is differentiating both sides of the equation and then substituting the original equation into the second-order system. From the general form of the deviation curvatures calculated using each method, the analytical results are obtained as (A), (B), and (C). (A) Equilibrium points are Jacobi unstable for both methods; however, the type of equilibrium points is different. In Method 1, the equilibrium point is a nonisolated fixed point. Conversely, the equilibrium point is a saddle point in Method 2. (B) When there is a Jacobi stable region, the size of the Jacobi stable region in the Method 1 is different from that in Method 2. Especially, the Jacobi stable region in Method 1 is always larger than that in Method 2. (C) When there are multiple equilibrium points, the Jacobi stable region always exists in the nonequilibrium region located between the equilibrium points. These results are confirmed numerically using specific dynamical systems, which are given by the logistic equation and its evolution equation with the Hill function. From the results of (A) and (B), differences in types of equilibrium points affect the size of the Jacobi stable region. From (C), the Jacobi stable regions appear as nonequilibrium regions where the equations cannot be linearized.
Preprint
Full-text available
In this work, we consider two SIR patterns with demography: the classical pattern and a modified pattern with a linear transmission coefficient of the infection. By reformulating of each first order differential systems as a system with two second-order differential equations, we investigate the nonlinear dynamics of the system from the Jacobi stability point of view by using the KCC geometric theory. We will study the intrinsic geometric properties of the systems by determining the geometric associated objects: the zero-connection curvature tensor, the nonlinear connection, the Berwald connection, and the five KCC invariants: the first invariant - the external force εi, the second invariant - the deviation curvature tensor Pji, the third invariant - the torsion tensor Pjki, the fourth invariant - the Riemann-Christoffel curvature tensor Pjkli, and the fifth invariant - the Douglas tensor Djkli. In order to obtain necessary and sufficient conditions for the Jacobi stability near the equilibrium points, the deviation curvature tensor will be determined at each equilibrium points. Furthermore, we will compare the Jacobi stability with the classical linear stability, inclusive by diagrams related to the values of parameters of the system.
Article
Full-text available
In this work, we will consider an autonomous three-dimensional quadratic system of first-order ordinary differential equations, with five parameters and with symmetry relative to the z-axis, which generalize the Hopf–Langford system. By reformulating the system as a system of two second-order ordinary differential equations and using the Kosambi–Cartan–Chern (KCC) geometric theory, we will investigate this system from the perspective of Jacobi stability. We will compute the five invariants of KCC theory which determine the own geometrical properties of this system, especially the deviation curvature tensor. Additionally, we will search for necessary and sufficient conditions on the five parameters of the system in order to reach the Jacobi stability around each equilibrium point.
Book
Full-text available
Pure theory of Finsler geometry with applications in Physics and biology..
Article
Full-text available
We consider the first variation of those curves on tangent manifold T M which have property that are parallel with respect to the canoni-cal metrical connection in a generalised Lagrange space. Accordingly we introduce and study the Jacobi fields on T M. Several particular cases are discussed.
Article
Full-text available
In this paper, singular perturbation theory is applied to the Antonelli/Kazarinoff limit cycle model of the Crown-of-Thorns starfish (COTS) predation of corals on the Great Barrier Reef. At the microscale of individual reefs, the previously published 2-patch dynamics based on aggregation of individual behaviour is extended to include social interactions, spatial diffusion, and advection currents. As a consequence, the parameters which characterize the well-known wave solutions (i.e., the Reichelt starfish waves) are reinterpreted in terms of individual behaviour parameters. Likewise, the analytical trophodynamics of coral community production (without starfish), involving geometric invariants of second-order ODEs is defined at the microscale by direct manipulation of the reefal structure. The aggregation method leads to new insights into Liapunov stability of production of the coral community as a whole.