This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

On the classification of planar constrained differential systems under topological equivalence

Otavio Henrique Perez and Paulo Ricardo da Silva São Paulo State University (Unesp), Institute of Biosciences, Humanities and Exact Sciences. Rua C. Colombo, 2265, CEP 15054–000. S. J. Rio Preto, São Paulo, Brazil. [email protected] [email protected]
Abstract.

This paper concerns the local study of analytic constrained differential systems (or impasse systems) of the form A(x)x˙=F(x)A(x)\dot{x}=F(x), x2x\in\mathbb{R}^{2}, where FF is a vector field and AA is a matrix valued function. Using techniques of resolution of singularities with weighted blow ups, we extend well-known results in the literature on topological classification of phase portraits of planar vector fields to the context of such class of systems. We also study and classify phase portraits of constrained systems near singular points of the so called impasse set Δ={x:detA(x)=0}\Delta=\{x:\det A(x)=0\}.

Key words and phrases:
Constrained differential systems, Implicit ordinary differential equations, Topological equivalence, Blow ups, Newton polygon
2020 Mathematics Subject Classification:
34A09, 34C05, 34C08.
.

1. Introduction, statement of the problem and main results

A constrained differential system (or simply constrained system) defined in an open set U2U\subseteq\mathbb{R}^{2} is given by

(1) A(𝐱)𝐱˙=F(𝐱),A(\mathbf{x})\dot{\mathbf{x}}=F(\mathbf{x}),

where 𝐱U2\mathbf{x}\in U\subseteq\mathbb{R}^{2}, A(𝐱)A(\mathbf{x}) is a matrix valued function and F:U2F:U\rightarrow\mathbb{R}^{2} is a vector field. We assume that these maps are analytic in the open set UU. A point 𝐱𝟎\mathbf{x_{0}} such that δ(𝐱𝟎)=0\delta(\mathbf{x_{0}})=0 is called impasse point, where δ:U\delta:U\rightarrow\mathbb{R} denotes the determinant function detA(𝐱)\det A(\mathbf{x}). The so called impasse set is the set Δ={𝐱U:δ(𝐱)=0}\Delta=\{\mathbf{x}\in U:\delta(\mathbf{x})=0\}.

Systems of the form (1) have been widely studied in the literature. We refer to [6, 7, 10, 15, 20, 21] for theoretical aspects on this subject. For applications in electric circuit theory, see [16, 19].

This paper deals with constrained systems such that the matrix AA does not have constant rank. We will further suppose that δ(𝐱)=0\delta(\mathbf{x})=0 only in a closed subset of UU. This implies that the adjoint matrix AA^{*} of AA is not identically zero and AA=AA=det(A)IAA^{*}=A^{*}A=\det(A)I.

Multiplying both sides of (1) by AA^{*}, it can be shown that γ\gamma is solution of (1) if, and only if, γ\gamma is a solution of

(2) δ(𝐱)𝐱˙=A(𝐱)F(𝐱).\delta(\mathbf{x})\dot{\mathbf{x}}=A^{*}(\mathbf{x})F(\mathbf{x}).

The constrained system written in the form (2) will be called diagonalized constrained system and, just as in [6], the vector field

(3) X(𝐱)=A(𝐱)F(𝐱)X(\mathbf{x})=A^{*}(\mathbf{x})F(\mathbf{x})

will be called adjoint vector field. Observe that, outside the impasse set, the constrained system (1) can be written as

𝐱˙=A1(𝐱)F(𝐱),\dot{\mathbf{x}}=A^{-1}(\mathbf{x})F(\mathbf{x}),

which is a classical ordinary differential equation.

At an impasse point 𝐱𝟎\mathbf{x_{0}}, since the matrix A(𝐱𝟎)A(\mathbf{x_{0}}) is not invertible, the classical results on the existence and uniqueness of the solutions break down. However, near such point we can describe the phase portrait as follows. In the open set where δ(𝐱)>0\delta(\mathbf{x})>0, the phase portrait is given by the adjoint vector field XX. On the other hand, in the open set where δ(𝐱)<0\delta(\mathbf{x})<0 the phase portrait is given by X-X.

The study of singularities of constrained systems whose matrix AA does not have constant rank was addressed in [20, 21], where the authors considered singularities near regular points of the impasse set Δ\Delta, such as tangencies between Δ\Delta and the adjoint vector field XX and equilibrium points of XX on Δ\Delta. By regular impasse point we mean that the determinant function δ\delta satisfies δ(𝐱)0\nabla\delta(\mathbf{x})\neq 0. In such papers was adopted the following notion of equivalence.

Definition 1.

Two constrained systems δi(𝐱)𝐱˙=Xi(𝐱)\delta_{i}(\mathbf{x})\dot{\mathbf{x}}=X_{i}(\mathbf{x}), i=1,2,i=1,2, are CkC^{k}-orbitally equivalent at the points piΔip_{i}\in\Delta_{i}, i=1,2i=1,2, if there are two open sets V1p1V_{1}\ni p_{1} and V2p2V_{2}\ni p_{2} and a CkC^{k} diffeomorphism Φ:V1V2\Phi:V_{1}\rightarrow V_{2} such that

  1. (1)

    Φ\Phi maps the impasse set Δ1\Delta_{1} to the impasse set Δ2\Delta_{2};

  2. (2)

    Φ\Phi maps the phase curves of X1X_{1} in V1\Δ1V_{1}\backslash\Delta_{1} to the phase curves of X2X_{2} in V2\Δ2V_{2}\backslash\Delta_{2}, not necessarily preserving orientation.

If Φ\Phi is a homeomorphism, we say that is a C0C^{0}-orbital equivalence or topological orbital equivalence. In the case where Φ\Phi preserves orientation, we say that Φ\Phi is an orientation preserving CkC^{k}-orbital equivalence.

For normal forms of planar constrained systems where the matrix AA has constant rank, see [7] for instance.

The main goal of this paper is to study singularities and classify phase portraits of constrained differential systems whose matrix AA does not have constant rank. Differently from [20, 21], we consider singular points of the impasse set Δ\Delta. Furthermore, the adjoint vector field XX can present equilibrium points on Δ\Delta more degenerated than those ones considered in these references. More precisely:

Definition 2.

We say that pΔp\in\Delta is a singularity of the constrained system (2) if one of the following conditions is satisfied:

  1. (1)

    pp is an equilibrium point of the adjoint vector field (3), that is, X(p)=0X(p)=0;

  2. (2)

    the impasse set Δ\Delta is not smooth at pp;

  3. (3)

    pp is a tangency point between XX and Δ\Delta, that is, dδ(X)=0d\delta(X)=0.

Otherwise, we say that pp is non-singular. See Figure 1.

Refer to caption
Refer to caption
Figure 1. Singularities of a planar constrained system. The rightmost phase portrait represents a non-singular point. The impasse set is highlighted in red.

For our purposes, the resolution of singularities using weighted blow ups is the main tool. Resolution of singularities of vector fields was considered in [5, 8, 12, 13] and we refer to [1, 9] for an introduction on this subject. The resolution of singularities of planar analytic constrained systems having impasse set was addressed in [6], where the authors studied constrained systems whose impasse set is given by an homogeneous polynomial. In this paper are considered constrained systems whose impasse set is defined by an analytic function.

Roughly speaking, in the process of resolution we replace a singularity of the constrained system by a compact set homeomorphic to 𝕊1\mathbb{S}^{1}, called exceptional divisor (that is, we blow-up a singularity). In this new compact set, the constrained system (possibly) presents less degenerated singularities. Continuing this reasoning at each singularity that appears in the divisor, in the end of the process there will be only elementary singularities in the divisor. By elementary singularity, we mean the following.

Definition 3.

The planar constrained differential system (2) is elementary at pp if one of the following conditions holds:

  1. (1)

    pp is a non singular point;

  2. (2)

    pp is a semi-hyperbolic equilibrium point of XX (that is, at least one of its eigenvalues is nonzero) and pΔp\not\in\Delta;

  3. (3)

    If pp is a semi-hyperbolic equilibrium point of XX and pΔp\in\Delta, then Δ\Delta coincides with a local separatrix of XX at pp.

See Figure 2.

Refer to caption
Figure 2. Elementary points of a constrained system on the divisor. The impasse set is highlighted in red.

Once there are only elementary singularities, we study the dynamics on the exceptional divisor in order to classify the phase portrait of the original constrained system near the blown-up singularity.

The resolution of singularities was applied in the study of topological classification of phase portraits of vector fields (that is, classification of phase portraits under topological equivalence) in [5, 8], and the results presented here can be seen as a generalization for planar constrained systems of those ones in these papers.

All along the paper a constrained system defined near a point pp will be denoted by a triple (X,δ,p)(X,\delta,p), where XX is the adjoint vector field (3) and δ\delta is an analytic irreducible function such that the impasse set is given by Δ={δ=0}\Delta=\{\delta=0\}. When pp is the origin, a constrained system is simply denoted by (X,δ)(X,\delta). After a suitable finite sequence of weighted blow ups, we obtain the so called strict transformed (X~,δ~,D)(\widetilde{X},\widetilde{\delta},D) of (X,δ,p)(X,\delta,p), where DD denotes the excepcional divisor.

In what follows we summarize the main results of the paper.

Firstly we give a condition based on the resolution of singularities to establish when two constrained differential systems are topologically orbitally equivalents (or C0C^{0}-orbitally equivalents) near a singularity. This condition is the so called elementary singularity scheme (see Section 3 for a precise definition). Roughly speaking, we prove that if two planar constrained systems have the same resolution of singularities, then they are equivalents. This result extends Theorem B of [8] to the context of constrained systems. More precisely:

Theorem A.

Consider two constrained systems (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}), where p1Δδp_{1}\in\Delta_{\delta} and p2Δγp_{2}\in\Delta_{\gamma} are singularities. If the strict transformed of both constrained systems have the same non-degenerated elementary singularity scheme, then (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}) are orientation preserving orbitally C0C^{0}-equivalents.

The Theorem A is a key theorem that we will use to prove the Theorems B and C. The next goal is to classify phase portraits of planar constrained systems having singular impasse curve. We define the set 𝒜\mathcal{A} as the set of analytic planar constrained systems (X,δ)(X,\delta) defined near the origin such that:

  1. (1)

    The adjoint vector field is constant, that is, it satisfies X(x,y)=(a,b)X(x,y)=(a,b), a,ba,b\in\mathbb{R} and a2+b20a^{2}+b^{2}\neq 0.

  2. (2)

    The function δ\delta is one of the maps given by Arnold’s ADE-classification [2] (see Table 1 in Section 4).

The Theorem B give conditions to assure when an analytic constrained system (X,δ)(X,\delta) is orbitally equivalent to a system in 𝒜\mathcal{A}, where the coefficients a1,0a_{-1,0} and b0,1b_{0,-1} in the expansion

(4) {δ(x,y)x˙=a1,0+a0,0x+a1,1y+,δ(x,y)y˙=b0,1+b0,0y+b1,1x+,\left\{\begin{array}[]{rcl}\delta(x,y)\dot{x}&=&a_{-1,0}+a_{0,0}x+a_{-1,1}y+...,\\ \delta(x,y)\dot{y}&=&b_{0,-1}+b_{0,0}y+b_{1,-1}x+...,\end{array}\right.

satisfy a1,02+b0,120a_{-1,0}^{2}+b_{0,-1}^{2}\neq 0 and the function δ\delta is one of the functions in Table 1.

Theorem B.

Consider an analytic planar constrained system (X,δ)(X,\delta) near the origin such that X(0)0X(0)\neq 0 and the origin is an ADE-type singular point of δ\delta. Then the system (4) is orientation preserving C0C^{0}-orbitally equivalent to a system in 𝒜\mathcal{A} if it satisfies one of the conditions in Table 2, Section 4.

In section 4 we present all the possible phase portraits of the systems that belong to 𝒜\mathcal{A}.

Concerning Theorem C, suppose that the origin 0 is a singularity of an analytic constrained system (X,δ)(X,\delta), where 0 is an equilibrium point of the adjoint vector field XX and a regular point of the impasse set Δ\Delta simultaneously. We associate a Newton polygon 𝒫\mathcal{P} to such constrained system and we show in Theorem C that the terms in the boundary 𝒫\partial\mathcal{P} of the polygon determine the phase portrait near the origin under topological orbital equivalence. In other words, the terms in the boundary 𝒫\partial\mathcal{P} define the so called principal part (XΩ,δ)(X_{\Omega},\delta) of the analytic constrained system (see Section 5 for a precise definition), and we prove that, under some non-degeneracy conditions, the original constrained system is orbitally equivalent to its principal part.

From a practical point of view, Theorem C says that if we want to study topological properties of a singularity in the regular part of Δ\Delta, one can “discard” the terms of the Taylor expansion that are not in the boundary 𝒫\partial\mathcal{P} of the polygon. This result extends Theorem A of [5] to the context of constrained systems with regular impasse set.

Theorem C.

Let (X,δ)(X,\delta) be an analytic planar constrained system such that Δ={δ=0}\Delta=\{\delta=0\} is a regular curve and the origin is an isolated Newton non-degenerated singularity. If XX has characteristic orbit (that is, the origin is neither a center nor a focus of XX), then (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) are orientation preserving C0C^{0}-orbitally equivalent.

The paper is organized as follows. In Section 2 we briefly introduce weighted blowing ups and the Theorem of resolution of singularities for planar constrained vector fields. In Section 3 is given the proof of Theorem A, where we use the resolution of singularities to establish when two constrained systems are C0C^{0}-orbitally equivalents. We apply this result in the following sections. In Section 4 we recall the construction of the Newton polygon and see how to define such mathematical object for planar constrained systems. Moreover, we study and classify phase portraits of constrained systems whose impasse curve has a singular point given by Arnold’s ADE-classification. Finally, in Section 5 we show that the terms in the boundary 𝒫\partial\mathcal{P} of the Newton polygon determines (under topological orbital equivalence) the phase portrait near a singularity of a constrained system in the case where the origin is an equilibrium point of XX and a regular point of Δ\Delta.

2. Preliminaries

The results presented in this paper strongly depends on the resolution of singularities. First of all we introduce weighted (or quasi-homogeneous) blow-ups and later we discuss some aspects concerning resolution of singularities of planar constrained systems. For an introduction on blow-ups for planar vector fields we refer [1, 9].

2.1. Weighted blow-ups

Let 0 be an isolated equilibrium point of the planar vector field XX. Given positive integers ω1\omega_{1} and ω2\omega_{2}, a weighted (or quasi homogeneous) polar blow-up is the map

ϕ(ω1,ω2):\phi_{(\omega_{1},\omega_{2})}: 𝕊1×\mathbb{S}^{1}\times\mathbb{R} \rightarrow 2\mathbb{R}^{2},
(θ,r)(\theta,r) \mapsto (rω1cosθ,rω2sinθ)(r^{\omega_{1}}\cos\theta,r^{\omega_{2}}\sin\theta).

Let X^\widehat{X} be the vector field defined in 𝕊1×\mathbb{S}^{1}\times\mathbb{R} that satisfies ϕ(ω1,ω2)(X^)=X\phi_{(\omega_{1},\omega_{2})*}(\widehat{X})=X. Define the vector field X¯\overline{X} as X¯=1rkX^\overline{X}=\frac{1}{r^{k}}\widehat{X}, where k1k\geq 1 is the biggest positive integer such that X¯\overline{X} is an analytic vector field. Observe that X¯\overline{X} is defined in a new ambient space and it is invariant in 𝕊1×{0}\mathbb{S}^{1}\times\{0\}. The set 𝕊1×{0}\mathbb{S}^{1}\times\{0\} is called exceptional divisor and the origin 0 is called blow up center.

In general, during the resolution of singularities it is necessary to iterate a finite number of blowing ups, and considering polar blow ups can lead us to complicated trigonometric expressions. For our purposes, it is way more practical to consider weighted (or quasi homogeneous) directional blow-ups. We define, respectively, the positive xx-directional and positive yy-directional blow ups as the maps

ϕx,(ω1,ω2)+:\phi_{x,(\omega_{1},\omega_{2})}^{+}: 0×\mathbb{R}_{\geq 0}\times\mathbb{R} \rightarrow 2\mathbb{R}^{2},
(x~,y~)(\tilde{x},\tilde{y}) \mapsto (x~ω1,x~ω2y~)(\tilde{x}^{\omega_{1}},\tilde{x}^{\omega_{2}}\tilde{y});
ϕy,(ω1,ω2)+:\phi_{y,(\omega_{1},\omega_{2})}^{+}: ×0\mathbb{R}\times\mathbb{R}_{\geq 0} \rightarrow 2\mathbb{R}^{2},
(x~,y~)(\tilde{x},\tilde{y}) \mapsto (y~ω1x~,y~ω2)(\tilde{y}^{\omega_{1}}\tilde{x},\tilde{y}^{\omega_{2}}).

Throughout this paper we only consider the computations in the positive directions because such calculations are analogous when we consider the negative directions. The exceptional divisor will be denoted by 𝔼x={x=0}\mathbb{E}_{x}=\{x=0\} and 𝔼y={y=0}\mathbb{E}_{y}=\{y=0\} in the xx and yy directions, respectively. Observe that the polar and directional blow ups are equivalent (see Figure 3).

(π2,π2)×(-\displaystyle\frac{\pi}{2},\displaystyle\frac{\pi}{2})\times\mathbb{R}2\mathbb{R}^{2}0×\mathbb{R}_{\geq 0}\times\mathbb{R}(0,π)×(0,\pi)\times\mathbb{R}2\mathbb{R}^{2}×0\mathbb{R}\times\mathbb{R}_{\geq 0}ϕ(ω1,ω2)\phi_{(\omega_{1},\omega_{2})}ϕx,(ω1,ω2)+\phi_{x,(\omega_{1},\omega_{2})}^{+}ξx\xi_{x}ϕ(ω1,ω2)\phi_{(\omega_{1},\omega_{2})}ϕy,(ω1,ω2)+\phi_{y,(\omega_{1},\omega_{2})}^{+}ξy\xi_{y}
Figure 3. Equivalence between polar and positive xx directional blow up (left), and polar and positive yy directional blow up (right).

2.2. Resolution of singularities for planar constrained systems

We emphasize that only isolated singularities of constrained systems are considered. Although a tangency point is a singularity, in this work we focus in the case where pp is an equilibrium point of XX and/or a singular point of the impasse set. Nevertheless, the theorem of resolution of singularities for 2-dimensional constrained differential systems (see Theorem 4) also deals with tangency points.

In order to state the theorem of resolution of singularities, one needs to work in a more general category of analytic manifolds with corners. Roughly speaking, a 2-dimensional manifold with corners is a topological space locally modeled by open sets of (0)2(\mathbb{R}_{\geq 0})^{2}. Just as we mentioned before, after a blow-up in a singularity we obtain a new ambient space, which is a manifold with corners.

A constrained differential system defined in a 2-dimensional manifold with corners is a triple (,𝒳,)(\mathcal{M},\mathcal{X},\mathcal{I}), where \mathcal{M} is a 2-dimensional real analytic manifold with corners, 𝒳={(Ui,Xi)}iI\mathcal{X}=\{(U_{i},X_{i})\}_{i\in I} is an 1-dimensional analytic oriented foliation defined on \mathcal{M} and ={(Ui,δi)}iI\mathcal{I}=\{(U_{i},\delta_{i})\}_{i\in I} is the principal ideal sheaf. On each open set UiU_{i} of the open covering of \mathcal{M}, we associate the diagonalized constrained differential system

δi(𝐱)𝐱˙=Xi(𝐱).\delta_{i}(\mathbf{x})\dot{\mathbf{x}}=X_{i}(\mathbf{x}).
Theorem 4.

Let (,𝒳,)(\mathcal{M},\mathcal{X},\mathcal{I}) be a 2-dimensional real analytic constrained differential system defined on a compact manifold with corners. Then there is a finite sequence of weighted blow ups

(~,𝒳~,~)=(n,𝒳n,n)ΦnΦ0(0,𝒳0,0)=(,𝒳,)(\widetilde{\mathcal{M}},\widetilde{\mathcal{X}},\widetilde{\mathcal{I}})=(\mathcal{M}_{n},\mathcal{X}_{n},\mathcal{I}_{n})\xrightarrow{\Phi_{n}}\cdots\xrightarrow{\Phi_{0}}(\mathcal{M}_{0},\mathcal{X}_{0},\mathcal{I}_{0})=(\mathcal{M},\mathcal{X},\mathcal{I})

such that (M~,X~,~)(\widetilde{M},\widetilde{X},\widetilde{\mathcal{I}}) is elementary.

A proof for Theorem 4 can be given by firstly applying the classical Bendixson–Seidenberg Theorem [4, 18] to reduce the singularities of the 1-dimensional foliation 𝒳\mathcal{X} and later applying results from algebraic geometry on resolution of singularities subordinated to foliations. Indeed, once the equilibrium points of the foliation are elementary (that is, the foliation is Log-Canonical), by [3] there is a resolution for the ideal sheaf \mathcal{I} that preserves the Log-Canonicity of the foliation. For a proof that does not require general results from algebraic geometry and uses weighted blow ups, we refer to [14].

The Theorem 4 is a global theorem, but the study carried throughout this paper is local. Without loss of generality, we always suppose that the constrained system is written in the diagonalized form (2). A constrained system defined near a point pp will be denoted as a triple (X,δ,p)(X,\delta,p), where XX is the adjoint vector field and δ\delta is an irreducible analytic function such that Δ={δ=0}\Delta=\{\delta=0\}.

Example 5.

Consider the following constrained differential equation

(5) xyx˙=y,xyy˙=x2.xy\dot{x}=y,\ xy\dot{y}=x^{2}.

The adjoint vector field is given by X(x,y)=(y,x2)X(x,y)=(y,x^{2}), and therefore there is a cusp singularity at the origin. Moreover, the impasse set is given by Δ={δ=xy=0}\Delta=\{\delta=xy=0\}. We consider the directional weighted blow-ups in the positive xx and yy directions

(x,y)=(x~2,x~3y~),(x,y)=(y¯2x¯,y¯3),(x,y)=(\tilde{x}^{2},\tilde{x}^{3}\tilde{y}),\ (x,y)=(\bar{y}^{2}\bar{x},\bar{y}^{3}),

respectively. Thus we obtain the constrained system

y~x~˙=x~y~2,y~y~˙=13y~22\tilde{y}\dot{\tilde{x}}=\displaystyle\frac{\tilde{x}\tilde{y}}{2},\ \tilde{y}\dot{\tilde{y}}=1-\displaystyle\frac{3\tilde{y}^{2}}{2}

in the positive xx direction and

x¯x¯˙=123x¯3,x¯y¯˙=x¯2y¯3\bar{x}\dot{\bar{x}}=1-\displaystyle\frac{2}{3}\bar{x}^{3},\ \bar{x}\dot{\bar{y}}=\displaystyle\frac{\bar{x}^{2}\bar{y}}{3}

in the positive yy direction. In Section we briefly discuss how to make a good choice of the weight vector. See Figure 4.

Refer to caption
Figure 4. Constrained system (5) after (left) and before (right) the weighted blow-up at the origin.

3. The construction of the orbital C0C^{0}-equivalence near the exceptional divisor

Let (X,δ,p1)(X,\delta,p_{1}) be a constrained differential system defined near an impasse point p1Δδp_{1}\in\Delta_{\delta}. Suppose that p1p_{1} is an equilibrium point of XX or a singular point of the impasse set Δδ\Delta_{\delta}. In this point is applied a suitable finite sequence of weighted blowing-ups, whose composition is simply denote by Φ\Phi. According to Theorem 4, at the end of the process, we obtain a constrained system (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) such that all the points in a neighborhood of the exceptional divisor D1=Φ1(p1)D_{1}=\Phi^{-1}(p_{1}) are elementary. Moreover, D1D_{1} is an invariant set of X~\widetilde{X}.

Analogously, let (Y,γ,p2)(Y,\gamma,p_{2}) be a planar constrained system where p2Δγp_{2}\in\Delta_{\gamma} is an equilibrium point of the adjoint vector field YY or a singular point of the impasse set Δγ\Delta_{\gamma}. After a finite number of weighted blowing-ups (whose composition we denote by Ψ\Psi), we obtain the strict transformed (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) where all the points near the exceptional divisor D2=Ψ1(p2)D_{2}=\Psi^{-1}(p_{2}) are elementary and D2D_{2} is an invariant set of Y~\widetilde{Y}.

The main goal is to use the resolution of singularities in the topological classification of phase portraits. In order to achieve this objective, firstly we give conditions based in the resolution of singularities to assure the equivalence between two constrained systems, that is, we study conditions to assure the existence of an orientation preserving orbital C0C^{0}-equivalence HH between (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}). The Proposition 6 will say that if there is an orientation preserving orbital C0C^{0}-equivalence H~\widetilde{H} between (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) in a neighborhood of the exceptional divisor, then (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}) are equivalents. See Figure 5.

(X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1})(Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2})(X,δ,p1)(X,\delta,p_{1})(Y,γ,p2)(Y,\gamma,p_{2})Φ\PhiΨ\PsiH~\widetilde{H}HH
Figure 5. Diagram of Proposition 6.

Let W1p1W_{1}\ni p_{1} and W2p2W_{2}\ni p_{2} be neighborhoods of the blow-up centers p1p_{1} and p2p_{2}, respectively. Observe that W~1=Φ1(W1\{p1})\widetilde{W}_{1}=\Phi^{-1}(W_{1}\backslash\{p_{1}\}) and W~2=Ψ1(W2\{p2})\widetilde{W}_{2}=\Psi^{-1}(W_{2}\backslash\{p_{2}\}) are open sets. Define the open sets U~=W~1D1\widetilde{U}=\widetilde{W}_{1}\cup D_{1} and V~=W~2D2\widetilde{V}=\widetilde{W}_{2}\cup D_{2} and suppose that there is a homeomorphism (possibly taking smaller neighborhoods W1W_{1} and W2W_{2} if it is necessary) H~:U~V~\widetilde{H}:\widetilde{U}\rightarrow\widetilde{V} satisfying the following conditions:

(H1):

H~(D1)=D2\widetilde{H}(D_{1})=D_{2};

(H2):

H~(Δ~δ)=Δ~γ\widetilde{H}(\widetilde{\Delta}_{\delta})=\widetilde{\Delta}_{\gamma};

(H3):

Let pU~p\in\widetilde{U} and denote by ϕX~(p,t)\phi_{\widetilde{X}}\big{(}p,t\big{)} the flow of X~\widetilde{X} through pp. Suppose that ϕX~(p,[0,s])U~\Δ~δ\phi_{\widetilde{X}}\big{(}p,[0,s]\big{)}\subset\widetilde{U}\backslash\widetilde{\Delta}_{\delta}, for some s>0s>0. Then there is s^>0\hat{s}>0 such that

H~(ϕX~(p,[0,s]))=ϕY~(H~(p),[0,s^]).\widetilde{H}\Bigg{(}\phi_{\widetilde{X}}\big{(}p,[0,s]\big{)}\Bigg{)}=\phi_{\widetilde{Y}}\Bigg{(}\widetilde{H}(p),[0,\hat{s}]\Bigg{)}.
\begin{overpic}[width=260.17464pt]{fig-hypotheses-h1-h2-h3} \put(50.0,30.0){$\widetilde{H}$} \put(0.0,46.0){$D_{1}$} \put(65.0,35.0){$D_{2}$} \put(30.0,15.0){$\widetilde{\Delta}_{\delta}$} \put(100.0,30.0){$\widetilde{\Delta}_{\gamma}$} \end{overpic}
Figure 6. Homeomorphism H~\widetilde{H} that satisfies the hypotheses (H1), (H2) and (H3). The impasse set is highlighted in red.
Proposition 6.

If there is a homeomorphism H~\widetilde{H} satisfying the conditions (H1), (H2) and (H3) above, then there are neighborhoods Up1U\ni p_{1} and Vp2V\ni p_{2} and an orientation preserving orbital C0C^{0}-equivalence H:UVH:U\rightarrow V between the constrained systems (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}).

Proof.

Denote U=Φ(U~)U=\Phi(\widetilde{U}) and V=Ψ(V~)V=\Psi(\widetilde{V}) and define the map H:UVH:U\rightarrow V as

H(p1)=p2;H(p)=ΨH~Φ1(p),pΦ(W~1).H(p_{1})=p_{2};\ H(p)=\Psi\circ\widetilde{H}\circ\Phi^{-1}(p),\ p\in\Phi(\widetilde{W}_{1}).

We claim that HH is an orientation preserving orbital C0C^{0}-equivalence between (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}). Recall that Φ\Phi restricted to W~1=Φ1(W1\{p1})\widetilde{W}_{1}=\Phi^{-1}(W_{1}\backslash\{p_{1}\}) is an analytic diffeomorphism that maps Δ~δ\widetilde{\Delta}_{\delta} into Δδ\Delta_{\delta}, and it maps phase curves of X~\widetilde{X} into phase curves of XX, preserving orientation. Analogous properties for Ψ\Psi restricted to W~2=Ψ1(W2\{p2})\widetilde{W}_{2}=\Psi^{-1}(W_{2}\backslash\{p_{2}\}) are true. Moreover, Φ(D1)=p1\Phi(D_{1})=p_{1}, H~(D1)=D2\widetilde{H}(D_{1})=D_{2} and Ψ(D2)=p2\Psi(D_{2})=p_{2}.

A straightforward computation shows that HH maps impasse set into impasse set. Furthermore, if ϕX(p,t)\phi_{X}\big{(}p,t\big{)} is the flow of XX through pUp\in U and ϕX(p,[0,s])U\Δδ\phi_{X}\big{(}p,[0,s]\big{)}\subset U\backslash\Delta_{\delta} for some s>0s>0, then there is s^>0\hat{s}>0 such that

H(ϕX(p,[0,s]))=ϕY(H(p),[0,s^]),H\Bigg{(}\phi_{X}\big{(}p,[0,s]\big{)}\Bigg{)}=\phi_{Y}\Bigg{(}H(p),[0,\hat{s}]\Bigg{)},

and therefore HH is an orientation preserving orbital C0C^{0}-equivalence. ∎

Since the existence of H~\widetilde{H} assures the existence of an orientation preserving orbital C0C^{0}-equivalence HH between (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}), now we will provide a condition based on the resolution of singularities that guarantees the existence of such homeomorphism H~\widetilde{H}. This condition is the so called elementary singularity scheme. The next theorem says that if (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) have the same elementary singularity scheme, then such triples are (orientation preserving) orbitally C0C^{0}-equivalent near the exceptional divisor. Such result and the Proposition 6 will assure that (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}) are orbitally C0C^{0}-equivalent near the blow-up center.

3.1. Basic points, basic singular interval and elementary singularity scheme

Let (X,δ,p1)(X,\delta,p_{1}) be a planar constrained system where p1Δδp_{1}\in\Delta_{\delta} is an equilibrium point of XX or a singular point of the impasse set. Consider its strict transformed (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) where all points in the exceptional divisor D1D_{1} are elementary.

We say that a coordinate system is adapted to the exceptional divisor if in such coordinate system the exceptional divisor is given by 𝔼x={x=0}\mathbb{E}_{x}=\{x=0\}, 𝔼y={y=0}\mathbb{E}_{y}=\{y=0\} or 𝔼xy={xy=0}\mathbb{E}_{xy}=\{xy=0\} (see Definition 5.1, [8]).

Define the following vector fields with restricted domain, where α,β{1,1}\alpha,\beta\in\{-1,1\}:

  1. (1)

    V1α,βV_{1}^{\alpha,\beta}: αxx+βyy,\alpha x\frac{\partial}{\partial x}+\beta y\frac{\partial}{\partial y},  y0y\geq 0. See Figure 7;

  2. (2)

    V2α,βV_{2}^{\alpha,\beta}: βx2x+αyy,\beta x^{2}\frac{\partial}{\partial x}+\alpha y\frac{\partial}{\partial y},  y0y\geq 0. See Figure 8;

  3. (3)

    V3αV_{3}^{\alpha}: αxxαyy,\alpha x\frac{\partial}{\partial x}-\alpha y\frac{\partial}{\partial y},  x0x\geq 0 and y0y\geq 0. See Figure 9;

  4. (4)

    V4αV_{4}^{\alpha}: αyy,\alpha y\frac{\partial}{\partial y},  0x10\leq x\leq 1 and y0y\geq 0. See Figure 9.

Analogously, define the following constrained differential systems with restricted domain, where α,β{1,1}\alpha,\beta\in\{-1,1\}:

  1. (1)

    C1α,βC_{1}^{\alpha,\beta}: xx˙=αx,xy˙=βy,x\dot{x}=\alpha x,x\dot{y}=\beta y,  y0y\geq 0. See Figure 10;

  2. (2)

    C2α,βC_{2}^{\alpha,\beta}: xx˙=βx2,xy˙=αy,x\dot{x}=\beta x^{2},x\dot{y}=\alpha y,  y0y\geq 0. See Figure 11;

  3. (3)

    C3αC_{3}^{\alpha}: xx˙=0,xy˙=α,x\dot{x}=0,x\dot{y}=\alpha,  0x10\leq x\leq 1 and y0y\geq 0. See Figure 12;

  4. (4)

    C4αC_{4}^{\alpha}: xx˙=0,xy˙=αy,x\dot{x}=0,x\dot{y}=\alpha y,  0x10\leq x\leq 1 and y0y\geq 0. See Figure 12.

Refer to caption
Figure 7. Basic singularities of the form V1α,βV^{\alpha,\beta}_{1}.
Refer to caption
Figure 8. Basic singularities of the form V2α,βV^{\alpha,\beta}_{2}.
Refer to caption
Refer to caption
Figure 9. Basic singularities of the form V3αV^{\alpha}_{3} (left) and V4αV^{\alpha}_{4} (right).
Refer to caption
Figure 10. Basic singularities of the form C1α,βC^{\alpha,\beta}_{1}. The impasse set is highlighted in red.
Refer to caption
Figure 11. Basic singularities of the form C2α,βC^{\alpha,\beta}_{2}. The impasse set is highlighted in red.
Refer to caption
Refer to caption
Figure 12. Basic singularities of the form C3αC^{\alpha}_{3} (left) C4αC^{\alpha}_{4} (right). The impasse set is highlighted in red.
Definition 7.

A point pD1p\in D_{1} is a basic singularity if in a neighborhood UU of pp there is a coordinate system adapted to the divisor such that the constrained system is orientation preserving orbitally C0C^{0}-equivalent to Viα,βV_{i}^{\alpha,\beta} or Cjα,βC_{j}^{\alpha,\beta}, where i,j=1,2,3i,j=1,2,3.

Definition 8.

A basic singular interval is an arc dD1d\subset D_{1} such that all the points in dd are equilibrium points and in a neighborhood UU of dd there is a coordinate system adapted to the divisor such that the constrained system is orientation preserving orbitally C0C^{0}-equivalent to V4αV_{4}^{\alpha} or C4αC_{4}^{\alpha}.

Since the excepcional divisor is a finite union of smooth arcs, one can enumerate them (without loss of generality) in the clockwise sense. We suppose that di+1d_{i+1} follows did_{i} and dn+1=d1d_{n+1}=d_{1}. With this orientation, one can also order the basic singularities and the basic singular intervals in the excepcional divisor. The arrangement of basic singularities and basic singular intervals in the divisor defines a finite word constructed with the alphabet

{Viα,β}i=12{Cjα,β}j=12{Viα}i=34{Cjα}j=34.\{V_{i}^{\alpha,\beta}\}_{i=1}^{2}\cup\{C_{j}^{\alpha,\beta}\}_{j=1}^{2}\cup\{V_{i}^{\alpha}\}_{i=3}^{4}\cup\{C_{j}^{\alpha}\}_{j=3}^{4}.

Such arrangement defines an equivalence relation in the set Σ\Sigma of triples (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) in the following way: two triples (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) are equivalent if, and only if, the word associated to (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) can be changed to the word associated to (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) by cyclic permutations.

Definition 9.

An equivalence class in Σ\Sigma is called elementary singularity scheme.

Example 10.

The elementary singularity scheme of the constrained system Example 5 is given by the word C31V11,1C31V11,1C31C31C^{1}_{3}V^{-1,1}_{1}C^{1}_{3}V^{-1,1}_{1}C^{1}_{3}C^{-1}_{3}.

The next theorem is true under the hypothesis that the elementary singularity schemes do not contain only words written with symbols of the form V3αV_{3}^{\alpha} or C3αC_{3}^{\alpha}. This is equivalent to require the following. Let (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) be the strict transformed of the constrained systems (X,δ,p1)(X,\delta,p_{1}) and (Y,γ,p2)(Y,\gamma,p_{2}), respectively, where p1Δδp_{1}\in\Delta_{\delta} and p2Δγp_{2}\in\Delta_{\gamma}. If p1p_{1} and p2p_{2} are equilibrium points of the adjoint vector field XX and YY, respectively, then both p1p_{1} and p2p_{2} have characteristic orbit. In other words, p1p_{1} and p2p_{2} are neither a center nor a focus. See Figure 13.

Definition 11.

An elementary singularity scheme is degenerated if the word associated to the triple (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) only contains symbols of the form V3αV_{3}^{\alpha} or C3αC_{3}^{\alpha}. Otherwise, we say that the elementary singularity scheme is non-degenerated.

Refer to caption
Figure 13. Singularities of a constrained system. The equilibrium point of the adjoint vector field does not have characteristic orbit.

3.2. Proof of Theorem A: The construction of the orbital C0C^{0}-equivalence

The next lemma is a well-known result from general topology and it will be useful in the proof of Proposition 13.

Lemma 12.

(The Pasting Lemma, Theorem 18.3, [11]) Let M,NM,N be topological spaces and let A1,A2MA_{1},A_{2}\subset M be open (or closed) subspaces. Denote A=A1A2A=A_{1}\cap A_{2}. Let f:A1f(A1)Nf:A_{1}\rightarrow f(A_{1})\subset N and g:A2g(A2)Ng:A_{2}\rightarrow g(A_{2})\subset N be homeomorphisms such that f|A=g|Af|_{A}=g|_{A}. Then there is a homeomorphism H:A1A2H(A1A2)NH:A_{1}\cup A_{2}\rightarrow H(A_{1}\cup A_{2})\subset N such that H|A1=fH|_{A_{1}}=f and H|A2=gH|_{A_{2}}=g.

Proposition 13.

Suppose that the triples (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) have the same non-degenerated elementary singularity scheme. Then there is an orientation preserving orbital C0C^{0}-equivalence HH between (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}) such that H(D1)=D2H(D_{1})=D_{2}.

Proof.

Given an arc ciD1c_{i}\subset D_{1}, we have its respective c^iD2\widehat{c}_{i}\subset D_{2}. The construction of the homeomorphism starts near a basic point pp (or basic singular interval) of (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) that is not of the form V3αV_{3}^{\alpha} or C3αC_{3}^{\alpha}. Hence there is an equivalent point p^\widehat{p} (or basic singular interval) of (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}). This means that there is adapted coordinates

h:U2,h^:U^2,h:U\rightarrow\mathbb{R}^{2},\ \ \widehat{h}:\widehat{U}\rightarrow\mathbb{R}^{2},

for (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}), respectively, where UpU\ni p and U^p^\widehat{U}\ni\widehat{p}. Denote h(U)=h^(U^)=Wh(U)=\widehat{h}(\widehat{U})=W. The neighborhoods UU and U^\widehat{U} are chosen in such way that there is only one basic point (or only one basic singular interval).

\begin{overpic}[width=216.81pt]{fig-dumortier} \put(10.0,28.0){$U$} \put(90.0,30.0){$\widehat{U}$} \put(38.0,38.0){$h$} \put(52.0,38.0){$\widehat{h}$} \put(68.0,18.0){$W$} \end{overpic}
Figure 14. Neighborhoods UU, U^\widehat{U} and WW.

If we start the construction of the homeomorphism from a basic point, we take the open set WW in such way that it is transversal to the coordinate axes. On the other hand, if the construction starts from a basic singular interval, the neighborhood WW is chosen in such way that it is transversal to the lines {x=0}\{x=0\} and {x=1}\{x=1\}.

Therefore, we have that h1(W)h^{-1}(\partial W) is transversal to D1UD_{1}\cap U. This is also true when we take (h^)1(W)(\widehat{h})^{-1}(\partial W) and D2U^D_{2}\cap\widehat{U}. Applying this reasoning to all basic points and basic singular intervals in the exceptional divisor, a sufficiently small neighborhood of the exceptional divisor D1D_{1} is divided into a finite number of sectors. If this sector is not locally given by V4αV_{4}^{\alpha}, we then subdivide such sector by the segment h1({x=0})h^{-1}(\{x=0\}). We apply the same reasoning for (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}).

These sectors are bounded by arcs of the excepcional divisor, phase curves, impasse curves or curves transversal to the divisor. Considering α{1,1}\alpha\in\{-1,1\}, these sectors are written in local coordinates as follows:

  • Sector S1αS_{1}^{\alpha}: α(x2+y2)x\alpha(x^{2}+y^{2})\frac{\partial}{\partial x}, where 1x1-1\leq x\leq 1 and y0y\geq 0. This sector is equivalent to the basic point V3αV^{\alpha}_{3};

  • Sector S2αS_{2}^{\alpha}: αxx+αyy\alpha x\frac{\partial}{\partial x}+\alpha y\frac{\partial}{\partial y}, where 1x1-1\leq x\leq 1 and y0y\geq 0;

  • Sector S3αS_{3}^{\alpha}: αyy\alpha y\frac{\partial}{\partial y}, where 1x1-1\leq x\leq 1 and y0y\geq 0;

  • Sector FαF^{\alpha}: αx\alpha\frac{\partial}{\partial x}, where 1x1-1\leq x\leq 1 and y0y\geq 0;

  • Sector SI1αSI_{1}^{\alpha}: xx˙=αx,xy˙=αyx\dot{x}=\alpha x,x\dot{y}=\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector SI2αSI_{2}^{\alpha}: xx˙=αx,xy˙=αyx\dot{x}=-\alpha x,x\dot{y}=-\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector SI3αSI_{3}^{\alpha}: xx˙=αx,xy˙=αyx\dot{x}=\alpha x,x\dot{y}=-\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector SI4αSI_{4}^{\alpha}: xx˙=αx,xy˙=αyx\dot{x}=-\alpha x,x\dot{y}=\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector SI5αSI_{5}^{\alpha}: xx˙=0,xy˙=αyx\dot{x}=0,x\dot{y}=\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector SI6αSI_{6}^{\alpha}: xx˙=0,xy˙=αyx\dot{x}=0,x\dot{y}=-\alpha y, where αx0\alpha x\geq 0 and y0y\geq 0;

  • Sector FIαFI^{\alpha}: xx˙=α,xy˙=0x\dot{x}=\alpha,x\dot{y}=0, where 1x1-1\leq x\leq 1 and y0y\geq 0.

We remark that the sector S1αS_{1}^{\alpha} is C0C^{0}-equivalent to V3αV_{3}^{\alpha}.

Refer to caption
Figure 15. Sectors S1αS_{1}^{\alpha} (left) and S2αS_{2}^{\alpha} (right).
Refer to caption
Figure 16. Sectors S3αS_{3}^{\alpha} (left) and FαF^{\alpha} (right).
Refer to caption
Figure 17. Sectors SI1αSI_{1}^{\alpha} (left) and SI2αSI_{2}^{\alpha} (right). The impasse set is highlighted in red.
Refer to caption
Figure 18. Sectors SI3αSI_{3}^{\alpha} (left) and SI4αSI_{4}^{\alpha} (right). The impasse set is highlighted in red.
Refer to caption
Figure 19. Sectors SI5αSI_{5}^{\alpha} (left) and SI6αSI_{6}^{\alpha} (right). The impasse set is highlighted in red.
Refer to caption
Figure 20. Sector FIαFI^{\alpha}. The impasse set is highlighted in red.

Consider once again the point pp that we took in the beginning of this proof. Recall that this point is neither V3αV_{3}^{\alpha} nor C3αC_{3}^{\alpha}. We know that (h^)1h:UU^(\widehat{h})^{-1}\circ h:U\rightarrow\widehat{U} is a homeomorphism such that h(p)=p^h(p)=\widehat{p}, it maps phase curves of X~\widetilde{X} into phase curves of Y~\widetilde{Y} and it sends impasse set into impasse set. Moreover, (h^)1h(\widehat{h})^{-1}\circ h sends excepcional divisor into excepcional divisor. Then we have constructed a homeomorphism for this first sector.

The adjacent sector (in the clockwise sense) must be of the form FαF^{\alpha} or FIαFI^{\alpha}. In this second sector it is defined a C0C^{0}-equivalence h^^\widehat{\widehat{h}} between (X~,δ~,D1)(\widetilde{X},\widetilde{\delta},D_{1}) and (Y~,γ~,D2)(\widetilde{Y},\widetilde{\gamma},D_{2}). Observe now that on U\partial U the homeomorphisms (h^)1h(\widehat{h})^{-1}\circ h and h^^\widehat{\widehat{h}} coincide.

Refer to caption
Figure 21. Gluing two adjacent sectors. The impasse set is highlighted in red.

We remark that each phase curve intersects U\partial U in only one point. Now, by the Pasting Lemma 12, there is an homeomorphism defined in both sectors that maps phase curves of X~\widetilde{X} into phase curves of Y~\widetilde{Y}, maps impasse set into impasse set and sends excepcional divisor into excepcional divisor.

Between each sector of the form SiαS_{i}^{\alpha} or SIjαSI_{j}^{\alpha} it may have a sector of the form FαF^{\alpha} or FIαFI^{\alpha}, and we need to glue all these sectors around the excepcional divisor. At the boundary of each sector, the Pasting Lemma 12 is applied.

At the end of the process, we must glue the last sector and the first sector in such a way that we “close” the construction of the homeomorphism, given that we are constructing such a homeomorphism around the exceptional divisor. Observe that, in general, such construction only can be closed if the basic point pp (or basic singular interval) that we took in the beginning of the proof is not of the form V3αV_{3}^{\alpha} or C3αC_{3}^{\alpha}. Geometrically, it means that we must avoid the center-focus case.

Thus we have constructed an homeomorphism HH such that H(D1)=D2H(D_{1})=D_{2}, HH maps phase curves of X~\widetilde{X} into phase curves of Y~\widetilde{Y} and HH sends Δ~δ\widetilde{\Delta}_{\delta} into Δ~γ\widetilde{\Delta}_{\gamma}. Then HH is the orbital C0C^{0}-equivalence desired. ∎

Combining the Propositions 6 and 13 we obtain the Theorem A, which gives a condition to assure the existence of an (orientation preserving) orbital C0C^{0}-equivalence between two constrained differential systems, and such condition is based on the resolution of singularities. Since the classification of phase portraits depends on the notion of equivalence adopted, this result plays an important role in what follows.

4. Constrained systems with singular impasse sets and the proof of Theorem B

The main goal of this section is to give a classification of phase portraits of planar constrained systems near singular points of the impasse curve. We recall from the introduction that the set 𝒜\mathcal{A} is the set of constrained systems (X,δ)(X,\delta) such that:

  1. (1)

    The adjoint vector field is constant, that is, it satisfies X(x,y)=(a,b)X(x,y)=(a,b), a,ba,b\in\mathbb{R} and a2+b20a^{2}+b^{2}\neq 0.

  2. (2)

    The function δ\delta is one of the curves given by Arnold’s ADE-classification (see Table 1).

Type Normal Form Codimension
AkA_{k}, k1k\geq 1 x2±yk+1x^{2}\pm y^{k+1} k1k-1
DkD_{k}, k4k\geq 4 x2y±yk1x^{2}y\pm y^{k-1} k1k-1
E6E_{6} x3±y4x^{3}\pm y^{4} 55
E7E_{7} y3yx3y^{3}-yx^{3} 66
E8E_{8} x3+y5x^{3}+y^{5} 77
Table 1. ADE-type singularities [2].

We chose a coordinate system such that δ\delta is one of the functions given by Table 1. Since the adjoint vector field and δ\delta are analytic, and assuming that X(0)0X(0)\neq 0, that is, the coefficients a1,0a_{-1,0} and b0,1b_{0,-1} in the expansion

(6) {δ(x,y)x˙=a1,0+a0,0x+a1,1y+,δ(x,y)y˙=b0,1+b0,0y+b1,1x+,\left\{\begin{array}[]{rcl}\delta(x,y)\dot{x}&=&a_{-1,0}+a_{0,0}x+a_{-1,1}y+...,\\ \delta(x,y)\dot{y}&=&b_{0,-1}+b_{0,0}y+b_{1,-1}x+...,\end{array}\right.

satisfy a1,02+b0,120a_{-1,0}^{2}+b_{0,-1}^{2}\neq 0, the Theorem B gives conditions to assure when the System (6) is orientation preserving C0C^{0}-orbitally equivalent to the system

(7) {δ(x,y)x˙=a1,0,δ(x,y)y˙=b0,1,\left\{\begin{array}[]{rcl}\delta(x,y)\dot{x}&=&a_{-1,0},\\ \delta(x,y)\dot{y}&=&b_{0,-1},\end{array}\right.

which belongs to 𝒜\mathcal{A}.

The proof of Theorem B is given by straightforward computations. Indeed, one must compare the process of resolution of singularities of both (6) and (7) in order to establish conditions to the coefficients of the Taylor expansion in such a way that, after a suitable finite sequence of weighted blow ups, the systems (6) and (7) have the same elementary singularity scheme. System (6) must satisfy one of the conditions presented in Table 2, where n0n_{0} is the first positive integer such that a1,n00a_{-1,n_{0}}\neq 0. Geometrically, the conditions below mean that we must avoid tangencies between the adjoint vector field of (6) and the components of the impasse set.

Curve δ\delta Conditions
AkA_{k} Case 11:  k>1k>1 and a1,00a_{-1,0}\neq 0
Case 22:  k=1k=1 and 0a1,0±b0,10\neq a_{-1,0}\neq\pm b_{0,-1}
Case 33:  a1,0=0a_{-1,0}=0 and k1<2n0k-1<2n_{0}
DkD_{k} Case 44:  a1,00a_{-1,0}\neq 0 and bm0,1=0b_{m_{0},-1}=0 for all integer m00m_{0}\geq 0
Case 55:  k=4k=4 and 0a1,0±b0,10\neq a_{-1,0}\neq\pm b_{0,-1}
Case 66:  a1,0=0a_{-1,0}=0 and k4<2n0k-4<2n_{0}
E7E_{7} Case 77:  b0,10b_{0,-1}\neq 0
Case 88:  a1,00a_{-1,0}\neq 0 and bm0,1=0b_{m_{0},-1}=0 for all integer m00m_{0}\geq 0
E6E_{6} or E8E_{8}
without requiring further assumptions in the
Taylor expansion of the adjoint vector field.
Table 2. The constrained system (6) must satisfy one of the conditions of this table in order to be equivalent to a system in 𝒜\mathcal{A}. The number n0n_{0} is the first positive integer such that a1,n00a_{-1,n_{0}}\neq 0.

In what follows we briefly describe the phase portraits of (7) near the exceptional divisor for each curve δ\delta of Table 1. We start this section reviewing the construction of the Newton polygon and how we can define such mathematical object for planar constrained systems.

4.1. The Newton polygon

The construction of the so called Newton polygon associated to a planar analytic vector field is well-known in the literature, and we refer to [1, 9, 13] for details. Afterwards we will see how to define such polygon for constrained systems.

Let XX be an analytic vector field. Just as in [12], we write XX in the so called logarithmic basis. More precisely, consider

X(x,y)=P(x,y)x+Q(x,y)y,X(x,y)=P(x,y)\displaystyle\frac{\partial}{\partial x}+Q(x,y)\displaystyle\frac{\partial}{\partial y},

where

P(x,y)=(am,nxmyn)x,Q(x,y)=(bm,nxmyn)y,P(x,y)=\Big{(}\sum a_{m,n}x^{m}y^{n}\Big{)}x,\ Q(x,y)=\Big{(}\sum b_{m,n}x^{m}y^{n}\Big{)}y,

with am,n,bm,na_{m,n},b_{m,n}\in\mathbb{R} and m,nm,n\in\mathbb{Z} satisfying:

  1. (1)

    For m<1m<-1 or n1n\leq-1, am,n=0a_{m,n}=0;

  2. (2)

    For m1m\leq-1 or n<1n<-1, bm,n=0b_{m,n}=0.

Let ω=(ω1,ω2)\omega=(\omega_{1},\omega_{2}) be a vector of positive integers. One can write the planar vector field XX as

(8) X(x,y)=d=1Xd(ω1,ω2)(x,y),X(x,y)=\displaystyle\sum_{d=-1}^{\infty}X_{d}^{(\omega_{1},\omega_{2})}(x,y),

where

(9) Xd(ω1,ω2)(x,y)=ω1r+ω2s=dxrys(ar,sxx+br,syy).X_{d}^{(\omega_{1},\omega_{2})}(x,y)=\displaystyle\sum_{\omega_{1}r+\omega_{2}s=d}x^{r}y^{s}\Big{(}a_{r,s}x\displaystyle\frac{\partial}{\partial x}+b_{r,s}y\displaystyle\frac{\partial}{\partial y}\Big{)}.

In other words, the vector field XX can be written as a sum of quasi-homogeneous components. We say that (8) is a (ω1,ω2)(\omega_{1},\omega_{2})-graduation of XX. Each Xd(ω1,ω2)X_{d}^{(\omega_{1},\omega_{2})} is called dd-level of the (ω1,ω2)(\omega_{1},\omega_{2})-graduation.

Given a (ω1,ω2)(\omega_{1},\omega_{2})-graduation of XX, we associate the monomials ar,sxrysa_{r,s}x^{r}y^{s} and br,sxrysb_{r,s}x^{r}y^{s} with nonzero coefficients to a point (r,s)(r,s) in the plane of powers, where each point (r,s)(r,s) is contained in a line of the form {ω1r+ω2s=d}\{\omega_{1}r+\omega_{2}s=d\}.

Definition 14.

The support 𝒬\mathcal{Q} of XX is the set

𝒬={(r,s)2:ar,s2+br,s20}.\mathcal{Q}=\{(r,s)\in\mathbb{Z}^{2}:a_{r,s}^{2}+b_{r,s}^{2}\neq 0\}.
Definition 15.

The Newton polygon 𝒫\mathcal{P} associated to the analytic vector field XX is the convex envelope of the set 𝒬++2\mathcal{Q}+\mathbb{R}^{2}_{+}.

The boundary 𝒫\partial\mathcal{P} of the Newton polygon 𝒫\mathcal{P} is the union of a finite number of segments. We enumerate them from the left to the right: γ0,γ1,,γn+1\gamma_{0},\gamma_{1},...,\gamma_{n+1}. Observe that γ0\gamma_{0} is vertical and γn+1\gamma_{n+1} is horizontal. Analogously, the non-regular points of 𝒫\partial\mathcal{P} are enumerated from the left to the right: v0,,vnv_{0},...,v_{n}.

Definition 16.

We say that v0,,vnv_{0},...,v_{n} are the vertices of the Newton polygon. The vertex v0=(r0,s0)v_{0}=(r_{0},s_{0}) is called main vertex and the segment γ1\gamma_{1} will be called main segment. The number s0s_{0} is the height of the Newton polygon. See Figure 22.

Refer to caption
Figure 22. Newton polygon of a vector field. The main vertex and main segment are highlighted in blue.

The main segment γ1\gamma_{1} is contained in an affine line of the form {rω1+sω2=R}\{r\omega_{1}+s\omega_{2}=R\}. Observe that the vector ω=(ω1,ω2)\omega=(\omega_{1},\omega_{2}) is normal with respect to {rω1+sω2=R}\{r\omega_{1}+s\omega_{2}=R\}. In the resolution of singularities of analytic vector fields, one chooses the vector ω\omega as the weight vector (see [1, 13]). We remark that the Newton polygon strongly depends on the coordinate system adopted.

4.2. The auxiliary vector field

The next goal is to relate a given constrained system to a Newton polygon. In order to achieve this objective, firstly we will define the so called auxiliary vector field. Recall that a diagonalized constrained system is written in the form

(10) δ(x,y)x˙=P(x,y),δ(x,y)y˙=Q(x,y),\delta(x,y)\dot{x}=P(x,y),\ \delta(x,y)\dot{y}=Q(x,y),

with δ(x,y)\delta(x,y) being an irreducible real-analytic function that can be written as

δ(x,y)=ck,lxkyl,\delta(x,y)=\sum c_{k,l}x^{k}y^{l},

where ck,lc_{k,l}\in\mathbb{Z} is such that ck,l=0c_{k,l}=0 when k<0k<0 or l<0l<0.

Definition 17.

The auxiliary vector field XAX_{A} associated to the constrained differential equation (10) is the vector field

(11) XA(x,y)=δ(x,y)(P(x,y)x+Q(x,y)y)=(ck,lxkyl)(xmyn(am,nxx+bm,nyy)),\scriptstyle X_{A}(x,y)=\delta(x,y)\Bigg{(}P(x,y)\frac{\partial}{\partial x}+Q(x,y)\frac{\partial}{\partial y}\Bigg{)}=\Bigg{(}\sum c_{k,l}x^{k}y^{l}\Bigg{)}\Bigg{(}\sum x^{m}y^{n}\big{(}a_{m,n}x\frac{\partial}{\partial x}+b_{m,n}y\frac{\partial}{\partial y}\big{)}\Bigg{)},

which is real analytic.

It is easy to sketch the phase portrait of the auxiliary vector field. Outside the impasse set, the auxiliary vector field (11) is obtained by multiplying the constrained differential system (10) by the positive function δ2\delta^{2}. On the other hand, the impasse set Δ\Delta is a curve of equilibrium points for (11).

Since the auxiliary vector field is analytic, all previous definitions and remarks concerning the Newton polygon remain true for (11). However, observe that the points on the plane of powers are of the form (k+m,l+n)(k+m,l+n), and therefore the support 𝒬\mathcal{Q} takes form

𝒬={(k+m,l+n)2:ck,l(am,n2+bm,n2)0}.\mathcal{Q}=\{(k+m,l+n)\in\mathbb{Z}^{2}:c_{k,l}(a_{m,n}^{2}+b_{m,n}^{2})\neq 0\}.

In other words, the support of the auxiliary vector field is obtained by the Minkowski Sum [17] between the points of the support of the adjoint vector field XX and the points of the support of δ\delta. Moreover, the levels of a (ω1,ω2)(\omega_{1},\omega_{2})-graduation are written in the form

XA,d(ω1,ω2)=(ω1k+ω2l=d1ck,lxkyl)(ω1m+ω2n=d2xmyn(am,nxx+bm,nyy)),X_{A,d}^{(\omega_{1},\omega_{2})}=\Bigg{(}\displaystyle\sum_{\omega_{1}k+\omega_{2}l=d_{1}}c_{k,l}x^{k}y^{l}\Bigg{)}\Bigg{(}\displaystyle\sum_{\omega_{1}m+\omega_{2}n=d_{2}}x^{m}y^{n}\Big{(}a_{m,n}x\displaystyle\frac{\partial}{\partial x}+b_{m,n}y\displaystyle\frac{\partial}{\partial y}\Big{)}\Bigg{)},

where d1+d2=dd_{1}+d_{2}=d. Finally, we remark that the Newton polygon of the adjoint vector field and the auxiliary vector field are not necessarily the same.

The process of resolution of singularities of real analytic 2-dimensional constrained differential systems with weighted blow-ups was discussed in details in [14], where one can also find examples. Such process can be summarized as follows. Given a constrained system, we write the system in its diagonalized form and then we define the auxiliary vector field XAX_{A}. The auxiliary vector field is an analytic vector field and it allows us to apply well known techniques of the literature, using the Newton polygon 𝒫XA\mathcal{P}_{X_{A}} to chose the weight vector of the blow up.

Now we are able to study constrained systems in the set 𝒜\mathcal{A}.

4.3. Curve AkA_{k}, k1k\geq 1

Firstly assume the conditions in the case 1. After a weighted blow up in the xx direction, the origin 0𝔼x0\in\mathbb{E}_{x} is a hyperbolic saddle and the impasse curve satisfies 1±y~k+1=01\pm\tilde{y}^{k+1}=0. In the yy direction there are no equilibrium points in 𝔼y\mathbb{E}_{y} and the impasse curve satisfies the equation x~2±1=0\tilde{x}^{2}\pm 1=0. This case is equivalente to the case 2. See Figures 23 and 24.

Now, consider the case 3. After a weighted blow-up in the xx direction, there are no equilibrium points in 𝔼x\mathbb{E}_{x}, and in the yy direction the origin 0𝔼y0\in\mathbb{E}_{y} is a hyperbolic saddle. The impasse set behaves just as the previous cases.

Geometrically, n0n_{0} measures the tangency order between the adjoint vector field XX and the coordinate axis {x=0}\{x=0\}, and kk measures the tangency order between δ\delta and {x=0}\{x=0\}. Therefore, the assumption k1<2n0k-1<2n_{0} sets a relation between such tangency orders. It is important to remark that, when k12n0k-1\geq 2n_{0}, the main segment of the Newton polygon of the auxiliary vector field of (6) has the point (1,n0)(1,n_{0}), and such point does not appear in the Newton polygon of the auxiliary vector field of (7) See figures 25 and 26.

Refer to caption
Figure 23. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta AkA_{k}-type, kk odd.
Refer to caption
Figure 24. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta AkA_{k}-type, kk even.
Refer to caption
Figure 25. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta AkA_{k}-type, kk odd.
Refer to caption
Figure 26. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta AkA_{k}-type, kk even.

4.4. Curve DkD_{k}, k4k\geq 4

Assume the hypotheses in case 4. After a weighted blow up in the xx direction, the origin 0𝔼x0\in\mathbb{E}_{x} is a saddle point of the adjoint vector field XX and there is a separatrix that coincides with a component of the impasse set {y~±y~k1=0}\{\tilde{y}\pm\tilde{y}^{k-1}=0\}. In the yy direction, there are no equilibrium points in 𝔼y\mathbb{E}_{y} and the impasse set is give by {x~2±1=0}\{\tilde{x}^{2}\pm 1=0\}.

We emphasize that the condition bm0,1=0b_{m_{0},-1}=0 for all integer m00m_{0}\geq 0 avoids tangency points between the adjoint vector field and the component {y=0}\{y=0\} of the impasse set, in which it would leads us to a different resolution of singularities from system (7). See Figures 27 and 28.

Now assume the hypotheses of the case 5. In the yy direction, there are no equilibrium points of XX in 𝔼y\mathbb{E}_{y}. On the other hand, in the xx direction the origin 0𝔼x0\in\mathbb{E}_{x} is a hyperbolic saddle of the adjoint vector field and an impasse point at the same time. Therefore, we must blow-up the origin once again. See Figures 29 and 30

Finally, assume the hypotheses of the case 6. In the xx direction the impasse curve is given by y~±y~k1=0\tilde{y}\pm\tilde{y}^{k-1}=0 and there are no equilibrium points of the adjoint vector field in 𝔼x\mathbb{E}_{x}. On the other hand, after a weighted blow up in the yy direction the impasse curve is given by x~2±1=0\tilde{x}^{2}\pm 1=0 and the origin 0𝔼y0\in\mathbb{E}_{y} is a hyperbolic saddle.

The number n0n_{0} is related to the tangency order between the adjoint vector field XX and the axis {x=0}\{x=0\}, and kk is related to the tangency order between δ\delta and the axis {x=0}\{x=0\}. Thus the assumption k4<2n0k-4<2n_{0} sets a relation between such tangency orders. Observe that in the case k42n0k-4\geq 2n_{0}, the main segment of the Newton polygon of the auxiliary vector field of (6) has the point (1,n0+1)(1,n_{0}+1), and such point obviously does not appear in the Newton polygon of the auxiliary vector field of (7). See Figures 31 and 32.

Refer to caption
Figure 27. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta DkD_{k}-type, kk even.
Refer to caption
Figure 28. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta DkD_{k}-type, kk odd.
Refer to caption
Figure 29. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta DkD_{k}-type, kk even.
Refer to caption
Figure 30. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta DkD_{k}-type, kk odd.
Refer to caption
Figure 31. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta DkD_{k}-type, kk even.
Refer to caption
Figure 32. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta DkD_{k}-type, kk odd.

4.5. Curve E6E_{6}

Suppose without loss of generality that δ=x3y4\delta=x^{3}-y^{4}. In the positive xx direction the impasse curve intersects the exceptional divisor at the points (0,±1)(0,\pm 1). For a1,00a_{-1,0}\neq 0 the origin 0𝔼x0\in\mathbb{E}_{x} is a hyperbolic saddle and if a1,0=0a_{-1,0}=0 we do not have equilibrium points of the adjoint vector field on the divisor. In the positive yy direction, the impasse curve intersects the exceptional divisor at (1,0)(1,0). If a1,00a_{-1,0}\neq 0, there are no equilibrium points of the adjoint vector field in 𝔼y\mathbb{E}_{y} and if a1,0=0a_{-1,0}=0 the origin is a hyperbolic saddle. See Figures 33 and 34.

Refer to caption
Figure 33. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta E6E_{6}-type.
Refer to caption
Figure 34. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta E6E_{6}-type.

4.6. Curve E7E_{7}

Suppose that a1,00a_{-1,0}\neq 0 and bm0,1=0b_{m_{0},-1}=0, for all integer m00m_{0}\geq 0. In the positive xx direction, the impasse curve intersects the 𝔼x\mathbb{E}_{x} at the points (0,1)(0,-1), (0,0)(0,0) and (0,1)(0,1). Moreover, the origin is a saddle point of the adjoint vector field and there is a separatrix that coincides with a component of the impasse set. In the positive yy direction, there are no equilibrium points of XX in 𝔼y\mathbb{E}_{y} and the impasse set intersects the exceptional divisor at (1,0)(1,0).

Once again we remark that the condition bm0,1=0b_{m_{0},-1}=0 avoids tangency points between the adjoint vector field XX and the component {y=0}\{y=0\} of the impasse set. Such tangency points would lead us to a different resolution of singularities.

If we consider b0,10b_{0,-1}\neq 0, in the positive xx direction there are no equilibrium points of XX in 𝔼x\mathbb{E}_{x}, and in the positive yy direction the origin 0𝔼y0\in\mathbb{E}_{y} is a saddle points of XX. The impasse curve behaves just as in the previous case.

See Figures 35 and 36.

Refer to caption
Figure 35. Dynamics when bm0,1=0b_{m_{0},-1}=0 for all m00m_{0}\geq 0, where δ\delta is E7E_{7}-type.
Refer to caption
Figure 36. Dynamics when b0,10b_{0,-1}\neq 0, for δ\delta E7E_{7}-type.

4.7. Curve E8E_{8}

In the positive xx direction, the impasse curve intersects the exceptional divisor at (0,1)(0,-1). If a1,00a_{-1,0}\neq 0, then the origin 0𝔼x0\in\mathbb{E}_{x} is a hyperbolic saddle of the adjoint vector field and if a1,0=0a_{-1,0}=0 there are no equilibrium equilibrium points of the adjoint vector field on the divisor. Finally, in the positive yy direction the impasse curve intersects the exceptional divisor 𝔼y\mathbb{E}_{y} at (1,0)(-1,0). If a1,00a_{-1,0}\neq 0, there are no equilibrium points of the adjoint vector field in 𝔼y\mathbb{E}_{y} and if a1,0=0a_{-1,0}=0 the origin is a hyperbolic saddle of the adjoint vector field. See Figures 37 and 38.

Refer to caption
Figure 37. Dynamics when a1,00a_{-1,0}\neq 0, for δ\delta E8E_{8}-type.
Refer to caption
Figure 38. Dynamics when a1,0=0a_{-1,0}=0, for δ\delta E8E_{8}-type.

5. Topological determination of a constrained system with smooth impasse curve via Newton polygon

Firstly, let us recall a classical result of the literature. Let XX be a planar analytic vector field defined near the origin such that X(0)=0X(0)=0, and denote its Newton polygon by 𝒫X\mathcal{P}_{X} (see Definition 15 below). In [5] the authors proved that, under some non degeneracy conditions, the terms of XX associated to the the boundary 𝒫X\partial\mathcal{P}_{X} of the Newton polygon determines the phase portrait of XX near the equilibrium point.

More precisely, given an analytic vector field XX, the terms of XX associated to the points in 𝒫X\partial\mathcal{P}_{X} define the so called principal part of XX, which is denoted by XΩX_{\Omega}. The Theorem A of [5] assures that XX and XΩX_{\Omega} are topologically equivalent.

Here we present a version of such result for analytic planar constrained differential systems near a non-singular point of the impasse curve. Our strategy is to work in a convenient coordinate system, in such a way that the classical result presented in [5] can be applied during the resolution of singularities.

During the resolution of singularities, one must apply a finite number of operations (weighted blow-ups) in order to obtain a “simpler” constrained system. This notion of “simple” comes from the notion of elementary constrained system (see Definition 3). In what follows we recall how to identify elementary points of the constrained system by means of the Newton polygon of the auxiliary vector field. We refer [14] for details.

Definition 18.

A planar constrained differential system is Newton elementary at pp if the Newton polygon 𝒫\mathcal{P} associated to the auxiliary vector field XAX_{A} satisfies one of the following:

  • The main vertex of 𝒫\mathcal{P} is (0,0)(0,0), (0,1)(0,-1) or (1,0)(-1,0) (that is, if the height is less or equal to zero);

  • The main segment γ1\gamma_{1} is horizontal.

Theorem 19.

(See [14]) A planar constrained system is elementary at pp if, and only if, it is Newton elementary at pp.

5.1. Favorable coordinates

Since our study is local, for simplicity sake we denote a planar real analytic constrained differential system defined near the origin as (X,δ)(X,\delta), where X=(P(x,y),Q(x,y))X=\Big{(}P(x,y),Q(x,y)\Big{)} is the adjoint vector field and δ\delta is a real irreducible analytic function. As usual, the auxiliary vector field will be denoted by XAX_{A}. The Newton polygons of the adjoint vector field and the auxiliary vector field will be denote by 𝒫X\mathcal{P}_{X} and 𝒫XA\mathcal{P}_{X_{A}}, respectively.

In this section, constrained differential systems with regular impasse set are considered. By “regular impasse set” we mean that the gradient vector δ\nabla\delta is nonzero. The main goal is to extend the Theorem A of [5] to the context of constrained differential systems, that is, the objective is to show that the terms associated to the boundary 𝒫XA\partial\mathcal{P}_{X_{A}} of the Newton Polygon of the auxiliary vector field determine (under topological equivalence) the phase portrait near the singularity of the constrained system. In order to achieve this objetive we will introduce a convenient coordinate system.

Definition 20.

The Newton polygon 𝒫\mathcal{P} associated to a vector field is controllable if the main vertex v0v_{0} is contained in {0}×\{0\}\times\mathbb{Z} or {1}×\{-1\}\times\mathbb{Z}.

By Lemma 9 in [14], we can always assume without loss of generality that the Newton polygon of an analytic vector field is controllable.

Lemma 21.

Let (X,δ)(X,\delta) be a constrained system such that the origin is an isolated equilibrium point and the impasse set is regular. Then there is a coordinate system such that:

  • The impasse set is given by Δ={y=0}\Delta=\{y=0\};

  • The Newton polygons 𝒫XA\mathcal{P}_{X_{A}} and 𝒫X\mathcal{P}_{X} of the auxiliary vector field XAX_{A} and the adjoint vector field XX, respectively, are controllable;

  • The polygon 𝒫XA\mathcal{P}_{X_{A}} is obtained by increasing one unity on the second coordinate of each point of the polygon 𝒫X\mathcal{P}_{X}.

Proof.

The first item is true due to the Implicit Function Theorem. For the second item, we can apply the change of coordinates

x=x¯+λy¯;y=y¯,x=\bar{x}+\lambda\bar{y};\ y=\bar{y},

with λ\lambda\in\mathbb{R} (see Lemma 9, [14]). Observe that after this change of coordinates, the impasse set is still of the form Δ={y=0}\Delta=\{y=0\}. Recall that the Newton polygon 𝒫XA\mathcal{P}_{X_{A}} of the auxiliary vector field XAX_{A} is obtained by the Minkowski sum between the Newton polygons of δ\delta and XX. Since in this coordinate system the function δ\delta is given by δ=y\delta=y, it follows that the Newton polygon 𝒫XA\mathcal{P}_{X_{A}} of the auxiliary vector field XAX_{A} is obtained by increasing one unity the second coordinate of each point of the Newton polygon 𝒫X\mathcal{P}_{X} of XX. See Figure 39. ∎

Refer to caption
Figure 39. Newton polygons 𝒫X\mathcal{P}_{X} (left) and 𝒫XA\mathcal{P}_{X_{A}} (right).
Definition 22.

The coordinate system given by Lemma 21 is called favorable.

Due to Lemma 21, without loss of generality we can always take favorable coordinates for (X,δ)(X,\delta). Furthermore, by Lemma 21, it follows that 𝒫XA\mathcal{P}_{X_{A}} and 𝒫X\mathcal{P}_{X} have the same number of segments. In addition, given a segment γi𝒫XA\gamma_{i}\subset\partial\mathcal{P}_{X_{A}}, the respective segment γi𝒫X\gamma_{i}^{\prime}\subset\partial\mathcal{P}_{X} has the same slope as γi\gamma_{i}. This implies that, if the main segment γ1𝒫XA\gamma_{1}\subset\partial\mathcal{P}_{X_{A}} is contained in a line of the form {rω1+sω2=R}\{r\omega_{1}+s\omega_{2}=R\}, then the segment γ1𝒫X\gamma_{1}^{\prime}\subset\partial\mathcal{P}_{X} is contained in a line of the form {rω1+sω2=S}\{r\omega_{1}+s\omega_{2}=S\}. Therefore, at each step of the resolution of singularities, the weights of the blow ups that we use in the desingularization of the constrained system and the vector field XX are the same.

Now we are ready to define the principal part of a planar constrained system with regular impasse set.

Definition 23.

The principal part of a 2-dimensional constrained system with regular impasse set is a pair (XΩ,δ)(X_{\Omega},\delta) in favorable coordinates, where XΩX_{\Omega} is the principal part of the adjoint vector field XX in the sense of [5].

We recall from [5] that the principal part of a real analytic vector field with respect to a fixed system of coordinates is given by

XΩ(x,y)=j=1nXj(x,y),X_{\Omega}(x,y)=\displaystyle\sum_{j=1}^{n}X_{j}(x,y),

where XjX_{j} is the vector field defined by the terms of XX that belong to the segment γj𝒫X\gamma_{j}\subset\partial\mathcal{P}_{X}, where γj\gamma_{j} is neither vertical nor horizontal.

Example. Consider the diagonalized constrained differential system

(12) yx˙=y3+x2y,yy˙=xy+x4;y\dot{x}=y^{3}+x^{2}y,\ y\dot{y}=xy+x^{4};

which is already written in favorable coordinates and whose auxiliary vector field is given by

(13) XA(x,y)=y(y3+x2y)x+y(xy+x4)y.X_{A}(x,y)=y\big{(}y^{3}+x^{2}y\big{)}\displaystyle\frac{\partial}{\partial x}+y\big{(}xy+x^{4}\big{)}\displaystyle\frac{\partial}{\partial y}.

The support of (13) is the set 𝒬={(1,4);(1,2);(4,0);(1,1)}\mathcal{Q}=\{(-1,4);(1,2);(4,0);(1,1)\}, and the terms associated to the boundary 𝒫XA\partial\mathcal{P}_{X_{A}} of the Newton polygon of XAX_{A} are y(y3)xy\big{(}y^{3}\big{)}\displaystyle\frac{\partial}{\partial x}, y(xy)yy\big{(}xy\big{)}\displaystyle\frac{\partial}{\partial y} and y(x4)yy\big{(}x^{4}\big{)}\displaystyle\frac{\partial}{\partial y}. Thus the principal part of (12) is

(14) yx˙=y3,yy˙=xy+x4.y\dot{x}=y^{3},\ y\dot{y}=xy+x^{4}.

5.2. Non-degeneracy conditions

In this subsection we discuss some non-degeneracy conditions required for the pair (X,δ)(X,\delta) in order to show the existence of an orbital C0C^{0}-equivalence between (X,δ)(X,\delta) and (XΩ,δ)(X_{\Omega},\delta). As usual, we consider (X,δ)(X,\delta) in favorable coordinates.

The first assumption is that the origin is an isolated equilibrium point of the adjoint vector field XX. This implies that 𝒫X\partial\mathcal{P}_{X} intersects the coordinate axes of the rsrs-plane. However, we remark that, in general, the Newton polygon 𝒫XA\partial\mathcal{P}_{X_{A}} of the auxiliary vector field only intersects the ss-axis.

The next non-degeneracy condition was introduced in [5].

Definition 24.

Let XX be an analytic vector field such that pp is an isolated equilibrium point and let XΩX_{\Omega} be its principal part. We say that XX is Newton non-degenerated at pp if any quasi homogeneous component XjX_{j} of XΩX_{\Omega} associated to a side γj𝒫X\gamma_{j}\subset\partial\mathcal{P}_{X} does not have singularities in (\{0})2(\mathbb{R}\backslash\{0\})^{2}, that is, if each XjX_{j} does not have singularities outside the coordinate axes.

This non-degeneracy condition implies that, during the resolution of singularities of analytic vector fields, the only point in the exceptional divisor that has positive height is the origin in the xx direction. In other words, all the equilibrium points in (𝔼x\{(0,0)})𝔼y\Big{(}\mathbb{E}_{x}\backslash\{(0,0)\}\Big{)}\cup\mathbb{E}_{y} will always be elementary (see Proposition B, [13]).

Moreover, such condition depends on the coordinate system adopted, that is, an analytic vector field can be Newton non-degenerated in a fixed coordinate system, but not be in other coordinate system (see [13]). However, being Newton non-degenerated is a generic property in the set of all analytic vector fields (see Proposition 6, [5]), and therefore this is a generic condition in the set Γ\Gamma of all planar analytic constrained systems defined near the origin in favorable coordinates.

We end this subsection with the following remark. Suppose that 𝒫X\mathcal{P}_{X} contains a point of the form (M,1)(M,-1). In other words, suppose that the component QQ of X=(P,Q)X=(P,Q) has a term of the form bM,1xMb_{M,-1}x^{M}, with bM,10b_{M,-1}\neq 0. Geometrically, this condition means that any phase curve of XX does not coincide with the impasse curve Δ={y=0}\Delta=\{y=0\}. By Lemma 21 it follows that 𝒫XA\mathcal{P}_{X_{A}} has a point of the form (M,0)(M,0). On the other hand, if 𝒫X\mathcal{P}_{X} does not contain a point of the form (M,1)(M,-1), then the impasse set Δ\Delta coincides with a phase curve. By Lemma 21 it follows that 𝒫XA\mathcal{P}_{X_{A}} does not intersect the rr-axis of the rsrs-plane.

5.3. Proof of Theorem C: The topological equivalence between a 2-dimensional constrained system and its principal part

Firstly, we consider the case where the boundary of the Newton polygon has just one side that is neither horizontal nor vertical, and afterwards we generalize such result. From now on, the pair (X,δ)(X,\delta) satisfies the hypotheses discussed in the previous subsection.

Lemma 25.

Let (X,δ)(X,\delta) be a 2-dimensional constrained system and take favorable coordinates. Suppose that

  1. (1)

    The vector field XX is Newton non-degenerated at the origin;

  2. (2)

    XΩX_{\Omega} = X1X_{1}, that is, the boundary 𝒫X\partial\mathcal{P}_{X} of the polygon of XX has just one segment that is neither vertical nor horizontal.

Then (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same elementary singularity scheme.

Proof.

We will study the case where the main vertex is of the form (0,N)(0,N). The case where the main vertex is (1,N)(-1,N) is analogous. The proof is based in directional weighted blow ups and we focus only in the positive xx and yy directions, provided that the computations are similar in the negative directions.

Since we are adopting favorable coordinates, the impasse curve {y=0}\{y=0\} only appears in the xx direction. This means that, in the yy direction, there are no impasse points in the exceptional divisor after a weighted blow-up, and therefore we are in the classical case of analytic vector fields. It follows that in the yy direction both pairs (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same equilibrium points, and the arrangement of such points is the same for both pairs. We emphasize that such equilibrium points in 𝔼y\mathbb{E}_{y} are semi-hyperbolic (and therefore elementary), because the constrained system is Newton non-degenerated.

After a blow up in the xx direction, the first graduation of the auxiliary vector field takes the form

X~A=y~(n=1Ny~n(am,nω1x~x~+(bm,nω2ω1am,n)y~y~)).\widetilde{X}_{A}=\tilde{y}\Bigg{(}\displaystyle\sum_{n=-1}^{N}\tilde{y}^{n}\Big{(}\frac{a_{m,n}}{\omega_{1}}\tilde{x}\displaystyle\frac{\partial}{\partial\tilde{x}}+(b_{m,n}-\frac{\omega_{2}}{\omega_{1}}a_{m,n})\tilde{y}\displaystyle\frac{\partial}{\partial\tilde{y}}\Big{)}\Bigg{)}.

Observe that the impasse curve is still of the form {y~=0}\{\tilde{y}=0\} and the equilibrium points of the strict transformed X~Ω\widetilde{X}_{\Omega} and X~\widetilde{X} are the same and they are semi-hyperbolic. We remark that this happens independently if the impasse set coincides or does not coincide with a separatrix of an equilibrium point. The difference is that, in the first case, the main segment of the polygon after the blow-up is horizontal, and in the second case, the polygon after the blow-up does not have positive height.

It follows that the pairs (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same elementary singularity scheme. ∎

It is interesting to remark that in Lemma 25, the process of resolution of singularities essentially desingularizes the vector field, since the impasse set is already “simple”. This idea will be important in the proof of the next proposition.

Proposition 26.

Let (X,δ)(X,\delta) be a 2-dimensional constrained system written in favorable coordinates. If the origin is Newton non-degenerated, then the pairs (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same elementary singularity scheme.

Proof.

Once again we consider weighted directional blow ups and study the case where the main vertex is of the form (0,N)(0,N). The case where the main vertex is (1,N)(-1,N) is analogous. We will compare the process of resolution of singularities of (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta). Since the segments of the Newton polygons 𝒫XA\mathcal{P}_{X_{A}} and 𝒫XΩ,A\mathcal{P}_{X_{\Omega,A}} have the same slope, at each step of the process we apply blow-ups with same weights for both constrained systems. The case were the boundary of the Newton polygon has just one segment that is neither vertical nor horizontal was treated in Lemma 25, so we will give a proof for the general case.

Just as in Lemma 25, the impasse set {y=0}\{y=0\} only appears in the xx direction. This means that in all points of the divisor 𝔼y\mathbb{E}_{y}, both pairs (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same equilibrium points, all of them are semi-hyperbolic and the arrangement of such equilibrium points are the same for both (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta), provided that in the positive yy direction we are in the classical case and the constrained system is Newton non-degenerated. Therefore it is sufficient to look at the origin of the xx direction. See Figure 40.

Refer to caption
Figure 40. Blow-up in the yy-direction.

Applying the first blow-up with weight ω=(ω1,ω2)\omega=(\omega_{1},\omega_{2}) in the xx direction, we obtain the strict transformed (XΩ1,δ1)(X^{1}_{\Omega},\delta_{1}) and (X1,δ1)(X^{1},\delta_{1}). It is straightforward to see that δ\delta is still of the form δ=y\delta=y, and therefore the origin is the only impasse point in the divisor. Concerning the adjoint vector field, with similar computations as in Lemma 25 on the divisor 𝔼x\mathbb{E}_{x} we have the same equilibrium points for both XX and XΩX_{\Omega}. We emphasize that all the points in 𝔼x\{(0,0)}\mathbb{E}_{x}\backslash\{(0,0)\} are elementary. Moreover, the arrangement of such equilibrium points in the divisor are the same. Observe that the origin is an equilibrium point and an impasse point at the same time.

Refer to caption
Figure 41. Blow-up in the xx-direction.

Due to Lemma 21, the Newton polygons of (XΩ1,δ1)(X^{1}_{\Omega},\delta_{1}) and (X1,δ1)(X^{1},\delta_{1}) are obtained by increasing one unity to the second coordinate of the points of 𝒫X1\mathcal{P}_{X^{1}}. We must go on with the resolution process, because the origin is not an elementary point. It is clear that, at each step, the boundaries of the Newton polygons are the same and hence we apply blowing-ups with same weights to both (XΩi,δi)(X^{i}_{\Omega},\delta_{i}) and (Xi,δi)(X^{i},\delta_{i}).

Essentially, we are only desingularizing equilibrium points of XX and XΩX_{\Omega} and preserving the impasse curve.

In the end of the process, we have three cases to consider at the impasse point (0,0)𝔼x(0,0)\in\mathbb{E}_{x}. The first two cases concern the case where the impasse set Δ\Delta does no coincide with a separatrix of the equilibrium point, that is, the Newton polygon 𝒫X\mathcal{P}_{X} has a point of the form (M,1)(M,-1). The third case concerns the case where 𝒫X\mathcal{P}_{X} does not have a point of the form (M,1)(M,-1) (which means that the impasse set coincides with a separatrix of the equilibrium point).

Case 1: The Newton polygons 𝒫XΩ,A\mathcal{P}_{X_{\Omega,A}} and 𝒫XA\mathcal{P}_{X_{A}} do not have a vertex of the form (m,1)(m,1). This means that the Newton polygon 𝒫X\mathcal{P}_{X} of XX does not have a vertex of the form (m,0)(m,0). By the coordinate system adopted in the hypothesis, we know that (M,1)(M,-1) is a vertex of 𝒫X\mathcal{P}_{X}. After the last weighted blow up in the xx direction at the origin, it follows by Lemma 25 that the Newton polygon of 𝒫X~\mathcal{P}_{\widetilde{X}} of the adjoint vector field X~\widetilde{X} will contain the point (0,1)(0,-1), which means that the origin is not an equilibrium point of X~\widetilde{X}. Concerning equilibrium points that may appear on the divisor, they will be the same for both (X~Ω,δ~)(\widetilde{X}_{\Omega},\widetilde{\delta}) and (X~,δ~)(\widetilde{X},\widetilde{\delta}) and the arrangement of such equilibria are the same for both pairs.

Moreover, when we look to 𝒫X~Ω,A\mathcal{P}_{\widetilde{X}_{\Omega,A}} and 𝒫X~A\mathcal{P}_{\widetilde{X}_{A}}, we see that these polygons have the point (0,0)(0,0), because the origin is an impasse point (see Figure 42). In brief, the origin is not an equilibrium point and it is an impasse point, so the origin is elementary. Observe that we applied the same number of blowing-ups with the same weight at each step in (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta). Therefore, both pairs have the same elementary singularity scheme.

Refer to caption
Figure 42. Main segment before and after the last blowing-up in the Case 1, where the Newton polygons 𝒫XΩ,A\mathcal{P}_{X_{\Omega,A}} and 𝒫XA\mathcal{P}_{X_{A}} do not have a vertex of the form (m,1)(m,1).

Case 2: The Newton polygons 𝒫XΩ,A\mathcal{P}_{X_{\Omega,A}} and 𝒫XA\mathcal{P}_{X_{A}} have a vertex of the form (m,1)(m,1). This means that the Newton polygon 𝒫X\mathcal{P}_{X} of XX has a vertex of the form (m,0)(m,0). By the coordinate system adopted in the hypothesis, we know that (M,1)(M,-1) is also a vertex of 𝒫X\mathcal{P}_{X}, with m<Mm<M. After a finite number of weighted blow ups in the xx direction, the Newton polygon 𝒫X~\mathcal{P}_{\widetilde{X}} of the adjoint vector field X~\widetilde{X} will have the point (0,0)(0,0), which means that the origin is a semi hyperbolic equilibrium point of X~\widetilde{X}. On the other hand, when we look at 𝒫X~Ω,A\mathcal{P}_{\widetilde{X}_{\Omega,A}} and 𝒫X~A\mathcal{P}_{\widetilde{X}_{A}}, we see that these polygons have the point (0,1)(0,1), given that the origin is a semi hyperbolic equilibrium point and an impasse point at the same time.

So, this case is different from Case 1 because we need to apply one more blowing up in the objects (X~Ω,δ)(\widetilde{X}_{\Omega},\delta) and (X~,δ)(\widetilde{X},\delta) in the xx direction. Observe that, since both 𝒫X~Ω,A\mathcal{P}_{\widetilde{X}_{\Omega,A}} and 𝒫X~A\mathcal{P}_{\widetilde{X}_{A}} have points of the form (0,1)(0,1) and (M^,0)(\widehat{M},0), we will apply the same weighted blow up to both (X~Ω,δ)(\widetilde{X}_{\Omega},\delta) and (X~,δ)(\widetilde{X},\delta). At the end of the process, by Lemma 25 we see that (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same elementary singularity scheme.

Refer to caption
Figure 43. Main segment before and after the last blowing-up in the Case 2, where the Newton polygons 𝒫XΩ,A\mathcal{P}_{X_{\Omega,A}} and 𝒫XA\mathcal{P}_{X_{A}} have a vertex of the form (m,1)(m,1).

Case 3: Suppose that the impasse set coincides with a separatrix of (0,0)(0,0). Following the ideas presented in the previous cases, one sees that the process of desingularization of the constrained system is, essentially, the process of the desingularization of the adjoint vector field. After a suitable sequence of weighted blow-ups, the equilibrium points in the exceptional divisor are the same for both vector fields XΩX_{\Omega} and XX. Furthermore, the arrangement of such equilibrium points are the same for XΩX_{\Omega} and XX and they are elementary. Since the impasse set coincides with a separatrix of (0,0)𝔼x(0,0)\in\mathbb{E}_{x}, it follows that (XΩ,δ)(X_{\Omega},\delta) and (X,δ)(X,\delta) have the same elementary singularity scheme. ∎

Combining Theorem A and Proposition 26, we obtain the Theorem C.

5.4. Example

Consider the diagonalized constrained system

(15) yx˙=y3+x2y+x4;yy˙=x3+xy2+y4;y\dot{x}=y^{3}+x^{2}y+x^{4};\ y\dot{y}=x^{3}+xy^{2}+y^{4};

which is already written in favorable coordinates. The support of its auxiliary vector field XAX_{A} is the set 𝒬={(1,4),(1,2),(3,1),(3,0),(0,4)}\mathcal{Q}=\{(-1,4),(1,2),(3,1),(3,0),(0,4)\}, and therefore the principal part of (15) is given by

(16) yx˙=y3+x2y;yy˙=x3+xy2.y\dot{x}=y^{3}+x^{2}y;\ y\dot{y}=x^{3}+xy^{2}.
Refer to caption
Figure 44. Newton polygon of the auxiliary vector field of (15).

Observe that the constrained systems (15) and (16) are under the hypotheses of the Theorem C. The main segment γ1𝒫XA\gamma_{1}\subset\partial\mathcal{P}_{X_{A}} is contained in the affine line {r+s=6}\{r+s=6\}, thus the weight vector is ω=(1,1)\omega=(1,1). It can be checked that, for both systems (15) and (16), in the positive yy direction the points (±1,0)𝔼y(\pm 1,0)\in\mathbb{E}_{y} are hyperbolic saddles of the adjoint vector field. On the other hand, in the positive xx direction the points (±1,0)𝔼x(\pm 1,0)\in\mathbb{E}_{x} are hyperbolic saddles and the origin is an impasse point. The systems (15) and (16) have the same elementary singularity scheme, and therefore they are orientation preserving C0C^{0}-orbitally equivalents. See the Figure 45.

Refer to caption
Figure 45. Constrained system (15) and its strict transformed.

6. Acknowledgments

The first author is supported by Sao Paulo Research Foundation (FAPESP) (grants 2016/22310-0 and 2018/24692-2), and by Coordenação de Aperfeiçoamento de Pessoal de Nível Superior - Brasil (CAPES) - Finance Code 001. The second author was financed by CAPES-Print and FAPESP. The authors are grateful for Daniel Cantergiani Panazzolo for many discussions and suggestions, and for LMIA-Université de Haute-Alsace for the hospitality during the preparation of this work.

References

  • [1] M. J. Álvarez, A. Ferragut and X. Jarque. A survey on the blow up technique. Int. J. Bifurcation and Chaos 21 (2011), 3103–3118.
  • [2] V.I. Arnold. Normal forms for functions near degenerate critical points, the Weyl groups of AkA_{k}; DkD_{k}; EkE_{k} and Lagrangian singularities. Functional Anal. Appl. 6 (1973), 254–272.
  • [3] A. Belotto. Global resolution of singularities subordinated to a 1-dimensional foliation. Journal of Algebra 447 (2016), 397–423.
  • [4] I. Bendixson. Sur les courbes définies par des équations différentielles. Acta Math. 24 (1901), 1–88.
  • [5] M. Brunella, M. Miari. Topological equivalence of a plane vector field with its principal part defined through Newton Polyhedra. J. Differ. Equations 85 (1990), 338–366.
  • [6] P. T. Cardin, P. R. da Silva, M. A. Teixeira. Implicit differential equations with impasse singularities and singular perturbation problems. Isr. J. Math. 189 (2012), 307–322.
  • [7] L. O. Chua, H. Oka. Normal forms for constrained nonlinear differential equations. I. Theory. IEEE Transactions on Circuits and Systems, vol. 35, no. 7 (1988), 881–901.
  • [8] F. Dumortier. Singularities of vector fields on the plane. J. Differ. Equations 23 (1977), 53–106.
  • [9] F. Dumortier, J. Llibre, J. C. Artés. Qualitative Theory of Planar Differential Systems. Universitext, Springer-Verlag Berlin Heidelberg (2006).
  • [10] B. D. Lopes, P. R. da Silva, M. A. Teixeira. Piecewise implicit differential systems. J. Dyn. and Differ. Equations, v.29-4 (2017), 1519–1537.
  • [11] J. Munkres. Topology. Prentice Hall, 2nd edition (2000).
  • [12] D. Panazzolo. Resolution of singularities of real-analytic vector fields in dimension three. Acta Math. 197 (2006), no. 2, 167–289.
  • [13] M. Pelletier. Éclatements quasi homogènes. Ann. Fac. Sci. Toulouse Math., vol 4 (1995), 879–937.
  • [14] O. H. Perez, P. R. da Silva. Resolution of singularities of 2-dimensional real analytic constrained differential systems. 2020, Submitted, available in https://arxiv.org/pdf/2012.00085.
  • [15] P. J. Rabier, W. C. Rheinboldt. Theoretical and numerical analysis of differential-algebraic equations, in P. G. Ciarlet et al. (eds.), Handbook of Numerical Analysis, North Holland/Elsevier, Vol. VIII (2002), 183–540.
  • [16] R. Riaza. Differential-Algebraic systems: Analytical aspects and circuit applications. World Scientific Publishing (2008).
  • [17] R. Schneider. Convex Bodies: The Brunn-Minkowski Theory. Encyclopedia of mathematics and its applications. 44. Cambridge University Press (1993).
  • [18] A. Seidenberg. Reduction of singularities of the differential equation Ady=BdxAdy=Bdx. Amer. J. Math. 90 (1968), 248–269.
  • [19] S. Smale. On the mathematical foundations of electrical circuit theory. J. Differ. Geom. 7 (1972), 193–210.
  • [20] J. Sotomayor, M. Zhitomirskii. Impasse singularities of differential systems of the form A(x)x=F(x)A(x)x^{\prime}=F(x). J. Differ. Equations 169 (2001), 567–587.
  • [21] M. Zhitomirskii. Local normal forms for constrained systems on 2-manifolds. Bol. Soc. Brasil. Mat. 24 (1993), 211–232.