lemthmm \aliascntresetthelem \newaliascntthmthmm \aliascntresetthethm \newaliascntcorthmm \aliascntresetthecor \newaliascntpropthmm \aliascntresettheprop \newaliascntdefithmm \aliascntresetthedefi \newaliascntremthmm \aliascntresettherem
Doubly stochastic Yule cascades (Part II): The explosion problem in the non-reversible case
Abstract
We analyze the explosion problem for a class of stochastic models introduced in [part1_2021], referred to as doubly stochastic Yule cascades. These models arise naturally in the construction of solutions to evolutionary PDEs as well as in purely probabilistic first passage percolation phenomena having a Markov-type statistical dependence, new for this context. Using cut-set arguments and a greedy algorithm, we respectively establish criteria for non-explosion and explosion without requiring the time-reversibility of the underlying branching Markov chain (a condition required in [part1_2021]). Notable applications include the explosion of the self-similar cascade of the Navier-Stokes equations in dimension and non-explosion in dimensions .
Keywords. Yule cascade, doubly stochastic Yule cascade, stochastic explosion, Navier-Stokes equations, KPP equation.
AMS subject classification. 60H30, 60J80.
1 Introduction
Largely motivated by the probabilistic method for the deterministic three-dimensional incompressible Navier-Stokes equations (NSE) introduced by Le Jan and Sznitman [lejan], relaxed in [chaos], the authors introduced a new class of stochastic branching models, referred to as doubly stochastic Yule cascades, in [part1_2021]. This class of branching models may also be viewed in the context of statistically dependent first passage percolation models on tree graphs, or in the context of certain cellular aging models in biology [DA_PS_1988, KB_PP_2010]. Roughly speaking, a doubly stochastic Yule cascade evolves from a single progenitor in which each particle waits for a random length of time depending on the particle’s position on the genealogical tree before giving birth to a new generation of particles, each evolving by the same rules as its parent. If the lengths of times between particle reproductions are sufficiently short then branching may produce infinitely many particles in a finite time, an event referred to as stochastic explosion. In this paper, we establish some sufficient conditions for explosion and non-explosion, and apply them for the doubly stochastic Yule cascades associated with several well-known differential equations and probabilistic models.
We begin by recalling, in the context of binary trees, some of the definitions introduced in [part1_2021] which contains more general definitions. With standard notations, debote by the indexed binary tree with the root .
Definition \thedefi (Yule cascade).
A tree-indexed family of independent random variables is said to be a Yule cascade with positive parameters (intensities) if for every .
In the above, we used the standard notation to denote that is exponentially distributed with intensity . If is a constant, this is the familiar Yule cascade with intensity . The special case is referred to as the standard Yule cascade.
Definition \thedefi (Doubly stochastic Yule cascade).
Consider a tree-indexed family of random variables and a tree-indexed family of positive random variables . is said to be a doubly stochastic Yule (DSY) cascade with random intensities if is, conditionally given , a Yule cascade with intensities .
Equivalently, is a DSY cascade with random intensities if and only if is a family of i.i.d. mean one exponentially distributed random variables (called holding times or clocks) independent of . Thus, a DSY cascade is completely specified once the family of random intensities is specified. For this reason, one can write a DSY cascade as . The associated counting process defined by
is referred to as the Yule process or doubly stochastic Yule process depending on the respective cascade model being considered. In the definition of , we have used standard notations for the length of a vertex and the truncation of of length , namely and for . By convention,
Definition \thedefi (Explosion time).
The explosion time of a DSY cascade is a -valued random variable defined by
The event of non-explosion is defined by . The cascade is said to be non-explosive if , and explosive if .
According to the definition, the explosion event has a positive probability in an explosive cascade. However, in all applications considered in this paper, explosion is in fact a event (Section 2). The DSY cascades introduced in Section 1 are quite general. To consider the explosion problem, we will further assume a certain branching Markov chain structure underlying the random intensities .
Using estimates on the spectral radius of an operator defined on the underlying branching Markov chains, the authors obtained in [part1_2021] criteria for non-explosion assuming a time-reversibility condition. These conditions were applied to the Bessel cascade, a particular cascade associated with the Navier-Stokes equations, and to the cascade of the Kolmogorov-Petrovski-Piskunov (KPP) equation in Fourier domain to show that they are both non-explosive. Other more purely probabilistic models were also analyzed there. However, time reversibility is too restrictive and does not apply to several examples of particular importance. These include the cascades induced by the scaling properties of the NSE, referred here as self-similar cascades associated with the NSE. The focus of this paper is to establish explosion and non-explosion criteria for a class of DSY cascades using probabilistic arguments that do not depend on the time-reversibility.
To establish a new non-explosion criterion (Section 2), we rely on the sub-criticality of nonhomogeneous Galton-Watson processes associated with the cascade, which is more intuitive and robust than the method of large deviation estimates used in [part1_2021]. The key idea to obtain non-explosion is to construct a sequence of recurring finite cut-sets through an inspection process. The construction of these cut-sets takes into consideration the uncountable set of paths in a binary tree, which involves a new notion of recurrence that naturally generalizes neighborhood recurrence along each path (Section 2). As an application, Section 2 implies the non-explosion of the self-similar cascade associated with NSE in dimensions and recovers the non-explosion result for the Bessel cascade of the 3-dimensional NSE and the cascade associated with the KPP equation proven in [part1_2021].
Remark \therem.
In spite of its intuitive appeal, it is the Markov dependence along tree paths, in place of independence, that makes the use of cut-sets technically challenging here. Similar challenges arise in computing the speeds of branching random walk particles with Markov-chain displacement paths, in place of independent displacements, and in computing the first passage percolation “flow”problems [HK_1986, GG_HK1984] if the i.i.d. passage times along paths are statistically dependent.
We establish a criterion for a.s. explosion (Section 2) based on a “fastest path” constructed by a greedy algorithm: at each time of branching, the fastest path chooses to follow the descendant vertex with the larger intensity . This criterion allows one to show the explosion of a cascade even if the explosion does not occur along any deterministic path , i.e.
As an application, we obtain the a.s. explosion of the self-similar cascade of the 3-dimensional Navier-Stokes equations. As shown in [athreya, alphariccati, smallness], the stochastic explosion of DSY cascades associated with certain evolution equations leads to both nonuniqueness and finite-time blowup of solutions to the initial-value problems in appropriate settings. In particular, the aforementioned explosion of the self-similar DSY cascade corresponding to the Navier-Stokes equations leads to a non-uniqueness result for an equation introduced by Montgomery-Smith [smith] as a model for possible Navier-Stokes finite-time blowup. Remarkably, our DSY framework also provides an alternative way to prove finite-time blowup of the smooth solutions to the Montgomery-Smith equations [smallness]. For the actual Navier-Stokes equations, a possible resolution of the outstanding nonuniqueness and finite-time blowup problems hinges on a better understanding of the connections between geometric structure of the nonlinear term and the branching structure of the associated DSY cascade.
2 Main results and organization of the paper
Motivated by the PDE and probabilistic models mentioned above, we are particularly interested in DSY cascades in which the random intensities are of the form where, for each , is a random variable taking values in a measurable space and is a measurable function. Throughout the paper, we assume the following two properties:
-
(A)
For any path , the sequence , , ,…is a time homogeneous Markov chain.
-
(B)
For any path , the stationary transition probability does not depend on .
In [part1_2021], DSY cascades satisfying conditions (A) and (B) are called DSY cascades of type (). We will refer to the family as the underlying branching Markov chain of the DSY cascade.
In the statements of the main results, the following conditions are sometimes required. We use the standard notation for concatenation of vertices .
-
(C)
For each , the two subfamilies and are conditionally independent of each other given and .
-
(D)
For each , the conditional joint distribution given does not depend on .
The next condition is a generalization of the neighborhood recurrence along a path to Markov branching process. First, for and , define
(2.1) |
Notice that does not depend on by virtue of Properites (A) and (B).
-
(E)
(Cut-set recurrence condition) There exists a set , a function and a number such that for all , and ,
While conditions (C) and (D) are intuitive, condition (E) is somewhat technical. It implies a kind of recurrence property for the branching Markov chain; in the language to be made clear later, the set is cut-set recurrent with respect to . The intuition is that the Markov chain returns to the set infinitely often in a uniform manner across all the paths (cf. Section 3 for the precise definition). Specifically, we have the following result.
Theorem \thethm.
Assume that the tree-indexed family of random variables satisfies (A), (B), (C), (E). Then, with and given by condition (E), we have:
-
(i)
The set is cut-set recurrent with respect to in the sense of Section 3.
-
(ii)
The cardinality of the first passage cut-set is a.s. finite and satisfies
Moreover, if is bounded on then the expected valued of the cardinality of each passage cut-set is bounded by
where
The proof will be given in Section 4. This theorem leads to a strategy to show the non-explosion of a DSY cascade by finding an appropriate value such that the set is cut-set recurrent. Next, we state the main criterion for non-explosion. We say that a function is locally bounded if it maps every bounded set into a bounded set.
Theorem \thethm.
Let be a branching Markov chain with values on a measurable state space and satisfy (A), (B), (C). Let be a DSY cascade with a measurable function. Suppose that condition (E) holds for for some and , where is locally bounded. Then the DSY cascade is non-explosive for any initial state .
The proof of this theorem is in Section 5. In our applications, the case is of particular interest. In this case, we have the following corollary.
Corollary \thecor.
Let be a DSY cascade in which has properties (A), (B), (C) on a measurable state space (and is the Borel -algebra). Suppose that the condition (E) holds for for some , with and both being locally bounded. Then the DSY cascade with is non-explosive for any
We will give in Section 6 two sufficient conditions (relatively simple to check) for condition (E) to be satisfied. An interesting class of DSY cascades is the self-similar DSY cascades. This class includes the cascade of the -Riccati equation, the cascade of the complex Burgers equation, and the cascade of the Navier-Stokes equations with a scale-invariant kernel (Section 12).
Definition \thedefi.
Let be a subgroup of . A DSY cascade is said to be self-similar if the branching Markov chain satisfies the following:
-
(i)
satisfies (A), (B), (C).
-
(ii)
For any initial state and , the family also satisfies (A), (B), (C) with the same transition probability as .
In Section 7, we provide a useful characterization of self-similar DSY cascades: these are exactly the DSY cascades in which the branching Markov chain satisfies (A), (B), (C) such that along each path , the sequence is a multiplicative random walk on (i.e. is an additive random walk on the additive group ). Multiplicative random walks on are natural examples of non-reversible Markov chains, which are of interest in this paper. Self-similar DSY cascades admit a rather simple criterion for non-explosion:
Proposition \theprop.
Let be a self-similar DSY cascade in which is a locally bounded function. Suppose for some . Then the cascade is non-explosive for any initial state .
The proof of this result is given in Section 7. The main criterion for a.s. explosion of DSY cascades is as follows.
Theorem \thethm.
Let be a DSY cascade in which satisfies (A), (B), (D). Put . If there exists a constant such that
then . In particular, the cascade is a.s. explosive for any initial state .
The proof this theorem is given in Section 8. In Section 12, it is shown that the explosion criterion is satisfied by the self-similar DSY cascades associated with the 3-dimensional Navier-Stokes equations. We also show that for large spatial dimensions (), the self-similar DSY cascades are non-explosive. The range remains inconclusive to us. Nevertheless, we show that the explosion event in a self-similar DSY cascade is a event. More generally, one has the following.
Theorem \thethm.
Let be a DSY cascade in which has property (A), (B), (C), (D). Then either for all or for all if one of the following conditions holds.
-
(i)
is a self-similar DSY cascade and is a multiplicative function, i.e. for all .
-
(ii)
Along each path , is an ergodic time-reversible Markov chain on a countable state space .
-
(iii)
Along each path , is time-reversible with respect to a probability measure (i.e. as measures on ) which satisfies for every .
The proof of this result is given in Section 9. Applications of the non-explosion and explosion criteria are given in the last three sections of the paper. In Section 10, we focus on examples originating from probability, whereas Section 11 gives examples originating from differential equations. The applications to NSE are of particular interest and are mentioned in a separated section (Section 12).
3 Preliminary notions
To prepare for the proofs of the main results, we introduce some preliminary notions on a general rooted tree (see also [lyons90, lyons92], [lyons, Sec. 2.5]).
Let be a connected locally finite random tree rooted at . Each vertex is connected to the root by a unique path, so one can identify a vertex with the path connecting it to the root. For a finite or infinite path , we denote by the path obtained by appending to :
The length of a path , or the genealogical height of vertex , is . For , the truncation of up to the ’th generation is . We use the convention that and . The boundary of is defined as the set of infinite paths:
Remark \therem.
One can interpret as the boundary of in the metric space endowed with the metric where is the standard Kronecker delta and
Definition \thedefi (cut-sets).
A finite set of vertices is called a cut-set of if for each path , there exists unique such that .
Intuitively, a cut-set is a set of vertices that separates the root from the boundary.
Definition \thedefi (Passage sets).
Let be a tree-indexed family of random variables independent of taking values on a measurable state space . For , we define a sequence of random sets depending on as follows.
for all . We call the ’th passage set of through .
Definition \thedefi (Cut-set recurrence).
Let be a tree-indexed family of random variables on a measurable state space . A set is said to be cut-set recurrent with respect to if each passage set , , , …is almost surely a cut-set. In this case, is called the ’th passage cut-set.
For the sake of simplicity, we will write the structure simply as .
4 Proof of the Section 2
With the terminology introduced in the previous section, the proof of Section 2 uses the decay of and an inspection process on whether to construct a sequence of passage cut-sets. Specifically at each vertex , starting from the root , we inspect each offspring of . If then passes the inspection. Otherwise, if , does not pass the inspection and the inspection process along the path containing stops at . For each offspring of that passes the inspection, we proceed to inspect its own offspring. For example, if and then we inspect the vertices , , , . If and , we only inspect and . In this manner, the inspection process either continues indefinitely or stops after finitely many steps. Note that we do not inspect the root: is considered passing the inspection whether or not .
For a vertex that passes the inspection, we denote by the number of offspring of that passes the inspection. The distribution of is given by
For , let so that . Note that this probability only depends on due to properties (A) and (B). Let
(4.1) |
which could be infinity. For each vertex with , the number of offspring passing the inspection has a distribution
(4.2) | |||||
(4.3) | |||||
(4.4) |
Invoking Properties (A) and (B) once again, one can write for Denote for . Let be the total number of individuals at generation that pass the inspection. Then and
where , , ,…are i.i.d copies of . The inspection process can be viewed as a nonhomogeneous Galton-Watson process with the offspring distribution given above. We will show this process stops eventually.
Lemma \thelem.
Under the assumption of Section 2, for every , the inspection process starting at almost surely stops after finitely many steps.
Proof.
We have
If then . In this case, the inspection process stops at the root. Consider the case . For , denote the two terms on the right hand side of (4.3) by and .
By Wald’s identity, . Applying this identity consecutively, we get
(4.5) |
If then for all , . Hence, the inspection process stops almost surely after generations. Consider the case . Put . By Fatou’s lemma,
Hence, a.s. Because ’s are nonnegative integers, a.s. for some random value of . ∎
To complete the proof of Part (i) of Section 2, we use Section 4 to show that the set is cut-set recurrent with respect to the branching Markov chain . By definition, the first passage set consists, for each , of the first vertex such that (see Figure 1). In other words, it consists of the vertices at which the inspection process stops. By Section 4, is, a.s, a finite set and, thus, must be a cut-set.
Suppose that for some , the passage set is a cut-set. For each , denote by the subtree of rooted at and write for . By conditions (A) and (B), we have that condition (E) is also satisfied by the branching Markov chain . Indeed, the analogue to (2.1) for this branching Markov process is
Given , by Section 4 the inspection process on must terminate a.s. Thus, for each path passing through , there exists a random integer , such that . Let be the smallest value of such ’s. Let
(4.6) |
corresponding to the first passage set of the tree rooted at through . Thus, by Section 4, is a.s. finite and nonempty and separates boundary paths in that pass through from . Note that, by definition, the ’st passage set is given by
(4.7) |
which is nonempty, finite almost surely. Since is a cut-set, is also a cut-set. We conclude that is cut-set recurrent thus completing the proof Part (i) of Section 2.

We begin the proof of Part (ii) of Section 2 by showing the estimate for the mean cardinality of , namely
(4.8) |
and use induction for the other estimate. With defined in (4.1), if then a.s. and holds. Consider the case . The mean total number of individuals passing the inspection process is
according to (4.5). Because each vertex in is an offspring of a vertex that passes the inspection process, card is at most twice the total number of vertices passing the inspection process. Therefore,
Consider now for some , so that Then, by properties (A) and (B) of the branching Markov chain we have
Recall that we now assume that is bounded on and that Then
(4.9) |
The proof of Section 2 is completed by noting that
5 Proof of Section 2
Thanks to Section 2, we can define a random sequence of nonnegative integers where and
(5.1) |
One can observe that for any and , there exists satisfying . In order to establish the non-explosive character of the DSY cascade, note that for and with , we have:
where with such that . In the last inequality we have used that so that since one has Since is independent of , in order to establish that the DSY cascade is not explosive it suffices to show that almost surely,
(5.2) |
For this, we define a random tree, referred as the “reduced tree”, consisting of the root and vertices in the passage cut-sets , , , …Throughout this section we will use
(5.3) |
denote the random number of vertices in the first cut-set of the tree rooted at For the reduced tree, the root has offspring, each of which is a vertex in the cut-set . Each vertex of the first generation of the reduced tree has offspring, each of which is a vertex in the cut-set , and so on (Figure 2). Thus, the first generation of the reduced tree consists of vertices in the first cut-set . The second generation of this tree consists of vertices in the second cut-set . In general, the ’th generation of the reduced tree consists of vertices in the ’th cut-set .

The set of all vertices of the random reduced tree is denoted by . Recall that the configuration of is independent of the clocks after a natural relabeling of the vertices (Figure 2). Now (5.2) becomes a non-explosion problem of a DSY cascade with all intensities equal to 1 on a random tree structure . The uniform bound obtained in Part (ii) of Section 2 on the expected number of offspring of each vertex turns out to be crucial for our approach. For further criteria for non-explosion of a DSY cascade on a random tree structure, see [part1_2021, Sec. 4].
To show (5.2), we will use a cut-set argument similar to the one used in the proof of Section 2. Let be a number to be chosen. Given the random tree , we start an inspection process of whether as follows. At each vertex , starting from the root , we inspect its offspring. If an offspring satisfies then it passes the inspection. Otherwise, it does not pass the inspection and the inspection process along the path containing stops at . For any vertex that has passed the inspection, we continue to inspect its own offspring. Note that we do not inspect the root: is considered passing the inspection whether or not . In this manner, the inspection process might keep going indefinitely or stop after finitely many steps. To show that the process stops almost surely after finite number of inspections we establish the following more general result.
Lemma \thelem.
Assume there exists such that
Let Then the inspection process almost surely stops after finitely many steps and is a cut-set recurrent with respect to
Proof.
The probability for a vertex to pass the inspection is
The number of offspring of the root that pass the inspection is
which is the sum of i.i.d. random variables with distribution Bernoulli() that are also independent of . Thus
For each , denote by be the total number of vertices of the ’th generation that pass the inspection. We have and . If we label the vertices of the ’th generation that pass the inspection by for then
Write . Since are independent from , we have
Moreover,
Thus
Because , . Put . By Fatou’s lemma, . Therefore, almost surely and hence, the inspection process for almost surely stops in finitely many steps.
To show that the interval is cut-set recurrent with respect to , construct the sequence of passage cut-sets by the procedure described in the proof of Section 2: The first passage cut-set consists of, for each , the first vertex such that . Starting at each vertex in , we start a new independent inspection process. The union of the vertices where one of these inspection processes stops gives the passage cut-set . Since each of these sets is finite, is cut-set recurrent as claimed. ∎
6 Criteria for cut-set recurrence
In this section, we give two criteria for a branching Markov chain to satisfy the condition (E) in terms of the transition probability density . These criteria will be used to show the non-explosion of the cascade of the KPP equation on the Fourier side (Section 11.2) and of the Bessel cascade of the three-dimensional Navier-Stokes equations (Section 11.4).
Proposition \theprop.
Let be a branching Markov chain on the state space satisfying (A) and (B) with the transition probability density . Suppose there exist a constant and functions such that:
-
(i)
for all .
-
(ii)
for some locally bounded function .
-
(iii)
, .
Then satisfies the condition (E).
Proof.
By Section 2, it is sufficient to show that there exist constants and such that , where is defined by
with . Fix an initial state and a path . For simplicity, we write the sequence , , ,…as , , ,…Since this is a time homogeneous Markov chain with transition probability density ,
By the condition (i), . We obtain the estimate
where . By choosing sufficiently large, we get . ∎
Proposition \theprop.
Let be a branching Markov chain on the state space satisfying (A), (B) and with following conditions:
-
(i)
For any path , the Markov chain , , ,…is time-reversible with the transition probability density and the invariant probability density .
-
(ii)
There exist and a locally bounded function such that
-
(iii)
There exists such that
-
(iv)
There exists a function such that
-
(v)
There exists such that
Then satisfies the condition (E) with , , and
Proof.
As noted in Section 2, it is sufficient to show , where is defined by (2.2). Fix an initial state and a path . Again, we write the sequence , , ,…as , , ,…Since this is a time homogeneous Markov chain with transition probability density ,
Consider a sequence of functions given by and
Then . Define a sequence of functions by and
(6.1) |
We now show by induction that . This is true for thanks to condition (ii). Suppose for some . By the induction hypothesis and condition (i),
which completes the proof by induction. We show by induction that
(6.2) |
Using the fact that , we can deduce from (6.1) that for all and . Because , (6.2) is true for . Suppose by induction that for some ,
We decompose given by (6.1) as follows.
The first term can be estimated using condition (c):
The second term can be estimated using condition (d):
By the inequality , we have
where . By the above estimates, . Therefore, (6.2) is true for every . We proceed to estimate . Put . If then
Each term can be estimated as follows:
Therefore, . If then
In this case, we have . ∎
7 Self-similar DSY cascades and proof of Section 2
We begin this section noting some properties of self-similar DSY cascades.
Lemma \thelem.
Let be a self-similar DSY cascade on . Then for any initial state and path , the ratios form an i.i.d. sequence with the common distribution .
Proof.
For , the ratio can be written as , where denotes the parent of . Now fix . By the definition of self-similar DSY cascades, the family satisfies (A), (B) with the transition probability . Since , and have the distribution . Therefore, also has the distribution .
Next, we show that the transition probability has the scaling property for all . This is obtained by expressing the cumulative distribution functions of in two ways. First, for any . Second, . By the change of variables in the last integral, one obtains .
Finally, we show that along each path , the ratios are independent. For any ,
where and are short notations for and . By the change of variables (with ) and the scaling property of , one has
∎
We now give the proof of Section 2. By Section 2 (taking into account Section 2), we only need to find , , and a locally bounded function on such that
where is an arbitrary path. For simplicity, we write the sequences and as and , respectively. Recall that is the distribution of . Define . For , we have
Therefore,
Next, we show by induction that
(7.1) |
where . For ,
Suppose (7.1) is true for some . Then
Therefore, (7.1) is also true for . We can now choose and .
8 Proof of Section 2
Starting from the root , we construct a random path in by recursively annexing one vertex at a time as follows. Suppose is the most recently annexed vertex. If then is the next to be annexed. Otherwise, is the next to be annexed. This random path, which we denote by , is a path with stepwise maximal intensities: at every branching step, the path follows the branch that has a larger intensity.

For , let . Then
By the independence of the clocks and the intensities,
(8.1) |
We only need to show that
(8.2) |
Indeed, once (8.2) is proved, one can infer from (8.1) that
which leads to a.s. Next, to show (8.2), it suffices to show that
(8.3) |
Fix and put . Then and . Because the joint distribution of given is the same as the joint distribution of given ,
Thus, . The proof of (8.3) is then completed by the law of total expectation,
9 Proof of Section 2
With the initial state , the explosion time can be expressed as where is the explosion time of the sub-cascade starting at vertex .
(9.1) | |||||
By Section 7, we can express , , as a product of i.i.d. ratios . Let us first assume that condition (i) is satisfied. We use the multiplicativity of to rewrite the explosion time as where
The event is the same as the event , whose probability does not depend on . Thus, . By (9.1), . Therefore, .
Next, assume that condition (iii) is satisfied. From (9.1), one has the estimate . Now note that along each path , is a stationary distribution of the Markov chain . By integrating the above inequality against , one obtains
which implies for -a.e. . Consider two cases.
for some :
Using the inequality , one has . Thus, for -a.e. . Because , for -a.e. . Now take an arbitrary . Since , for -a.e. . By the inequality , we have
Therefore, .
for all :
In this case, for -a.e. . For any , for -a.e. . By the inequality , we have
Therefore, .
Finally, assume that condition (ii) is satisfied. Using the same technique as in Part (iii) and noting that is fully supported on in the discrete topology, one has for all . Let . If then . Otherwise, for any , we have , which leads to . This implies that contains all the states that can be reached from an element of in one step. By the irreducibility of the Markov chain, .
10 DSY cascades resulting from probabilistic models
In this section, we give four examples of non-explosive cascades motivated by probabilistic considerations.
10.1 A pure birth process
The classical Yule process is a pure process starting from a single progenitor in which a particle survives for a mean (deterministic constant) exponentially distributed time before being replaced by two offspring independently evolving in the same manner. The population is finite at every finite time (i.e. non-explosive) if and only if
(10.1) |
The classical Yule cascades are the simplest case of DSY cascades where . The branching Markov chain can be chosen arbitrarily. We can of course choose (deterministic). It is well-known that the classical Yule cascades are non-explosive. One can apply Section 2 with the choice to obtain an alternative proof by cut-set theory. In fact, Section 5 implies the non-explosion of a more general pure birth process:
Proposition \theprop.
Let be a positive number. Consider a branching process starting with a single progenitor in which a particle, independently of all others, survives for a mean-one exponentially distributed time before being replaced by a random number of offspring. The offspring distributions are not necessarily the same across all the particles. Suppose that the expected number of offspring of each particle is less than . Then the total population is finite at any finite time.
In Section 10.1, if the offspring distributions are the same across all the particles then the pure birth process is a continuous-time Galton-Watson process in which the offspring distribution has a finite expected value. The non-explosion problem for the case of infinite expected offspring number was studied in [amini2013].
10.2 A mean-field cascade (dependent first passage percolation on a tree)
Consider a DSY cascade on the state space such that along each path , is an i.i.d. sequence of random variables. The branching Markov chain clearly has properties (A), (B), (C), (D). Choose sufficiently large such that . Then
Therefore, the condition (E) holds for and . By Section 2, the cascade is non-explosive.
10.3 A cascade with geometric-like sequence along each path
Consider a DSY cascade on the state space in which the branching Markov chain has properties (A), (B), (C) and the transition probability density is given by
Intuitively, can be as large as but with a small probability. This allows the sequence to behave like a geometric sequence up to any prescribed index. It is the geometric growth of intensities along each path that causes the explosion in the -Riccati equation for (Section 11.1). We show that the present cascade is in fact non-explosive. Denote
If then for all . In this case, . If then
Because
we have . Using this inequality repeatedly, we get
Therefore, the condition (E) holds for and . By Section 2, the cascade is non-explosive.
10.4 Birth-death branching Markov chain
Consider a DSY cascade on the state space in which the branching Markov chain has properties (A) and (B) with the transition probabilities given by
where , and for Along each path , the sequence , , , …is the birth-death process on with reflection at and birth-death rates . This is an ergodic time-reversible Markov chain (see [RB_EW2009], Theorem 3.1(b), p. 241) with the invariant probability
(10.2) |
provided that
In particular, this is the case when . Since each state is visited infinitely often, the pathwise total waiting time is infinite:
However, it will be shown below that the cascade can still be explosive.
Proposition \theprop.
Let and where . Suppose that for each , and are conditionally independent given . Then the cascade is a.s. explosive for any initial state .
Proof.
One can see that condition (D) is satisfied. Put . Then . First, consider the case . Put
By Section 2, it suffices to show that for all . For , . For ,
By Section 2, the cascade is non-explosive. Next, we consider the case . Take and denote
We proved in the first case that . Observe that and thus, . This implies a.s. ∎
Proposition \theprop.
Let . Then for any function , a DSY cascade with property (C) is non-explosive for any initial state .
Proof.
Any function from to is locally bounded. By Section 2, we only need to find and a function such that
As shown below, one can choose and . Fix a path and write in lieu of Denote
With the initial state , can be viewed as the first passage time of a simple random walk on with probability of going to the left , and probability of going to the right . One has the estimates ([RB_EW2009, p. 11])
with and with the convention that if or is not an integer. Because , the Markov chain is recurrent on . Thus, . It is known that as a consequence of Sterling’s formula . Thus,
We obtain
∎
Remark \therem.
Section 10.4 and Section 10.4 do not give a conclusion about the explosion or non-explosion in the case and , . However, by Section 2 (ii), we know that it must be a event.
11 DSY cascades resulting from differential equations
Our first example serves as a precursor to the general method of associating a branching cascade structure to a quasilinear evolutionary partial differential equation. This method goes back to McKean’s treatment for the Fisher-KPP equation in the physical space [mckean, MB1978] and Le Jan and Sznitman’s treatment for the Navier-Stokes equations in the Fourier space [lejan].
11.1 The -Riccati equation
For , consider the -Riccati equation
(11.1) |
This equation can be viewed as a toy model for self-similar Navier-Stokes equations (see [alphariccati]), It also appears from purely probabilistic models (see [athreya, DA_PS_1988]). After rewriting the equation in the mild formulation
we can interpret as the expected of a stochastic functional defined implicitly by
where and are two independent copies of . Thus, the stochastic functional is defined over the stochastic structure . In the notation of the present paper, this is a self-similar DSY cascade with and (deterministic). The explosion problem of the -Riccati equation, especially in the connection with the existence and uniqueness of solutions, was studied in detail in [athreya, alphariccati]. The cascade was known to be explosive if and only if . Section 2 provides an alternate justification. Indeed, the ratios along each path are . With , . We have
with . The non-explosion of the cascade of the -Riccati equation for can be inferred from the non-explosion of the standard Yule cascade (see [complexburgers, yule]). If , Section 2 provides an alternative justification. Indeed, with sufficiently large, . Note that the case results in a classical Yule process, which is non-explosive.
In the case , the stochastic non-explosion was exploited in [complexburgers] to prove existence and uniqueness results as well as long-time behavior of solutions to (11.1). In the case , the stochastic explosion was used to prove a nonuniqueness result in [athreya] for the case and [alphariccati] developed a framework for using stochastic explosion and distribution of branches of the underlying cascade to prove both non-uniqueness and finite-time blowup of the solutions for more general initial data.
DSY cascades associated with evolutionary PDEs in Fourier space
The next examples concern the DSY cascades originating from partial differential equations. A common feature of these equations is that they define a dissipative dynamical system in which, when formulated in the Fourier space, the linear term determines the intensity of the random waiting time and the quadratic nonlinear term leads to a random binary tree. In all of these examples, the equations can be written in the Fourier space as
(11.2) |
where , are radially symmetric positive functions, and is a bilinear map. The functions , , will be determined by the specific PDE under consideration. A key step in the probabilistic reformulation of (11.2) is to find a function such that
(11.3) |
is a probability density function on . Once is identified, we introduce the new unknown , which satisfied the normalized equation
(11.4) |
Equation (11.4) leads to a family of wave numbers satisfying , for all , and conditionally given , and are each distributed as . For , one gets a DSY cascade . In most cases, the holding times between branchings only depends on the magnitudes of the random wave numbers which, in turn, have a well-behaved branching Markov structure. For example, for the Navier-Stokes equations, the choice of turns out to be more efficient than the choice of .
Similarly to the -Riccati equation (Section 11.1), the solution can be expressed as the expected value of a “solution” stochastic functional defined over the DSY cascade by
(11.5) |
where and are (conditionally) independent copies of . The stochastic explosion or non-explosion of the associated DSY cascades has direct implications to the existence and uniqueness of global-in-time solutions of these equations [chaos, alphariccati, smallness].
We will illustrate the aforementioned generic scheme in greater detail for the Fisher-KPP equation (on the Fourier side) in the next subsection. Note that the stochastic structure on the Fourier side (namely, DSY cascade) is very different from that on the physical side (namely, branching Brownian motion) which was identified in [mckean, MB1978].
11.2 The Fourier-transformed KPP equation
The KPP equation introduced by Fisher, Kolmogorov, Petrovsky and Piskunov, also referred to as the F-KPP equation, has the general form
where and are positive constant. McKean constructed a cascade model for this equation in the physical domain [mckean, MB1978]. We now construct a cascade for this equation in the Fourier domain. By rescaling the time and space variables, we can assume that . Then by introducing , we get
Taking Fourier transform with respect to , we get
In the integral form,
We normalize by where is a positive function to be determined. Function satisfies the equation
(11.6) |
where . For to be a probability density function, must satisfy the equation
The function satisfies , which has a solution . It follows that
(11.7) |

Then the representation (11.6) is equivalent to where is a stochastic functional defined recursively by
(11.8) |
where and are i.i.d. copies of . This definition, when applied recursively, leads to a family of exponential mean-one clocks and a family of frequencies described as follows. For and , we start a branching process with a particle located at . The particle is time away from the horizon. The clock runs for an exponential time . If then the branching process stops (no branching occurs). Otherwise, the particle dies and is replaced by two particles. A random variable is then sampled according to the p.d.f. . The first particle is placed at and the second particle is placed at . Each particle is now away from the time horizon. These particles continue to evolve by the same rule, independently of each other. Note that . Therefore, we can associate the equation (11.6) with a DSY cascade where , and . The corresponding explosion time is
The stochastic functional is a.s. well-defined by (11.8) for all if and only if a.s. We will apply Section 2 and Section 6 to show the non-explosion of the cascade. The same cascade was analyzed by the authors in [part1_2021, Example 5.4] and Orum in [orum, Sec. 7.9]. The non-explosion was proved by using a large deviation method together with a spectral radius technique in [part1_2021], and by exploiting the uniqueness of solutions to the KPP equation in [orum].
We start the proof of the non-explosion by observing that the function given by (11.7) is even and decreasing on . According to Section 11.2 below, is logarithmically concave on . Thus,
Now fix a number . Using the fact that is bounded from above by , we have
where . Therefore, where and . Since , decays exponentially as . Thus, Section 6 is satisfied with .
Lemma \thelem.
Let . Then for all , .
Proof.
Put . By basic differentiations, . Using the fact that for all , we have
Therefore, is concave on . ∎
11.3 The complex Burgers equation
Consider the one-dimensional Burgers equation
where is a complex-valued function. In the analysis of self-similar solutions in the Fourier domain [complexburgers], the equation is associated with a self-similar DSY cascade in which and the ratios are uniformly distributed on . Since , the cascade is non-explosive according to Section 2.
11.4 Bessel cascade of the three-dimensional NSE
The -dimensional incompressible Navier-Stokes equations are given by
(NSE) |
In the general formulation (11.4) for the Navier-Stokes equations, and the branching distribution of wave numbers is given by
(11.9) |
The function satisfies the equation and is called a standard majorizing kernel [rabi]. In three dimensions, the Bessel cascade is the DSY cascade corresponds to the choice of (see e.g. [rabi, chaos, lejan], [orum, Prop. 3.8]). In the analysis of the explosion problem, the choice turns out to be more efficient than the choice . The corresponding function is . With and , the branching distribution of given is (see [chaos]):
Along each path , the sequence is a time-reversible Markov chain with the invariant probability density
The non-explosion of the Bessel cascade was proved in [part1_2021, Example 5.2] via a large deviation method. Here we present an alternative proof using Section 2 and Section 6. The conditions (i)–(v) in Section 6 are satisfied for the choice of
Therefore, the DSY cascade is non-explosive.
12 Cascade for the Navier-Stokes equations with scale-invariant kernel
This section explores the explosion/non-explosion problem of a self-similar DSY cascade naturally associated with the -dimensional incompressible Navier-Stokes equations with . Based on the general scheme (11.4), the present cascade corresponding to and the choice of standard majorizing kernel , called the scale-invariant kernel. The probability density function is
(12.1) |
In contrast to the Bessel kernel discussed above, this family of kernels scales according to the natural scaling of (NSE), namely
Choose the branching Markov chain . The corresponding intensity function is . In the case , the ratios have the dilogarithmic distribution with density
(12.2) |
and the corresponding self-similar DSY cascade is referred to as the dilogarithmic cascade [chaos]. It was shown in [chaos] that, for , the event of non-explosion is a event and that a.s. explosion does not occur along any deterministic path , i.e. (see [chaos, Cor. 5.1]). Section 2 (i) implies that the non-explosion is a event for any dimension . We will show that non-explosion occurs with probability one for (Section 12.1), and with probability zero for (Section 12.2).
Remark \therem.
In the construction of solutions to the 3-dimensional incompressible Navier-Stokes equations, Le Jan and Sznitman circumvented the problem of stochastic explosion by introducing a clever, albeit adhoc, critical coin tossing device. This technique assures the a.s. termination of the branching after finitely many steps. Subcritical coin tosses could also be implemented for the same effect (see [rabi]). The elimination of the coin-toss device naturally leads to an intrinsic explosion problem for the branching cascade itself. This problem depends on the Markov transition probabilities of the wave numbers , which is determined by the choice of . In the construction of global or local-in-time solutions, Bhattacharya et al [rabi] relaxed the condition on to only for some constants and (called the majorizing kernel). Majorizing kernels determine the space of initial data suitable for the construction of solutions. While our focus here is on zero forcing, nonzero forcing can also be accommodated in this framework (see [chaos]).
We will derive the branching distribution of given as follows. Write , and and let be the angle between and . Then . Put . Choose the spherical coordinates in by where , the unit sphere in . Then . Therefore,
(12.3) |
For any smooth function ,
Thus, the distribution of given has a density
(12.4) |
where
(12.5) |
The ratios have a density
(12.6) |
which is independent of and . Therefore, for any , is a self-similar DSY cascade with the ratio distribution given by (12.6).
Remark \therem.
One can check with basic calculations that for , the distribution of is symmetric with respect to the group identity on , i.e. the median of is equal to 1. If , one can show by using the change of variable . Thus, the median of is less than 1, which better facilitates the non-explosion of the cascade. At this heuristic level, the non-explosion is more likely to happen for than for . One may suspect that for very large , the self-similar cascade of the Navier-Stokes equations is non-explosive. The problem will be further discussed in the next section.
Besides the spherical coordinates, one can also express in the coordinates where is the angle between and . Using the sine rule in the triangle made by , , , we get
Then changing variables in (12.3) from to , noting the Jacobian , we obtain
(12.7) |
where . This is the triangle formed by , and . As is randomized according to , and are values of the corresponding random variables and . The joint distribution density of the random angles conditionally given is
(12.8) |
where is given by (12.5). It is easy to see that is uniformly distributed in the triangle if and only if .
12.1 Non-explosion of the self-similar Navier-Stokes cascades in high dimensions
As , given by (12.8) approaches the uniform distribution on the line segment . Geometrically, the triad , , tend to form a right triangle with sides with hypotenuse . Consequently, and become bounded by 1 in the limit . The explosion time of the self-similar cascade of the Navier-Stokes becomes bounded by that of the classical Yule cascade, or equivalently the -Riccati cascade with , which is non-explosive. Hence, it is quite conceivable that the self-similar DSY cascade of the Navier-Stokes equations is non-explosive if is sufficiently large. The following result affirms that this is the case.
Proposition \theprop.
For , the self-similar cascade of the Navier-Stokes equations in is a.s. non-explosive for any initial state .
Proof.
As shown above, has a density function given by (12.6). By Section 2, it is sufficient to show that . With the change of variable , one gets
Put . We show for all , . By the fact that and by integration by parts, one obtains
Thus, for all . The next step is to evaluate for some values of . By the sine rule in the triangle with edges (Figure 5), . Thus,
We obtain the numerical approximations
This implies that for all . ∎
Remark \therem.
Although an explicit formula for is not needed for the above proof, one can obtain it with the aid of Mathematica:
Remark \therem.
As mentioned before, (12.8) implies that, as , the random vector converges in distribution to a vector distributed uniformly on the line segment . The corresponding random variables and would then represent the sides of the right triangle with hypothenuse 1 and adjacent angles , . This configuration gives raise to a self-similar DSY cascade that can be viewed as the limit case of the cascades associated with (12.1), corresponding to . The mean-field model of this cascade for yields , which is exactly the DSY cascade of the -Riccati equation with . The special case has several critical properties in the context of -Riccati cascades. First, this continuous-time Markov process is a Poisson process with unit intensity. Second, the infinitesimal generator for the Markov process defined by the set of vertices alive at time is a bounded operator if and only if (see [alphariccati, yule]).
12.2 Explosion of the self-similar Navier-Stokes cascade in dimension
In the case , the self-similar DSY cascade (or dilogarithmic cascade) has several special properties. First, the ratios have a dilogarithmic distribution with density given by (12.2). Second, conditionally given , the angle between and and the angle between and have a joint uniform distribution in
We will use Section 2 to show that the dilogarithmic cascade is a.s. explosive. Consider the random variable . Recall that and . By the triangular inequality, . Our first step is to compute the distribution of .
Lemma \thelem.
The random variable has a probability density
Proof.

We observe that
Thus,
By the chain rule of differentiation,
(12.9) |
Since , the second term on the right hand side is zero. To compute , we note that on one hand
(12.10) |
and on the other hand (the chain rule),
(12.11) |
Observe that (Figure 5)
(12.12) |
By (12.10), (12.11), (12.12), we obtain
Now substitute this result into (12.9):
∎
Remark \therem.
If we change variable , then the distribution density of is given by
which is a tilted dilogarithmic distribution. Recall that and individually have the dilogarithmic distribution.
Proposition \theprop.
The self-similar cascade of the Navier-Stokes equations in is a.s. explosive for any initial state .
Proof.
Our explosion criterion (Section 2) is not satisfied for dimensions . The following proposition provides another way to obtain the expectation in (12.13) for and also shows that is the critical case for our method.
Proposition \theprop.
For , we have
The sequence is strictly increasing with , and .
Proof.
By the sine rule in the triangle with edges (Figure 5), and . On the triangle , we have . Thus,
where is given by (12.8) and . One can express in terms of the hypergeometric function as follows.
where . One can now compute either with the aid of Mathematica or directly by induction on even and odd :
Using the asymptotic approximation as , one gets as . To show the monotonicity, we use Kershaw’s inequality for all (see [kershaw]).
∎
Remark \therem.
As mentioned in the introduction, there is a strong connection between the stochastic explosion of the DSY cascade underlying a certain evolution PDE and the well-posedness problem for the PDE itself. In particular, if we simplify the mild-type formulation (11.2) corresponding to the Navier-Stokes equations in Fourier space by replacing with a scalar function and replacing the sophisticated product-like structure in (coming from the Fourier transform of ) with the a simple product of scalars, we will obtain a non-linear scalar PDE with exactly the same DSY cascade structure as the Navier-Stokes equations (NSE). This scalar PDE was first considered by Montgomery-Smith [smith] as a model for a finite-time blowup. As shown in [smallness], the simplified product structure of allows to recover the blowup results from [smith] directly from the DSY cascade. Also, the explosion of the self-similar DSY cascade (Section 12.2) yields a nonuniqueness result for the initial value problem of the Montgomery-Smith equation [smallness].
In the case of 3-dimensional Navier-Stokes equations, the problem of existence and uniqueness of global solutions naturally involves possibility of cancellations in (11.5) due to the geometric structure of the bilinear product . While the present paper focussed on the time evolution of the DSY cascade process itself, it remains to be determined whether the additional geometric structure emanating from the nonlinear term corresponding to (NSE) provides sufficient cancellations in some smooth initial data to negate the impact of explosion of the self-similar DSY cascade.