
We describe the generators of the vector fields tangent to the bifurcation diagrams and caustics of simple quasi boundary singularities. As an application, submersions on the pair (G,B), which consists of a cuspidal edge G in R3 that contains a distinguishing regular curve B, are classified. This classification was used as a means to investigate the contact that a general cuspidal edge G equipped with a regular curve B⊂G has with planes. The singularities of the height functions on (G,B) are discussed and they are related to the curvatures and torsions of the distinguished curves on the cuspidal edge. In addition to this, the discriminants of the versal deformations of the submersions that were accomplished are described and they are related to the duality of the cuspidal edge.
Citation: Fawaz Alharbi, Yanlin Li. Vector fields on bifurcation diagrams of quasi singularities[J]. AIMS Mathematics, 2024, 9(12): 36047-36068. doi: 10.3934/math.20241710
[1] | Yaning Wang, Hui Wu . Invariant vector fields on contact metric manifolds under $\mathcal{D}$-homothetic deformation. AIMS Mathematics, 2020, 5(6): 7711-7718. doi: 10.3934/math.2020493 |
[2] | Süleyman Şenyurt, Filiz Ertem Kaya, Davut Canlı . Pedal curves obtained from Frenet vector of a space curve and Smarandache curves belonging to these curves. AIMS Mathematics, 2024, 9(8): 20136-20162. doi: 10.3934/math.2024981 |
[3] | Ibrahim Al-Dayel, Meraj Ali Khan . Ricci curvature of contact CR-warped product submanifolds in generalized Sasakian space forms admitting nearly Sasakian structure. AIMS Mathematics, 2021, 6(3): 2132-2151. doi: 10.3934/math.2021130 |
[4] | Ayşe Yavuz, Melek Erdoǧdu . Non-lightlike Bertrand W curves: A new approach by system of differential equations for position vector. AIMS Mathematics, 2020, 5(6): 5422-5438. doi: 10.3934/math.2020348 |
[5] | Yanlin Li, Nasser Bin Turki, Sharief Deshmukh, Olga Belova . Euclidean hypersurfaces isometric to spheres. AIMS Mathematics, 2024, 9(10): 28306-28319. doi: 10.3934/math.20241373 |
[6] | Amira Ishan . On concurrent vector fields on Riemannian manifolds. AIMS Mathematics, 2023, 8(10): 25097-25103. doi: 10.3934/math.20231281 |
[7] | Ibrahim Al-Dayel, Sharief Deshmukh, Olga Belova . Characterizing non-totally geodesic spheres in a unit sphere. AIMS Mathematics, 2023, 8(9): 21359-21370. doi: 10.3934/math.20231088 |
[8] | Mohammed Guediri, Sharief Deshmukh . Hypersurfaces in a Euclidean space with a Killing vector field. AIMS Mathematics, 2024, 9(1): 1899-1910. doi: 10.3934/math.2024093 |
[9] | Quanxiang Pan . Real hypersurfaces in complex space forms with special almost contact structures. AIMS Mathematics, 2023, 8(11): 27200-27209. doi: 10.3934/math.20231391 |
[10] | Nasser Bin Turki, Sharief Deshmukh, Olga Belova . A note on closed vector fields. AIMS Mathematics, 2024, 9(1): 1509-1522. doi: 10.3934/math.2024074 |
We describe the generators of the vector fields tangent to the bifurcation diagrams and caustics of simple quasi boundary singularities. As an application, submersions on the pair (G,B), which consists of a cuspidal edge G in R3 that contains a distinguishing regular curve B, are classified. This classification was used as a means to investigate the contact that a general cuspidal edge G equipped with a regular curve B⊂G has with planes. The singularities of the height functions on (G,B) are discussed and they are related to the curvatures and torsions of the distinguished curves on the cuspidal edge. In addition to this, the discriminants of the versal deformations of the submersions that were accomplished are described and they are related to the duality of the cuspidal edge.
One of the core tools in singularity theory is to classify functions on a certain space equipped with a distinguished hyperspace in it. The infinitesimal level problems of this kind require finding diffeomorphisms of the ambient space such that this hypersurface is preserved. In order to construct these diffeomorphisms, it is necessary to provide a description of the generators of vector fields that are parallel to the hypersurface. Many authors have studied algorithm and algebraic aspects of such vector fields (see for example [1,2,3]) to classify singularities of maps (functions) between two manifolds that can be constructed from the differential geometry point of view (see e.g. [4,5]. Further motivations of the topics can be found in various relevant papers with differential geometry [6,7,8] and submanifolds theory [9,10,11]. The classification can help study manifolds via other functions such as the height function and distance squared function. In many cases, this hypersurface appears as a discriminant (or bifurcarion diagram) or caustics of versal deformation of classes with respect to a standard equivalence relation.
In a series of papers [12,13,14], a new non-standard equivalence relation, on a space Rn equipped with a variety Γ, are studied, and, consequently, simple classes were obtained. Classification of projections of Lagrangian manifolds endowed with a hypersurface Γ is accomplished through the utilization of these classes. As a result of the classification, the bifurcation diagrams and caustics of versal unfolding of simple classes were described in [15], which were conduct in a different manner. In particular, let G(z,u)=˜G(z,u)+u0, with z∈Rn and u=(u0,u1…,us) as parameters, be a versal unfolding of the simple g(z)=G(z,0) with respect to the quasi equivalence relation. Then, the respective bifurcation diagram in the space of parameters consists of two components W0, which is the standard discriminant given by the equations G=0 and ∂G∂z=0 and W1, which is contained in W0 and it is determined by constraints that define Γ. The caustics is in the unfolding base ˜u=(u1,…,us) (which does not include λ0), and consists of two parts Σ0 which represents the singular set image of W0 under the projection π:u→˜u and Σ1=π(W1). The preceding construction yields that the bifurcation diagrams is a pair W=(W0,W1), where W0 is a hypersurface in Rsu and W1⊂W0, while the caustics is the union Σ∗=Σ0∪Σ1 with dim(Σ0)=dim(Σ1).
In this work, in Section 2, we calculate the generators of the vector fields that are parallel to the quasi bifurcation diagrams and caustics, obtained in [15]. This implies that, for the bifurcation diagrams, we seek vector fields that preserve not only W0 but also the points of W0, and for the caustics, we seek vector fields that preserve both Σ0 and Σ1.
Singularity theory techniques and differential geometry tools can be employed to comprehend the geometry of an object by examining its interaction with planar objects, such as planes or lines. In order to determine the former, it is necessary to analyze the singularities of the height functions along particular directions, which define the object's contact with the plane orthogonal to that direction. Many authors have investigated the contact with planes of singular surfaces, including the cross-cap [16], the swallowtail [17], the cuspidal edge [18], and the folded cuspidal edges [19].
Thus, as an application, in Section 3, we consider a cuspidal edge G equipped with a distinguished regular curveB in it. The object appears as a bifurcation diagram of the quasi boundary class B3. We then apply the module of vector fields obtained in Section 2 to classify submersions on the pair (G,B). Then, we use such classification to study the contact of a general cuspidal edge equipped with a regular curve in it by studying the singularities of height function on (G,B). There are two distinguished regular curves, ΣG (the singular set) and B. Finally, we examine the duality of the two curves by describing the versal deformation of the generic submersions that are obtained.
Let K denote the real number R or the complex numbers C with local coordinates z. The set of all smooth function germs from (Kn,0) to K is denoted by En (or Ez), and the maximal ideal in this set is denoted by Mn. Let θn represent the module over En consisting of all vector fields formed on (Kn,0). Let K[z] be the polynomial ring or formal power series over K.
Let V⊂(Kn,0) be an analytic variety. The ideal of germs that vanish on V is denoted by I(V).
Definition 1. If ξ(I(V))⊆I(V), then a vector field ξ∈θn is considered to be tangent to V or to preserve V. The module of vector fields of this nature is represented by Der(−logV).
It is important to note that if ξ∈Der(−logV), then it can be integrated to generate a diffeomorphism φ:(Kn,0)→(Kn,0) that preserves V, i.e., φ(V)⊆V.
Also note that Definition (1) implies that if V=(h1,h2,…,hs), where hi∈Mn, then
Der(−logV)={ξ∈θn:∃fij∈Ensuch that ξ(hi)=s∑i=1fijhi,j=1,…,s}. |
Definition 2. Let ζ be a vector field on (Kp) and f:(Kn,0)→(Kp) be a smooth map germ. Then, ζ is said to be liftable over f if there exist a vector field η on (Kn) such that df∘η=ζ∘f. In this case, η is said to be lowerable.
The concepts of liftable and tangent vector fields on the discriminant are identical for stable map germs when K=C (see [20]). In fact, Arnold, in [2], showed that there are liftable vector fields that are not tangent vector fields when K=R.
Let V1 be an R−analytic variety in (Rr,0) in local coordinates (u0,…,ur−1). Assume that V0⊆V1 of codimension 1. Denote by ˜V the pair consisting of V1 and a distinguished sub-variety V0 in it and set ˜V=(V1,V0). Thus, we may assume that the pair represents a variety equipped with a boundary.
Definition 3. A diffeomorphism ϕ:(Rr,0)→(Rr,0) will be said to preserve ˜V if and only if ϕ(V1)⊆V1 and ϕ(V0)⊆V0.
Definition 4. A vector field ξ∈θr is considered tangent to ˜V if and only if the following conditions are fulfilled.
(1) ξ(I(V1))⊆I(V1),
(2) ξ(I(V0))⊆I(V0).
The module of all vector fields satisfying the given conditions will be represented as Der(−log˜V) over the Er-module, that is
Der(−log˜V)={ξ=r−1∑i=0˙u∂∂ui∈θr:ξg1∈I(V1),ξg2∈I(V0)∀g1∈I(V1),g2∈I(V0)}, |
and it is commonly referred to as the stationary algebra of ˜V.
Remark 1. If ξ belongs to Der(−log˜V), then ζ conserves ˜V and, as a result, is tangent to it. Furthermore, Der(−log˜V) is the Lie algebra associated with the group of diffeomorphisms that preserve (˜V,0) in the space (Rn,0).
Consider the coordinate space Rn in local coordinates z=(x,y1,…,yn−1) equipped with a smooth hypersurface Γ={x=0}, which is referred to as a boundary.
Recall from [12] that on the Γ, every simple function germ g can be stably transformed via the quasi equivalence relation into one of the subsequent germs:
Bk:g1(x,y1)=±xk±y21,wherek≥2, |
or
Fp,k:g2(x,y1)=±(x±yp1)2±yk1, wherek>p≥2. |
The tangent space to the quasi boundary equivalence singularity of g at g is
TQΓg={∂g∂x(xA+n∑i=1∂g∂ziBi)+n−1∑j=1∂g∂yjEj:A,Bi,Ej∈Ez}. |
Let G(z,u) be a deformation of g:(Rn,0)→(R,0), where u=(u0,u1,…,ur−1)∈Rr are parameters. Set Gu(z)=G(z,u); so that, G0=g.
The initial speeds of G are defined by
˙Gi=∂G∂ui(z,0),∀i∈{0,1,2,…,r−1}. |
The subsequent result is an adaptation of Theorem 3 from [16].
Proposition 1. A deformation G of a function g is considered versal in regard to the quasi equivalence if and only if
TQΓg+R{˙G0,…˙Gr−1}=Ez. |
Assume that the elements ω0,…,ωr−1∈Ez form a basis of the quotient space Ez/TQΓg. Then, Proposition 1 implies that a miniversal deformation of a function germ g may take the form:
G(z,u)=g(z)+r−1∑i=0uiωi(z). | (3.1) |
Therefore, the formulas of quasi boundary versal deformations of g1∈Bk and g2∈Fp,k are
Gk(z,u)=±y21±xk+k−1∑i=0uixi |
and
Gp,k(z,u)=±(x±yp1+p+k−2∑j=k−1ujyj−(k−1)1)2±yk1+k−2∑i=0uiyi1, |
respectively.
Remark 2. The versal deformation of the class F2,3 can be written equivalently as
G(x,y1,u)=±x2±y31+u0+u1x+u2y+u3xy1. |
Definition 5. [15] The quasi bifurcation diagram of a germ g with G(z,u) being its quasi versal deformation, is the pair W(g)=(W1,W0), where
W1={u:G=∂G∂zi=0}, |
and
W0={u:G=∂G∂zi=x=0}. |
Note that W0 is contained in W1 and it satisfies the constraint x=0. Thus, in particular, the bifurcation diagrams of the classes Bk is W(Bk)=(W1,W0), where
W1={(u0,…,uk−1):u0=(±1∓k)xk−k−1∑i=2(i−1)uixi,u1=∓kxk−1−k−1∑j=2jujxj−1,x∈R}, |
and
W0={(u0,…,uk−1):u0=0,u1=0}, |
On the other hand, the bifurcation diagram of the class F2,3 is W(F2,3)=(W1,W0), where
W1={(u0,u1,u2,u3):u0=±x2±2y31+u3xy1,u1=∓2x−u2y1,u2=∓3y21−u3x,x,y1,u3∈R}, |
and
W0={(u0,u1,u2,u3):u0=±2y31,u1=−u2y1,u2=∓3y21,y1,u3∈R}. |
Theorem 3.1. The stationary algebra of W(Bk), for k=2,3,4, and W(F2,3) is described as follows.
(1) Der(−logW(B2)) is generated by
ξ1=0∂∂u0+(u0−4u21)∂∂u1,ξ2=2u0∂∂u0+u20∂∂u1,ξ3=2u1∂∂u0+u0u1∂∂u1. |
(2) Der(−logW(B3)) is generated by
ξ1=3u0∂∂u0+2u1∂∂u1+u2∂∂u2,ξ2=u0u1∂∂u0+2u0u2∂∂u1+3u0∂∂u2,ξ3=3u0u2∂∂u0+18u0∂∂u1+(12u1−3u22)∂∂u2,ξ4=(u21−3u0u2)∂∂u0+(3u1−u22)∂∂u2. |
(3) Der(−logW(B4)) is generated by
ξ1=4u0∂∂u0+3u1∂∂u1+2u2∂∂u2+u3∂∂u3,ξ2=u0u1∂∂u0+2u0u2∂∂u1+3u0u3∂∂u2+4u3∂∂u3,ξ3=2u0u3∂∂u0+24u0∂∂u1+(18u1−2u2u3)∂∂u2+(12u2−4u23)∂∂u3,ξ4=4u0u2∂∂u0+18u0u3∂∂u1+(24u0−4u22+12u1u3)∂∂u2+(18u1−2u2u3)∂∂u3,ξ5=(3u21−8u0u2)∂∂u0+(9u1u3−4u23)∂∂u2+(12u1−2u2u3)∂∂u3. |
(4) Der(−logW(F2,3)) is generated by
ξ1=6u0∂∂u0+3u1∂∂u1+4u2∂∂u2+u3∂∂u3,ξ2=(3u21−u2u23)∂∂u0+(6u1−u33)∂∂u1+6u1u3∂∂u2+6u3∂∂u3,ξ3=(−16u1u2u3+6u0u23)∂∂u0+(−16u2u3−5u1u33)∂∂u1+(24u21−4u2u23)∂∂u2+(48u1+u33)∂∂u3,ξ4=(−16u1u22−18u0u2u3)∂∂u0+(−32u22+3u1u2u3−24u0u23)∂∂u1+(72u0u1−12u22u3)∂∂u2+(144u0+u2u23)∂∂u3,ξ5=(−64u22+48u1u2u3−18u0u23)∂∂u0+(−32u2u3+23u1u23)∂∂u1+(288u0−72u21+12u2u23)∂∂u2+5u33∂∂u3,ξ6=(9u31−8u22u3+3u1u2u23)∂∂u0+(72u0−2u2u23)∂∂u1+(36u0u3+9u21u3+3u2u33)∂∂u2+(8u2−6u1u3)∂∂u3. |
Proof. Let p1,p2∈Er. Assume that p1 is the defining equation of W1 and p1,p2 are the defining equations of W0. Let I(W1) be the ideal generated by p1 and I(W0) is the ideal generated by p1 and p2.
Let Θ(W1) be the module of all vector fields ξ=r∑i=1ξi∂∂ui on Rr such that ξ(I(W1))⊆I(W1). To find Θ(W1), we have to solve the equation
r∑i=1ξi∂p1∂ui=qp1, |
for ξi and q∈Er. Now consider the map ϕ:Er+1r→R, given by
Φ(ξ,q)=r∑i=1ξi∂h1∂ui−qh1, |
where ξ=(ξ1,…,ξr)∈Err and q∈E1r. Let K=kerΦ and π:Er+1r→Err be defined by π(ξ,q)=ξ. Then Θ(W1)=K. Using the syzygies that are supplied in the Singular software package, we are able to obtain the K.
Next, we are looking for the module Θ(W1) of all vector fields such that ξ(I(W0))⊆I(W0). This implies that, for each j=1,2, we have to solve
r∑i=1ξi∂pj∂ui=2∑i=1qipi, |
for ξ=r∑i=1ξi∂∂ui and qi.
For j=1,2, let Φj:Er+2r→R be the map that is defined as
Φj(ξ,˜q)=r∑i=1ξi∂pj∂ui−2∑i=1qipi, |
where ξ=(ξ1,…,ξr)∈Err and ˜q=(q1,q2)∈E2r. Let Kj=kerΦj. Set π:Er+2r→Err be defined by π(ξ,˜q)=ξ. Let Sj=π(Kj). Therefore,
Θ(W0)=S1⋂S2. |
Again we use the syzygies to obtain the Ki. Finally, we have
Θ(W)=Θ(W1)⋂Θ(W0). |
All of the vector fields ξ that are created by this approach can be verified to be liftable. This means that there is a vector field η on Rr−1 that is such that dpi(η)=ξ∘pi, and as a result, they are tangent to W.
In (3.1), if we set ω0=1 and ωi∈Mz, then the space Rr−1 in the local coordinates u1,…,ur−1 is known as the principle of a shortened quasi-boundary versal deformation of g.
Consider the projection map π:Rr→Rr−1,(u0,u1,…ur−1)↦(u1,…ur−1).
Definition 6. [15] The caustic of a a versal deformation of a function g with respect to the quasi equivalence relation is a hypersurface in Rr−1 which is a union
Σ1∪Σ0, |
which will be denoted by Σ∗, where Σ10=the π -image of the singular points of W1 and Σ0=π(W0).
Recall from [15] (applying Definition [6]) that the caustic of Bk is a union of two hypersurfaces which are tangent to each other, the first one is the set
Σ1={(u1,…,uk−1:u1=∓kxk−1−k−1∑i=2juixi−1,u2=∓k(k−1)2xk−2−k−1∑j=3j(j−1)2ujxj−2x∈R}, |
that is a cylindrical generalized swallow tail over a line. The second one of them is the submanifold of the greatest dimension that crosses the edge of Σ0., in particular:
Σ0={u1,…,uk−1:u1=0}. |
Furthermore, the caustic of the class F2,3 is a union of the smooth surface
Σ1={(u1,u2,u3):u1=∓2x−u3312,u2=∓3144u43−u3x,x,u3∈R}, |
and Whitney Umbrella
Σ0={(u1,u2,u3):u1=−u3y1,u2=∓3y21,y1,u3∈R}. |
Theorem 3.2. The module Der(−log(Σ∗) is generated by the vector fields
(1) For the caustic of B3
ξ1=2u1∂∂u1+u2∂∂u2,ξ2=2u1u2∂∂u1+3u1∂∂u2. |
(2) For the caustic of B4
ξ1=3u1∂∂u1+2u2∂∂u2+u3∂∂u3,ξ2=(−4u22+9u1u2)∂∂u2+(12u1−2u2u3)∂∂u2,ξ2=(8u1−3u2u3)∂∂u2+(16u2−6u23)∂∂u2. |
(3) For the caustic of F2,3
ξ1=(48u1−12u2u3)∂∂u1+(24u2+2u33)∂∂u2+0∂∂u3,ξ2=(6u22−2u1u23)∂∂u1+(−4u1u3−2u2u23)∂∂u2+12u2∂∂u3,ξ3=(12u2u3)∂∂u1+(12u2−2u33)∂∂u2+12u3∂∂u3,ξ4=(3u22u3+u1u33)∂∂u1+(6u22+2u1u23)∂∂u2+0∂∂u3. |
Proof. Assume that Θ(Σ1) and Θ(Σ0) are the modules of vector fields that are that are tangential to Σ1 and Σ0, respectively.
To determine Θ(Σi),i=0,1, we employ similar processes as those used in the proof of Theorem 3.1 to determine Θ(W1).
Hence, we have,
Θ(Σ∗)=Θ(Σ1)⋂Θ(Σ0). |
All such vector fields ξ∈Θ(Σ∗) are liftable and as a result, they are tangent to Σ∗.
Next, we investigate whether or not W and Σ∗ are considered to be free divisors. Recall that a hypersurface V={h=0}⊂(Kn,0) with a reduced defining ideal I(V), is called free divisor, in the sense of Saito, if Der(−logV) is a free En-module, necessarily, so its rank is equal to n. The following criterion was established also by K. Saito and is now commonly referred to after him.
Proposition 2. [21][Saito's Criterion] Let h∈K[z] be reduced. Then, h defines a free divisor if and only if there exists n×n matrix H with entries in K[z] such that
det(H)=hand(∇h)H≡0mod(h). |
Here, ∇h=(∂h∂z1,…,∂h∂zn) and the last condition just expresses that each entry of the (row) vector (∇h)H is divisible by h in K[z]. The columns of H can then be viewed as the coefficients of a basis, with respect to the partial derivatives ∂∂zi, of the logarithmic vector fields along the divisor h=0. The matrix H is called a discriminant or Saito matrix of h.
The above discussion and Theorem 3.1 implies that the bifurcation diagrams of the classes Bk,k=2,3,4 and class F2,3 are not free divisors due to the cardinality of a basis of Der(−logW(Bk)), and Der(−logW(F2,3)), respectively. Moreover,
Proposition 3. The caustics of Bi,i=3,4 are free divisors.
Proof. The caustic of the B3 class is a union of the parabola 3u1−u22=0 and the line u1=0. Hence, the defining equation of Σ∗⊂R2 for B3 is h1=u1(3u1−u22). On the other hand, by similar consideration we find that the defining equation of the caustic Σ∗⊂R3 for the class B4 is h2=u1(108u21+32u32−108u1u2u3−9u22u23+27u1u33). Applying Saito's Criterion and Theorem 3.2, one can easily show that the caustics of Bi,i=3,4 are free divisors.
Let G be a general cuspidal edge in R3 with local coordinates z=(u,v,w), having the following parametrization at the origin as in [18]:
f:U⊆R2,0→R3, |
(t,s)↦(t,a(t)+12s2,b1(t)+s2b2(t)+s3b3(t,s)), |
where U is an open set in R2 and
a(t)=12a20t2+16a30t3+124a40t4+O(5);b1(t)=12b20t2+16b30t3+124b40t4+O(5);b2(t)=12b12t+16b22t2+124b32t3+O(4);b3(t,s)=16b03+16b13t+124b04s+124b23t2+124b14ts+1120b05s2+O(3), |
and aij,bij∈R. Let B⊆G be a distinguished smooth curve which is parametrized by:
γ1(t)=f(t,t)=(t,a(t)+t22,b1(t)+t2b2(t)+t3b3(t,t)), |
i.e., the image of the line s=t. The pair ˜G=(G,B) will be called a geometric cuspidal edge with a smooth boundary. Note that ˜G is diffeomorphic to W(B3).
The set of critical points of G is {s=0}, and hence the singular set of G is the curve:
γ0(t)=f(t,0)=(t,a(t),b1(t)), |
which will be denoted by Σ.
Denote by (κΣ(0),κB(0)) and (τΣ(0),τB(0)) the pairs of the curvatures and torsions of the pair of space curves (Σ,B), respectively.
Proposition 4. (1) (κΣ(0),κB(0))=(√a220+b220,√(a20+1)2+b220).
(2) (τΣ(0),τB(0))=(a20b30−b20a30κ2Σ(0),(a20+1)(b30+b03+3b12)−b20a30κ2B(0)).
(3) The osculating planes of Σ and B at the origin are orthogonal to the vectors (0,−b20,a20) and (0,−b20,a20+1), respectively.
Proof. The results are obtained via the standard rules of calculating curvature and torsion on space curves.
Note that the two curves Σ and B have a common tangent line, and hence the tangential direction, which will be denoted by Td at 0, is parallel to the vector (1,0,0). Note that w=0 is the plane that represents the tangent cone Ld to ˜G.
Consider the pair (˜f,˜γ1) which consists of the A-normal form of a cuspidal edge:
˜f:U⊆R2,0→R3, |
(t,s)↦(t,12(s2−t2),s3+t3−2ts2), |
and the smooth curve ˜γ1(t)=˜f(t,t)=(t,0,0). Let V1=˜f(U⊆R2) and V0=˜γ1(R). Then, the pair V=(V1,V0) will be considered as a model of a cuspidal edge with a smooth curve, which serves as a boundary. (See Figure 1).
Recall that the bifurcation diagrams of the B3 class is the pair W=(W1,W0), where W1 is the cuspidal edge, which is parameterized by
(t,s)↦(t,−3s2−2ts,−2s3−ts2), |
and hence it is A-equivalent to ˜f, via the right change of coordinates t↦t,s↦s−13t, followed by normalising the coefficients. Hence, the pair V is diffeomorphic to W.
The defining equation of V1 is
K(z)=w2−v3+4uvw+u2v2+2u3w+u4v=0, |
while the defining equations of V0 are K(z)=0,v=0 and w=0.
Definition 7. We say that g1,g2∈Ez are R(V)-equivalent whenever g2=g1∘Φ where Φ:(R3,0)→(R3,0) is diffeomorphism-germ and Φ preserves V, i.e., Φ(V1)⊆V1 and Φ(V0)⊆V0.
Define Θ(V) as the module over Ez of vector fields in R3 that is tangential to V and define
Θ0(V)={ξ∈Θ(V):ξ(0)=0}. |
Then, the tangent space and the expanded tangent space to the orbit of g at g are
LR(V).g={ξ(g):ξ∈Θ0(V)}, |
and
LeR(V).g={ξ(g):ξ∈Θ(V)}, |
respectively.
The R+e-codimension of g is defined as d(g,R+e(V))=dimR(Mz/LeR(V).g).
Theorem 4.1. The Ez-module Θ(V) is generated by
ξ1=u∂∂u+2v∂∂v+3w∂∂w,ξ2=v∂∂u−2uv∂∂v+(u2v−2v2)∂∂w,ξ3=(3w−u3)∂∂u−(6uw+2u2v)∂∂v−6vw∂∂w,ξ4=7u2∂∂u+(6w+20uv)∂∂v+(9v2−12vu2)∂∂w. |
Proof. By following the same procedures as that to prove Theorem 3.1.
Corollary 1. For g∈Ez, the tangent space to the orbit of g atg with respect to the R(V)-equivalence relation is defined as
LR(V).g={[uA1+vA2+(3u−u3)A3+7u2A4]∂g∂u+[2vA1−2uvA2−(6uw+2u2v)A3+(6w+20uv)A4]∂g∂v+[3wA1+(u2v−2v2)A2−6vwA3+(9v2−12vu2)A4]∂g∂w:Ai∈Ez}. |
We proceed to classify submersion-germs g:(R3,0)→(R,0). g from (R3,0) to (R,0) in accordance with the R(V)-equivalence relation, where d(g,R+e(V))≤2. The classification method and prenormal forms are described in the following lemma and relies on Arnold's spectral sequence. At first, we go over several concepts from [22].
Let us assume that we have a certain appropriate Newton diagram Γ that is a subset of the non-negative integers Zn≥0. Each face Γi of Γ corresponds to a specific quasihomogeneity type αi=(αi1,αi2,…,αin). In this type, the monomials xk=xk11xk22…xknn with exponents lying on Γi have a degree of one, meaning that ⟨k,αi⟩=αi1k1+⋯+αinkn=1.
A monomial, denoted as xk, is considered to have a Newton degree of d if d is the minimum value obtained from the inner product of k and αj. The monomials of Newton degree d are precisely those whose exponents are contained in the diagram dΓ, which is created by scaling Γ by a factor of d.
The smallest of the Newton degrees of the monomials that appear in a power series is known as the Newton order d. An ideal Sj in the ring Ex is formed by the series of order at least d. The Newton filtration in Ex is generated by the ideals Sj. More precisely, S0=Ex, and Sk⊆Sl whenever k>l.
The principal part of a power series g of order d is the sum of the terms of Newton degree d.
Let g∈Ez. We may decompose g into its principal portion g0 of Newton degree being N and greater order elements ˜g as g=g0+˜g. It is assumed that the R+e-codimension of g0 is finite, meaning that d(g0,R+e(V))<∞. The subsequent result is a rendition of Lemma 8.1 in [12] and Lemma 2.10 in [13].
Lemma 1. Consider a monomial basis of the linear space Ez/LR(V).g0 and let ρ1(z),ρ2(z), …,ρs(z) be the subset of generators that have Newton degrees greater than N.
Assume that for every ω∈Sβ∖S>β, β>N:
(1) There is a vector field ξ=˙u∂∂u+˙v∂∂v+˙w∂∂w∈Θ(V), such that
ω=∂g0∂u˙u+∂g0∂v˙v+∂g0∂w˙w+ˆω+s∑i=1ciρi(z), |
where ˆω∈S>β and ci∈R.
(2) Furthermore, given any δ, if N<δ<β, and for every ψ∈Sδ, the following statement
E(ψ,ω)=∂ψ∂u˙u+∂ψ∂v˙v+∂ψ∂w˙w, |
belongs to Sβ.
Then any germ g=g0+˜g is R(V)-equivalent to a germ g0+s∑i=1diρi, where di∈R.
Remark 3. By eliminating the prerequisite that g0 has a finite codimension, the proof of Lemma 1 demonstrates that any function g=g0+˜g is R(V)-equivalent to a comparable form:
g0+∑diρi+Λ, |
where Λ is a member of a relatively large power of the maximal ideal. $
Definition 8. The families of germs of functions H1,H2:(R3×Rl,(0,0))→(R,0) are referred to as P-R+(V).-equivalent if
H2(z,u)=H1∘Φ(z,u)+C(u), |
where Φ:(R3×Rl,(0,0))→(R3×Rl,(0,0)) is a germ of diffeomorphism with Φ(z,u)=(φ(z,u),χ(u)), and C:(Rl,0)→R is a function germ.
Consider G(z,u) as a deformation of g∈Ez.Then, G is referred to as versal with regard to the R+(V)-equivalence relation if, for whatever other form of deformation H of g, there exists a map germ Φ(z,u) (which may not necessarily be a diffeomorphism) and C as described above, satisfying
H(z,u)=G∘Φ(z,u)+C(u). |
Proposition 5. A deformation G of g on V is R+(V)-versal provided
LR(˜X).g+R{1,˙G1,…˙Gl}=Ez, |
where ˙Gi are the initial speeds of G.
The result that follows is a description of the classification of germs of submersions in accordance with the R(V)-equivalence relation.
Theorem 4.2. Let V be a pair consisting of the cuspidal edge V1, which is represented by the parametrization ˜f(t,s)=(t,12(s2−t2),s3+t3−2ts2), and a distinguished smooth curve V0 within it, represented by the parametrization ˜γ(t)1=˜f(t,t)=(t,0,0). Then, any function germ g at 0 that has a R+(V)-codimension of not more than 2 (with moduli) is equivalent to a function germ mentioned in the Table 1.
Equivalent Germ | Constraints | d(g,R+(V)) | mini-versal unfolding |
±u | – | 0 | ±u |
±v+ϵu2,1 | ϵ≠0,±1 | 1 | ±v+ϵu2+λ1u |
±v+ϵu3,1 | ϵ≠0 | 2 | ±v+ϵu3+λ1u+λ2u2 |
±w+ϵu2,1 | ϵ≠0 | 2 | ±w+ϵu2+λ1u+λ2v |
*Note: The symbol ϵ represents a modulus, while the codimension refers to the dimension of the stratum. |
Proof. The linear transformations of coordinates derived via the integration of the 1-jets of the vector fields in Θ(V) are:
φ1(z)=(ec1u,e2c1v,e3c1w);φ2(z)=(u+c2v,v,w);φ3(z)=(u+c3w,v,w);φ3(z)=(u,v+c4w,w); |
where ci∈R.
Let g be decomposed into its 1-jet g0=au+bv+cw, where a,b,c∈R and ˜g∈M2z. Using φi, one can show that the orbits of the space of 1-jets are ±u, ±v, and ±w.
The subsequent conclusions may be established by applying Lemma 1 and Remark 3.
Let g0=±u. Then, the tangent space to the orbit of g0 at g0 is
LR(V).g0={uA1+vA2+(3w−u3)A3+7u2A4:Ai∈Ez}. |
Clearly, we have modLR(V).g0: u≡0,v≡0 and w≡0. Hence, Lemma 3 implies that g is R(V)-equivalent to its principal part g0=±u.
Next, consider the principal part g0=±v. Then,
LR(V).g0={2vA1−3uvA2−(6uw+vu2)A3+(6w+20uv)A4:Ai∈Ez}. |
Therefore, we have modLR(V).g0: v≡0 and w≡0. It follows that
Ez/LR(V).g0≃{q(u):q∈Eu}. |
Using Remark 3 and taking into account the constraints on ˜g, the germ g is reduced to the form h=±v+˜h(u), where ˜h∈M2u. Let ˜h=d2u2+d3u3+…, di∈R. If d2≠0, then h is R(V)-equivalent to the germ ±v+ϵu2, where 0,±1≠ϵ∈R (modulus) and its mini-versal deformation may be taken as ±v±u2+λ1u. Next, if d2=0 but d3≠0, then h is R(V)-equivalent to the germ ±v+ϵu3, where 0≠ϵ∈R (modulus) and its mini-versal deformation may be taken as ±v+ϵu3+λ1u+λ2u2.
Finally, consider the 1-jet g0=±w. Then,
LR(V).g0={3wA1+(u2v−2v2)A2−6vwA3+(9v2−12vu2)A4:Ai∈Ez}. |
Therefore, we have modLR(V).g0:
w≡0, | (4.1) |
u2v−2v2≡0, | (4.2) |
and
9v2−12vu2≡0. | (4.3) |
Clearly, relations (4.2) and (4.3) are linearly independent. Hence, v2≡0 and vu2≡0. Consequently
Ez/LR(V).g0≃{a1v+a2uv+q(u):q∈Eu,ai∈R}. |
Using Remark 3 and taking into account the constraints on ˜g, the germ g is reduced to the form h=±w+a2uv+˜h(u), where ˜h∈M2u. If ˜h contains ˜d3u2, where 0≠˜d3∈R, then h is equivalent to ±w+ϵu2, 0≠ϵ∈R (modulus) and its mini-versal deformation may be taken as ±w+ϵu2+λ1u+λ2v. If ˜d3=0, then in the most degenerate case h has codimension greater than 2. The proof of the theorem is now complete.
Let F:(R3×R2,0)→R;(z,λ)↦F(z,λ) be a deformation of a germ h(z) on V and consider the family P(s,t,λ)=F(˜f(t,s),λ), where ˜f(s,t)=(t,12(s2−t2),s3+t3−2ts2). Then, we define the following types of discriminants:
(1) The discriminant of the family P, everywhere:
D1={(λ,P):∂P∂t=∂P∂s=0at(t,s,λ)}, |
(2) The discriminant of P, restricted to Σ:
D2={(λ,P):∂P∂t=0at(t,0,λ)}, |
and
(3) The discriminant of P, restricted to the boundary V0:
D3={(λ,P):∂P∂t=0at(t,t,λ)}. |
We shall calculate Di,i=1,2,3 for the mini-versal deformations F(z,λ) of the submersions g(z)=F(z,0) in Table 1.
(1) g(z)=u. We have F(z,λ)=u, and hence P=t. Note that the fiber g=0 is transverse for both Td and Ld. Clearly, Di,i=1,2, are all empty sets.
(2) g(z)=±v+ϵuk,k=2,3. Note that the tangent plane to the fiber g=0 contains Td but is transverse Ld.
● For k=2, we have F(z,λ)=±v+ϵu2+λ1u, and hence P=±(s2−t2)+ϵt2+λ1t. The D1 set is a smooth surface which is parametrized by
(t,λ2)↦(2(±1−ϵ)t,λ2,(±1−ϵ)t2). |
The D2 set coincides with D1. On the other hand, on the boundary we have P=F(˜f(t,t),λ)=ϵt2+λ1t. Therefore, the D3 set is also a smooth surface which is parametrized by:
(t,λ2)↦(−2ϵt,λ2,−ϵt2). |
Note that D1=D2 and D3 are tangent along the λ2-axis.
● For k=3, we have F(z,λ)=±v+ϵu3+λ1u+λ2u2. The D1 and D2 sets are coinciding cuspidal edge, which are parameterized by:
(t,λ2)↦(−3ϵt2−2(λ2∓1)t,λ2,(±1−λ2)t2−2ϵt3). |
The D3 set is also cuspidal edge, that is parametrized by:
(t,λ2)↦(−3ϵt2−2λ2t,λ2,−λ2t2−2ϵt3). |
Note that D1=D2 intersects D3 along a curve.
(3) g=±w+ϵu2. We have F=±w+ϵu2+λ2u+λ2v. We may consider the versal deformation:
F(z,λ)=±w+ϵu2+λ1u+λ2v. |
The tangent plane to g=0 includes both Td and Ld in this scenario. Now, we have
P=±(s3+t3−2ts2)+ϵt2+λ1t+λ2(s2−t2). |
Note that ∂P∂s=0 if and only if s=0 or λ2=32s−2t. Moreover, ∂P∂t=0 if and only if λ1=2s2∓3t2−2(ϵ+λ2)t. Hence, the D1 set consists of two components, the first one is a cuspidal edge which is parametrized by
(t,λ2)↦(∓3t2−2(ϵ+λ2)t,λ2,∓2t3−(ϵ+λ2)t2), |
and the second one is a smooth surface which is parametrized by
(t,s)↦(λ1,λ2,(−32±1)s3+(∓2−6)t3−ϵt2+2ts2−32st2), |
where λ1=±2s2+(∓3−4)t2−(2ϵ+3s)t and λ2=32s−2t. The D2 set coincides with the first part of D1. The D3 is a regular surface that may be described by a parametrization
(t,λ2)↦(−2ϵt,λ2,−ϵt2). |
We condense the above calculation in the following.
Proposition 6. (1) The discriminants D1, D2 and D3 of the singularity g=u are empty sets.
(2) The discriminants D1 and D2 of the singularity g=±v+ϵu2 are coincident smooth surface, and the D3 is also a smooth surface that is tangent to D1 along a curve (Figures 2 and 3).
(3) The discriminants D1 and D2 of the singularity g=±v+ϵu3 are coincident cuspidal edges, and the D3 is a different cuspidal edge that is tangent to D1 along a curve (Figure 4).
(4) The discriminants D1 of the singularity g=±w+ϵu2 is a combination of two components: a cuspidal edge and a regular surface (Figure 5). The D2 is a cuspidal edge and coincides with one of the components of the D1. The D3 is a smooth surface (Figure 6).
Let H:˜G×S2→R,H((t,s),η)=Hη(t,s)=f(t,s)⋅η, be a family of height functions on ˜G, where S2 is the 2-sphere and η=(η1,η2,η3)∈R3. Then, Hη measures the contact of ˜G with the plane πp which is orthogonal to the vector η at the point p∈˜G. Generically, the submersions g, which are obtained in Theorem 4.2, describe explicitly such contact. The contact between the fiber g=0 and the model V is equivalent to the contact between ˜G and πp.
We discuss the contact of πp with ˜G along Σ and B at the origin.
Note that the restriction of Hη along Σ and B are
Hη(t,0)=tη1+12(a20η2+b20η3)t2+16(a30η2+b30η3)t3+O(4), |
and
Hη(t,t)=tη1+12[(a20+1)η2+b20η3]t2+16[a30η2+(b30+3b12+b03)η3]t3+O(4), |
respectively.
Clearly, Hη is singular when η1=0. Further, the contact of ˜G with πp is measured by the zero of g=u with V at the origin when πp is transverse to Td. The following is a description of the remaining cases in which πp is a part of the pencil family of planes that is not transverse to Td:
Theorem 4.3. If πp is not the tangential cone to ˜G, then the contact of πp with ˜G is equivalent to that of the zero of g=±v+ϵuk (where k=2,3 and ϵ≠0,±1) with the representation V. Moreover,
(1) the plane πp has an A1-contact with Σ and B if and only if πp is not the osculating of neither Σ nor B.
(2) πp has an A1-contact with Σ and an A2-contact with B if and only if and only if πp is not the osculating of Σ but πp coincides with the osculating plane of B and τB(0)≠0.
Proof. Among the submersions listed in Theorem 4.2, the tangent plane to the zero fiber of g=±v+ϵuk, (where k=2,3), contains Td and is transverse to the tangent cone of ˜G. As a result, the contact of πp with ˜G is equivalent to that of g=0 with the representation V.
(1) Let k=2. Then, the contact of the tangential line along Σ of and B is measured by the type of
g(˜f(t,0))=(∓1+ϵ)t2, |
and
g(˜f(t,t)=ϵt2, |
respectively. So, it is of type A1 along the two curves. Now, consider the restrictions Hη(t,0) and Hη(t,t) on Σ and B, respectively. Then, the plane πp has an A1-contact with Σ0 if and only if η2a20+η3b20≠0, which implies that (η2,η3)≠(−b20,a20). On the other hand, πp has an A1-contact with B if and only if η2(a20+1)+η3b20≠0, which implies that (η2,η3)≠(−b20,a20+1). Geometrically, this means that πp is not the osculating plane of neither Σ nor B at p.
(2) Let k=3. Then, the contact of the tangential line along Σ and B is measured by the type of
g(˜f(t,0))=∓t2+ϵt3, |
and
g(˜f(t,t))=ϵt3, |
respectively. So, it is of type A1 along Σ and of type A2 along B. Now, consider the restriction Hη(t,t) along B. Then, the plane πp has an A2-contact with B if and only if
(a20+1)η2+b20η3=0, | (4.4) |
and
a20η2+(b30+3b12+b03)η3≠0. | (4.5) |
The constraint (4.4) implies that (η2,η3)=(−b20,a20+1), which means that πp is the osculating plane of B. On the other hand, the constraint (4.5) becomes
−a20b20+(b30+3b12+b03)(a20+1)≠0, | (4.6) |
which implies that τB(0)≠0.
Theorem 4.4. If πp is the tangent cone to ˜G, then the contact of πp at ˜G is equivalent to that of the zero fiber of g=±w+ϵu2 (ϵ≠0,±1) with the representation V. Furthermore, the plane πp has an A1-contact with both Σ and B, and it is not the osculating of neither Σ nor B.
Proof. The tangent to the zero of g=±w+ϵu2, (ϵ≠0,±1), is the same as the tangent cone of ˜G for the submersions shown in Theorem 1. Hence, the contact of πp with ˜G is the same as that of g=0 with the model V. Note here that η=(0,0,1). On the other hand, the contact of the tangential line along Σ and B is measured by the singularity of
g(˜f(t,0))=±t2+ϵt2, |
and
g(˜f(t,t))=ϵt2, |
respectively, where ϵ≠0.±1. So, it is of type A1. The corresponding height function restricted to Σ and B has an A1 singularity if and only if η3b20≠0, which means that b20≠0, and hence does not coincide with the osculating plane of both Σ and B.
The discriminants may be used for the examination of the dual of the cuspidal edge equipped with a smooth curve as explained below.
As pointed out in [23], an oriented plane in R3 in local coordinates z=(u,v,w) is characterized by a unit vector η and a real number c. The equation of the plane can be expressed as z⋅η=c, where ⋅ represents the scalar product. It is important to observe that the pairs (η,c) and (−η,−c) represent the same plane, but with opposing orientations. A unit space curve γ(t) can be associated with an oriented tangent plane at t0∈I⊂R by a unit vector η that is perpendicular to the tangent vector T(t) of γ(t) at t0. The equation of the tangent plane is given by z⋅η=γ(t0)⋅η. The collection of all oriented tangent planes to the curve γ(t) is referred to as "the dual" of γ(t). Consequently, it is associated with the set defined as follows:
{(η,c)∈S2×R:c=γ(t)⋅η,T(t)⋅η=0}. |
Define the following families:
D1(H)={(η,Hη(t,s))∈S2×R:∂H∂t=∂H∂s=0at(t,s,η)}, |
D2(H)={(η,Hη(t,0))∈S2×R:∂H∂t=0at(t,0,η)}, |
and
D3(H)={(η,Hη(t,t))∈S2×R:∂H∂t=0at(t,t,η)}. |
Then, in accordance with [23], if the contact of ˜G with πη is characterized by that of the fiber g=0 with V, where g is defined in Theorem (Classification of germs of submersions), then Di(H) is diffeomorphic to Di(F), where F is a R+(V)-versal unfolding of g with 2-parameters. Therefore, we have the following:
Proposition 7. Let ˜G be a pair of a geometric cuspidal edge G in R3 equipped with a smooth B in it. Then, the calculations and figures in Section 4.2 give the models, up to diffeomorphisms, of Di(H), i=1,2,3.
The above result implies that if πp is tangent to Td but it is transverse Ld, then D1(H)=D2(H) and D3(H) describe locally the dual of the curve Σ and B, respectively. On the other hand, if πp coincides with Ld and it is tangent to Td, then D1(H) composed of two parts: one is D2(H) which is the dual of Σ) and the second is the proper dual of G away from points of Σ, whereas the set D3(H) describes locally the dual of the curve B.
In this paper, we calculated the generators of the vector fields that are tangent to the bifurcation diagrams and caustics of the classes Bk, k=2,3,4 and F2,3 with respect to the quasi equivalence which is a non-standard equivalence relation. Consequently, we considered for application the generators of the B3-class in which case the bifurcation diagram consists of two components: a cuspidal edge in R3 and a smooth curve in it, which serves as a boundary and denoted it by V=(V1,V0). Then, we classified the submersion on V with codimension less or equal 2. This model and classifications were used to study the geometry of the pair ˜G=(G,B) of the geometric cuspidal edge G equipped with a distinguished curve B in it. Apart from the standard structure, ˜G contains two curves: the singular pints (the ridge) Σ and the smooth curve B. Thus, we discussed and described the contact of ˜G with the plane πp at p∈˜G along the curves Σ and B via the height function on ˜G, using the zero fibers of the submesrion obtained on V. In particular, we distinguished two cases. First, if πp is the tangent cone to ˜G, then the contact is of type A1 along both Σ and B if and only if πp is not the osculating plane of neither Σ nor B, and of type A1 along Σ and A2 along B if and only if πp is not the osculating plane of Σ but it coincides with the osculating plane of B and τB(0)≠0 (the torsion of B at 0). Second, if πp is not the tangent cone to ˜G, then the contact is of type A2 along both Σ and B if and only if πp is not the osculating plane of neither Σ nor B.
Subsequent study extending beyond this work may involve examining the height function on other singular hypersurfaces in R3 characterized by a smooth or singular boundary. When the hypersurface is equipped with a distinguished singular curve, it is more intriguing as it may involve two transversal tangential directions, such as the situation of the cuspidal edge with a singular curve (cusp) in it.
Yanlin Li: Conceptualization, investigation, methodology, writing-review and editing; Fawaz Alharbi: Conceptualization, investigation, methodology, writing-review and editing. All authors have read and approved the final version of the manuscript for publication.
The authors declare they have not used Artificial Intelligence (AI) tools in the creation of this article.
We wish to thank the anonymous reviewers for their insightful suggestions and careful reading of the manuscript.
The first author would like to express his gratitude to Raul Oset Sinha from the Universitat de València for the valuable discussion.
The authors declare no conflicts of interest.
[1] |
V. M. Zakalyukin, Reconstructions of fronts and caustics depending on a parameter and versality of mappings, J. Math. Sci., 27 (1984), 2713–2735. https://doi.org/10.1007/BF01084818 doi: 10.1007/BF01084818
![]() |
[2] |
V. Arnold, Wave front evolution and equivariant Morse lemma, Commun. Pur. Appl. Math., 29 (1976), 557–582. https://doi.org/10.1002/cpa.3160290603 doi: 10.1002/cpa.3160290603
![]() |
[3] |
J. W. Bruce, Vector fields on discriminants and bifurcation varieties, Bull. London Math. Soc., 17 (1985), 257–262. https://doi.org/10.1112/blms/17.3.257 doi: 10.1112/blms/17.3.257
![]() |
[4] |
A. Alghanemi, A. Alghawazi, The λ-point map between two Legendre plane curves, Mathematics, 11 (2023), 977–997. https://doi.org/10.3390/math11040997 doi: 10.3390/math11040997
![]() |
[5] |
T. Fukui, M. Hasegawa, Singularities of parallel surfaces, Tohoku Math. J., 64 (2012), 387–408. https://doi.org/10.2748/tmj/1347369369 doi: 10.2748/tmj/1347369369
![]() |
[6] |
Y. Li, E. Guler, Right conoids demonstrating a Time-like axis within minkowski Four-Dimensional space, Mathematics, 12 (2024), 2421. https://doi.org/10.3390/math12152421 doi: 10.3390/math12152421
![]() |
[7] |
Y. Li, H. Abdel-Aziz, H. Serry, F. El-Adawy, M. Saad, Geometric visualization of evolved ruled surfaces via alternative frame in Lorentz-Minkowski 3-space, AIMS Math., 9 (2024), 25619–25635. https://doi.org/10.3934/math.20241251 doi: 10.3934/math.20241251
![]() |
[8] |
Y. Li N. Turki, S. Deshmukh, O. Belova, Euclidean hypersurfaces isometric to spheres, AIMS Math., 9 (2024), 28306–28319. https://doi.org/10.3934/math.20241373 doi: 10.3934/math.20241373
![]() |
[9] |
Y. Li, M. S. Siddesha, H. A. Kumara, M. M. Praveena, Characterization of Bach and Cotton Tensors on a Class of Lorentzian Manifolds, Mathematics, 12 (2024), 3130. https://doi.org/10.3390/math12193130 doi: 10.3390/math12193130
![]() |
[10] |
Y. Li, S. Bhattacharyya, S. Azami, Li-Yau type estimation of a semilinear parabolic system along geometric flow, J. Inequal Appl., 131 (2024). https://doi.org/10.1186/s13660-024-03209-y doi: 10.1186/s13660-024-03209-y
![]() |
[11] |
Y. Li, A. K. Mallick, A. Bhattacharyya, M. S. Stankovic, A conformal η-Ricci soliton on a Four-Dimensional lorentzian Para-Sasakian manifold, Axioms, 13 (2024), 753. https://doi.org/10.3390/axioms13110753 doi: 10.3390/axioms13110753
![]() |
[12] |
F. Alharbi, V. Zakalyukin, Quasi corner singularities, P. Steklov I. Math., 270 (2010), 1–14. https://doi.org/10.1134/S0081543810030016 doi: 10.1134/S0081543810030016
![]() |
[13] |
F. Alharbi, Quasi cusp singularities, J. Sing., 12 (2015), 1–18. https://doi.org/10.5427/jsing.2015.12a doi: 10.5427/jsing.2015.12a
![]() |
[14] |
F. Alharbi, S. Alsaeed, Quasi semi-border singularities, Mathematics, 7 (2019), 495. https://doi.org/10.3390/math7060495 doi: 10.3390/math7060495
![]() |
[15] |
F. Alharbi, Bifurcation diagrams and caustics of simple quasi border singularities, Topo. Appl., 9 (2012), 381–388. https://doi.org/10.1016/j.topol.2011.09.011 doi: 10.1016/j.topol.2011.09.011
![]() |
[16] | J. W. Bruce, J. M. West, Functions on cross-caps, Math. Proc. Cambridge, 123 (1988), 19–39. |
[17] |
A. P. Francisco, Functions on a swallowtail, arXiv Prep., 53 (2023), 52–74. https://doi.org/10.48550/arXiv.1804.09664 doi: 10.48550/arXiv.1804.09664
![]() |
[18] | R. O. Sinha, F. Tari, On the geometry of the cuspidal edge, Osaka J. Math., 55 (2018), 393–421. |
[19] |
R. O. Sinha, K. Saji, On the geometry of folded cuspidal edges, Rev. Mat. Complut., 31 (2018), 627–650. https://doi.org/10.1007/s13163-018-0257-6 doi: 10.1007/s13163-018-0257-6
![]() |
[20] |
J. Damon, A-equivalence and the equivalence of sections of images and discriminants, Singular. Theory Appl., 1462 (1991), 93–121. https://doi.org/10.1007/BFb0086377 doi: 10.1007/BFb0086377
![]() |
[21] | D. Mond, R. Buchweitz, Linear free divisors and quiver representations, London Math. Soc. Lecture Note Ser., 324 (2005), 18–20. |
[22] | V. Arnold, Singularities of caustics and wave fronts, Dordrecht: Kluwer Academic Publishers, 1990. |
[23] | J. W. Bruce, P. J. Giblin, Curves and singularities: A geometrical introduction to singularity theory, Cambridge University Press, 1984. |
Equivalent Germ | Constraints | d(g,R+(V)) | mini-versal unfolding |
±u | – | 0 | ±u |
±v+ϵu2,1 | ϵ≠0,±1 | 1 | ±v+ϵu2+λ1u |
±v+ϵu3,1 | ϵ≠0 | 2 | ±v+ϵu3+λ1u+λ2u2 |
±w+ϵu2,1 | ϵ≠0 | 2 | ±w+ϵu2+λ1u+λ2v |
*Note: The symbol ϵ represents a modulus, while the codimension refers to the dimension of the stratum. |