A complete characterization of sharp thresholds to spherically symmetric multidimensional pressureless Euler-Poisson systems
Abstract.
The Euler-Poisson system describes the dynamic behavior of many important physical flows including charge transport, plasma with collision and cosmological waves. We prove sharp threshold conditions for the global existence/finite-time-breakdown of solutions to the multidimensional pressureless Euler-Poisson (EP) system with or without background and general initial data. In particular, the initial data could include points where velocity is negative, that is, the flow is directed towards the origin. Obtaining threshold conditions for such systems is extremely hard due to the coupling of various local/nonlocal forces. Remarkably, we are able to achieve a sharp threshold for the zero background case and most importantly, the positive background case, which is quite delicate due to the oscillations present in the solutions. We discover a completely novel nonlinear quantity that helps to analyze the system. In the case of positive background, if the initial data results in a global-in-time solution, then we show that the density is periodic along any single characteristic path. We use the Floquet Theorem to prove periodicity.
Key words and phrases:
Critical thresholds, global regularity, shock formation, Euler-Poisson system2020 Mathematics Subject Classification:
35A01; 35B30; 35B44; 35L451. Introduction
A general system of pressureless Euler-Poisson (EP) equations has the following form,
(1.1a) | |||
(1.1b) | |||
(1.1c) |
with smooth initial data . The constant parameters are the forcing coefficient and background state respectively. The sign of the forcing coefficient signifies the type of particles being modeled and its magnitude gives a measure of the strength between them. When the force in-between particles is repulsive, for example in the case of charge flow, then . Within the pressureless setup (1.1), is relevant in the case of interstellar clouds where the pressure gradient becomes negligible compared to the gravitation forces, see [11]. In the pressureless setup with same charge particles (), the background state is, in practicality, a profile, that is, a function of the spatial variable, , see [13]. The background models the doping profile for charge flow in semiconductors. However, the sheer complicacy of the system has restricted researchers to consider the background as a constant. To our knowledge, [3] is the only work that considers background as a profile in obtaining critical thresholds.
A locally well-posed PDE system exhibiting critical threshold phenomena is the one wherein existence of global-in-time solutions is dependent on whether the initial data crosses a certain threshold manifold. This critical threshold manifold divides the phase space of initial data into two mutually exclusive regions or sets. If the initial data lies completely in one of the sets (also called subcritical region), there is global solution. However, if some part of initial data lies outside this set, or in other words, in the supercritical region, a breakdown occurs and solution loses its smoothness in finite time.
As one can expect, the question of proving the existence and subsequently finding the critical threshold manifold is simpler in one dimension (). In this case, the threshold is a curve on the plane and the subcritical region is given by,
A vast amount of literature exists for critical thresholds to (1.1) and other similar systems. The existence of such curve was first identified and analyzed in [9] for EP systems. The authors analyzed the one dimensional and multidimensional (with spherical symmetry) cases. A series of works then followed for EP as well as other systems, [2, 3, 4, 5, 7, 12, 16, 17, 18, 19, 20, 21, 22, 24, 25, 26, 27, 28, 29] among many others. As is known for general hyperbolic conservation laws that for any initial data, singularities develop as one moves forward in time, [15]. This is due to the convective forces present. Addition of source terms however, can result in a set of initial data that leads to global solutions. To extract a set of initial data (subcritical region) for which there is global well-posedness, is an interesting territory to be explored. The ‘goodness’ of the external forces can balance or even outweigh the ‘bad’ convective forces and result in a ‘large’ subcritical region. An example where the external forces completely obliterate the convective forces is the strongly singular Euler-Poisson-alignment (EPA) system. In [14], the authors show that the system (1.1) with and an additional nonlocal alignment force in the momentum equation results in global-in-time solutions for any initial data. This result is remarkable and in stark opposition to the conventional results for EP systems.
For threshold results in one dimensional EP and EPA systems, one can refer to [2, 3, 4, 5, 8, 14, 26, 28]. In [26], the authors include pressure and exclude background. It is quite difficult to obtain thresholds for the full EP system (with pressure) due to the strict hyperbolicity resulting in two characteristic flow paths. Critical threshold for EP systems with pressure and is largely an open problem.
For pressureless EP systems, the key is to obtain bounds on the gradient of velocity. Owing to the local existence result, global-in-time solutions to (1.1) can be ‘extended’ upon a priori smooth local solutions. The local existence for such systems is well-known, see [23] for a proof. We state the result here.
Theorem 1.1 (Local wellposedness).
As a result, in 1D, one needs to bound to guarantee global well-posedness. However, things are different in higher dimensions where the gradient of velocity is in fact an matrix. One finds out that an ODE along the characteristic path can be obtained for the divergence of the gradient matrix. However, analyzing the divergence is not enough. One needs to control the spectral gap (sum of all the absolute differences of eigenvalues), see for example [12], in order to guarantee smooth solutions for all time. This is the tricky part. Given the inherent property of hyperbolic balance laws, it is in general easier to obtain sufficient conditions for breakdown of solutions as compared to conditions for global existence. In [6, 7], the authors obtain bounds on the supercritical region for (1.1) for .
To obtain bounds on subcritical region is more involved. Several workarounds have been used by various researchers by simplifying the EP system. In [16, 21], the authors study and obtain thresholds for the restricted EP equations in two and three dimensions respectively.
Another derivative of the EP system that has been extensively studied is (1.1) with spherical symmetry. It can be shown that if the initial data to (1.1) is such that for smooth functions , then for any fixed time and as long as the solution is smooth, the symmetry extends, that is, the solutions are . Applying this simplification to (1.1), we obtain the following system of equations,
(1.2a) | |||
(1.2b) | |||
(1.2c) |
with and subject to smooth initial density and velocity
Even though the system is now simplified as compared to the general EP system, it turns out that it is still tricky to analyze for thresholds, especially when . (1.2) was first studied in [9] with and expanding flows (). Sufficient conditions on global existence and finite-time-breakdown were obtained for and a sharp condition was obtained for . Later, a sharp condition was derived in [29] however, once again, only for zero background and expanding flows. A sufficient blow-up condition was derived in [30] for , however, no conclusion was made with regards to global existence.
The dynamics of (1.2) is quite different for compared with . One major difference is that in the former, the density is a perturbation around the constant background. The mass is infinite and the following holds,
which is different from the zero background state where the mass is finite and conserved. A major step forward to mitigate the restrictiveness of the past results was taken in [28]. The author there reduced (1.2) to a ODE system along a characteristic path, and subsequently proved the existence of critical thresholds. This ODE system derived by the author is crucial in our analysis as well. An intriguing discovery by the author was that the Poisson forcing in (1.2b) is enough to avoid concentration at the origin, even if there are points where the initial velocity points towards the origin. To our knowledge, it is the first work with regards to thresholds of spherically symmetric EP system wherein the assumption of expanding flows was dropped. It was also noted that is the critical case and the analysis needs to be done fairly differently as compared to the case when . The author obtained partial results for thresholds for , that is, bounds on the subcritical and supercritical regions were presented. However, no results were presented for citing the highly chaotic dynamics in such a scenario.
In this paper, we present sharp thresholds for both as well as . Several steps are needed to arrive at the precise thresholds. Special care has to be taken with regards to finding the subcritical region for . We essentially start out by characterizing the supercritical region and narrowing down the possible initial configurations that might allow for global solutions, eventually extracting out the subcritical region.
Our main results can be stated non-technically as follows:
-
•
For the nonzero background case (), we show that the EP system admits a global-in-time smooth solution if and only if the initial data is smooth, lying in a subcritical region, . Theorem 2.1 contains such critical threshold result. The explicit definition of is stated after the result.
-
•
For the zero background case (), we show that the EP system admits a global-in-time smooth solution if and only if the initial data is smooth, and lying in certain subcritical region, . Theorem 2.6 contains the precise thresholds for dimensions greater than or equal to three. Theorem 2.8 contains the threshold results in dimension two. The explicit definition of is stated after each of the results.
We discover a completely novel nonlinear quantity that helps to analyze the system. At , this is given by
(1.3) |
In fact, such quantity is very crucial in analyzing and simplifying the representation of subcritical/supercritical regions. The full motivation and usage of this quantity will be thoroughly discussed in the Sections to come. It is found that a simpler breakdown condition can be obtained using this expression, the corresponding result for dimensions greater than or equal to three is provided in Theorem 2.4.
1.1. A roadmap of the case
Before stating our main results, we give a short roadmap of how threshold regions are identified. Following is a list of key points:
-
•
The full dynamics of (1.2) can be reduced to a weakly coupled system of four equations (along characteristic path ) as,
with
The initial data being respectively, which depends on , but we do not specify it explicitly as we will be analyzing one characteristic path at a time. A thorough analysis of this ODE system becomes the main task. We restrict to the case , and .
-
•
The system is decoupled and admits a closed form of trajectory curves, expressed as,
with,
(1.6) This ensures that the system admits a global bounded solution if and only if . Moreover, the solutions are periodic and the trajectories rotate clockwise on the plane as time progresses. We also have
where are the minimum/maximum values of attained. In addition, is shown to be bounded along with,
-
•
Another key transformation of form,
leads to a new system,
This is a linear system with time-dependent yet bounded coefficients. This way, the key to existence of global solution to is equivalent to ensuring for all .
-
•
A nonlinear quantity of the form,
is shown to be bounded. and in (1.3) are essentially the same and we will see more about this in Section 5. More precisely, for , we obtain
We are able to conclude that will surely be zero at some positive time if does not change sign. This is in fact the case if either , or with
where is the unique positive root of ,
-
•
Therefore, to precisely identify the possible initial configurations for global existence, it is necessary to require
The key idea is to construct two nonnegative functions to precisely demarcate the subcritical region. We will show that if a solution, starts in between these functions at , then it remains so for all except discrete times (where it is positive), thereby providing a condition to ensure positivity of . The converse also holds. The two constructed functions form beads and the strictly positive solutions are contained within those beads, see Figure 5. More precisely, we show that if satisfies the above inclusion, then there exists two nonnegative functions so that the following holds. For the case , we show that if
then,
The converse also holds. Quite similarly, for , we have that if
then,
The functions are the key components through which the threshold curves will be defined.
2. Main Results
This section is devoted to stating precise versions of our main results. We start by specifying that for (1.2), the finite-time-breakdown is manifested in the following form,
for some and finite . The first blowup (shock formation) can be concluded from Theorem 1.1. The density blowing up at the same time/position is a consequence of weak hyperbolicity of the pressureless Euler-Poisson systems and is a well-known phenomena for such systems. In the next section, our analysis will allow us to conclude the same. Now we state our results and follow it up with an interpretation. As mentioned in the previous section that the case is critical and needs to be handled separately. Consequently, the representation of subcritical regions is also different for than those for . In this section, we also provide the subcritical region sets.
Theorem 2.1 (Sharp threshold condition).
We will describe the subcritical region set with the help of the quantity (1.3). Motivated by this, we set, for a point ,
(2.1) |
Definition 2.2.
A point () if and only if the following holds,
-
•
For
(2.2)
Definition 2.3.
A point if and only if the following holds,
-
•
For
(2.3)
Here, are the roots of the equation,
and
with
The explicit expression for is in (1.6), also in (4.7). The functions
are defined via a second order linear IVP and are positive for all except at infinitely many discrete times.
We now give an interpretation of Theorem 2.1. Similar to existing threshold results, Theorem 2.1 enables us to check pointwise whether or not the given initial data lies in the subcritical region. Not only this, we can in fact construct the subcritical region piece by piece. Our result enables us to divide the space of into a space and a plane, that is, (say ) and (say ). For each point in , we can construct the subricitical region on through . Let be any point in . Note that can now be calculated using and hence, one can now explicitly find the interval given in Definition 2.2. For any in that interval, (2.1) outputs a line in . In other words, by fixing an , we have a linear relation between . For this value of , we can find the two constants, . Indeed for fixed , the functions are known functions. Finally, (2.2) gives the desired portion of the line on that forms part of the subcritical region, as and according to whether is zero or not. We can do this for all in the above interval and obtain the complete part of the subcritical region for the fixed point in . Carrying out this procedure for each point in gives the entire subcritical region.
The next two results give a complete picture of the zero background case with dimension greater than or equal to three.
Theorem 2.4 (Sufficient condition for blow-up for ).
Remark 2.5.
This result gives an easy criteria to check breakdown. Moreover, it shows the criticality at . For , there is a flat cutoff that bounds the subcritical region from one side. However, this flat cutoff is absent in where can take any negative value and still possibly lie in the subcritical region.
In the next theorem, we give a full picture of what happens for initial data which satisfies for all .
Theorem 2.6 (Global solution for with zero background).
Suppose and in (1.2). If for all , the set of points
where is as in Definition 2.7, then there is global solution.
Moreover, if there exists an such that
then there is finite-time-breakdown.
Definition 2.7.
A point (for ) if and only if one of the following holds,
-
•
For ,
(2.4) -
•
For ,
(2.5) -
•
For ,
(2.6) -
•
For ,
(2.7)
Here, . The explicit expression for is in (4.7) (with ). The functions
are defined later via linear IVPs.
Theorem 2.8 (Global solution for with zero background).
Suppose and in (1.2). If for all , the set of points
where is as in Definition 2.9, then there is global solution.
Moreover, if there exists an such that
then there is finite time breakdown.
Definition 2.9.
A point if and only if one of the following holds,
-
•
For ,
(2.8) -
•
For ,
(2.9) -
•
For ,
(2.10) -
•
For ,
(2.11)
Here, . The explicit expression for is in (4.7) (with ).
This paper is arranged as follows. Section 3 entails some preliminary calculations showing how to reduce the full dynamics to a weakly coupled system of four ODEs along characteristics. Section 4 is devoted to the analysis of the system of two of the four ODEs, that are decoupled from the other two, obtaining a closed form of trajectory curves with related properties established for later use. Section 5 entails conditions under which the velocity gradient/density blow up. Section 6 is devoted to constructions of the precise subcritical region leading to global solutions, proving Theorem 2.1. The zero-background case is analysed in Section 7. Finally concluding remarks are given in Section 8.
3. Preliminary calculations
From Theorem 1.1, we note that to extend the local solutions, we need to obtain bounds on the gradient of velocity. The following result, that was derived by the authors in [25], designates the quantities that need to be bounded for (1.2).
Lemma 3.1.
Suppose be a radially symmetric function. Then for , has exactly two eigenvalues: and . Also, is an eigenvector corresponding to and is the eigenspace for the other eigenvalue.
Though the proof has already been given by the authors in [25], we include it here for the sake of completion.
Proof.
Let . Taking gradient of , we have
Taking gradient again,
Note that and for any . This completes the proof of the Lemma. ∎
Through this Lemma, we conclude that the Hessians of radial functions are diagonalizable using the same matrix of eigenvectors. From Theorem 1.1, we know that for solutions to persist for all time, the gradient matrix of velocity should be bounded. Given the velocity is radial vector field, we have the following relation for , in (1.1b), (1.2b) respectively,
Hence, the two eigenvalues of will be and . Therefore, to ensure that the gradient of velocity is bounded, we need to control the two quantities: .
Inspired by this, we will obtain ODE along the characteristic path,
(3.1) |
for the desired quantities. Rearranging (1.2a), we obtain,
(3.2) |
Note that by (1.2c),
Taking spatial derivative of (1.2b) and using the above equation we obtain,
(3.3) |
Next using (1.2b),
(3.4) |
Upon integrating (1.2c) from to ,
(3.5) |
Local well-posedness requires a boundary condition at the origin. The convention is to assume zero boundary conditions, that is, the above limit is zero. Similarly the density flux () approaches zero, which signifies that there is no loss of material at the origin. Taking time derivative of (3.5),
As a consequence, we get . We use this and the previously obtained expression for in the below calculation.
(3.6) |
Set
(3.7) |
Equations (3.2), (3.3), (3.4) and (3.6) can be used to obtain an ODE system,
(3.8a) | |||
(3.8b) | |||
(3.8c) | |||
(3.8d) |
with initial data respectively. Here, ′ indicates differentiation along the path (3.1). Note that we use the notation for the density as in (1.2a), which is a function of time and the spatial variable, as well as for the solution to the ODE (3.8a), which is another function of time only (for a fixed parameter ). Similarly for the notation . However, it will be clear from context whether we are referring to as the solution of the main PDE system (1.2) or as the unknown in (3.8a).
4. The system
Note that (3.8c), (3.8d) are decoupled from (3.8a), (3.8b). In fact, a closed form for the trajectory curve for equations (3.8c), (3.8d) can be obtained. We have the following Lemma which describes the behavior of the system (3.8c),(3.8d).
Lemma 4.1.
Global solution to (3.8c), (3.8d):
-
(1)
exist for all time and are uniformly bounded if and only if . In particular, if , then in finite time.
-
(2)
When , the solutions lie on bounded trajectory curves satisfying,
(4.1) (4.2) -
(3)
When , the solutions are periodic and the trajectories rotate clockwise on the plane as time progresses.
We will time and again use the variable,
(4.3) |
and (4.4b) instead of and (3.8d) as the situation demands since this will make calculations simpler and easy to understand. The equivalent system is,
(4.4a) | |||
(4.4b) |
Proof of Lemma 4.1: It seems intuitive to use instead of . We first prove the only if part of the first assertion through a contradiction argument. From (4.4b), one can see that maintains sign as long as exists. Now suppose . This implies as long as exists. For the sake of contradiction, suppose remains bounded for all time. From (4.4a) we have,
Since is bounded above by a strictly negative constant, there exists some finite time, , such that . Once is negative, the dynamics of is that of a Ricatti equation,
and therefore, for some . This is a contradiction.
Now we prove the if part of the first assertion along with the second assertion. We derive the equation of trajectory assuming , which in turn implies for all as long as the solutions to (4.4) exist. Dividing (4.4a) by (4.4b) we have the following,
Using the integrating factor ,
Owing to the last term, we see that case has to be handled separately before integrating the above equation. First, we move on with case. Integrating the equation above, we obtain
For , upon integration,
From the trajectory equation for , we have for ,
Hence, solutions are uniformly bounded. Very similarly, bounds on can be derived for as well. This completes the proof to the first and second assertions.
Note that trajectory equation is invariant under the transformation . Putting this together with the fact that linearized system around the only critical point, , has imaginary eigenvalues, we obtain that the solution trajectories starting at any are closed curves around the critical point. Hence, the solutions are periodic. The direction of motion of trajectory as time progresses is clear from (4.4b) since for and for . ∎

Corollary 4.2.
Proof.
Remark 4.3.
This result simply asserts that in multidimension, Poisson forcing is enough to avoid concentrations at the origin, irrespective of how negative the initial velocity is. This is not the case in one dimension. A more detailed discussion comparing 1D and multi-D cases is included in Section 8 (conclusion).
At this point, we set up the notation for the trajectory curve as obtained in Lemma 4.1. It will be used in this as well as in later sections.
(4.7) |
From Lemma 4.1,
Since is a constant, we will use the notation to denote it as a single constant or for a function of . The next Lemma pertaining to (4.4) will be useful in proving the blowup/global existence results in Section 6.
Lemma 4.4.
Proof.
We give a proof for only. Very similar arguments apply for . Note that for , the statement is equivalent to saying,
(4.8) |
where , .
Since the solution trajectories, , are bounded periodic orbits around , and lie on either side of implying that . We now analyze the following function,
The function is essentially the left-hand-side of (4.2) with . The aim is to show that has exactly two roots.
goes to infinity as or . We also have,
Setting the derivative as zero, we obtain that the minimum is unique and is obtained at . Hence, .
Given the structure of , we have that the algebraic equation,
has exactly two roots if the constant, . These roots correspond to and . Clearly, if , the second inequality in (4.8) stands true since nonpositive reals are not in the domain of . In other words, we only need to prove the inequality for or equivalently, when . We achieve this by making use of the sign of the third derivative of ,
Since,
we have that for all , . Hence,
Consequently, for any ,
The second inequality is a result of the fact that . Integrating the obtained inequality with respect to with zero as the lower limit,
Consequently, for the same shift from the function attains a higher value on the left than on the right. Therefore, the two points, , in the level set will be such that
This completes the proof. ∎
We also include a short lemma to derive a different relationship between and which will be helpful in the later sections. To this end, we first define the following useful quantity,
(4.9) |
Lemma 4.5.
Proof.
From Corollary 4.2, we have the all time existence of and strict positivity of . Note that . We can divide this by (4.4b) to obtain,
Upon integrating, we conclude (4.10). Uniform boundedness of follows immediately from the uniform boundedness of . Moreover, since is periodic with period , we have that for any ,
Therefore,
To obtain the last equality, we used the fact that has period . ∎
5. Blow up of solutions
Now we move onto the analysis of (3.8a) and (3.8b). In this section, our aim is to entail conditions under which blow up. From Corollary 4.2, we already know that are uniformly bounded, periodic. Hence, the quantities in (3.8) that can blowup are or .
If , then from (3.8a), as long as exist. From (3.8b),
leading to Ricatti-type blow up of . Consequently, zero density always leads to blowup. Hence, further onwards we will assume . This in turn implies for as long as exist. With this, we move onto the simplification of the system (3.8a), (3.8b). This simplification is along the lines of [28]. From (3.8a), (3.8b), we obtain that,
Noticing that the coefficients of in their respective ODEs are the same, we can multiply the above ODEs by integrating factor to obtain,
Setting
(5.1) |
we obtain a new system,
(5.2a) | |||
(5.2b) |
We label the initial data as .
Remark 5.1.
Owing to Corollary 4.2 and Lemma 4.5, the coefficients in the linear ODE system (5.2) are uniformly bounded. Hence, remain bounded and well-defined for all . Consequently, the key to existence of global solution is to ensure for all . From (5.1), this means that are both bounded for all . Conversely, if there is a finite time, , at which becomes zero, then
Moreover, since is bounded for all times,
Therefore, at the time of breakdown, blow up together.
The above remark results in the following key proposition.
Proposition 5.2.
Next, we have one of the key contributions of this paper in the form of a nonlinear quantity. This quantity will be instrumental in analyzing the system (3.8a),(3.8b). Set,
(5.3) |
Using (3.8c),(3.8d) and (5.2),
(5.4) |
We have,
We assume . As evident in the calculations below, the case has to be handled separately and we will tackle it at the end of the section. Upon integration and setting ,
Finally,
(5.5) |
Here, we have abused the notation and by assigning both of them to two functions. is a function of time and a function of as in (5.5). is the argument of the function as well as a function itself as in (4.9) and takes positive values. It will, however, be clear from context which functions we are referring to.
Remark 5.3.
Next, we state a sufficient condition for taking zero value in finite time.
Proposition 5.4.
If is nonnegative (or nonpositive) for all time, then there exists a time such that .
Proof.
Suppose for all time. At the time where is maximum, from (3.8d) we have and,
From Lemma 4.4, . Hence, . This gives us the existence of a time with such that .
The proof is similar for situation. ∎
Remark 5.5.
At the times where achieves its extrema, is a priori known and essentially depends only on . Remarkably, a simple computation gives at these times. Indeed when achieves max/min (say at ), . Hence, from (5.3), . Moreover, since are periodic with same period , .
Lemma 5.6.
Suppose .
-
•
If , then for all .
-
•
If for all , then for all .
-
•
If for all , then for all .
Proof.
The first assertion follows from the fact that if , then from (5.5), for all time.
Note that if , then it can be readily seen from (5.5) that has exactly two real roots, and
(5.6) |
Moreover, for nonnegative arguments in (5.5), for and for .
Now suppose the hypothesis of the second assertion holds. Straightforward calculations then imply for all . Therefore, for all attainable values of . This proves the second assertion.
The third assertion is similar, only that here the hypothesis implies . Therefore, . ∎
Corollary 5.7.
Remark 5.8.
Note that all the conditions in the hypothesis depend only on the initial data, thereby, will lead to provide a characterization of the supercritical region for (1.2).
Proof.
If the first condition holds, then by the first assertion of Lemma 5.6, for all . Proposition 5.4 leads to the conclusion.
As mentioned, (5.7) is equivalent to . Hence, from Lemma 4.4, the two roots of (5.7) are the maximum and minimum attainable values of . Now suppose the hypothesis of the second assertion holds. There could be two situations:
-
(1)
or,
-
(2)
,
for all . Suppose holds. Using Lemma 4.5,
Hence, for all . Using the second assertion of Lemma 5.6, we obtain that for all time. Then by Proposition 5.4, we conclude the result. Very similar arguments hold for (2) as well. ∎
We complete this section by presenting the case. Upon integrating (5.4) with , we have,
(5.8) |
For the initial data to , we use the same notation, . Proposition 5.4 holds as it is for in place of . Analogous to Lemma 5.6, we have the following result.
Lemma 5.9.
Suppose .
-
•
If for all , then for all .
-
•
If for all , then for all .
The proof is very similar to that of Lemma 5.6 and fairly straightforward given the simplicity of (5.8).
Remark 5.10.
Finally, we have the result analogous to Corollary 5.7.
Corollary 5.11.
Suppose . If the two roots, , of the equation,
(5.9) |
are such that,
then there is a finite time such that . Here, is the constant right-hand-side in (4.7).
Once again, the proof is very similar.
Before we begin the new section, we would like to introduce some notations. will be used to denote a time when are zero respectively. Owing to their periodicity, there are infinitely many such times. In view of this, we will also use the notations for to refer to more than one such times as and when it is required.
6. Global Solution
Up to this point, we have narrowed down the class of possible initial data for (3.8) that could ensure , or equivalently, for all . This section is devoted to precisely identifying the possible initial configurations, from the set remaining from Section 5, that lead to the all-time positivity of . Therefore, from this point onwards, we will assume the negation of the hypothesis in Corollary 5.7. Hence, for this section, we assume the following for ,
(6.1) |
with as in (5.6) and are the minimum and maximum attainable values of , which are also the roots of (5.7). For , we set and being the minimum and maximum attainable values of , which are the roots of (5.9). The first assumption in (6.1) is a vacuous statement for . Under these notational assumptions, the analysis and results of this section for the case is the same as that of with replaced by . Hence, in this section, we do not differentiate between the and cases as there is no critical behaviour with respect to the system (5.2).
Owing to the first assumption in (6.1), has a unique positive root equal to . From (4.10), the second assumption in (6.1) implies,
where ∘ denotes the interior of the set. Since is the root of , from (5.5) we conclude that changes sign. Recall that this is essentially the negation of the hypothesis of Proposition 5.4 and, therefore, must necessarily change sign to hope for a situation wherein for all . See Figure 2 for a visualization of this situation.

From the expression for in (5.5), we see that the leftmost point on the green line in Figure 2 corresponds to the time when attains its minimum and the rightmost point is when it attains the maximum. Using this fact in (4.10) and (5.5), we make an important conclusion that will be quite helpful throughout this section. For a ,
(6.2) |
In most parts of this section, we will be analyzing the system (5.2). Our aim will be to obtain a necessary and sufficient condition to ensure for all . Recall from Remark 5.5 that at the times when (or equivalently ) attains extrema, the values of are a priori known. That is to say, if is such that , then from (5.3),
(6.3) |
In particular, when , we have . From Lemma 4.4, we have . Using this in (6.2), we have . From (6.3), this implies . Likewise, at the time when achieves maximum, and switch signs ensuring that is again positive.
Keeping this in mind, we need to ensure that remains strictly positive at other times apart from the ones where attains its extreme values. To this end, we will construct two functions that form a cloud around . This will enable us to prove the positivity of .
We ensure this as follows. Suppose are given. This fixes the functions , which are unknowns of a closed ODE system, whose properties have been analyzed. Fixing also establishes a linear relation between . Given these three functions, we will derive a condition for exactly one of either or . Each of these conditions will ensure the positivity of . Once a condition on (or ) is imposed, it automatically implies a condition on (or ) through as in (5.3). For example, if , then for given , and using (5.3) we have,
which gives the subsequent bounds on as well. Therefore, it remains to find the appropriate conditions.
From Sections 4 and 5, we have that if the initial data is given, then functions are all known, periodic and uniformly bounded. Then (5.2) is an inhomogeneous, linear system with uniformly bounded coefficients and, therefore, solutions exist for all . We now state and prove a few lemmas.
Lemma 6.1.
Consider the first order linear ODE,
(6.4) |
Let . For all intervals with , suppose and satisfy (6.4) and for some . Then following statements hold,
-
(1)
satisfy
(6.5) for all ,
-
(2)
for all ,
-
(3)
for all ,
-
(4)
for all . In particular, changes sign at only and all .
Proof.
Taking derivative of (6.4), and using (3.8c), (3.8d), (5.4) results in (6.5), which is linear with bounded coefficients. Note that since is discrete, satisfy (6.5) for all by continuity. Hence, the first statement holds.
Consequently, are well-defined for all times. Therefore, the limit in (6.4),
must exist. Hence, for any . The arguments to the fact that are the same as were used in the paragraph following (6.3). This completes the proof to the second assertion.
We will prove the third assertion by contradiction. To that end, assume for some . Note that satisfy (6.4) as well as (6.5). Consider the IVP (6.5) along with initial data . This IVP has a unique solution. Consequently, . However, this is a contradiction since . Hence, the assertion stands.
For , follows by the second assertion and a uniqueness of ODE argument as above. Consequently, if for for sufficiently small, then from the second and third assertions, we have . To see this,
Hence, changes sign at . This completes the proof. ∎
Figure 3 below gives an illustration of the Lemma.

Lemma 6.2.
Suppose satisfies (6.4) on the intervals as specified in Lemma 6.1 and . Let be the first time (if it exists) when . Let be the smallest interval such that and , . could possibly be negative. Let be the unique time when . Then it must be that and for .
In particular, if , then there is no such and for all . Or if , then for all .
Note that from (4.4b) and (6.2), and cannot be zero at the same time. Moreover, from (4.10) and (5.5) there is a unique time with . Therefore, exists and is unique. Also, could be nonpositive. In that case, for .
Proof.
From Lemma 6.1, . Hence, for some small enough such that , we have for . satisfies (6.4) in the interval . Multiplying (6.4) with the appropriate integrating factor, we have,
Note that, and have opposite signs in the interval and same sign in . Indeed, if is such that , then from (4.10) and (5.5), for and strictly positive in . From assertion three of Lemma 4.1, in and hence, signs of are opposite. Same holds if .
Therefore, from the ODE above, the quantity is decreasing in the interval and increasing in the interval . Since , we conclude that . for follows directly from the above ODE.
Lastly, if , then there is no because if it is so, then which is a contradiction. Hence, for all and is positive at by Lemma 6.1. Also, if , then from (6.4) we have and (6.5) then implies that . Hence, is a local minima and is positive in a neighbourhood of . However, on the right, have same sign and the ODE above implies after . ∎
We will now move on to define two functions through an IVP using the linear differential equation (6.5) along with appropriate initial conditions. These functions will form a cloud around the solution, .
Let with be the first two times when . Define the two functions as follows,
(6.6) |
Now we state an important result for these functions as well as .
Proposition 6.3.
Proof.
Part 1: The homogeneous system.
To prove this, we will use the Floquet Theorem, [10]. To use the result, we will first study the homogeneous form of (5.2) as follows,
(6.13) |
Let be the period of and therefore, the period of the matrix . By Floquet Theorem, a fundamental matrix to (6.13) is of the form , where is a -periodic, matrix which is nonsingular for all times and is a constant matrix. could possibly be complex but if one of them is real, then the other will be since is real.
The substitution reduces (6.13) to
(6.14) |
To see this, note that
Also, since is a fundamental matrix of (6.13), we have,
Substituting this for above and multiplying by , we obtain (6.14).
Next, we define a quantity for (6.13) analogous to (5.3). Define . Using (6.13), (3.8c), (3.8d),
Therefore, . We used (4.10) for the last equality. If is a time where (or equivalently ) achieves maximum, then for ,
(6.15) |
Since is a constant matrix, the general solution form of (6.14) is known. Consequently, from (6.14) and that , we obtain that a general solution to (6.13) depending on whether the eigenvalues of are distinct or repeated,
or,
where are the eigenvalues (possibly complex) of , are linearly independent, periodic vectors and are arbitrary constants. We will show that is periodic. Suppose the eigenvalues are repeated and latter formula is the general solution, then we have,
where are the first elements of respectively. This should be a general solution form for any constant . Using (6.15),
Assume . If Re, then left-hand-side tends to zero as but right-hand-side does not. If Re, then left-hand-side is an exponential function of while the right-hand-side is linear and therefore, the equation cannot hold for all . Lastly, if is purely imaginary, then the left-hand-side changes periodically with but the right-hand-side does not. This leaves us with the only possibility that the repeated eigenvalue must be zero. In this case,
However, note that since and the vectors are periodic, it must be that,
for some periodic function . This could only hold if , which is a contradiction to being nonsingular. Consequently,
for . Again, using (6.15),
It is clear that the real part of both eigenvalues is zero because if not, then a contradiction is obtained as . Consequently, are purely imaginary. Taking difference with in (6.15) with itself,
Since are arbitrary,
As argued before, both cannot be zero together. So, at least one of the . In fact, both these terms are zero, because if (WLOG), then , in which case is no longer arbitrary, hence, a contradiction. Therefore,
where are new periodic functions obtained after dissolving the exponential terms, which had the same period.
Part 2: Periodicity.
Owing to the fact that , we have eventually proved that a fundamental matrix to (6.13), is of the form,
We will now show that is periodic. Very similar arguments apply for as well as any solution to (5.2a), . The important fact is that all of these three functions satisfy (6.4) in the intervals as mentioned in the hypothesis of Lemma 6.1 and hence, the assertions of the Lemma hold.
has the following closed form expression,
Through routine calculations, one can check that in the above expression indeed satisfies (6.6) for . We can further evaluate this expression as follows,
(6.20) | ||||
(6.25) | ||||
(6.30) | ||||
(6.33) |
where are T-periodic functions defined by,
Periodicity of follows from (4.10) as also has period . Since is a solution to a ODE system, proving periodicity is equivalent to proving equality at a single point. In particular, if we prove the following for some , then we obtain periodicity.
From (6.20) and periodicity of ,
Set
If , then is identically zero. Indeed, if are both zero together at the same time then from nonsingularity of , we have that , and hence, , thereby proving that is periodic. Our aim is to find such a . We begin with some calculations. For and , and owing to periodicity of ,
Using this, we can obtain,
(6.34) |
Similar calculation for leads to
(6.35) |
Note that from the second assertion of Lemma 6.1, we have , where is such that . From (6.34), . However, may not be equal to zero. In a similar way, observing (6.4) and applying it to (6.35), we also have that , where is such that , whereas may not be zero. The key is to find a such that both are zero.
We will consider two cases. The first is when and are zero at the same times and the other when they are never zero together. The two situations are exhaustive, in the sense that there can never be a scenario when are zero together at some times and there is the existence of some other times when one of them is zero and the other is not. sharing zeros is only dependent on the initial data . Indeed from (4.10), is zero at the times when and from (5.5), is zero at times . Hence, are zero at the same times if and only if , which only depends on the . Conversely, if , then there is no time when are both zero.
Taking note of this discussion, we first prove periodicity of for the situation when for any . Since is periodic, it is enough to consider the restricted function , where is such that . From second assertion of Lemma 6.1 and (6.34), we have , where such that . In a similar way, observing (6.4) and applying it to (6.35), we also have that , where are the two unique times in the interval where . Now suppose there is a point such that . Then firstly from (6.35), . Secondly, from (6.4) and (6.34),
Therefore, . Consequently, . Hence, if is not identically zero, then it must be that it is, without loss of generality, strictly increasing in the intervals and strictly decreasing in . This implies that has exactly one maximum and it occurs at . Therefore, from (6.34), we have the existence of a natural number large enough so that,
implying that for all . However, this is a contradiction since from (6.4),
Therefore, is identically zero.
Now we prove periodicity for the scenario when are zero at the same times. Firstly, we argue that being uniformly bounded implies periodicity. This can be seen by plugging in a special sequence of times in (6.34) and (6.35). To this end, let be a sequence of times such that . We obtained this sequence by applying Rolle’s Theorem to in the intervals . Rewrite as,
By compactness, there exists a convergent subsequence . Let as . Plugging in in (6.35), we obtain,
Dividing by and letting , we have that
Similarly using (6.34),
If is uniformly bounded in time, then dividing by and letting results in,
As a result of this, it is enough to show that if are zero at the same times, then is uniformly bounded. Without loss of generality, assume it is not bounded above and attains a positive maximum value in the interval . Let,
Let be a time when . By continuous dependence of ODE solutions on initial data, we can choose a solution to (6.5), , with initial data such that,
and small enough so that,
However, by previous discussion, is periodic. Therefore, it must be that for a ,
hence, a contradiction. This completes the full proof. ∎


Proposition 6.4.
For each ,
In particular, and ’s are distinct.
Proof.
We will show that for any satisfying the condition (6.1), there exists a strictly positive solution to (6.4). The proposition statement can then be proved through the following argument. From Lemma 6.2, for . Hence, it can only be zero in . If just touches zero in , then by (6.4), that time must be and by (6.5) and uniqueness of ODE, . If crosses zero, then by Lemma 6.2, must be in between the two roots. Similar statement holds for as well. Both these cases are illustrated in Figure 4. In any case, Lemma 6.1 implies that there can never be a strictly positive solution. Hence, it must be that for each is positive everywhere on except at .
In view of the above discussion, we prove the statement mentioned at the beginning of the proof. Let,
Claim 1: is non-empty.
We check that is a positive solution to (6.5). is positive since is. Note from (4.4),
The last equality follows from (4.10). Consequently belongs to .
Claim 2: is open.
If , then for , (6.4) has a strictly positive solution. By continuous dependence of solutions on initial data, there is a neighbourhood around for which there is strictly positive solution.
Claim 3: is closed in the set where (6.1) holds.
Let be a sequence with and belongs to the set where (6.1) holds. By well-posedness of ODEs, there must be solution to (6.4) for satisfying . If is strictly positive then . If not, then from Lemma 6.2 and (6.4), it is positive everywhere except, or or both. If for both , then by uniqueness, is -periodic. Note that , the period of . If this is true, then by (6.5),
for all . Since , we can obtain an interval on which a closed form of is obtained given by above formula. This is a contradiction as it can be readily checked that it does not satisfy (6.5).
Therefore, (WLOG) and , . But by uniqueness, . By definition of and Lemma 6.1, it must be such that , . However, if this is the case, then from Lemma 6.1 there has to be a strictly positive solution to (6.4) squeezed between and . Hence, and consequently, is closed.
This finishes the proof. ∎
The key results are as follows,
Proposition 6.5.
Proof.
Firstly, note that satisfy (6.4) along with . Indeed, if a function satisfies (6.4) with , then
and hence, by first assertion of Lemma 6.1 and uniqueness of ODE, . Similarly for . Also, the solution to (5.2a), , satisfies (6.4). This is true directly from the definition of as in (5.3) and (5.2a). Consequently, all the three functions, pairwise satisfy hypothesis of Lemma 6.1 with . The result follows from Lemma 6.1. Indeed at each , all cross each other, hence, maintaining that will be contained in between if and only if it was so initially. ∎
An illustration of the situation is provided in Figure 5.

Proposition 6.6.
Suppose . If
then for any ,
Conversely, if
then
Proof.
Corollary 6.7.
for all if and only if one of the following holds,
-
•
If then,
-
•
If then,
Proof.
Using the results developed above, we move on to proving Theorem 2.1.
Proof of Theorem 2.1: We will prove for as is very similar. Suppose initial data satisfies the hypothesis of the Theorem. Along the characteristic path (3.1), this translates to the condition that for all ,
We will now analyze a single characteristic path and replace the initial data notations with . Under the transformation (3.7), we now turn to the unknowns of the ODE system (3.8), with initial data . Global-in-time existence of these variables is equivalent to the global-in-time existence of the original variables. Note that if , then as argued, through simple calculations, at the beginning of Section 5, there is blowup of density, . Hence, we can safely assume . Using the transformations (5.1), we work with the unknowns with initial data . Through (3.7) and (5.1), we see that indeed . Set to obtain as in (5.3) satisfying (5.4). Turning to the Definition 2.2, we use (4.3) and as in (5.7) to get,
Rearranging this result in condition (6.1). Also, if and only if . In (2.2), this is equivalent to whether is zero or not. If (equivalently ), then through the transformation (5.1), (2.2) is equivalent to,
which is exactly the hypothesis of Proposition 6.5. On the other hand, if (equivalently ), then using (3.7) and (5.1), (2.2) reduces to
From and (5.3), the above inclusion becomes,
which is exactly the hypothesis of Proposition 6.6. Note that on a single characteristic path, are known functions because are fixed. In particular, can first be evaluated and then it can be checked that one of the above hypothesis is satisfied depending on whether or not. By Corollary 6.7, we obtain the all-time positivity of . By Proposition 5.2, all unknowns in (3.8) exist for all time. Since the above analysis holds for all characteristic paths, then by Lemma 3.1 and Theorem 1.1 we have global-in-time solution to (1.2).
Conversely, suppose there is a characteristic path corresponding to some parameter such that,
Without loss of generality, we assume . Then there could be two situations. Either the inclusion in Definition 2.2 does not hold or the inclusion holds but violates (2.2). Suppose the first case is true. Similar to how we argued above, we have,
Using (5.6) in the above expression, and subsequently with the help of Corollary 5.7, we obtain that there is blow up of density. On the other hand suppose the above inclusion holds and (2.2) does not. Then very similar to the way as we argued for the global existence result, we have that for ,
and for ,
From Corollary 6.7, we obtain that there is finite time with . From Proposition 5.2, there is breakdown at and the solution ceases to be smooth. This completes the proof of the Theorem. ∎
7. The zero background case
The zero background case has been analyzed by several researchers. Most notably, the authors in [29] give a sharp threshold condition, however, assuming that the flow is expanding (). A more refined analysis was done by the author in [28], however, the threshold condition was not sharp. In this section, we present a sharp characterization of the subcritical and supercritical regions for general velocity. To this end, we consider (3.8) with ,
(7.1a) | |||
(7.1b) | |||
(7.1c) | |||
(7.1d) |
with the same notation for initial data. Also recall the system (5.2). With zero background, it reduces to,
(7.2a) | |||
(7.2b) |
We will make use of the same quantity as in (5.3). Using similar computations, one finds that the expression for is exactly the same as in (5.5) for . Also for , the expression for the corresponding quantity is the same as in (5.8).
A robust analysis of the system in this case has been carried out by the author in [28]. We only state the important results which will be used directly.
Proposition 7.1 ([28][Theorem 3.5, Lemma 3.4, 3.7).
Once again, the Poisson forcing is enough to avoid concentrations at the origin. In particular, Corollary 4.2 holds and . We state it as a Lemma below.
We now move on to some key results. We first give a sufficient condition for achieving zero in finite time.
Proposition 7.3.
Suppose . If
then there is a time such that .
Proof.
Since satisfies (6.4), we have,
(7.4) |
We will use the convergence estimates from Proposition 7.1. Suppose is a time so that the convergence rates are valid for all . Throughout our calculations are constants that may change from step to step but only depend on . From (7.3), we have for ,
(7.5) |
Integrating (7.4), we have,
From (7.5), I is uniformly bounded. Owing to (5.5) and (4.10), we can find convergence rates for as well. Using (7.3b) in (5.5) along with (4.10), we have for ,
Consequently, from (7.3a),
Since are positive and is negative, we have for ,
II | |||
Finally, for , we have,
Therefore, there exists a such that . This completes the proof. ∎
We now move on to characterizing the subcritical region. We first define a function through an IVP using (6.4),
(7.6) |
where is a time (if it exists) such that .
Proposition 7.4.
Consider the case. Suppose . Given , we have that for all if one of the following conditions is satisfied,
-
•
,
-
•
and .
Additionally, if and , then there is a time such that .
Proof.
Firstly, note that the hypothesis and Proposition 7.1 imply are positive for all times. Hence, from (7.1d), is monotonically decreasing to zero. Owing to (4.10), is strictly decreasing and tends to zero as . Moreover, if , then,
where is the root of as in (5.6). From (5.5), for all . Consequently, from (7.4), can never be zero in finite time if it was initially positive.
Now suppose the second hypothesis holds. Here, we have that . Owing to the properties of as listed above, we obtain that there is a unique time such that . Moreover, for and for . Owing to (7.4) once again, remains greater than zero if it is so at . In fact, serves as a lower bound for at each time, see Figure 6. Since satisfy the same first order ODE we have a comparison principle. Indeed on taking difference, we obtain,
From (7.6), we have . Also, upon taking derivative of (7.6), one can check that and, therefore, is indeed the unique time where the minimum of is attained. Hence, for all . Conversely, if , then since remains below , there is a time , such that . ∎

Proposition 7.5.
Consider the case. Suppose and . Given , we have that for all if and only if .
Note that is in the limit sense in (7.6) since . We know that this limit exists because satisfies (6.5) (with ), which is an inhomogeneous second order linear ODE with bounded coefficients. Also, the assumption implies .
Proof.
Now we present the result for . We take derivative of (7.6) and state two second order IVPs for ,
(7.7) |
where is such that . There need not be two such times. In that case, we only consider . Note that by uniqueness, in (7.6) is the same function as above. Also, similar to (4.2), we obtain the trajectory for this case as,
(7.8) |
From this, one can directly note the maximum attained value of , which is when , is
(7.9) |
Proposition 7.6.
Consider the case. Suppose . Given , we have the following.
-
•
Suppose . Then for all if and only if and .
-
•
Suppose . Then for all if and only if .
Proof.
Suppose . From the first assertion of Proposition 7.1 and (5.3), we conclude that at ,
Therefore, a necessary condition for to be positive is that . From (4.10) and (5.5), the condition is equivalent to,
Here, we must keep in mind that the dynamics of of is such that increases until then decreases monotonically and approaches zero as . If the above inequality holds, then there are two positive times, , when . Also, . The very same arguments as in Proposition 6.4 allow us to conclude that for . Since ’s also satisfy (7.6), the all-time-positivity of is guaranteed once again by arguments as in Lemma 6.1 if it was in between and at . In particular, assertions 1-4 of Lemma 6.1 are valid with a slight modification that the set has only one element, see Figure 7 for a visualization. Conversely, if , then from Lemma 6.1 becomes zero in finite time.
If , then . From Proposition 7.1, increases to a maximum and then decreases to zero. Once again, making use of the relation (4.10) and (5.5), we have that is zero only once, at . From (7.6), we have,
We conclude that for , since are both negative in this interval. Since from Lemma 6.1, for , we conclude that for all since is nonnegative and serves as a lower bound for in this domain. Conversely, if , then for all and hence, it must be zero in a finite time, , since . ∎

Proposition 7.7.
Consider the case. Suppose be given. If
then there is a time such that .
We state this proposition and its proof separately because the technique used in the proof of Proposition 7.3 does not apply. In fact, it turns out that it is much easier to work in the variables instead of the variables.
Proof of Proposition 7.7: Using (5.3) and in the expression of as in (5.5), we obtain,
Substituting for using (5.1), we obtain,
As a result, we can find in terms of the other variables as,
We can divide by because from (7.3a), is eventually positive and decays accordingly. We will consider sufficiently large times so that this holds. Plugging this in the ODE of , (7.1a), we get,
For all sufficiently large times, the rates in Proposition 7.1 hold. We can assume has not already blown up, because if it has then the proof is done. Using these rates, we conclude that , are bounded. Therefore, there occurs a Riccati-type blow up of density, and the blow up is aggravated by the in the denominator.
∎
Propositions 7.3, 7.4, 7.5, 7.6 and 7.7 enable us to put together the picture for . Now, we move onto the critical case . We will omit the repetitive parts in the proofs of these results. Note the quantity in (5.8) analogous to . Also note that there always exist a positive root of , , no matter the sign of , which is unlike the case for . could lie on or either side of depending on the sign of . For the case, we will consider the function as in (7.6) and the functions ’s as in (7.7) with replaced by in the definition.
Proposition 7.8.
Suppose and consider as in (7.6). Suppose . Given , we have that for all if one of the following conditions is satisfied,
-
•
,
-
•
and .
Additionally, if and , then there is a time such that .
The proof is very similar to that of Proposition 7.4.
Proposition 7.9.
Consider the case. Suppose . Given , we have that for all if and only if .
The proof is very similar to that of Proposition 7.5.
Proposition 7.10.
Consider the case. Suppose . Given , we have the following.
-
•
Suppose . Then for all if and only if and .
-
•
Suppose . Then for all if and only if .
Once again, the proof is very similar to that of Proposition 7.6. The expression for is different from since the trajectory equation in the critical case is different. Similar to the way we obtained (7.8), we can obtain the trajectory for the critical case ,
(7.10) |
where . From this, one concludes,
Finally, we analyze the case when the initial density is zero, that is, . Firstly, from (7.1a), is equivalent to , as long as in (7.1b) exists. Therefore, as long as exists, (7.1b) reduces to,
(7.11) |
Proposition 7.11.
Proof.
Consider the quantity . From (7.11), (7.1c) and (7.1d), we have,
Consequently, as long as exists, maintains sign and we have,
(7.12) |
Suppose and . From the first assertion of Proposition 7.1, we have that for all time. For the sake of contradiction, suppose that breaks down in finite time. Since, is uniformly bounded, it is clear from (7.11) that at the time of breakdown, . Since are uniformly bounded quantities, it must be that at a certain time before breakdown, the left-hand-side in (7.12) is negative but the right-hand-side is non-negative. This is a contradiction and hence, is finite for all times. In particular, it decays with a lower bound, , the decay rate for which can be obtained directly from Proposition 7.1.
Now suppose but . We first assume . The critical case when needs separate treatment in this regard and will be analyzed at the end. Note that if has not broken down, the following holds,
From Proposition 7.1, the integral on the left-hand-side is well-defined. The second inequality is a direct result of (7.12). The first inequality can be shown as follows. Suppose for the sake of contradiction, there is a such that . Then using (7.11), we obtain that for some ,
Consequently, from (7.11), we conclude that a Riccati-type blowup occurs and at some time greater than .
Therefore, we can assume that for all . Choose large enough so that the convergence estimates of Proposition 7.1 hold. Using these estimates, we can rewrite the above bounds on as follows,
(7.13) |
are appropriate positive constants whose values may change through the proof but they depend only on . Once again from Proposition 7.1, and (7.13), we obtain
Now we analyze the right-hand-side of (7.12).
We used Lemma 4.5 to obtain the equality. We used (7.13) to conclude that is integrable and hence, the inequality holds. Combining the above two inequalities for the two sides of the equation (7.12), we have for sufficiently large times,
which is a contradiction. Therefore, must blow up in finite time, that is, for some .
Now suppose . Firstly, note that because if not, then a Riccati-type blowup occurs. This implies
Hence, for all time. Now from Proposition 7.1, we know that after a sufficiently large time , and follow the convergence rates. Hence, the same arguments as above apply which lead to a finite-time-breakdown of .
We now analyze the case. Just as for case, we need to prove a contradiction when and . All the other arguments are the same. To this end, we assume and . Assuming exists for all times, we have the following,
For ( such that the rates of Proposition 7.1 hold), we have
(7.14) |
for all . We focus on the integral above. A substitution changes it into,
This expression can be represented using a well-known exponential integral function given by
Moreover, it has the following bounds, see [1, Page 229],
Using the above lower bound in (7.14), we obtain the following bounds,
(7.15) |
Therefore, we have,
for all times sufficiently large. On the other hand, if we analyze the right-hand-side of (7.12), we see that for ,
Once again, this contradicts (7.12) since for sufficiently large times,
This completes the proof. ∎
Proof of Theorem 2.4: Suppose the hypothesis holds. Then by using (5.1) in Proposition 7.3, one immediately obtains the blow up of density in finite time if .
If , then the result follows from Proposition 7.7. ∎
Proof of Theorem 2.6: Suppose initial data satisfies the hypothesis of the Theorem. Along the characteristic path (3.1), this translates to the condition that for all ,
We will now analyze a single characteristic path and replace the initial data notations with . Under the transformation (3.7), we now turn to the unknowns of system (7.1), . Global-in-time existence of these variables is equivalent to the global-in-time existence of the original variables. If , then Proposition 7.11 gives the all-time existence of and hence, the solution is global.
Next, we prove global existence for the case when . Turning to the Definition 2.7, we will use the equivalence of and (5.3), and analyze the conditions (2.4), (2.5), (2.6), (2.7) one by one. To this end, let it be such that,
We thus have fulfilment of (2.4). Rearranging (2.4), and noting the transformations (3.7), (5.1), we obtain for ,
and for ,
Also, if and only if . The first inequality then is the hypothesis to the second assertion of Proposition 7.4 and the second assertion of Proposition 7.6 (depending on whether or ). The second inequality is the hypothesis to Proposition 7.5. As a result, through (5.1), we obtain the all time existence of in (7.1a), (7.1b).
Next, we suppose that . Then (2.5) reduces to for and there are no extra conditions if . Note that cannot be equal to zero because that would be a violation to . These two scenarios form the hypothesis of the first assertion of Proposition 7.4 and the first assertion of Proposition 7.6. Therefore, we have the all time existence of .
Now, suppose
Then using (7.9) with and rearranging this gives,
with as in (5.6). Note that , which in the definition of initially, was not given explicitly is now given explicitly by (7.9). Clearly it only depends on . Also note that since , directly from its formula, (5.6). From (2.6), we obtain,
if , and no extra conditions if . Note that cannot be equal to zero because that would be a violation to . Therefore, all time existence of follows from the first assertion of Proposition 7.6 (if ) and first assertion of Proposition 7.4 (if ).
Lastly, we assume
Similar to how we argued in the previous case, here the above inequality is equivalent to saying that . The conditions (2.7) imply that . We can then apply the first assertion of Proposition 7.4 to obtain the all time existence of .
Putting all the cases together and applying Proposition 5.2, we have obtained that solutions to (7.1) exist for all time if any single characteristic lies in any of the above situations. In particular, if all the characteristics satisfy the above conditions, then an application of Lemma 3.1 and Theorem 1.1 gives the existence of global-in-time solutions to (1.2) with .
Conversely, suppose there is a characteristic path corresponding to some parameter such that,
If , then a direct application of Proposition 7.11 gives the finite-time-breakdown of .
Now, we suppose . Then it could be that . Finite time breakdown is then a direct result of Propositions 7.3 or 7.7.
If , then negation of one of the conditions among (2.4), (2.5), (2.6), (2.7) has to be true as and according to the value of . Once again, we can check each condition one by one. All the analysis is a repetition of above except from the fact that, instead of the all-time existence results, we use the finite-time-breakdown results of Propositions 7.4, 7.5 and 7.6. Consequently, solutions to (1.2) cease to be smooth.
This completes the proof of the Theorem. ∎
8. Conclusion
The techniques developed in this paper work for the one-dimensional case as well. However, there is one main difference in the 1D and multi-dimensional scenarios. As pointed out by the author in [28], that in multi-D, the Poisson forcing is enough to avoid flow concentration at the origin. In particular, by Corollary 4.2, we know that no matter how large (in absolute value) the initial velocity is, there are no concentrations at the origin. Let’s do similar calculations for the 1D case. In 1D, system (3.8c), (3.8d) for reduce to,
with . Using a transformation, , we obtain a simple linear ODE system,
We can analytically solve this to obtain,
From this, one can conclude that (or equivalently, ) for some if
which in the original variables is equivalent to,
Hence, for a large initial velocity (absolute value), there are concentrations at the origin. For , the result is one sided, that is, a sufficiently large negative initial velocity (flow pointing towards origin) would lead to concentrations at the origin. This concentration is completely avoided for , wherein the Poisson forcing turns out to be sufficient.
In this work, we make a subsequent important discovery that, in multi-D, the subcritical region can contain arbitrarily large initial velocities. In other words, no matter how large the initial velocity is (positive or negative), one can have a region in the phase plane of initial density and gradient of velocity, that corresponds to all-time-existence of the solution.
Moreover, we are hopeful that our techniques can be applied to other systems such as Euler-Poisson-alignment and Euler-Poisson with swirl. However, we leave that for future studies.
Acknowledgments
This research was partially supported by the National Science Foundation under Grant DMS1812666.
References
- [1] M. Abramowitz, I.A. Stegun. Handbook of mathematical functions. National Bureau of Standards, Appl. Math. Series, Vol. 55, 1964.
- [2] M. Bhatnagar and H. Liu. Critical thresholds in one-dimensional damped Euler-Poisson systems. Math. Mod. Meth. Appl. Sci., 30(5): 891–916, 2020.
- [3] M. Bhatnagar and H. Liu. Critical thresholds in 1D pressureless Euler-Poisson systems with variable background. Physica D: Nonlinear Phenomena, 414: 132728, 2020.
- [4] J. A. Carrillo, Y.-P. Choi, E. Tadmor, and C. Tan. Critical thresholds in 1D Euler equations with non-local forces. Math. Mod. Meth. Appl. Sci., 26:185–206, 2016.
- [5] J.A. Carrillo, Y.P. Choi, E. Zatorska. On the pressureless damped Euler-Poisson equations with quadratic confinement. Math. Mod. Meth. Appl. Sci., 26: 2311-2340, 2016.
- [6] D. Chae and E. Tadmor. On the finite time blow-up of the Euler-Poisson equations in . Commun. Math. Sci., 6(3): 785–789, 2008.
- [7] B. Cheng and E. Tadmor. An improved local blow-up condition for Euler–Poisson equations with attractive forcing. Physica D, 238: 2062–2066, 2009.
- [8] T. Do, A. Kiselev, L. Ryzhik, C. Tan. Global regularity for the fractional Euler alignment system. Arch. Rat. Mech. Anal., 228(1): 1–37, 2018.
- [9] S. Engelberg, H. Liu, E. Tadmor. Critical thresholds in Euler-Poisson equations. Indiana University Math. Journal, 50:109–157, 2001.
- [10] G. Floquet. Sur les équations différentielles linéaires à coefficients périodiques. Annales Scientifiques de l’École Normale Supérieure, 12: 47–88, 1883.
- [11] L. Hartman. Accretion processes in star formation. Cambridge University Press, 32, 2009.
- [12] S. He, E. Tadmor. Global regularity of two-dimensional flocking hydrodynamics. C. R. Math., 355(7): 795–805, 2017.
- [13] J.D. Jackson. Classical electrodynamics. Wiley, 1975.
- [14] A. Kiselev, C. Tan. Global regularity for 1D Eulerian dynamics with singular interaction forces. SIAM J. Math. Anal., 50(6):6208–6229, 2018.
- [15] P. Lax. Development of singularities of solutions of nonlinear hyperbolic partial differential equations. Journal of Math. Phys., 5, 611, 1964.
- [16] Y. Lee and H. Liu. Thresholds in three-dimensional restricted Euler-Poisson equations. Physica D: Nonlinear Phenomena, 262: 59–70, 2013.
- [17] T. Li, H. Liu. Critical Thresholds in a relaxation model for traffic flows. Indian Univ. Math. J., 57:1409–1431, 2008.
- [18] T. Li, H. Liu. Critical thresholds in hyperbolic relaxation systems. J. Diff. Eqns., 247:33–48, 2009.
- [19] H. Liu, E. Tadmor. Critical thresholds in a convolution model for nonlinear conservation laws. SIAM J. Math. Anal., 33(4): 930–945, 2001.
- [20] H. Liu and E. Tadmor. Spectral dynamics of the velocity gradient field in restricted flows. Commun. Math. Phys., 228: 435–466, 2002.
- [21] H. Liu, E. Tadmor. Critical Thresholds in 2-D restricted Euler-Poisson equations. SIAM J. Appl. Math., 63(6):1889–1910, 2003.
- [22] H. Liu, E. Tadmor. Rotation prevents finite-time breakdown. Physica D, 188:262–276, 2004.
- [23] A. Majda. Compressible Fluid Flow and Systems of Conservation Laws in Several Space Variables. Springer Science+Business Media, Vol. 53, 1984.
- [24] E. Tadmor, C. Tan. Critical thresholds in flocking hydrodynamics with non-local alignment. Phil. Trans. R. Soc. A., 372: 20130401, 2014.
- [25] E. Tadmor, C. Tan. Critical threshold for global regularity of Euler-Monge-Ampère system with radial symmetry. SIAM J. Math. Anal., 54(4): 4277–4296, 2022.
- [26] E. Tadmor, D. Wei. On the global regularity of subcritical Euler-Poisson equations with pressure. J. Eur. Math. Soc., 10:757–769, 2008.
- [27] C. Tan. On the Euler-alignment system with weakly singular communication weights. Nonlinearity, 33(4): 1907–1924, 2020.
- [28] C. Tan. Eulerian dynamics in multi-dimensions with radial symmetry. SIAM J. Math. Anal., 53(3): 3040–3071, 2021.
- [29] D. Wei, E. Tadmor, H. Bae. Critical Thresholds in multi-dimensional Euler-Poisson equations with radial symmetry. Commun. Math. Sci., 10(1):75–86, 2012.
- [30] M. Yuen. Blowup for the Euler and Euler-Poisson equations with repulsive forces. Nonlinear Analysis 74:1465–1470, 2011.