On the classification of planar constrained differential systems under topological equivalence
Abstract.
This paper concerns the local study of analytic constrained differential systems (or impasse systems) of the form , , where is a vector field and 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 .
Key words and phrases:
Constrained differential systems, Implicit ordinary differential equations, Topological equivalence, Blow ups, Newton polygon2020 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 is given by
(1) |
where , is a matrix valued function and is a vector field. We assume that these maps are analytic in the open set . A point such that is called impasse point, where denotes the determinant function . The so called impasse set is the set .
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 does not have constant rank. We will further suppose that only in a closed subset of . This implies that the adjoint matrix of is not identically zero and .
Multiplying both sides of (1) by , it can be shown that is solution of (1) if, and only if, is a solution of
(2) |
The constrained system written in the form (2) will be called diagonalized constrained system and, just as in [6], the vector field
(3) |
will be called adjoint vector field. Observe that, outside the impasse set, the constrained system (1) can be written as
which is a classical ordinary differential equation.
At an impasse point , since the matrix 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 , the phase portrait is given by the adjoint vector field . On the other hand, in the open set where the phase portrait is given by .
The study of singularities of constrained systems whose matrix does not have constant rank was addressed in [20, 21], where the authors considered singularities near regular points of the impasse set , such as tangencies between and the adjoint vector field and equilibrium points of on . By regular impasse point we mean that the determinant function satisfies . In such papers was adopted the following notion of equivalence.
Definition 1.
Two constrained systems , are -orbitally equivalent at the points , , if there are two open sets and and a diffeomorphism such that
-
(1)
maps the impasse set to the impasse set ;
-
(2)
maps the phase curves of in to the phase curves of in , not necessarily preserving orientation.
If is a homeomorphism, we say that is a -orbital equivalence or topological orbital equivalence. In the case where preserves orientation, we say that is an orientation preserving -orbital equivalence.
For normal forms of planar constrained systems where the matrix 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 does not have constant rank. Differently from [20, 21], we consider singular points of the impasse set . Furthermore, the adjoint vector field can present equilibrium points on more degenerated than those ones considered in these references. More precisely:
Definition 2.
We say that is a singularity of the constrained system (2) if one of the following conditions is satisfied:
-
(1)
is an equilibrium point of the adjoint vector field (3), that is, ;
-
(2)
the impasse set is not smooth at ;
-
(3)
is a tangency point between and , that is, .
Otherwise, we say that is non-singular. See Figure 1.


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 , 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 if one of the following conditions holds:
-
(1)
is a non singular point;
-
(2)
is a semi-hyperbolic equilibrium point of (that is, at least one of its eigenvalues is nonzero) and ;
-
(3)
If is a semi-hyperbolic equilibrium point of and , then coincides with a local separatrix of at .
See Figure 2.

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 will be denoted by a triple , where is the adjoint vector field (3) and is an analytic irreducible function such that the impasse set is given by . When is the origin, a constrained system is simply denoted by . After a suitable finite sequence of weighted blow ups, we obtain the so called strict transformed of , where 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 -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 and , where and are singularities. If the strict transformed of both constrained systems have the same non-degenerated elementary singularity scheme, then and are orientation preserving orbitally -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 as the set of analytic planar constrained systems defined near the origin such that:
-
(1)
The adjoint vector field is constant, that is, it satisfies , and .
- (2)
The Theorem B give conditions to assure when an analytic constrained system is orbitally equivalent to a system in , where the coefficients and in the expansion
(4) |
satisfy and the function is one of the functions in Table 1.
Theorem B.
In section 4 we present all the possible phase portraits of the systems that belong to .
Concerning Theorem C, suppose that the origin is a singularity of an analytic constrained system , where is an equilibrium point of the adjoint vector field and a regular point of the impasse set simultaneously. We associate a Newton polygon to such constrained system and we show in Theorem C that the terms in the boundary of the polygon determine the phase portrait near the origin under topological orbital equivalence. In other words, the terms in the boundary define the so called principal part 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 , one can “discard” the terms of the Taylor expansion that are not in the boundary of the polygon. This result extends Theorem A of [5] to the context of constrained systems with regular impasse set.
Theorem C.
Let be an analytic planar constrained system such that is a regular curve and the origin is an isolated Newton non-degenerated singularity. If has characteristic orbit (that is, the origin is neither a center nor a focus of ), then and are orientation preserving -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 -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 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 and a regular point of .
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 be an isolated equilibrium point of the planar vector field . Given positive integers and , a weighted (or quasi homogeneous) polar blow-up is the map
, | |||
. |
Let be the vector field defined in that satisfies . Define the vector field as , where is the biggest positive integer such that is an analytic vector field. Observe that is defined in a new ambient space and it is invariant in . The set is called exceptional divisor and the origin 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 -directional and positive -directional blow ups as the maps
, | |||
; |
, | |||
. |
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 and in the and directions, respectively. Observe that the polar and directional blow ups are equivalent (see Figure 3).
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 is an equilibrium point of 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 . 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 , where is a 2-dimensional real analytic manifold with corners, is an 1-dimensional analytic oriented foliation defined on and is the principal ideal sheaf. On each open set of the open covering of , we associate the diagonalized constrained differential system
Theorem 4.
Let 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
such that 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 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 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 will be denoted as a triple , where is the adjoint vector field and is an irreducible analytic function such that .
Example 5.
Consider the following constrained differential equation
(5) |
The adjoint vector field is given by , and therefore there is a cusp singularity at the origin. Moreover, the impasse set is given by . We consider the directional weighted blow-ups in the positive and directions
respectively. Thus we obtain the constrained system
in the positive direction and
in the positive direction. In Section we briefly discuss how to make a good choice of the weight vector. See Figure 4.

3. The construction of the orbital -equivalence near the exceptional divisor
Let be a constrained differential system defined near an impasse point . Suppose that is an equilibrium point of or a singular point of the impasse set . In this point is applied a suitable finite sequence of weighted blowing-ups, whose composition is simply denote by . According to Theorem 4, at the end of the process, we obtain a constrained system such that all the points in a neighborhood of the exceptional divisor are elementary. Moreover, is an invariant set of .
Analogously, let be a planar constrained system where is an equilibrium point of the adjoint vector field or a singular point of the impasse set . After a finite number of weighted blowing-ups (whose composition we denote by ), we obtain the strict transformed where all the points near the exceptional divisor are elementary and is an invariant set of .
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 -equivalence between and . The Proposition 6 will say that if there is an orientation preserving orbital -equivalence between and in a neighborhood of the exceptional divisor, then and are equivalents. See Figure 5.
Let and be neighborhoods of the blow-up centers and , respectively. Observe that and are open sets. Define the open sets and and suppose that there is a homeomorphism (possibly taking smaller neighborhoods and if it is necessary) satisfying the following conditions:
- (H1):
-
;
- (H2):
-
;
- (H3):
-
Let and denote by the flow of through . Suppose that , for some . Then there is such that
Proposition 6.
If there is a homeomorphism satisfying the conditions (H1), (H2) and (H3) above, then there are neighborhoods and and an orientation preserving orbital -equivalence between the constrained systems and .
Proof.
Denote and and define the map as
We claim that is an orientation preserving orbital -equivalence between and . Recall that restricted to is an analytic diffeomorphism that maps into , and it maps phase curves of into phase curves of , preserving orientation. Analogous properties for restricted to are true. Moreover, , and .
A straightforward computation shows that maps impasse set into impasse set. Furthermore, if is the flow of through and for some , then there is such that
and therefore is an orientation preserving orbital -equivalence. ∎
Since the existence of assures the existence of an orientation preserving orbital -equivalence between and , now we will provide a condition based on the resolution of singularities that guarantees the existence of such homeomorphism . This condition is the so called elementary singularity scheme. The next theorem says that if and have the same elementary singularity scheme, then such triples are (orientation preserving) orbitally -equivalent near the exceptional divisor. Such result and the Proposition 6 will assure that and are orbitally -equivalent near the blow-up center.
3.1. Basic points, basic singular interval and elementary singularity scheme
Let be a planar constrained system where is an equilibrium point of or a singular point of the impasse set. Consider its strict transformed where all points in the exceptional divisor 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 , or (see Definition 5.1, [8]).
Define the following vector fields with restricted domain, where :
Analogously, define the following constrained differential systems with restricted domain, where :








Definition 7.
A point is a basic singularity if in a neighborhood of there is a coordinate system adapted to the divisor such that the constrained system is orientation preserving orbitally -equivalent to or , where .
Definition 8.
A basic singular interval is an arc such that all the points in are equilibrium points and in a neighborhood of there is a coordinate system adapted to the divisor such that the constrained system is orientation preserving orbitally -equivalent to or .
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 follows and . 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
Such arrangement defines an equivalence relation in the set of triples in the following way: two triples and are equivalent if, and only if, the word associated to can be changed to the word associated to by cyclic permutations.
Definition 9.
An equivalence class in is called elementary singularity scheme.
Example 10.
The elementary singularity scheme of the constrained system Example 5 is given by the word .
The next theorem is true under the hypothesis that the elementary singularity schemes do not contain only words written with symbols of the form or . This is equivalent to require the following. Let and be the strict transformed of the constrained systems and , respectively, where and . If and are equilibrium points of the adjoint vector field and , respectively, then both and have characteristic orbit. In other words, and 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 only contains symbols of the form or . Otherwise, we say that the elementary singularity scheme is non-degenerated.

3.2. Proof of Theorem A: The construction of the orbital -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 be topological spaces and let be open (or closed) subspaces. Denote . Let and be homeomorphisms such that . Then there is a homeomorphism such that and .
Proposition 13.
Suppose that the triples and have the same non-degenerated elementary singularity scheme. Then there is an orientation preserving orbital -equivalence between and such that .
Proof.
Given an arc , we have its respective . The construction of the homeomorphism starts near a basic point (or basic singular interval) of that is not of the form or . Hence there is an equivalent point (or basic singular interval) of . This means that there is adapted coordinates
for and , respectively, where and . Denote . The neighborhoods and are chosen in such way that there is only one basic point (or only one basic singular interval).
If we start the construction of the homeomorphism from a basic point, we take the open set 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 is chosen in such way that it is transversal to the lines and .
Therefore, we have that is transversal to . This is also true when we take and . Applying this reasoning to all basic points and basic singular intervals in the exceptional divisor, a sufficiently small neighborhood of the exceptional divisor is divided into a finite number of sectors. If this sector is not locally given by , we then subdivide such sector by the segment . We apply the same reasoning for .
These sectors are bounded by arcs of the excepcional divisor, phase curves, impasse curves or curves transversal to the divisor. Considering , these sectors are written in local coordinates as follows:
-
•
Sector : , where and . This sector is equivalent to the basic point ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and ;
-
•
Sector : , where and .
We remark that the sector is -equivalent to .






Consider once again the point that we took in the beginning of this proof. Recall that this point is neither nor . We know that is a homeomorphism such that , it maps phase curves of into phase curves of and it sends impasse set into impasse set. Moreover, 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 or . In this second sector it is defined a -equivalence between and . Observe now that on the homeomorphisms and coincide.

We remark that each phase curve intersects in only one point. Now, by the Pasting Lemma 12, there is an homeomorphism defined in both sectors that maps phase curves of into phase curves of , maps impasse set into impasse set and sends excepcional divisor into excepcional divisor.
Between each sector of the form or it may have a sector of the form or , 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 (or basic singular interval) that we took in the beginning of the proof is not of the form or . Geometrically, it means that we must avoid the center-focus case.
Thus we have constructed an homeomorphism such that , maps phase curves of into phase curves of and sends into . Then is the orbital -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 -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 is the set of constrained systems such that:
-
(1)
The adjoint vector field is constant, that is, it satisfies , and .
-
(2)
The function is one of the curves given by Arnold’s ADE-classification (see Table 1).
Type | Normal Form | Codimension |
---|---|---|
, | ||
, | ||
We chose a coordinate system such that is one of the functions given by Table 1. Since the adjoint vector field and are analytic, and assuming that , that is, the coefficients and in the expansion
(6) |
satisfy , the Theorem B gives conditions to assure when the System (6) is orientation preserving -orbitally equivalent to the system
(7) |
which belongs to .
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 is the first positive integer such that . 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 | Conditions | ||
---|---|---|---|
Case : and | |||
Case : and | |||
Case : and | |||
Case : and for all integer | |||
Case : and | |||
Case : and | |||
Case : | |||
Case : and for all integer | |||
or |
|
In what follows we briefly describe the phase portraits of (7) near the exceptional divisor for each curve 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 be an analytic vector field. Just as in [12], we write in the so called logarithmic basis. More precisely, consider
where
with and satisfying:
-
(1)
For or , ;
-
(2)
For or , .
Let be a vector of positive integers. One can write the planar vector field as
(8) |
where
(9) |
In other words, the vector field can be written as a sum of quasi-homogeneous components. We say that (8) is a -graduation of . Each is called -level of the -graduation.
Given a -graduation of , we associate the monomials and with nonzero coefficients to a point in the plane of powers, where each point is contained in a line of the form .
Definition 14.
The support of is the set
Definition 15.
The Newton polygon associated to the analytic vector field is the convex envelope of the set .
The boundary of the Newton polygon is the union of a finite number of segments. We enumerate them from the left to the right: . Observe that is vertical and is horizontal. Analogously, the non-regular points of are enumerated from the left to the right: .
Definition 16.
We say that are the vertices of the Newton polygon. The vertex is called main vertex and the segment will be called main segment. The number is the height of the Newton polygon. See Figure 22.

The main segment is contained in an affine line of the form . Observe that the vector is normal with respect to . In the resolution of singularities of analytic vector fields, one chooses the vector 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) |
with being an irreducible real-analytic function that can be written as
where is such that when or .
Definition 17.
The auxiliary vector field associated to the constrained differential equation (10) is the vector field
(11) |
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 . On the other hand, the impasse set 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 , and therefore the support takes form
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 and the points of the support of . Moreover, the levels of a -graduation are written in the form
where . 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 . 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 to chose the weight vector of the blow up.
Now we are able to study constrained systems in the set .
4.3. Curve ,
Firstly assume the conditions in the case 1. After a weighted blow up in the direction, the origin is a hyperbolic saddle and the impasse curve satisfies . In the direction there are no equilibrium points in and the impasse curve satisfies the equation . 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 direction, there are no equilibrium points in , and in the direction the origin is a hyperbolic saddle. The impasse set behaves just as the previous cases.
Geometrically, measures the tangency order between the adjoint vector field and the coordinate axis , and measures the tangency order between and . Therefore, the assumption sets a relation between such tangency orders. It is important to remark that, when , the main segment of the Newton polygon of the auxiliary vector field of (6) has the point , and such point does not appear in the Newton polygon of the auxiliary vector field of (7) See figures 25 and 26.




4.4. Curve ,
Assume the hypotheses in case 4. After a weighted blow up in the direction, 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 direction, there are no equilibrium points in and the impasse set is give by .
We emphasize that the condition for all integer avoids tangency points between the adjoint vector field and the component 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 direction, there are no equilibrium points of in . On the other hand, in the direction the origin 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 direction the impasse curve is given by and there are no equilibrium points of the adjoint vector field in . On the other hand, after a weighted blow up in the direction the impasse curve is given by and the origin is a hyperbolic saddle.
The number is related to the tangency order between the adjoint vector field and the axis , and is related to the tangency order between and the axis . Thus the assumption sets a relation between such tangency orders. Observe that in the case , the main segment of the Newton polygon of the auxiliary vector field of (6) has the point , and such point obviously does not appear in the Newton polygon of the auxiliary vector field of (7). See Figures 31 and 32.






4.5. Curve
Suppose without loss of generality that . In the positive direction the impasse curve intersects the exceptional divisor at the points . For the origin is a hyperbolic saddle and if we do not have equilibrium points of the adjoint vector field on the divisor. In the positive direction, the impasse curve intersects the exceptional divisor at . If , there are no equilibrium points of the adjoint vector field in and if the origin is a hyperbolic saddle. See Figures 33 and 34.


4.6. Curve
Suppose that and , for all integer . In the positive direction, the impasse curve intersects the at the points , and . 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 direction, there are no equilibrium points of in and the impasse set intersects the exceptional divisor at .
Once again we remark that the condition avoids tangency points between the adjoint vector field and the component of the impasse set. Such tangency points would lead us to a different resolution of singularities.
If we consider , in the positive direction there are no equilibrium points of in , and in the positive direction the origin is a saddle points of . The impasse curve behaves just as in the previous case.


4.7. Curve
In the positive direction, the impasse curve intersects the exceptional divisor at . If , then the origin is a hyperbolic saddle of the adjoint vector field and if there are no equilibrium equilibrium points of the adjoint vector field on the divisor. Finally, in the positive direction the impasse curve intersects the exceptional divisor at . If , there are no equilibrium points of the adjoint vector field in and if the origin is a hyperbolic saddle of the adjoint vector field. See Figures 37 and 38.


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 be a planar analytic vector field defined near the origin such that , and denote its Newton polygon by (see Definition 15 below). In [5] the authors proved that, under some non degeneracy conditions, the terms of associated to the the boundary of the Newton polygon determines the phase portrait of near the equilibrium point.
More precisely, given an analytic vector field , the terms of associated to the points in define the so called principal part of , which is denoted by . The Theorem A of [5] assures that and 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 if the Newton polygon associated to the auxiliary vector field satisfies one of the following:
-
•
The main vertex of is , or (that is, if the height is less or equal to zero);
-
•
The main segment is horizontal.
Theorem 19.
(See [14]) A planar constrained system is elementary at if, and only if, it is Newton elementary at .
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 , where is the adjoint vector field and is a real irreducible analytic function. As usual, the auxiliary vector field will be denoted by . The Newton polygons of the adjoint vector field and the auxiliary vector field will be denote by and , respectively.
In this section, constrained differential systems with regular impasse set are considered. By “regular impasse set” we mean that the gradient vector 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 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 associated to a vector field is controllable if the main vertex is contained in or .
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 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 ;
-
•
The Newton polygons and of the auxiliary vector field and the adjoint vector field , respectively, are controllable;
-
•
The polygon is obtained by increasing one unity on the second coordinate of each point of the polygon .
Proof.
The first item is true due to the Implicit Function Theorem. For the second item, we can apply the change of coordinates
with (see Lemma 9, [14]). Observe that after this change of coordinates, the impasse set is still of the form . Recall that the Newton polygon of the auxiliary vector field is obtained by the Minkowski sum between the Newton polygons of and . Since in this coordinate system the function is given by , it follows that the Newton polygon of the auxiliary vector field is obtained by increasing one unity the second coordinate of each point of the Newton polygon of . See Figure 39. ∎

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 . Furthermore, by Lemma 21, it follows that and have the same number of segments. In addition, given a segment , the respective segment has the same slope as . This implies that, if the main segment is contained in a line of the form , then the segment is contained in a line of the form . 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 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 in favorable coordinates, where is the principal part of the adjoint vector field 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
where is the vector field defined by the terms of that belong to the segment , where is neither vertical nor horizontal.
Example. Consider the diagonalized constrained differential system
(12) |
which is already written in favorable coordinates and whose auxiliary vector field is given by
(13) |
5.2. Non-degeneracy conditions
In this subsection we discuss some non-degeneracy conditions required for the pair in order to show the existence of an orbital -equivalence between and . As usual, we consider in favorable coordinates.
The first assumption is that the origin is an isolated equilibrium point of the adjoint vector field . This implies that intersects the coordinate axes of the -plane. However, we remark that, in general, the Newton polygon of the auxiliary vector field only intersects the -axis.
The next non-degeneracy condition was introduced in [5].
Definition 24.
Let be an analytic vector field such that is an isolated equilibrium point and let be its principal part. We say that is Newton non-degenerated at if any quasi homogeneous component of associated to a side does not have singularities in , that is, if each 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 direction. In other words, all the equilibrium points in 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 of all planar analytic constrained systems defined near the origin in favorable coordinates.
We end this subsection with the following remark. Suppose that contains a point of the form . In other words, suppose that the component of has a term of the form , with . Geometrically, this condition means that any phase curve of does not coincide with the impasse curve . By Lemma 21 it follows that has a point of the form . On the other hand, if does not contain a point of the form , then the impasse set coincides with a phase curve. By Lemma 21 it follows that does not intersect the -axis of the -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 satisfies the hypotheses discussed in the previous subsection.
Lemma 25.
Let be a 2-dimensional constrained system and take favorable coordinates. Suppose that
-
(1)
The vector field is Newton non-degenerated at the origin;
-
(2)
= , that is, the boundary of the polygon of has just one segment that is neither vertical nor horizontal.
Then and have the same elementary singularity scheme.
Proof.
We will study the case where the main vertex is of the form . The case where the main vertex is is analogous. The proof is based in directional weighted blow ups and we focus only in the positive and directions, provided that the computations are similar in the negative directions.
Since we are adopting favorable coordinates, the impasse curve only appears in the direction. This means that, in the 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 direction both pairs and have the same equilibrium points, and the arrangement of such points is the same for both pairs. We emphasize that such equilibrium points in are semi-hyperbolic (and therefore elementary), because the constrained system is Newton non-degenerated.
After a blow up in the direction, the first graduation of the auxiliary vector field takes the form
Observe that the impasse curve is still of the form and the equilibrium points of the strict transformed and 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 and 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 be a 2-dimensional constrained system written in favorable coordinates. If the origin is Newton non-degenerated, then the pairs and 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 . The case where the main vertex is is analogous. We will compare the process of resolution of singularities of and . Since the segments of the Newton polygons and 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 only appears in the direction. This means that in all points of the divisor , both pairs and have the same equilibrium points, all of them are semi-hyperbolic and the arrangement of such equilibrium points are the same for both and , provided that in the positive 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 direction. See Figure 40.

Applying the first blow-up with weight in the direction, we obtain the strict transformed and . It is straightforward to see that is still of the form , 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 we have the same equilibrium points for both and . We emphasize that all the points in 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.

Due to Lemma 21, the Newton polygons of and are obtained by increasing one unity to the second coordinate of the points of . 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 and .
Essentially, we are only desingularizing equilibrium points of and and preserving the impasse curve.
In the end of the process, we have three cases to consider at the impasse point . The first two cases concern the case where the impasse set does no coincide with a separatrix of the equilibrium point, that is, the Newton polygon has a point of the form . The third case concerns the case where does not have a point of the form (which means that the impasse set coincides with a separatrix of the equilibrium point).
Case 1: The Newton polygons and do not have a vertex of the form . This means that the Newton polygon of does not have a vertex of the form . By the coordinate system adopted in the hypothesis, we know that is a vertex of . After the last weighted blow up in the direction at the origin, it follows by Lemma 25 that the Newton polygon of of the adjoint vector field will contain the point , which means that the origin is not an equilibrium point of . Concerning equilibrium points that may appear on the divisor, they will be the same for both and and the arrangement of such equilibria are the same for both pairs.
Moreover, when we look to and , we see that these polygons have the point , 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 and . Therefore, both pairs have the same elementary singularity scheme.

Case 2: The Newton polygons and have a vertex of the form . This means that the Newton polygon of has a vertex of the form . By the coordinate system adopted in the hypothesis, we know that is also a vertex of , with . After a finite number of weighted blow ups in the direction, the Newton polygon of the adjoint vector field will have the point , which means that the origin is a semi hyperbolic equilibrium point of . On the other hand, when we look at and , we see that these polygons have the point , 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 and in the direction. Observe that, since both and have points of the form and , we will apply the same weighted blow up to both and . At the end of the process, by Lemma 25 we see that and have the same elementary singularity scheme.

Case 3: Suppose that the impasse set coincides with a separatrix of . 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 and . Furthermore, the arrangement of such equilibrium points are the same for and and they are elementary. Since the impasse set coincides with a separatrix of , it follows that and have the same elementary singularity scheme. ∎
5.4. Example
Consider the diagonalized constrained system
(15) |
which is already written in favorable coordinates. The support of its auxiliary vector field is the set , and therefore the principal part of (15) is given by
(16) |

Observe that the constrained systems (15) and (16) are under the hypotheses of the Theorem C. The main segment is contained in the affine line , thus the weight vector is . It can be checked that, for both systems (15) and (16), in the positive direction the points are hyperbolic saddles of the adjoint vector field. On the other hand, in the positive direction the points 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 -orbitally equivalents. See the Figure 45.

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 ; ; 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 . 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 . 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.