In this paper, we investigate Zermelo's navigation problem for some special rotation surfaces. In this respect, we find some Randers-type metrics for these rotation surfaces. Furthermore, we get the H-distortion for the metric induced by surfaces.
Citation: Yanlin Li, Piscoran Laurian-Ioan, Lamia Saeed Alqahtani, Ali H. Alkhaldi, Akram Ali. Zermelo's navigation problem for some special surfaces of rotation[J]. AIMS Mathematics, 2023, 8(7): 16278-16290. doi: 10.3934/math.2023833
[1] | Jia Tang, Yajun Xie . The generalized conjugate direction method for solving quadratic inverse eigenvalue problems over generalized skew Hamiltonian matrices with a submatrix constraint. AIMS Mathematics, 2020, 5(4): 3664-3681. doi: 10.3934/math.2020237 |
[2] | Limin Li . Normwise condition numbers of the indefinite least squares problem with multiple right-hand sides. AIMS Mathematics, 2022, 7(3): 3692-3700. doi: 10.3934/math.2022204 |
[3] | Jamilu Sabi'u, Ibrahim Mohammed Sulaiman, P. Kaelo, Maulana Malik, Saadi Ahmad Kamaruddin . An optimal choice Dai-Liao conjugate gradient algorithm for unconstrained optimization and portfolio selection. AIMS Mathematics, 2024, 9(1): 642-664. doi: 10.3934/math.2024034 |
[4] | Qian Li, Qianqian Yuan, Jianhua Chen . An efficient relaxed shift-splitting preconditioner for a class of complex symmetric indefinite linear systems. AIMS Mathematics, 2022, 7(9): 17123-17132. doi: 10.3934/math.2022942 |
[5] | André M. McDonald, Michaël A. van Wyk, Guanrong Chen . The inverse Frobenius-Perron problem: A survey of solutions to the original problem formulation. AIMS Mathematics, 2021, 6(10): 11200-11232. doi: 10.3934/math.2021650 |
[6] | Man Chen, Huaifeng Chen . On ideal matrices whose entries are the generalized $ k- $Horadam numbers. AIMS Mathematics, 2025, 10(2): 1981-1997. doi: 10.3934/math.2025093 |
[7] | Haixia Chang, Chunmei Li, Longsheng Liu . Generalized low-rank approximation to the symmetric positive semidefinite matrix. AIMS Mathematics, 2025, 10(4): 8022-8035. doi: 10.3934/math.2025368 |
[8] | Yinlan Chen, Lina Liu . A direct method for updating piezoelectric smart structural models based on measured modal data. AIMS Mathematics, 2023, 8(10): 25262-25274. doi: 10.3934/math.20231288 |
[9] | Ximing Fang . The simplified modulus-based matrix splitting iteration method for the nonlinear complementarity problem. AIMS Mathematics, 2024, 9(4): 8594-8609. doi: 10.3934/math.2024416 |
[10] | Luca Giorgetti, Wei Yuan, XuRui Zhao . Separable algebras in multitensor C$ ^* $-categories are unitarizable. AIMS Mathematics, 2024, 9(5): 11320-11334. doi: 10.3934/math.2024555 |
In this paper, we investigate Zermelo's navigation problem for some special rotation surfaces. In this respect, we find some Randers-type metrics for these rotation surfaces. Furthermore, we get the H-distortion for the metric induced by surfaces.
In an affine space A3, we will consider the following proper affine surfaces of revolution by (u,v)∈R2, local coordinates. With Ψ:Ω⊂R2→R3 a nondegenerate Blaschke immersion, we can consider a convenient coordinate system (x1,x2,x3) of the following surfaces (for more details, see [1]).
Elliptic type: If Ψ is of elliptic type, in local coordinates there is the following representation:
Ψ(u,v)=(f(u)cosv,f(u)sinv,g(u)), | (1.1) |
Ψ(u,v)=(f(u)cosv,f(u)sinv,u), | (1.2) |
Ψ(u,v)=(ucosv,usinv,g(u)). | (1.3) |
Hyperbolic type: If Ψ is of hyperbolic type, in local coordinates there is the following representation:
Ψ(u,v)=(f(u)coshv,f(u)sinhv,g(u)), | (1.4) |
Ψ(u,v)=(f(u)coshv,f(u)sinhv,u), | (1.5) |
Ψ(u,v)=(ucoshv,usinhv,g(u)). | (1.6) |
Remark 1.1. Also, there is the parabolic type of surface, but in this paper, we will choose just elliptic and hyperbolic surface types.
In some recent papers[1,2,3,4,5,6,7,8,9], Zermelo's problem and the generalized Zermelo's problem were investigated. Under the assumption that the wind is a time-independent mild breeze for a Riemannian manifold (M,h), they found that the paths minimizing travel time are exactly the geodesics of a Randers metric.
F(x,y)=α(x,y)+β(x,y)=√λ|y|2+W20λ−W0λ, |
where W=Wi∂∂xi is the wind velocity, |y|2=h(y,y), λ=1−|W|2, W0=h(W,y). According to [10], the Randers metric F is said to solve the Zermelo navigation problem in the case of a mild breeze, which means h(W,W)<1.
The condition h(W,W)<1 ensures that F is a positive definite metric.
For a positive function f:R→R+, in paper [10], the following rotation surface was considered.
M={(f(u)cosv,f(u)sinv,u),u∈I,0<v≤2π}, |
with
ds2=(1+f′2)du2+f2dv2. |
The coefficients of the first fundamental form are
E=1+f′2,F=0,G=f2. |
The geodesics equation can be found by
¨xk+∑i,jΓkij˙xi˙xj=0, | (1.7) |
that is,
{d2uds2+f′f″2(1+f′2)(duds)2−ff′2(1+f′2)(dvds)2=0,d2vds2+2f′fdudsdvds=0. |
A function f:[0,+∞)→[0,+∞), defined in [10], is constructed as a rotational Randers metric on M, putting W=μ∂∂v in the h-orthogonal system (∂∂u,∂∂v) of the tangent space Tx(M). One obtains W=(W1,W2)=(0,μ)⇒
h(W,W)=h(μ∂∂v,μ∂∂v)=μ2h(∂∂v,∂∂v). |
The navigation data (h,W) gives new data:
aij=λhij+WiWjλ2,bi=−Wiλ, |
with Wi=hijWi, λ=1−h(W,W)=1−μ2f2>0.
Finally, after computations (see [10]), we can obtain
aij=(11−f2μ200f2(1−f2μ2)2),bi=(0−μf21−μ2). |
Here, i,j=1,2. It can be easily observed that α(b,b)=aijbibj=h(W,W)<1, and this ensures that F is a positive Randers metric (see [10]).
The Zermelo navigation problem, as the authors in [11] have remarked, is an important problem in Finsler geometry because it aims to find the minimum time trajectories in a Riemannian manifold (B,h) under the influence of a drift (wind), represented by the vector field W. In [12], authors proved that the trajectories that minimize the travel time are exactly the geodesics of the Randers metric which has the norm:
F(x,v)=√aij(x)vivj+bi(x)vi, |
where v∈TxB.
In fact, it is known that every Randers metric F could be obtained as a solution to a Zermelo navigation problem defined by the data {hij,Wi}. This can be observed by (as we saw in [11]) the following
hij=λ(aij−bibj),λ=1−aijbibj, |
Wi=−biλ,bi=aijbj, |
hij=λaij+bibjλ2, |
Then, it can be observed that there is a natural identification of Randers metrics with solutions to the Zermelo problem:
|W|2=hijWiWj=aijbibj≡|b|2. |
The Finsler condition |b|2≤1 ensures that F is positive and the metric is convex, i.e., ∂vi∂vj(F22) is positive definite for all non-zero v.
Also, it is well known from the literature that the geodesic of a Finsler norm is obtained by taking F as the Finsler Lagrangian. If F(x,v) represents a homogeneous Finsler Lagrangian of degree one, a Lagrangian could be defined as L=12F2. The Hamiltonian will be on degree two in the momenta pi=∂L∂vi=F∂F∂vi, and using [11], we can remark that we can choose a function G(x,p) such that H=12G2, G(x,p)=F(x,v). For the Randers metric F(x,v)=√aij(x)vivj+bi(x)vi, we have
pi=F(ni+bi),ni=aijvj√akrvkvr. |
Since, aijninj=1, we get G=√hijpipj−Wipi, with {hij,Wi}, the associated Zermelo data. Thus, the Legendre transformation maps the Randers (Lagrangian) to the Zermelo data (Hamiltonian). We can see G as the sum of the two moment maps G=G0+G1, G0=√hijpipj, G1=−Wipi.
Also, we can observe that these flows commute {G0,G1}=0 if and only if W is a Killing vector field. We continue to present some other aspects from the paper [11]. The geodesic flow of the Randers metric can be seen as the null geodesic flow in a stationary space-time with a generic form
gμνdxμdxν=−V2(dt+widxi)2+γijdxidxj, |
so that
gij=γij−V2wiwj. |
As is remarked in [11], the Fermat principle arises from the Randers structure given by
aij=V−2γij,aij=V2γij,bi=−wi. |
Thus, the form ds2=V2[−(dt−bidxi)2+aijdxidxj], is called the Randers form of a stationary metric. According to [11], the Randers data
hij=λ(aij−bibj),λ=1−aijbibj, |
Wi=−biλ,bi=aijbj, |
are equivalent to
hij=11+V2grswrwsgijV2,Wi=V2gijwj, |
with
γij=gij+V2wiwj, |
γij=gij−V2wiwj1+V2grswrws, |
wi=gijwj,Wi=V2γijwj1−V2grswrws, |
1−V2γijwiwj=11+V2gijwiwj. |
With the above notations, the generic metric
gμνdxμdxν=−V2(dt+widxi)2+γijdxidxj |
could be put (according to [11]) as follows:
ds2=V21−hijWiWj[−dt2+hij(dxi−Widt)(dxj−Wjdt)], |
which represents the Zermelo form of stationary spacetime. The metric in square brackets from above is of Painlevé-Gullstrand form.
In this part of the paper, we recall some classical results regarding the Frobenius (Hilbert-Schmidt) norm of a matrix. For more results, please see [13]. The Frobenius norm of a matrix A=(aij) is defined as follows:
‖A‖HS=√n∑i=1n∑j=1a2ij. |
Some of its properties are
‖A⋅B‖HS≤‖A‖HS⋅‖B‖HS, |
‖A‖HS≤√rσmin(A), |
where σmin(A) denotes the minimum singular value of A, and r is the rank of A.
Using a similar approach as in paper [10], we will try to solve Zermelo's navigation problem for the following two surfaces:
M1={(f(u)cosv,f(u)sinv,g(u)),u∈I,v∈[0,2π]}, | (2.1) |
M2={(f(u)coshv,f(u)sinhv,g(u)),u∈I,v∈[0,2π]}. | (2.2) |
These two surfaces represent two revolution surfaces of elliptic type and hyperbolic type, respectively. We will start with surface (2.1). We will consider the Randers metric
F(x,y)=α(x,y)+β(x,y)=√λ|y|2+W20λ−W0λ, |
with W=Wi∂∂xi the wind velocity. Here, |y|2=h(y,y), λ=1−|W|2, W0=h(W,y).
For two definite positive functions f,g:R→R+, surfaces (2.1) and (2.2) represent revolution surfaces.
For surface (2.1), after computations, we get the following coefficients for the first fundamental form: E=f′2+g′2, F=0, G=f2. From this, we get
ds2=(f′2+g′2)du2+f2dv2. |
The coefficients of the second fundamental form, L,M,N, are
L=f′g″−f″g′√f′2+g′2,M=0,N=g′f√f′2+g′2 |
Finally, the Gauss curvature of this surface can be obtained as follows:
K=LN−M2EG−F2=g′f′(f′g″−f″g′). |
The Christoffel symbols for this surface are
Γ111=f′f″+g′g″f′2+g′2, |
Γ112=0=Γ211=Γ222, |
Γ212=f′f, |
Γ122=ff′f′2+g′2. |
Remark 2.1. The geodesic equations for this surface are given by
{d2uds2+f′f″+g′g″f′2+g′2(duds)2−ff′f′2+g′2(dvds)2=0,d2vds2+2f′fdudsdvds=0. |
Next, inspired by [10], we will construct a rotational Randers metric on M1, choosing W=μ∂∂v which is in the h-orthogonal coordinate system (∂∂u,∂∂v) on TxM1. Also, we have W=(W1,W2)=(0,μ), λ=1−h(W,W)=1−μf2. Therefore, we can obtain
h(W,W)=h(μ∂∂v,μ∂∂v)=μ2(∂∂v,∂∂v)=μ2f2<1. |
The navigation data (h,W) induce the following:
aij=λhij+WiWjλ2,bi=−Wiλ, |
W1=h11W1+h12W2=0, |
W2=h21W1+h22W2=μf2, |
with h11=f′2+g′2 and h22=f2. Next, one obtains the following:
Lemma 2.1. The Riemannian metric (aij) and the functions (bi) obtained through Zermelo's navigation problem from h and W, for the surface (2.1), are given by
aij=(f′2+g′21−f2μ200f2(1−f2μ2)2),bi=(0−μf21−μ2). | (2.3) |
We remark that α(b,b)=h(W,W)<1.
Remark 2.2. We can observe that in a particular case, for g(u)=u, one can obtain the same metric as in Lemma 3.1 from paper [10].
Let us now choose another coordinates system: (∂∂u,∂∂v) on TxM1. Next, W=μ∂∂u+ϵ∂∂v⇒W=(W1,W2)=(μ,ϵ).
λ=1−h(W,W)=1−h(μ∂∂u+ϵ∂∂v,μ∂∂u+ϵ∂∂v)=1−μ2h(∂∂u,∂∂u)−ϵ2h(∂∂v,∂∂v)=1−μ2(f′2+g′2)−ϵ2f2. |
Therefore, we get
h(W,W)=μ2(f′2+g′2)+ϵ2f2<1. |
Let f and g are bounded positive real-valued functions defined as f,g:R→R+, then there exist positive values μ>0 and ϵ>0 such that f(z)<1ϵ and √f′(z)2+g′(z)2<1μ. Using the previous statement we can formulate the following
Lemma 2.2. The Riemannian metric (aij) and the functions (bi) obtained through Zermelo's navigation problem from h and W, for the surface (2.1), are given by
aij=(f′2+g′21−f2ϵ2−μ2(f′2+g′2)00f21−f2ϵ2−μ2(f′2+g′2)),bi=(0−f2ϵ2−μ2(f′2+g′2)1−f2ϵ2−μ2(f′2+g′2)). | (2.4) |
We remark that α(b,b)=h(W,W)<1.
Now, let us construct the rotational Randers metric on surface M2 given in (2.2), choosing W=μ∂∂v which is in the h-orthogonal coordinate system (∂∂u,∂∂v) on TxM1. Also, we have W=(W1,W2)=(0,μ), λ=1−h(W,W)=1−μf2.
h(W,W)=h(μ∂∂v,μ∂∂v)=μ2h(∂∂v,∂∂v)=μ2f2<1. |
Let us recall that the surface M2 is given by
M2={(f(u)coshv,f(u)sinhv,g(u)),u∈I,v∈[0,2π]}. |
This means that for the first fundamental form, we get the following results:
E=f′2cosh(2v)+g′2,F=ff′sinh(2v),G=f2(1+2sinh(v)2)=f2cosh(2v). |
Thus, we get
ds2=(f′2cosh(2v)+g′2)du2+2ff′sinh(2v)dudv+f2cosh(2v)dv2. |
The coefficients of the second fundamental form, L,M,N, are
L=f′g″−f″g′√f′2+g′2(cosh(2v))2,M=0,N=g′f√f′2+g′2(cosh(2v))2. |
Finally, the Gauss curvature of this surface can be obtained as follows:
K=LN−M2EG−F2=g′(f′g″−f″g′)f′(f′2+g′2(cosh(2v))2). |
Using [10], following the same approach, we will construct a rotational Randers metric on M2, (2.2), choosing W=μ∂∂v, which is in the h-orthogonal coordinates system (∂∂u,∂∂v) on TxM1. Also, we have W=(W1,W2)=(0,μ), λ=1−h(W,W)=1−μf2.
h(W,W)=h(μ∂∂v,μ∂∂v)=μ2h(∂∂v,∂∂v)=μ2f2<1, |
for positive function f:R→R+, f(z)<1μ. Now, we can give the following lemma:
Lemma 2.3. The Riemannian metric (aij) and the functions (bi), obtained through Zermelo's navigation problem from h and W, for surface (2.2), are the following ones:
aij=(f′2cosh(2v)+g′21−μ2f22ff′sinh(2v)1−μ2f22ff′sinh(2v)1−μ2f2f2(1−μ2f2)2),bi=(0−μf21−μ2f2). | (2.5) |
Now, using Lemma 2.1, we will give the main result theorem for this paper.
Theorem 2.1. The Zermelo data for a stationary space-time with respect to metric (2.3) take the following form:
ds2=V21−(f′2+g′2)μ−ϵf2[−dt2+(f′2+g′2)(dx1−μdt)2+f2(dx2−ϵdt)2]. |
Proof. Before we begin to prove this theorem, we can remark that the metric in square brackets has the Painlevé-Gullstrand form, and this can be defined up to a conformal factor due to the fact that Zermelo's problem is interested only in the null geodesic flow. From Lemma 2.1, we deduce the following:
h11=f′2+g′2,h22=f2,W1=0,W2=μf2, |
W1=μ,W2=ϵ,g11=μV2w1,g11=V2w1μ, |
g22=ϵV2w2,g11=V2w2ϵ. |
Using all these above relations and also the fact that the form of Randers metric in stationary space-time is given by
gμνdxμdxν=−V2(dt+widxi)2+γijdxidxj, |
the following form for the Randers metric:
ds2=V21−hijWiWj[−dt2+hij(dxi−Widt)(dxj−Wjdt)]. |
Thus, we get
ds2=V21−(f′2+g′2)μ−ϵf2[−dt2+(f′2+g′2)(dx1−μdt)2+f2(dx2−ϵdt)2]. |
Then, the proof is done.
Remark 2.3. The metric in the square represents the Painlevé-Gullstrand form, and the physical-geometrical interpretation is that the Zermelo data for the metric (2.3) can be encoded in a stationary space-time just by writing this metric in these coordinates. In quantum physics, these kinds of coordinates can be used for various metrics to describe the process of gravitational collapse. In this respect, the link with the Zermelo navigation problem is not an easy but a very challenging problem and could be considered an open problem.
We are ready now to give another main theorem for this paper but this time using the Frobenius norms.
Theorem 2.2. The following inequality takes place for the metric (2.3)
√rσmin(aij)≥1√2√f′2+g′21−f2μ2+f2(1−f2μ2)2. |
Here, σmin(aij) represents the minimum singular eigenvalue for the matrix (aij) and r represents the rank of the matrix.
Proof. Starting with the matrix (2.3),
aij=(f′2+g′21−f2μ200f2(1−f2μ2)2), |
we can easily compute the Frobenius (or Hilbert-Schmidt) norm for this matrix as follows:
‖aij‖HS=√(f′2+g′2)2(1−f2μ2)2+f4(1−f2μ2)4. |
Using the well-known inequality
√A2+B2≥A+B√2, |
we get
√((f′2+g′2)(1−f2μ2))2+(f2(1−f2μ2)2)2≥1√2√f′2+g′21−f2μ2+f2(1−f2μ2)2, |
and this concludes the proof of the theorem.
Now, we will investigate some properties involving the Frobenius (Hilbert-Schmidt) norm for the above-constructed Riemannian metrics. As we know, the Frobenius norm is computed for the matrix A=(aij) as follows:
‖A‖F=√∑i,j|aij|2. |
In [14], we have introduced the notion of H-distortion for pseudo-Riemannian manifolds. From a geometrical point of view, the H-distortion is important because it is directly linked with the classical distortion in Finsler's geometry. We already know the importance of distortion not just in Finsler geometry but also in physics because the distortion could be linked with Tchebychev's potential. In this respect, please see [15].
Definition 2.1. [14] For a pseudo-Riemannian manifold, we will denote by
σF,∇F=σF(x)σ∇F(x)=√det(gij)det(∇2gf), |
the H-distortion, if and only if σ2F,∇F(x)=constant.
Using the same approach as in [14], we will compute for the constructed Riemannian metrics (2.3), (2.4), and the H-distortion for the surface (2.1). As we well know, for this surface, we already obtained the Christoffel coefficients:
Γ111=f′f″+g′g″f′2+g′2, |
Γ112=0=Γ211=Γ222=Γ121, |
Γ212=f′f=Γ221, |
Γ122=ff′f′2+g′2. |
For a function h:M1→R, we know from [14], the following relations:
h,11=∂2h∂u2−(Γ111∂h∂u+Γ212∂h∂v), |
h,12=∂2h∂u∂v−(Γ112∂h∂u+Γ212∂h∂v), |
h,21=∂2h∂v∂u−(Γ121∂h∂u+Γ221∂h∂v), |
h,22=∂2h∂v2−(Γ121∂h∂u+Γ222∂h∂v). |
With the above Christoffel coefficients obtained before, we get the Hessian matrix
∇2h=(∂2h∂u2−(f′f″+g′g″f′2+g′2∂h∂u+f′f∂h∂v)∂2h∂u∂v−f′f∂h∂v∂2h∂u∂v−f′f∂h∂v∂2h∂v2). |
Using this matrix we will try to find the Hessian matrix for the function h(u,v)=u2+v2, which represents a paraboloid. One obtains: ∂h∂u=2u, ∂h∂v=2v, ∂2h∂u2=2, ∂2h∂v2=2. Finally, we get
∇2h=(2−(f′f″+g′g″)2uf′2+g′2+2vf′f−2vf′f−2vf′f2). |
From this, we get
det(∇2h(u,v))=4(1−(f′f″+g′g″)uf′2+g′2), |
det(aij)=(f′2+g′2)f2(1−μ2f2)3. |
Finally, we get the H-distortion for the function h(u,v), as follows:
σF,∇F=f(f′2+g′2)2(1−μ2f2)√1(1−μ2f2)(f′2+g′2−(f′f″+g′g″)u). |
Imposing the condition from Definition 2.1, we get the following two conditions to have H-distortion for the function h(u,v):
{f(f′2+g′2)2(1−μ2f2)=constant,(1−μ2f2)(f′2+g′2−(f′f″+g′g″)u)=constant. |
Using these equalities, we can construct easily two functions that respect these conditions. For example, f(u)=c=constant and g(u)=u, are two functions that fulfill these conditions. The H-distortion is directly linked with the classical distortion in Finsler's geometry. For this reason, the geometrical significance is huge. The distortion in Finsler geometry is directly linked with the S-curvature because in Finsler geometry the S-curvature was introduced to measure the rate of distortion along geodesics:
S(y)=ddt[τ(˙γ(t))]t=0, |
where γ(t) represent the geodesic with ˙γ(0)=y.
As we have proved in this paper. we successfully obtained some important results regarding the Zermelo navigation problem for some special manifolds. We have also obtained the conditions for the H-distortion for this kind of manifold. The importance of H-distortion and the distortion in Finsler geometry is huge because is in direct link with the S-curvature. The S-curvature is in Finsler geometry an important and unique quantity and geometrically, the S-curvature represents the study of the rate of change of distortion along geodesics. It is known that S-curvature is a non-Riemannian quantity, and this means that every Riemannian manifold has a vanishing S-curvature. For this reason, the results obtained by us in this paper underline the role of these metrics that we found for the rotations surfaces considered. The H-distortion could be computed for all these metrics and also we can compute the S-curvature for all these metrics. This will be done in a future paper. Another important result established by us in this paper was to express one of the constructed metrics in Painlevé-Gullstrand form and we have explained the physical-geometrical interpretation. The Zermelo data for that metric can be encoded in a stationary space-time just by writing this metric in this Painlevé-Gullstrand coordinates. In quantum physics, these kinds of coordinates can be used to describe the formation of the gravitational collapse. Therefore, we recommend to the readers new problems in [15,16,17,18,19,20,21,22,23] and references therein. In this respect, the link with the Zermelo navigation problem is not an easy but a very challenging problem and could be considered an open problem.
The author (Ali. H. AlKhaldi) extends his appreciation to the Deanship of Scientific Research at King Khalid University for funding this work through the Research Groups Program under grant number R. G. P2/199/43. This research was funded by the National Natural Science Foundation of China (Grant No. 12101168) and the Zhejiang Provincial Natural Science Foundation of China (Grant No. LQ22A010014).
The authors declare no conflict of interest.
[1] |
H. Es, Affine rotation surfaces of elliptic type in affine 3-space, Therm. Sci., 24 (2020), 399–409. http://dx.doi.org/10.2298/TSCI200405260E doi: 10.2298/TSCI200405260E
![]() |
[2] | D. Bao, S. Chern, Z. Shen, An introduction to Riemann-Finsler geometry, New York: Springer-Verlag, 2000. http://dx.doi.org/10.1007/978-1-4612-1268-3 |
[3] |
D. Brody, D. Meier, Solution to the quantum Zermelo navigation problem, Phys. Rev. Lett., 114 (2015), 100502. http://dx.doi.org/10.1103/PhysRevLett.114.100502 doi: 10.1103/PhysRevLett.114.100502
![]() |
[4] |
P. Kopacz, On generalization of Zermelo navigation problem on Riemannian manifolds, Int. J. Geom. Methods M., 16 (2019), 1950058. http://dx.doi.org/10.1142/S0219887819500580 doi: 10.1142/S0219887819500580
![]() |
[5] |
B. Russell, S. Stepney, Zermelo navigation and a speed limit to quantum information processing, Phys. Rev. A, 90 (2014), 012303. http://dx.doi.org/10.1103/PhysRevA.90.012303 doi: 10.1103/PhysRevA.90.012303
![]() |
[6] |
B. Russell, S. Stepney, Zermelo navigation in the quantum brachistochrone, J. Phys. A: Math. Theor., 48 (2015), 115303. http://dx.doi.org/10.1088/1751-8113/48/11/115303 doi: 10.1088/1751-8113/48/11/115303
![]() |
[7] |
R. Yoshikawa, S. Sabau, Kropina metrics and Zermelo navigation on Riemannian manifolds, Geom. Dedicata, 171 (2014), 119–148. http://dx.doi.org/10.1007/s10711-013-9892-8 doi: 10.1007/s10711-013-9892-8
![]() |
[8] |
N. Aldea, P. Kopacz, Generalized Zermelo navigation on Hermitian manifolds under mild wind, Differ. Geom. Appl., 54 (2017), 325–343. http://dx.doi.org/10.1016/j.difgeo.2017.05.007 doi: 10.1016/j.difgeo.2017.05.007
![]() |
[9] |
N. Aldea, P. Kopacz, Generalized Zermelo navigation on Hermitian manifolds with a critical wind, Results Math., 72 (2017), 2165–2180. http://dx.doi.org/10.1007/s00025-017-0757-6 doi: 10.1007/s00025-017-0757-6
![]() |
[10] |
R. Hama, P. Chitsakul, S. Sabau, The geometry of a Randers rotational surface, Publ. Math. Debrecen, 87 (2015), 473–502. http://dx.doi.org/10.5486/PMD.2015.7395 doi: 10.5486/PMD.2015.7395
![]() |
[11] |
G. Gibbons, C. Herdeiro, C. Warnick, M. Werner, Stationary metrics and optical Zermelo-Randers-Finsler geometry, Phys. Rev. D, 79 (2009), 044022. http://dx.doi.org/10.1103/PhysRevD.79.044022 doi: 10.1103/PhysRevD.79.044022
![]() |
[12] |
D. Bao, C. Robles, Z. Shen, Zermelo navigation on Riemannian manifolds, J. Differential Geom., 66 (2004), 377–435. http://dx.doi.org/10.4310/jdg/1098137838 doi: 10.4310/jdg/1098137838
![]() |
[13] |
A. Böttcher, D. Wenzel, The Frobenius norm and the commutator, Linear Algebra Appl., 429 (2008), 1864–1885. http://dx.doi.org/10.1016/j.laa.2008.05.020 doi: 10.1016/j.laa.2008.05.020
![]() |
[14] |
P. Laurian-Ioan, A. Ali, B. Catalin, A. Alkhaldi, The χ-Hessian quotient for Riemannian metrics, Axioms, 10 (2021), 69. http://dx.doi.org/10.3390/axioms10020069 doi: 10.3390/axioms10020069
![]() |
[15] |
C. Guo, T. Liu, Q. Wang, B. Qin, A. Wang, A unified strong spectral Tchebychev solution for predicting the free vibration characteristics of cylindrical shells with stepped-thickness and internal-external stiffeners, Thin Wall. Struct., 168 (2021), 108307. http://dx.doi.org/10.1016/j.tws.2021.108307 doi: 10.1016/j.tws.2021.108307
![]() |
[16] |
Y. Li, S. Liu, Z. Wang, Tangent developables and Darboux developables of framed curves, Topol. Appl., 301 (2021), 107526. http://dx.doi.org/10.1016/j.topol.2020.107526 doi: 10.1016/j.topol.2020.107526
![]() |
[17] |
Y. Li, K. Eren, K. Ayvacı, S. Ersoy, The developable surfaces with pointwise 1-type Gauss map of Frenet type framed base curves in Euclidean 3-space, AIMS Mathematics, 8 (2023), 2226–2239. http://dx.doi.org/10.3934/math.2023115 doi: 10.3934/math.2023115
![]() |
[18] |
Y. Li, Z. Chen, S. Nazra, R. Abdel-Baky, Singularities for timelike developable surfaces in Minkowski 3-space, Symmetry, 15 (2023), 277. http://dx.doi.org/10.3390/sym15020277 doi: 10.3390/sym15020277
![]() |
[19] |
Y. Li, M. Aldossary, R. Abdel-Baky, Spacelike circular surfaces in Minkowski 3-space, Symmetry, 15 (2023), 173. http://dx.doi.org/10.3390/sym15010173 doi: 10.3390/sym15010173
![]() |
[20] |
Y. Li, A. Abdel-Salam, M. Saad, Primitivoids of curves in Minkowski plane, AIMS Mathematics, 8 (2023), 2386–2406. http://dx.doi.org/10.3934/math.2023123 doi: 10.3934/math.2023123
![]() |
[21] |
Y. Li, O. Tuncer, On (contra)pedals and (anti)orthotomics of frontals in de Sitter 2-space, Math. Meth. Appl. Sci., 1 (2023), 1–15. http://dx.doi.org/10.1002/mma.9173 doi: 10.1002/mma.9173
![]() |
[22] |
Y. Li, M. Erdoğdu, A. Yavuz, Differential geometric approach of Betchow-Da Rios soliton equation, Hacet. J. Math. Stat., 52 (2023), 114–125. http://dx.doi.org/10.15672/hujms.1052831 doi: 10.15672/hujms.1052831
![]() |
[23] |
Y. Li, A. Abolarinwa, A. Alkhaldi, A. Ali, Some inequalities of Hardy type related to Witten-Laplace operator on smooth metric measure spaces, Mathematics, 10 (2022), 4580. http://dx.doi.org/10.3390/math10234580 doi: 10.3390/math10234580
![]() |
1. | Yanlin Li, Dhriti Sundar Patra, Nadia Alluhaibi, Fatemah Mofarreh, Akram Ali, Geometric classifications of k-almost Ricci solitons admitting paracontact metrices, 2023, 21, 2391-5455, 10.1515/math-2022-0610 | |
2. | Yanlin Li, Erhan Güler, A Hypersurfaces of Revolution Family in the Five-Dimensional Pseudo-Euclidean Space E25, 2023, 11, 2227-7390, 3427, 10.3390/math11153427 | |
3. | Ali H. Hakami, Mohd Danish Siddiqi, Significance of Solitonic Fibers in Riemannian Submersions and Some Number Theoretic Applications, 2023, 15, 2073-8994, 1841, 10.3390/sym15101841 | |
4. | A. A. Abdel-Salam, M. I. Elashiry, M. Khalifa Saad, On the equiform geometry of special curves in hyperbolic and de Sitter planes, 2023, 8, 2473-6988, 18435, 10.3934/math.2023937 | |
5. | Sümeyye Gür Mazlum, On the Gaussian curvature of timelike surfaces in Lorentz-Minkowski 3-space, 2023, 37, 0354-5180, 9641, 10.2298/FIL2328641G | |
6. | Nasser Bin Turki, A Note on Incompressible Vector Fields, 2023, 15, 2073-8994, 1479, 10.3390/sym15081479 | |
7. | Yanlin Li, Manish Kumar Gupta, Suman Sharma, Sudhakar Kumar Chaubey, On Ricci Curvature of a Homogeneous Generalized Matsumoto Finsler Space, 2023, 11, 2227-7390, 3365, 10.3390/math11153365 | |
8. | Yanlin Li, Huchchappa A. Kumara, Mallannara Siddalingappa Siddesha, Devaraja Mallesha Naik, Characterization of Ricci Almost Soliton on Lorentzian Manifolds, 2023, 15, 2073-8994, 1175, 10.3390/sym15061175 | |
9. | Mohammad Nazrul Islam Khan, Fatemah Mofarreh, Abdul Haseeb, Mohit Saxena, Certain Results on the Lifts from an LP-Sasakian Manifold to Its Tangent Bundle Associated with a Quarter-Symmetric Metric Connection, 2023, 15, 2073-8994, 1553, 10.3390/sym15081553 | |
10. | Yanlin Li, Aydin Gezer, Erkan Karakaş, Some notes on the tangent bundle with a Ricci quarter-symmetric metric connection, 2023, 8, 2473-6988, 17335, 10.3934/math.2023886 | |
11. | Ibrahim Al-Dayel, Meraj Ali Khan, Impact of Semi-Symmetric Metric Connection on Homology of Warped Product Pointwise Semi-Slant Submanifolds of an Odd-Dimensional Sphere, 2023, 15, 2073-8994, 1606, 10.3390/sym15081606 | |
12. | Yanlin Li, Fatemah Mofarreh, Rashad A. Abdel-Baky, Kinematic-geometry of a line trajectory and the invariants of the axodes, 2023, 56, 2391-4661, 10.1515/dema-2022-0252 | |
13. | Yanlin Li, Abdussamet Çalışkan, Quaternionic Shape Operator and Rotation Matrix on Ruled Surfaces, 2023, 12, 2075-1680, 486, 10.3390/axioms12050486 | |
14. | Ibrahim Al-Dayel, Estimation of Ricci Curvature for Hemi-Slant Warped Product Submanifolds of Generalized Complex Space Forms and Their Applications, 2023, 15, 2073-8994, 1156, 10.3390/sym15061156 | |
15. | Jing Li, Zhichao Yang, Yanlin Li, R.A. Abdel-Baky, Khalifa Saad, On the curvatures of timelike circular surfaces in Lorentz-Minkowski space, 2024, 38, 0354-5180, 1423, 10.2298/FIL2404423L | |
16. | Amin Jafarimoghaddam, Manuel Soler, The applicability of Zermelo's equation to indirect and direct optimization of commercial aircraft flights, 2024, 135, 0307904X, 457, 10.1016/j.apm.2024.07.004 | |
17. | Esra Erkan, Concircular Vector Fields on Radical Anti-Invariant Lightlike Hypersurfaces of Almost Product-like Statistical Manifolds, 2023, 15, 2073-8994, 1531, 10.3390/sym15081531 | |
18. | Yanlin Li, Sujit Bhattacharyya, Shahroud Azami, Apurba Saha, Shyamal Kumar Hui, Harnack Estimation for Nonlinear, Weighted, Heat-Type Equation along Geometric Flow and Applications, 2023, 11, 2227-7390, 2516, 10.3390/math11112516 | |
19. | Yanlin Li, Kemal Eren, Soley Ersoy, On simultaneous characterizations of partner-ruled surfaces in Minkowski 3-space, 2023, 8, 2473-6988, 22256, 10.3934/math.20231135 |