1 Introduction

1.1 Aim of the paper

In this paper we study the long-time behavior of the coagulation equation with injection:

$$\begin{aligned} \partial _{t}f\left( t, x\right) =\mathbb {K}[f](t,x)+\eta \left( x\right) \end{aligned}$$
(1.1)

where

$$\begin{aligned}{} & {} \mathbb {K}[f](t,x):=\frac{1}{2}\int _{0}^{x}K\left( x-y,y\right) f\left( t, x-y\right) f\left( t, y\right) d y\nonumber \\{} & {} \quad -\int _{0}^{\infty }K\left( x,y\right) f\left( t, x\right) f\left( t, y\right) d y, \end{aligned}$$
(1.2)

and where \(\eta \ge 0\). We will assume in all of what follows that \(\eta (x) \not \equiv 0\), and that it is either compactly supported or it decays fast enough with the cluster size.

The study of problems with this form arises naturally in problems of aerosols and atmospheric science ( [12, 18, 21]). In this context, the function f(tx) denotes the density of clusters with size x at time t.

The collision operator \(\mathbb {K}[f](t,x)\) in (1.2) was introduced by Smoluchowski (see [19]). The kernel K(xy) encodes information about the mechanism driving the coagulation of clusters. In this paper we are interested in kernels arising in atmospheric science applications. A common feature of these kernels is the homogeneity property. Indeed, in many cases the rate of aggregation scales like a power law with the cluster size. This means that the coagulation kernel K(xy) satisfies

$$\begin{aligned} K(a x,a y)=a^{\gamma }K(x,y) for a >0 and (x,y)\in (0,\infty )^{2} \end{aligned}$$
(1.3)

for some \(\gamma \in \mathbb {R}\).

On the other hand, since the coagulation process does not depend on the order in which the clusters of size x and y are chosen, we have the symmetry property

$$\begin{aligned} K(x,y)=K(y,x) for (x,y)\in (0,\infty )^{2}. \end{aligned}$$
(1.4)

We will assume that the coagulation kernel K satisfies

$$\begin{aligned} c_2 \left[ \frac{x^{\gamma +\lambda }}{y^{\lambda }}+\frac{y^{\gamma +\lambda }}{x^{\lambda }} \right] \le K\left( x,y\right) \le c_1 \left[ \frac{x^{\gamma +\lambda }}{y^{\lambda }}+\frac{y^{\gamma +\lambda }}{x^{\lambda }} \right] and \gamma +2\lambda \ge 0,\nonumber \\ \end{aligned}$$
(1.5)

with \(0<c_{2}\le c_{1}<\infty \), where \(\gamma \) is the homogeneity parameter introduced in (1.3) and \(\lambda \in \mathbb {R}\). The polynomial bounds (1.5) are satisfied by many of the most relevant collision kernels arising in aerosol and atmospheric science, such as the diffusive coagulation kernel and the free molecular kernel (see for instance [12]).

Notice that the condition \(\gamma +2\lambda \ge 0\) in (1.5) does not imply any loss of generality. This can be seen from the fact that the function \(x^{\alpha }y^{\beta }+x^{\beta }y^{\alpha }\) can be written as \(\left[ \frac{x^{\gamma +\lambda }}{y^{\lambda }}+\frac{y^{\gamma +\lambda }}{x^{\lambda }} \right] \) with \(\gamma +\lambda =\max \{\alpha ,\beta \}\) and \(-\lambda =\min \{\alpha ,\beta \}\).

Since we will consider solutions of (1.1), (1.2) in which \(f(t, \cdot )\) is a Radon measure, it is convenient to impose the following condition on the kernel K:

$$\begin{aligned} K\in C ((0,\infty )^{2}). \end{aligned}$$
(1.6)

It is well known that part or all the mass of the solutions of (1.1), (1.2) can escape towards \(x=\infty \) in finite, or even zero time. This phenomenon is known as gelation. For a more detailed analysis on this matter, see for example [3, 4]. In order to guarantee that gelation does not take place, we will assume in all of what follows that

$$\begin{aligned} \gamma<1 and \gamma +\lambda <1. \end{aligned}$$
(1.7)

See [5] for a more detailed discussion on the gelation regimes.

We expect that gelation does not take place under the weaker assumptions \(\gamma \le 1, \gamma +\lambda \le 1\) (see [3]). However, in the critical cases \(\gamma =1\) or \(\gamma +\lambda =1\), the self-similar solutions are not defined using power laws to scale the particle sizes but most likely using exponential functions. Given that the analysis of these solutions would require arguments different from the ones in this paper, we will not consider this case here.

The existence of solutions for Eq. (1.1), (1.2) has been considered in [6, 7]. In this paper we study the long time behavior of the solutions of Eq. (1.1). Due to the presence of the source we can expect the solutions to (1.1), (1.2) to converge to a stationary non-equilibrium solution \(\overline{f}=\overline{f}(x)\) as \(t\rightarrow \infty ,\) i.e. to a solution of

$$\begin{aligned} \mathbb {K}[f](x)+\eta (x)=0. \end{aligned}$$
(1.8)

However, it turns out that if \(\eta \not \equiv 0\) and \(\gamma +2\lambda \ge 1,\) a solution for (1.8) does not exist. In [13], the discrete stationary coagulation model

$$\begin{aligned} \sum _{k=1}^{n-1}K_{k,n-k}f_{k}f_{n-k}- \sum _{k=1}^{\infty }K_{k,n}f_{k}f_{n}+\overline{\delta }_{k,n}=0, k\ge 1 \end{aligned}$$
(1.9)

has been studied for the explicit coagulation kernel

$$\begin{aligned} K_{k,n}=k^{\gamma + \lambda }n^{-\lambda }+k^{-\lambda }n^{\gamma + \lambda }, k,n\in \mathbb {N}, \end{aligned}$$
(1.10)

when \(|\gamma + \lambda |<1\), \(|\lambda |<1 \), \(|\gamma |< 1\). Formal asymptotics for the large size behavior of the solutions \(f_k \) of (1.9) has been obtained in [13]. The results in that paper indicate that a solution of (1.9) exists if and only if \(\gamma +2\lambda <1\).

In the case of general kernels satisfying the assumptions (1.4), (1.5), (1.6), and source terms \(\eta \) decreasing fast enough, it has been proved in [10] that the solutions of (1.8) (as well as its discrete counterpart) exist, if and only if \(\gamma +2\lambda <1.\)

It is worth to remark that it has been proved in [10] that the solutions of (1.8) can be estimated, up to a multiplicative constant, from above and below by the power law \(x^{- \frac{3+\gamma }{2}}\) for large values of x.

Since in the case \(\gamma + 2 \lambda \ge 1 \) a stationary solution of (1.8) does not exist, we cannot expect the solutions to (1.1), (1.2) to behave as the stationary solution \(\overline{f}\) of (1.8) as \(t\rightarrow \infty \) for x of order one. It is then natural to ask what is the long time asymptotics of the solutions to (1.1), (1.2) for large values of t and x. The scaling hypothesis that has been extensively used in the study of coagulation equations suggests that the mass of the particle distributions \(f\left( t,x\right) \) is concentrated in cluster sizes x of order \(t^{p}\) for a suitable exponent p that would be determined from dimensional considerations, which take into account the way in which the mass rescales in time. In the case of kernels \(K\left( x,y\right) =x^{\gamma + \lambda }y^{-\lambda }+y^{-\lambda }x^{\gamma + \lambda }\), \( -1< \lambda <0\) with \(0\le \gamma + \lambda <1\) and \(\gamma <1\), it was suggested in [5], using a combination of matched asymptotics and numerical simulations, that the long time behavior of the solutions of (1.1), (1.2) is given by self-similar solutions with the form

$$\begin{aligned} f_s\left( t,x\right) =\frac{1}{t^{\frac{3+\gamma }{1-\gamma }}}\Phi \left( \xi \right) ,\ \ \xi =\frac{x}{t^{\frac{2}{1-\gamma }}}. \end{aligned}$$
(1.11)

The approximation (1.11) can be expected to be valid for large cluster sizes, i.e. \(x\gg 1\). We will use from now the notation with \(x\gg 1\) to indicate large cluster sizes x. In the case considered in [5] we have that \(\gamma +2\lambda <1\) and therefore stationary solutions \(f_{s}\) solving (1.8) exist. In this case, (1.8) and (1.11) suggest that \(\Phi \left( \xi \right) \) behaves for small values of \(\xi \) as \(K\xi ^{-\frac{3+\gamma }{2}}\), for a suitable constant \(K>0.\) More precisely, plugging (1.11) in (1.1), (1.2), it follows that \(\Phi \) solves

$$\begin{aligned} -\frac{2}{1-\gamma }\xi \Phi _{\xi }-\frac{3+\gamma }{1-\gamma }\Phi =\mathbb {K} \left[ \Phi \right] , \end{aligned}$$
(1.12)

where \(\Phi \) satisfies the following boundary condition at \(\xi \rightarrow 0\) which guarantees that there is a constant flux of particles from the origin:

$$\begin{aligned} \lim _{R\rightarrow 0}\int _{0}^{R}d \xi \int _{R-\xi }d \eta K\left( \xi ,\eta \right) \xi \Phi \left( \xi \right) \Phi \left( \eta \right) =J, \end{aligned}$$
(1.13)

with \(J=\int _{0}^{\infty }x\eta \left( x\right) d x.\) The existence of solutions of Eq. (1.12) satisfying the constant flux solution condition at \(\xi =0\), (1.13), has been rigorously proved in [9] for kernels K satisfying (1.4), (1.5), (1.6), (1.7) with \(\gamma + 2 \lambda <1\).

The picture described above, which combines the stationary behavior \(\overline{f}\), (cf. (1.8)) for cluster sizes x of order one, and the self-similar behavior (1.11) for large cluster sizes, provides a rather complete description of the long time behavior of the solutions to (1.1), (1.2) in the case \(\gamma +2\lambda <1.\) However, the same scenario cannot yield a description of the long time asymptotics of the solutions to (1.1), (1.2) if \(\gamma +2\lambda \ge 1,\) because, as explained above, in this case a solution of (1.8) does not exist.

Notice that the existence/non-existence of solutions to (1.8) is related to the existence of stationary solutions of (1.8) yielding a constant flux of particles with the form of a power law, i.e. \(\overline{f}(x)= cx^{- \frac{\gamma + 3 }{2}}\). These solutions exist for \(\gamma + 2 \lambda < 1 \) and they do not exist for \(\gamma + 2 \lambda \ge 1 \). In the framework of the wave turbulence it would be stated that the kernels K with \(\gamma + 2 \lambda < 1 \) satisfy the locality property, while the kernels K with \(\gamma + 2 \lambda \ge 1 \) do not have the locality property, see [22].

In this paper we are interested in the study of the long-time behavior of the solutions to (1.1), (1.2) in the case \(\gamma + 2 \lambda \ge 1 \) (and the non-gelling regime \(\gamma < 1 \) and \(\gamma + \lambda < 1\)) that is described using formal asymptotics arguments. The kernel we work with is homogeneous, hence it can be expressed as

$$\begin{aligned} K\left( x,y\right) =\left( x+y\right) ^{\gamma } F \left( \frac{x}{x+y}\right) ,\ \ F \left( s\right) =F\left( 1-s\right) \ \ \text{ for } s\in \left( 0,1\right) . \nonumber \\ \end{aligned}$$
(1.14)

Notice that (1.5) implies that \({\overline{c}}_1 \le F(s) \le {\overline{c}}_2 \) for some \({\overline{c}}_1, {\overline{c}}_2 >0\). We will assume in what follows a condition that is more restrictive than (1.5), namely

$$\begin{aligned} \lim _{s\rightarrow 0^{+}}\left[ s^{\lambda }F \left( s\right) \right] =1. \end{aligned}$$
(1.15)

In order to prove that in the case \(\gamma + 2 \lambda \ge 1\) there are no solutions of (1.8), the main idea used in [10] is based on the fact that, for this range of exponents, the transfer of clusters of size x of order one towards very large cluster sizes is so fast that the concentration of clusters with size of order one would become zero.

In the case of time dependent solutions having initially finite mass, this increases linearly due to the fact that

$$\begin{aligned} \partial _t \left( \int _0^\infty x f(t, x) d x \right) = \int _0^\infty x \eta (x) d x. \end{aligned}$$

For large times, due to the increase of the average cluster size, we might expect that, if \(\gamma +2\lambda \ge 1\), there should be a fast transport of the newly injected clusters of order one towards much larger cluster sizes. This almost instantaneous transport results in small concentrations of clusters of order one for large times.

We now remark that the part of the coagulation operator which describes the coagulation between particles of different sizes can be approximated by means of a transport operator in the space of cluster sizes. These arguments, that will be described in detail using formal asymptotics in Sect. 2, show that the solutions of (1.1), (1.2) for large cluster sizes can be approximated by means of the following equation if \(\gamma +2\lambda \ge 1\):

$$\begin{aligned} \partial _{t}f\left( t,x\right) +\frac{\partial _{x}\left( x^{\gamma +\lambda }f\left( t,x\right) \right) }{\int _{0}^{\infty }z^{\gamma +\lambda }f\left( t,z\right) d z}=\mathbb {K}\left[ f\right] (t,x),\ \ \text {for }x\gg 1. \end{aligned}$$
(1.16)

We emphasize that the non-local transport term \(\frac{\partial _{x}\left( x^{\gamma +\lambda }f\left( t,x\right) \right) }{\int _{0}^{\infty } z^{\gamma +\lambda }f\left( t,z\right) d z}\) is a consequence of the presence of the source \(\eta \left( x\right) \) in (1.1). The fact that the contributions of the coagulation operator \(\mathbb {K}\left[ f\right] \) which are due to the aggregation of particles with very different sizes can be approximated by a differential operator has been extensively used in the literature of coagulation equations (cf. [12]). The resulting first order terms are often referred to represent heterogeneous condensation (cf. [12]). On the other side, the term \(\mathbb {K}\left[ f\right] \) in (1.16) describes the aggregation of clusters of comparable sizes.

The solutions of (1.16) are expected to describe the asymptotic behavior of the solution of Eq. (1.1) both when \(\gamma + 2 \lambda > 1\) and when \(\gamma + 2 \lambda =1 \), even if in these two cases we have two slightly different scenarios. Namely, when \(\gamma +2\lambda >1\), we will have that if x is of order 1, then \(f(t, x) \rightarrow 0 \) as \(t \rightarrow \infty \), while when \(\gamma + 2 \lambda =1 \) we will have that if x is of order 1, then \(f(t, x) \rightarrow f_s(x)\) as \(t \rightarrow \infty \), where \(f_s \) is a solution of

$$\begin{aligned} \mathbb K[f](x) + \eta (x)- x^{-\lambda } f(x)=0 \end{aligned}$$
(1.17)

(compare with (1.8)). Due to the presence of the term \(x^{-\lambda }f(x)\), Eq. (1.17) might have solutions when \(\gamma +2\lambda =1\).

The scaling properties of (1.16) suggest that this equation is compatible with the existence of self-similar solutions for Eq. (1.16) with the form

$$\begin{aligned} f_s\left( t, x \right) =\frac{1}{t^{\frac{3+\gamma }{1-\gamma }}}\Phi \left( \frac{x}{t^{\frac{2}{1-\gamma }}}\right) . \end{aligned}$$
(1.18)

Here the self-similar profile \(\Phi \) satisfies the following equation, obtained by substituting equality (1.18) in Eq. (1.16) and using the self-similar change of variables \(\xi =\frac{x}{t^{\frac{2}{1-\gamma }}}\),

$$\begin{aligned}&{} -\frac{3+\gamma }{1-\gamma }\Phi \left( \xi \right) -\frac{2}{1-\gamma }\xi \partial _\xi \Phi \left( \xi \right) +\frac{1}{\int _{0}^{\infty }\eta ^{\gamma +\lambda }\Phi \left( \eta \right) d \eta }\nonumber \\{}&{} \quad \times \frac{\partial }{\partial \xi }\left( \xi ^{\gamma +\lambda }\Phi \left( \xi \right) \right) =\mathbb {K} \left[ \Phi \right] \left( \xi \right) , \quad \xi >0. \end{aligned}$$
(1.19)

The main result of this paper is to determine the range of exponents \(\gamma \) and \(\lambda \) satisfying \(\gamma +2\lambda \ge 1\) and the non-gelation conditions (1.7) for which self-similar solutions of (1.16) with the form (1.18) exist (see Fig. 1 for a classification of these exponents). Specifically, we will prove the following: suppose that \(\gamma +2\lambda \ge 1\) and that (1.7) holds. Then

  • If \(\gamma >-1,\) there exists at least one self-similar solution of (1.16) with the form (1.18).

  • If \(\gamma \le -1\) and \(\gamma +2\lambda >1\), no solutions of (1.16) with the form (1.18) exist (See Fig. 1).

  • If \(\gamma \le -1\) and \(\gamma +2\lambda =1\), we prove that there are no self-similar solutions \(f_{s}\) of the form (1.18) such that \(\int _0^1 x^{-\lambda } \Phi (x) d x < \infty \).

The meaning of the condition \(\int _0^1 x^{-\lambda } \Phi (x) d x < \infty \) is that the number of clusters removed by the coagulation process in any bounded time interval is finite. We therefore do not exclude the existence of a self-similar solution of Eq. (1.16) with \(\int _0^1 x^{-\lambda } \Phi (x) d x = \infty \).

A remarkable property of the self-similar solutions of (1.16), that we constructed in this paper, is that they vanish identically for \(0<\xi \ < \rho (M_{\gamma +\lambda }):=\left( \frac{1-\gamma }{2 M_{\gamma +\lambda }}\right) ^{\frac{1}{1-\gamma -\lambda } }\), where

$$\begin{aligned} M_{\gamma +\lambda } :=\int _{(0,\infty )}\xi ^{\gamma +\lambda }\Phi (\xi ) d \xi . \end{aligned}$$

The fact solutions \(\Phi \) of (1.19) vanish in an interval \((0,\rho (M_{\gamma +\lambda }))\) means that, for large times t,  the injected particles are transferred almost instantaneously to clusters with sizes \(x\ge \rho (M_{\gamma +\lambda }) t^{\frac{2}{1-\gamma }}.\) Moreover, the fraction of clusters with sizes \(x< \rho (M_{\gamma +\lambda }) t^{\frac{2}{1-\gamma }}\) becomes negligible for very long times. The existence of this “minimal” cluster size for large times is a remarkable feature that, to our knowledge, has not been observed in the literature on self-similar solutions for the coagulation equation (see for instance [7, 8, 11, 17, 20]). In particular, it is worth to notice that this behavior of the self-similar solutions is very different from the one exhibited by the self-similar solutions obtained in the case \(\gamma +2\lambda <1\) in [9].

The results that we obtain in this paper in the critical case \(\gamma +2\lambda =1\) are more fragmentary than those obtained for \(\gamma +2\lambda >1\). In the case \(\gamma +2\lambda =1\), we only prove the existence of a solution \(\Phi \) of (1.19) for \(\gamma >-1\) with \(\Phi =0\) on \((0,\rho (M_{\gamma +\lambda }))\), but we do not prove that any solution vanishes in the interval \((0,\rho (M_{\gamma +\lambda }))\).

On the other hand, we will prove in this paper that the function \(\Phi \) which describes the self-similar profile in (1.18) decreases exponentially as \(\xi \rightarrow \infty \) in the same manner as the self-similar solutions constructed in [9] and as it usually happens for the self-similar solutions of coagulation equations in problems without injection, see [7, 11].

As indicated above, if \(\gamma \le -1\) and \(\gamma + 2\lambda >1\), there are no self-similar solutions of (1.16) with the form (1.18) and, if \(\gamma \le -1\) and \(\gamma + 2\lambda =1\), there are no self-similar solutions of (1.16) with the form (1.18) satisfying \(\int _0^1 x^{-\lambda } \Phi (x) d x <\infty \). It is natural to ask what is the long time asymptotics of the solutions of (1.1), (1.2) in this case. This question will be the subject of study of a future work.

Fig. 1
figure 1

Coefficients for existence and non-existence of self-similar profiles in the case \(\gamma +2\lambda >1\)

In both cases, the coagulation term \(\mathbb {K}\left[ f\right] \) in (1.16) which describes the aggregation of particles with comparable sizes is negligible for large times, and the long time behavior of the distribution of clusters is determined by the coagulation of particles of size x of order one with large particles. The main difference between the cases \(\gamma +\lambda \ge 0\) and \(\gamma +\lambda < 0\) arises from the fact that in the second case the transfer of clusters (not monomers) from the region where x is of order one to \(x\gg 1\), heterogeneous condensation, is relevant.

There are several results in the physical literature which are related, and are consistent with the ones in this paper; see for instance [2, 14,15,16]. In [2] a coagulation model with kernels satisfying (1.5) with parameters \(\gamma , \lambda \) such that \(\gamma + 2 \lambda \ge 1 \) and including also a source term and a removal of particles term has been studied. In that problem, the cluster concentrations as \(t \rightarrow \infty \) are determined by the coalescence of particles of very different sizes. It is then possible to approximate the coagulation-removal model for large clusters by means of the equation

$$\begin{aligned} \partial _t f(t, x)= & {} - \partial _x \left( x^{\gamma + \lambda } f(t, x) \right) +\int _1^{x/2 } y^{1-\lambda } f(t, y ) d y \\{} & {} - f(t,x) \int _x^K y^{\gamma + \lambda } f(t, y) d y, \end{aligned}$$

where K is the maximum particle size in the system. The numerical simulations in [2] show that the concentration of cluster sizes of order one approach to a stationary solution that converges to zero if K is sent to infinity.

In [14] a coagulation model with injection and with kernels satisfying (1.5) with \(\gamma =0\) and \(\lambda \in (1/2, 1] \) has been considered. Numerical simulations and formal computations in [14] suggest that the cluster concentration for clusters of order one tends to zero as \(t \rightarrow \infty \). In addition, it is suggested that the concentration for large clusters are described by a self-similar solution.

In [15, 16] coagulation equations with injection are considered with kernels satisfying (1.5) with \(-\lambda =\gamma \le -1 \). It is seen there that for \(\gamma + 2 \lambda \ge 1 \), the solutions of the corresponding coagulation equation behave in a non self-similar manner and decay logarithmically. This is in agreement with the non-existence of self-similar solutions that we obtained in this paper for \(\gamma \le -1\).

1.2 Notation and plan of the paper

We use the notation \(\mathbb R_*:=(0, \infty )\) and \(\mathbb R_+:=[0, \infty ). \) Given an interval \(I \subset \mathbb R_+ \) we denote with \(C_c(I)\) the Banach space of the functions on I that are continuous and compactly supported. We endow the space \(C_c(\mathbb R_*)\) with the supremum norm denoted with \(\Vert \cdot \Vert _\infty .\) For a function \(\phi \), if \(\phi (x)=0\), for all x in an interval I, we will denote it by \(\phi (I)=0\). We keep the same notation to mean that the support of a measure is outside the interval I. We denote with \({\mathcal {M}}_+ (I)\) the space of the non-negative Radon measures on I. Given a measure \(\mu \in {\mathcal {M}}_+(I)\) we denote with \(\Vert \mu \Vert _{TV}\) the total variation norm of \(\mu .\) For a compact interval \(I\subseteq \mathbb {R}_{*}\) and two bounded measures \(\mu ,\nu ,\) we denote the Wasserstein metric by \(W_{1},\) namely

$$\begin{aligned} W_{1}(\mu ,\nu )=\sup _{||\varphi ||_{Lip }\le 1}\int _{I}\varphi (x)(\mu -\nu )(d x), \end{aligned}$$
(1.20)

where the supremum is taken over the Lipschitz functions and where \(\Vert f \Vert _{Lip }=\Vert f\Vert _\infty + [f]_{Lip }\), with \([f]_{Lip }=\sup _{\underset{x\ne y}{x,y\in I}}\frac{|f(x)-f(y)|}{|x-y|}\).

To keep the notation lighter, we denote with C and c constants that might change from line to line in the computations. In addition, we use the notation \(f\lesssim g,\) for two functions fg, to mean that there exists a constant \(C>0\) such that \(f\le C g\).

Moreover, given a measure \(\mu \), we denote with \(M_\alpha (\mu )\) the \(\alpha \) moment of \(\mu \), i.e.

$$\begin{aligned} M_{\alpha } (\mu ):=\int _{\mathbb R_*} x^\alpha \mu (d x ). \end{aligned}$$

To simplify the notation, in some cases, we write \(M_\alpha \) instead \(M_\alpha (\mu ) \) if the choice of the measure \(\mu \) is clear in the argument.

We use the notation \(f \sim g \) as \(x \rightarrow x_0\) to indicate that \(\lim _{x \rightarrow x_0} \frac{f(x)}{g(x)} =1\), while we use the notation \(f \approx g \) to say that there exists a constant \(M>0\) such that

$$\begin{aligned} \frac{1}{M } \le \frac{f}{g} \le M. \end{aligned}$$

As previously mentioned, we use the notation \(x\gg 1\) for large cluster sizes x. Additionally, for cluster sizes xy, we denote \(x\gg y\) or \(y\ll x\) to mean that x is much larger than y. For two terms, A and B, we use the notation \(A\simeq B\) to mean informally that A can be approximated in terms of B in the region under consideration in the respective formula.

The rest of this paper is organized as follows: in Sect. 2 we present a heuristic motivation to study the existence of the self-similar solutions considered in this paper. In Sect. 3 we present the main results of the paper regarding the existence and non-existence of a self-similar profile and its properties. In Sect. 4 we explain the main ideas behind the proofs of existence and non-existence, skipping the technical difficulties of the proofs. Section  5 deals with the proof of the existence of a self-similar solution for Eq. (1.19) when \(\gamma > -1\). We also prove in this section that the solution decays exponentially for large values. Section  6 deals with the non-existence of self-similar solutions for Eq. (1.19) when \(\gamma \le -1\).

2 Asymptotic description of the long time behavior

2.1 The case \(\gamma + 2 \lambda > 1\)

In this section we describe the long time asymptotics of the solutions of (1.1), (1.2) if \(\gamma +2\lambda > 1\) using formal asymptotic arguments.

As indicated in the Introduction, we expect \(f\left( t, x\right) \) to converge to zero as \(t\rightarrow \infty \) for x of order one. Therefore, the contribution due to the term \(\frac{1}{2}\int _{0}^{x}K\left( x-y,y\right) f\left( t,x-y\right) f\left( t,y\right) d y\) can be expected to be negligible in this region since this term is quadratic in f and we can expect the linear term \(\int _{0}^{\infty }K\left( x,y\right) f\left( t,y\right) d y f\left( t,x\right) \) to give a larger contribution. We will check that these assumptions are self-consistent, in the sense that they will predict an asymptotic behavior for f for which the assumptions made hold.

We examine the asymptotic behavior of the linear term \(\int _{0}^{\infty }K\left( x,y\right) f\left( t,y\right) d y f\left( t,x\right) \) when x is of order one. Due to the effect of the coagulation, we expect the distribution \(f\left( t,y\right) \) to be concentrated for long times in the larger cluster sizes y as \(t\rightarrow \infty .\) Using the assumptions (1.14) and (1.15), we obtain the following asymptotic behavior of \(K\left( x,y\right) \) for \(x\ll y\):

$$\begin{aligned} K\left( x,y\right) \simeq y^{\gamma +\lambda }x^{-\lambda }. \end{aligned}$$
(2.1)

We then expect to have the following asymptotics as \(t\rightarrow \infty ,\) due to the concentration of f in the large cluster sizes

$$\begin{aligned} \int _{0}^{\infty }K\left( x,y\right) f\left( t,y\right) d y\sim x^{-\lambda }\int _{0}^{\infty }y^{\gamma +\lambda }f\left( t,y\right) d y\ \ \text {as\ \ } t\rightarrow \infty \end{aligned}$$

for x of order one. Then, considering the dominant terms in (1.1), (1.2) for x of order one, we obtain the following equation:

$$\begin{aligned} \partial _{t}f\left( t,x\right) =-x^{-\lambda }f\left( t,x\right) \int _{0}^{\infty }y^{\gamma +\lambda }f\left( t,y\right) d y+\eta \left( x\right) . \end{aligned}$$
(2.2)

As explained in the introduction, if \(\gamma +2\lambda > 1,\) since steady states describing the cluster concentrations with x of order one do not exist, we expect to have \(f\left( t,x\right) \rightarrow 0\) as \(t\rightarrow \infty .\) This suggest that we should have \(M_{\gamma +\lambda }=\int _{0}^{\infty } y^{\gamma +\lambda }f\left( t,y\right) d y\rightarrow \infty \) as \(t\rightarrow \infty .\) Suppose that \(\partial _{t}M_{\gamma +\lambda }\ll M_{\gamma +\lambda }\) as \(t\rightarrow \infty \) (something that would happen if \(M_{\gamma +\lambda }\) behaves like a power law, as we will see to be the case). Then (2.2) implies the following asymptotic behavior for \(f\left( t,x\right) \):

$$\begin{aligned} f\left( t,x\right) \sim \frac{x^{\lambda }\eta \left( x\right) }{\int _{0}^{\infty }y^{\gamma +\lambda }f\left( t,y\right) dy}\ \ \text {as\ \ } t\rightarrow \infty \end{aligned}$$
(2.3)

for x of order one. We notice that the higher order contributions to (2.3) due to the contributions of the term \(\frac{1}{2}\int _{0} ^{x}K\left( x-y,y\right) f\left( t,x-y\right) f\left( t,y\right) d y\) are of order \(\frac{1}{ \left( M_{\gamma +\lambda }\right) ^{2}}\) or smaller. Therefore, these contributions will be neglected in what follows.

Equation (2.3) yields an approximate formula for the concentration of clusters with x of order one. We now approximate the part of the coagulation operator which is due to the aggregation of particles with size one with very large particles. To this end we introduce a characteristic length \(L=L\left( t\right) \gg 1\) such that we can approximate \(f\left( t,y\right) \) by means of (2.3) for \(y\le L.\) We attempt to approximate the evolution of the distribution \(f\left( t,x\right) \) for \(x\gg L.\) Then, the coagulation operator in (1.2) can be approximated, using the symmetry properties of the first term on the right of (1.2), as follows:

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right)&=\frac{1}{2}\int _{0} ^{x}K\left( x-y,y\right) f\left( t,x-y\right) f\left( t,y\right) d y\\&\quad -\int _{0}^{\infty }K\left( x,y\right) f\left( t,x\right) f\left( t,y\right) d y\\&=\left[ \int _{0}^{L}K\left( x-y,y\right) f\left( t,x-y\right) f\left( t,y\right) d y\right. \\&\quad \left. -\int _{0}^{L}K\left( x,y\right) f\left( t,x\right) f\left( t,y\right) d y\right] \\&\quad +\left[ \frac{1}{2}\int _{L}^{x-L}K\left( x-y,y\right) f\left( t,x-y\right) f\left( t,y\right) d y \right. \\&\quad \left. -\int _{L}^{\infty }K\left( x,y\right) f\left( t,x\right) f\left( t,y\right) d y\right] . \end{aligned}$$

We can rewrite this formula as

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right){} & {} =\left[ \int _{0}^{L}\left[ K\left( x-y,y\right) f\left( t,x-y\right) -K\left( x,y\right) f\left( t,x\right) \right] f\left( t,y\right) d y\right] \nonumber \\{} & {} \quad +\mathbb {K}\left[ f\chi _{\left[ L,\infty \right) }\right] \left( t,x\right) , \end{aligned}$$
(2.4)

where \(\chi _{\left[ L,\infty \right) }\) denotes the characteristic function of the interval \(\left[ L,\infty \right) .\) We now approximate \(\mathbb {K} \left[ f\right] \left( t,x\right) \) for large values of x and \(t\rightarrow \infty \) and more precisely for \(x\gg L.\) To this end we use (2.3) and we assume also that K and f are sufficiently regular for large values of x. Then, using the fact that \(y\ll x\) we obtain the approximation

$$\begin{aligned} K\left( x-y,y\right) f\left( t,x-y\right) -K\left( x,y\right) f\left( t,x\right) \simeq -y\frac{\partial }{\partial x}\left[ K\left( x,y\right) f\left( t,x\right) \right] . \nonumber \\ \end{aligned}$$
(2.5)

Moreover, (2.1) yields an approximation for \(K\left( x,y\right) \) if \(x\ll y.\) Exchanging the roles of x and y we obtain \(K\left( x,y\right) \sim x^{\gamma +\lambda }y^{-\lambda }\) as \(\frac{x}{y}\rightarrow \infty \) and plugging this formula into (2.5) we obtain

$$\begin{aligned} K\left( x-y,y\right) f\left( t,x-y\right) -K\left( x,y\right) f\left( t,x\right) \simeq -y^{1-\lambda }\frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f\left( t,x\right) \right] . \end{aligned}$$

Using this approximation in (2.4) we then obtain

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right) {=}{-}\left[ \int _{0} ^{L}y^{1{-}\lambda }f\left( t,y\right) d y\right] \left[ \frac{\partial }{\partial x}\left[ x^{\gamma {+}\lambda }f\left( t,x\right) \right] \right] {+}\mathbb {K}\left[ f\chi _{\left[ L,\infty \right) }\right] \left( t,x\right) \end{aligned}$$

for \(x\gg 1\) and \(t\rightarrow \infty .\) We can now use (2.3) to derive a formula for \(\int _{0}^{L}y^{1-\lambda }f\left( t,y\right) d y.\) We then obtain the approximation

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right) \simeq -\frac{\int _{0}^{\infty } y\eta \left( y\right) d y}{\int _{0}^{\infty }y^{\gamma +\lambda }f\left( t,y\right) d y}\frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f\left( t,x\right) \right] +\mathbb {K}\left[ f\chi _{\left[ L,\infty \right) }\right] \left( t,x\right) . \nonumber \\ \end{aligned}$$
(2.6)

Notice that we use that \(\int _{0}^{L}y\eta \left( y\right) d y\simeq \int _{0}^{\infty }y\eta \left( y\right) d y\) since by assumption \(\eta \left( y\right) \) decreases sufficiently fast for large values of y.

We will denote as \(f_{out}\) the distribution of particles in the region where \(x\gg 1.\) More precisely we write \(f_{out}=f\chi _{\left[ L,\infty \right) }.\) Combining (1.1), (1.2) with (2.6) we obtain the following evolution equation for \(f_{out}\):

$$\begin{aligned} \partial _{t}f_{out}\left( t,x\right) +\frac{\int _{0}^{\infty }y\eta \left( y\right) d y}{\int _{0}^{\infty }y^{\gamma +\lambda }f_{out}\left( t,y\right) d y}\frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f_{out}\left( t,x\right) \right] =\mathbb {K}\left[ f_{out}\right] \left( t,x\right) . \nonumber \\ \end{aligned}$$
(2.7)

Notice that we use the approximation \(\int _{0}^{\infty }y^{\gamma +\lambda }f\left( t,y\right) d y\simeq \int _{0}^{\infty }y^{\gamma +\lambda } f_{out}\left( t,y\right) d y\) that might be expected because \(f\left( t,y\right) \rightarrow 0\) for \(y\le L\) as \(t\rightarrow \infty \) (cf. (2.3)).

In the rest of the paper we will study the properties of the self-similar solutions associated to Eq. (2.7). It is worth to remark that the transport term on the left of (2.7) is the way in which the injection of particles with size x of order one affects the outer distribution of clusters \(f_{out}\). Indeed, multiplying (2.7) by x and integrating we obtain

$$\begin{aligned}&\partial _{t}\left( \int _{0}^{\infty }xf_{out}\left( t,x\right) d x\right) +\frac{\int _{0}^{\infty }y\eta \left( y\right) d y}{\int _{0}^{\infty } y^{\gamma +\lambda }f_{out}\left( t,y\right) d y}\int _{0}^{\infty } x\frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f_{out}\left( t,x\right) \right] d x \nonumber \\&\qquad =\int _{0}^{\infty }x\mathbb {K}\left[ f_{out}\right] \left( t,x\right) d x. \end{aligned}$$
(2.8)

The mass conservation property associated to the coagulation kernel yields

$$\begin{aligned} \int _{0}^{\infty }x\mathbb {K}\left[ f_{out}\right] \left( t,x\right) d x=0. \end{aligned}$$

On the other hand, integrating by parts in the second term on the left of (2.8) we obtain \(\int _{0}^{\infty }x\frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f_{out}\left( t,x\right) \right] d x=-\int _{0}^{\infty }x^{\gamma +\lambda }f_{out}\left( t,x\right) d x.\) Combining these results we obtain

$$\begin{aligned} \partial _{t}\left( \int _{0}^{\infty }xf_{out}\left( t,x\right) d x\right) =\int _{0}^{\infty }x\eta \left( x\right) d x. \end{aligned}$$
(2.9)

The identity (2.9) states that the total mass of the clusters in the outer region is equal to the injection rate. This result, that holds for long times, could be expected because in the regime described in this section, the injected particles are transferred instantaneously to large cluster sizes. This is also consistent with \(f(t,x)\rightarrow 0\) as \(t\rightarrow \infty \) when \(x\approx 1\).

2.2 The case \(\gamma +2\lambda =1\)

In the case \(\gamma +2\lambda =1\), the approximation of the concentrations of clusters \(f(t, \cdot )\) in the region where x is of order one must be obtained in a different manner. The reason is that in this case we cannot expect the moment \(M_{\gamma +\lambda }=\int _0^\infty x^{\gamma +\lambda } f(t,x) d x \) to converge to infinity as \(t\rightarrow \infty .\) Indeed, suppose that most of the mass of the monomers is distributed in a characteristic length L(t) that increases as \(t\rightarrow \infty .\) Then, if we denote as \(M_{0}\) and \(M_{1}\) the moments of f of order zero and one, respectively, (i.e. \(\int _{0}^{\infty }f\left( t,x\right) d x\) and \(\int _{0}^{\infty }xf\left( t,x\right) d x \), respectively), we have that \(M_{1}=M_{0} L(t)\) and \(M_{\gamma +\lambda } =M_{0}L(t)^{\gamma +\lambda }.\)

Assuming that \(\int _{0}^{\infty } x\eta \left( x\right) d x=1\) and hence that \(M_{1}\simeq t\) as \(t\rightarrow \infty \), we deduce that \(M_{0}L(t)=t\). The rescaling properties of (1.1), (1.2) suggest that \(\frac{M_{0}}{t} \simeq \left( M_{0}\right) ^{2}L(t)^{\gamma }.\) Therefore, plugging the identity \(M_{0}=\frac{t}{L(t)}\) in this formula we obtain \(L(t)= t^{\frac{2}{1-\gamma }}.\) Hence

$$\begin{aligned} M_{\gamma +\lambda }=M_{0}L(t)^{\gamma +\lambda }=tL(t)^{\gamma +\lambda -1}= t^{1+\frac{2}{1-\gamma }\left( \gamma +\lambda -1\right) }=1. \end{aligned}$$
(2.10)

It then follows that the self-similar rescaling ansatz implies that \(M_{\gamma +\lambda }\) remains of order one for large times. A consequence of this is that we cannot approximate (1.1), (1.2) for clusters of order one by means of Eq. (2.2). Instead of this we will use a different approximation by splitting f in an outer part which describes the cluster distribution for x of order L and an inner part that describes the cluster distribution for x of order one. More precisely, we write

$$\begin{aligned} f\left( t,x\right) =f_{inner}\left( t, x \right) +f_{outer}\left( t, x\right) , \end{aligned}$$
(2.11)

where \(f_{outer} = f \chi _{[L, \infty ) }\), while \(f_{inner} =f \chi _{(0, L] }\), for a constant \(L>0.\)

Using (1.1), (1.2) we would then obtain the following evolution equation for \(f_{inner}\):

$$\begin{aligned} \partial _{t}f_{inner}\left( t,x\right){} & {} =\mathbb {K}\left[ f_{inner}\right] \left( t,x\right) -f_{inner}\left( t,x\right) \int _{0}^{\infty }K\left( x,y\right) f_{outer}\left( t,y\right) d y\nonumber \\{} & {} +\eta \left( x\right) . \end{aligned}$$
(2.12)

Here we have used the decomposition (2.11) in the loss term of the coagulation operator \(\mathbb {K}\left[ f\right] \). We use also the fact that in order to compute the gain term for x of order one we need to use only \(f_{inner}.\)

By assumption, the main contribution of \(f_{outer}\left( t,y\right) \) is due to clusters with size \(L \gg 1.\) On the other hand, for x of order one, we have the approximation

$$\begin{aligned} K\left( x,y\right) f_{outer}\left( t,y\right) \simeq y^{\gamma +\lambda }x^{-\lambda }f_{outer}\left( t,y\right) \end{aligned}$$

and (2.12) becomes

$$\begin{aligned} \partial _{t}f_{inner}\left( t,x\right){} & {} =\mathbb {K}\left[ f_{inner}\right] \left( t,x\right) -f_{inner}\left( t,x\right) x^{-\lambda }\int _{0} ^{\infty }y^{\gamma +\lambda }f_{outer}\left( t,y\right) d y\nonumber \\{} & {} \quad +\eta \left( x\right) . \end{aligned}$$
(2.13)

The scaling argument above, (2.10), suggests that \(\int _{0}^{\infty }y^{\gamma +\lambda }f_{outer}\left( t,y\right) d y\) approaches to a positive constant as \(t\rightarrow \infty \) if \(f_{outer}\) behaves in a self-similar manner and \(\gamma +2\lambda =1.\) We will write

$$\begin{aligned} M_{\gamma +\lambda }=\int _{0}^{\infty }y^{\gamma +\lambda }f_{outer}\left( t,y\right) d y. \end{aligned}$$
(2.14)

On the other hand, we will assume that the function \(f_{inner}\), which is described by means of (2.13), approaches to a stationary solution for large times. We then obtain the following equation which would be expected to describe the behavior of the solutions of (2.13) for long times

$$\begin{aligned} \mathbb {K}\left[ f_{inner}\right] \left( x\right) -M_{\gamma +\lambda }f_{inner}\left( x\right) x^{-\lambda }+\eta \left( x\right) =0. \end{aligned}$$
(2.15)

We can now derive an equation describing the evolution of \(f_{outer}.\) To this end we argue as in the case of \(\gamma +2\lambda >1\) in order to approximate the effect in \(f_{outer}\) due to the collisions of clusters with size L with clusters with size of order one. Using approximations analogous to the ones used for the derivation of (2.7) (cf. (2.4), (2.5)) we obtain the following approximation for \(\mathbb {K}\left[ f\right] \left( t,x\right) \):

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right){} & {} \simeq -\left[ \int _{0}^{\infty }y^{1-\lambda }f_{inner}\left( t,y\right) d y\right] \left[ \frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f_{outer}\left( t,x\right) \right] \right] \nonumber \\{} & {} \quad +\mathbb {K}\left[ f_{outer}\right] \left( t,x\right) \end{aligned}$$
(2.16)

for x of order L.

On the other hand, multiplying (2.15) by x and integrating in \(\left( 0,\infty \right) \) we obtain, since \(f_{inner}\) is zero for \(x>L\), that

$$\begin{aligned} M_{\gamma +\lambda }\int _{0}^{\infty }y^{1-\lambda }f_{inner}\left( y\right) d y=\int _{0}^{\infty }y\eta \left( y\right) d y. \end{aligned}$$

Using this formula to eliminate \(\int _{0}^{\infty }y^{1-\lambda }f_{inner} \left( y\right) d y\) in (2.16) we obtain

$$\begin{aligned} \mathbb {K}\left[ f\right] \left( t,x\right) \simeq -\frac{\int _{0}^{\infty }y\eta \left( y\right) d y}{M_{\gamma +\lambda }}\left[ \frac{\partial }{\partial x}\left[ x^{\gamma +\lambda }f_{outer}\left( t,x\right) \right] \right] +\mathbb {K}\left[ f_{outer}\right] \left( t,x\right) . \end{aligned}$$

Therefore, we obtain that \(f_{outer}\) satisfies Eq. (2.7).

The whole asymptotic behavior derived here relies on the existence of solutions of Eq. (2.15). In this equation, the value of \(M_{\gamma +\lambda }\) is chosen as one associated to a self-similar solution of (2.7). Equation (2.15) can then be interpreted as a stationary solution for a coagulation equation with source \(\eta \) and a removal term \(-M_{\gamma +\lambda }f_{inner}\left( x\right) x^{-\lambda }.\) The results in [10] imply that no solutions of (2.15) exist if \(M_{\gamma +\lambda }=0\) and \(\gamma +2\lambda =1\), but when \(M_{\gamma +\lambda } >0\) the existence/non-existence of stationary solutions for Eq. (2.15) is still an open problem.

3 Setting and main results

The main results of this paper concern the existence and the non-existence of solutions to the following equation:

$$\begin{aligned}&{} -\frac{3+\gamma }{1-\gamma }\Phi \left( \xi \right) -\frac{2}{1-\gamma }\xi \Phi _{\xi }\left( \xi \right) +\frac{1}{\int _{0}^{\infty }\eta ^{\gamma +\lambda }\Phi \left( \eta \right) d \eta }\nonumber \\{}&{} \quad \times \frac{\partial }{\partial \xi }\left( \xi ^{\gamma +\lambda }\Phi \left( \xi \right) \right) =\mathbb {K}\left[ \Phi \right] \left( \xi \right) ,\ \ \xi >0. \end{aligned}$$
(3.1)

We now write precisely what we mean by a solution of Eq. (3.1) and then we state the main theorems on the existence of a self-similar profile under certain assumptions on the parameters \(\gamma \) and \(\lambda \) (Theorem 5.1), on its properties (Theorem 6.1) and on the non-existence of the self-similar solutions under different assumptions on the parameters \(\gamma \) and \(\lambda \) (Theorem 3.5).

Definition 3.1

Let K be a homogeneous symmetric coagulation kernel satisfying (1.5), (1.6), with homogeneity \(\gamma < 1 \) and with \(\gamma + \lambda < 1\) and \(\gamma + 2 \lambda \ge 1 \). A self-similar profile of Eq. (2.7) with respect to the kernel K, is a measure \(\Phi \in {\mathcal {M}}_+(\mathbb R_*) \) such that

$$\begin{aligned} 0< \int _{\mathbb R_*}x^{\gamma +\lambda }\Phi (d x)< \infty \text { and } \int _{(0,1)}x^{1-\lambda }\Phi (d x) < \infty , \end{aligned}$$
(3.2)

and such that it satisfies the equation

$$\begin{aligned}&\int _{\mathbb R_*} \varphi '(x) \left[ \frac{2}{1-\gamma }x - \frac{x^{\gamma + \lambda }}{\int _{\mathbb R_*}z^{\gamma +\lambda }\Phi (d z)} \right] \Phi (d x)\nonumber \\&\qquad - \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (x) \Phi (d x)\nonumber \\&\quad =\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K(x,y) \left[ \varphi (x+y)- \varphi (x) - \varphi (y) \right] \Phi (d x) \Phi (d y) \end{aligned}$$
(3.3)

for every test function \(\varphi \in C^1_c(\mathbb R_*).\)

Remark 3.2

For every test function \(\varphi \in C^1_c(\mathbb R_*)\) (hence such that \(\varphi =0\) near zero) all the integrals in Eq. (3.3) are finite. The integrals in the left-hand side of Eq. (3.3) are bounded due to the fact that \(\varphi \) is compactly supported and that \(\Phi \) is a Radon measure. We analyse now the right hand side. Since \(\gamma + 2 \lambda \ge 0 \)

$$\begin{aligned}&\int _{\mathbb R_*} \int _{\mathbb R_*} K(x,y) \left| \varphi (x+y)- \varphi (x) - \varphi (y) \right| \Phi (d x) \Phi (d y) \\&\quad \le 4 c_1 \int _{\mathbb R_*} \int _{(0, y]} x^{- \lambda }y^{\gamma + \lambda } \left| \varphi (x+y)- \varphi (x) - \varphi (y) \right| \Phi (d x) \Phi (d y) \\&\quad \le c \int _{\mathbb R_*} \int _{(0, y]} x^{- \lambda }y^{\gamma + \lambda } \left( \left| \varphi (x+y) - \varphi (y) \right| + | \varphi (x)| \right) \Phi (d x) \Phi (d y) \\&\quad \le c ||\varphi '||_{\infty }\int _{\mathbb R_*} \int _{(0, 1)} x^{1 - \lambda }y^{\gamma + \lambda } \Phi (d x) \Phi (d y ) \\&\qquad + c \int _{\mathbb R_*} \int _{[1, y]} x^{ - \lambda }y^{\gamma + \lambda } \left| \varphi (x+y) - \varphi (y) \right| \Phi (d x) \Phi (d y ) \\&\qquad + c \int _{\mathbb R_*} y^{\gamma + \lambda } \Phi (d y) \int _{\mathbb R_*} x^{- \lambda } |\varphi (x)| \Phi (d x). \end{aligned}$$

Using (3.2), the fact that \(- \lambda \le \gamma + \lambda \), as well as the fact that \(\varphi \) is compactly supported the desired conclusion follows.

Theorem 3.3

(Existence of the self-similar profiles) Let K be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that

$$\begin{aligned} -1 < \gamma , \quad \gamma +2\lambda \ge 1. \end{aligned}$$

Then there exists a self-similar profile \(\Phi \) as in Definition 3.1. Moreover, \(\Phi \) is such that \(\Phi ((0, \rho (M_{\gamma +\lambda })))=0\) for

$$\begin{aligned} \rho (M_{\gamma +\lambda }):= \left( \frac{1-\gamma }{2 \int _{\mathbb R_*} x^{\gamma + \lambda } \Phi (d x ) }\right) ^{\frac{1}{1-\gamma -\lambda }}. \end{aligned}$$
(3.4)

Additionally, \(\Phi \) it is such that

$$\begin{aligned} \int _{\mathbb R_*} e^{Lx } \Phi (d x) < \infty \end{aligned}$$

for some \(L>0\) and it is absolutely continuous with respect to the Lebesgue measure. Then \(\Phi (d x )=\phi (x) d x \) and the density \(\phi \) is such that

$$\begin{aligned} \limsup _{x \in \mathbb R_*} \phi (x) e^{M x} < \infty \end{aligned}$$

for a positive constant M.

Remark 3.4

In this paper we do not prove the uniqueness of the self-similar profiles. Therefore, it makes sense to understand if the proven properties for the self-similar profile constructed in Theorem 5.1 hold for each self-similar profile as in Definition 3.1.

When \(\gamma + 2 \lambda >1 \) we prove that each self-similar profile as in Definition 3.1 is zero in the set \((0, \rho (M_{\gamma +\lambda }))\), where \(\rho (M_{\gamma +\lambda })\) is given by (3.4), see Theorem 6.1 for more details. In contrast, when \(\gamma + 2 \lambda =1 \), we only prove that the self-similar profile constructed in the proof of Theorem 3.3 is such that \(\Phi ((0, \rho (M_{\gamma +\lambda })))=0\). However, we do not know if this property holds for every self-similar profile as in Definition 3.1.

Theorem 3.5

(Non-existence of the self-similar profiles) Let K be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds.

  1. 1.

    If \( \gamma \le -1, \gamma + 2\lambda > 1\), then a self-similar profile \(\Phi \) as in Definition 3.1 does not exist.

  2. 2.

    If \( \gamma \le -1, \gamma + 2\lambda =1 \), then a self-similar profile \(\Phi \) as in Definition 3.1 with the additional property

    $$\begin{aligned} \int _{(0,1]} x^{-\lambda } \Phi (d x ) < \infty \end{aligned}$$
    (3.5)

    does not exist.

Remark 3.6

Notice that if \(\gamma + 2 \lambda > 1\) we prove that self-similar solutions as in Definition 3.1 do not exist when \(\gamma \le -1 \). Instead, when \(\gamma + 2 \lambda =1 \) and \(\gamma \le -1\), we do not exclude the existence of a self-similar solution \(\Phi \) as in Definition 3.1 with

$$\begin{aligned} \int _{(0,1]} x^{-\lambda } \Phi (d x ) = \infty . \end{aligned}$$

4 Main ideas of the proofs

In this section we explain the main ideas for the proofs of existence/non-existence of self-similar solutions. Both in the case \(\gamma + 2 \lambda >1\) and \(\gamma + 2 \lambda =1 \), to prove that a self-similar solution exists, we find an invariant region for the evolution equation corresponding to (3.1), namely the following equation:

$$\begin{aligned} \partial _t \Phi (t, \xi ) - \frac{3+\gamma }{1-\gamma }\Phi (t, \xi ) - \frac{2\xi }{1-\gamma } \partial _\xi \Phi (t, \xi ) + \frac{\partial _\xi \left( \xi ^{\gamma +\lambda } \Phi (t, \xi )\right) }{\int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (t, d x ) } =\mathbb K [\Phi ](t, \xi ) . \end{aligned}$$
(4.1)

By Tychonoff fixed point theorem the existence of an invariant region implies that there exists a solution of Eq. (3.1).

We prove that the set

$$\begin{aligned} P=\left\{ H: \quad \begin{aligned}&\int _0^\infty x H(x) d x =1,\ \frac{1}{C_{2}} \le \int _0^\infty x^{\gamma + \lambda } H (x) d x \le C_1, \\&H((0, \rho (C_1) ))=0, \ \int _0^\infty x^{2-\gamma - \lambda } H (x) d x \le C_{2} \end{aligned} \right\} \end{aligned}$$

is invariant when \(\gamma + 2 \lambda \ge 1 \) and \(\gamma > -1\) for suitable constants \(C_1, C_2>0 \) and \(\rho (C_1)\) given by

$$\begin{aligned} \rho (C_1):= \left( \frac{1-\gamma }{2 C_1} \right) ^{\frac{1}{1-\gamma -\lambda }}. \end{aligned}$$
(4.2)

To prove that the set P is invariant we proceed as follows:

  1. 1.

    We prove that \(\int _0^\infty x\Phi _0(x) d x =1 \) implies \(\int _0^\infty x\Phi (t,x) d x =1 \), for every \(t >0\). This is done multiplying by x Eq. (4.1), integrating from zero to infinity and then studying the ODE for the first order moment obtained in this manner.

  2. 2.

    As a second step we prove that there exists an upper bound for \(M_{\gamma + \lambda }\). To this end we multiply both sides of Eq. (3.1) by \(x^{\gamma + \lambda }\) and we integrate over x in \((0, \infty )\) to obtain an ODE for the \(\gamma + \lambda \) moment. Using Grönwall’s lemma, the fact that \(\gamma + 2 \lambda \ge 1 \) and that \(\gamma > -1\), the desired conclusion follows.

  3. 3.

    As a third step we use the fact that \(M_{\gamma + \lambda } \le C_1\) to prove that \(\Phi (t, (0, \rho (C_{1}))=0.\) Indeed, the evolution described by Eq. (4.1) is driven by two mechanisms: coagulation, which increases the average size of the particles in the system, and the growth term,

    $$\begin{aligned} \frac{\xi ^{\gamma + \lambda }}{\int _0^\infty x^{\gamma + \lambda } \Phi (t,x ) d x } - \frac{2}{1-\gamma } \xi , \end{aligned}$$

    which is positive for every \(\xi < \rho (C_1) \). Hence, if we start from an initial data \(\Phi _0\) such that \(\Phi _0((0, \rho (C_1) ))=0\), then we will have that \(\Phi (t,(0, \rho (C_1) ))=0\), for every \(t >0.\)

  4. 4.

    Using the fact that \(\gamma + \lambda <1\), \(\gamma > -1\), \(\Phi (t,(0, \rho (C_1)))=0\), as well as the fact that \(\int _0^\infty x^{\gamma +\lambda } \Phi (t, d x) \le C_1 \), we prove that \( \int _0^\infty x^{2-\gamma -\lambda } \Phi (t,x) d x \le C_{2}\).

  5. 5.

    Finally, from the upper bound for \(M_{2-\gamma -\lambda }\), we derive a lower bound for \(M_{\gamma + \lambda }\). Indeed, Cauchy-Schwarz inequality implies that

    $$\begin{aligned} 1= \left( \int _{\mathbb R_*} x \Phi (t, x) d x \right) ^2\le \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (t, x) d x \int _{\mathbb R_*} x^{2-\gamma - \lambda } \Phi (t, x) d x . \end{aligned}$$
    (4.3)

    Hence \(\frac{1}{C_{2}} \le M_{\gamma + \lambda }. \)

In order to prove non-existence we proceed by contradiction. Due to the contribution of the coagulation operator in Eq. (1.16), we expect the zeroth moment of f to decay in time. However, assuming the self-similar change of variable (1.18), we have that

$$\begin{aligned} \int _0^\infty f(t, x) d x = t^{- \frac{1+\gamma }{1-\gamma } } \int _0^\infty \Phi (x) d x. \end{aligned}$$

If \(\gamma \le -1 \), then \(\int _0^\infty f(t, x) d x \) is constant or increasing in time and this gives a contradiction. Hence we cannot expect self-similar solutions to exist. To make the argument rigorous we will have to prove that \(0< \int _0^\infty \Phi (d x ) < \infty \). When \(\gamma + 2 \lambda > 1 \) we do this by proving that for each self-similar profile there exists a \(\delta >0\) such that \(\Phi ((0, \delta ))=0\) and that \( \Phi \) tends to zero sufficiently fast as \(x \rightarrow \infty \). Instead, when \(\gamma + 2 \lambda =1 \), the methods used in this paper do not allow to prove that \(\Phi \) is equal to zero near the origin, hence we use (3.5) as well as (3.2) to prove that \(0< \int _0^\infty \Phi (d x ) < \infty \).

5 Existence of a self-similar profile

We aim to prove the existence of a solution of Eq. (4.1). Namely, we will prove the following theorem, which is just a reformulation of Theorem 3.3:

Theorem 5.1

Let K be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that

$$\begin{aligned} -1 < \gamma , \quad \gamma +2\lambda \ge 1. \end{aligned}$$

Then there exists a self-similar profile \(\Phi \) as in Definition 3.1. Moreover, \(\Phi \) is such that \(\Phi ((0, \rho (M_{\gamma +\lambda })))=0\) with \(\rho (M_{\gamma +\lambda })\) given by (3.4).

To this end we prove the existence of a (time-dependent) solution for a suitably truncated and regularized version of Eq. (4.1):

$$\begin{aligned}&\partial _{t} \Phi \left( t, \xi \right) - \frac{3+\gamma }{1-\gamma }\Phi (t, \xi ) - \frac{2\xi }{1-\gamma } \partial _\xi \Phi (t, \xi ) + \frac{\partial _\xi \left( \xi ^{\gamma +\lambda } \Phi (t, \xi )\right) }{\int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (t, d x ) }\\ {}&=\mathbb K_R [\Phi ](t, \xi ). \nonumber \end{aligned}$$
(5.1)

The operator \(\mathbb K_R\) is the truncated operator of parameter \(R>0\) defined as

$$\begin{aligned} \mathbb K_R[\Phi ](t,x):=&\frac{1}{2} \int _{0}^{x} K_R\left( x-y,y\right) \Phi \left( t, x-y\right) \Phi \left( t, y\right) d y\nonumber \\&-\int _{0}^{\infty }K_R\left( x,y\right) \Phi \left( t, x\right) \Phi \left( t, y\right) d y, \end{aligned}$$
(5.2)

where the kernel \(K_R\) is a truncated kernel, i.e. it is a continuously differentiable, bounded and symmetric function \(K_R: \mathbb R_*^2 \rightarrow \mathbb R_+\), such that for every \((x,y ) \in \mathbb R_*^2 \) we have that \(K_R(x,y) \le K(x,y)\) and

$$\begin{aligned} {\left\{ \begin{array}{ll} K_R(x,y)=0, &{}\text { if } x>R \text { or } y > R; \\ | K_R(x,y) - K(x,y) |\le e^{-R}, &{}\text { if } (x,y) \in \left[ \frac{1}{2 R}, \frac{R}{4}\right] ^2. \end{array}\right. } \end{aligned}$$
(5.3)

Notice that the kernel \(K_R\) can be obtained starting from the kernel K by means of standard truncations and mollifying arguments.

Using Tychonoff fixed point theorem we prove the existence of a stationary solution \(\Phi _{R} \) for Eq. (5.1) and we will prove that there exists the limit \(\Phi \) of \(\{ \Phi _{R} \}_{R}\) as R tends to infinity. To conclude we will prove that the measure \(\Phi \) satisfies Eq. (1.19).

5.1 Existence of a time dependent solution for the truncated equation

Since in this section we work only with the truncated equation, we omit the label R in \(\Phi _{R}\). We will reintroduce the label R in Sect. 5.3. We start this section by introducing a definition of solutions for Eq.  (5.1).

Definition 5.2

Let \(K_R\) be the truncated kernel defined as in (5.3) as a function of the homogeneous symmetric coagulation kernel K satisfying (1.5), (1.6), with homogeneity \(\gamma < 1 \) and with \(\gamma + \lambda < 1\) and \(\gamma + 2 \lambda \ge 1 \). A function \(\Phi \in C^1([0,T]; {\mathcal {M}}_{+,b}(\mathbb R_*))\) is a solution of Eq. (5.1) if

$$\begin{aligned} 0< \inf _{ t \in [0, T] }\int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (t, d x ) \ \text { and } \ \sup _{ t \in [0, T] }\int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (t, d x ) < \infty \end{aligned}$$

and if \(\Phi \) satisfies

$$\begin{aligned}&\int _{\mathbb R_*} \varphi (x) \dot{\Phi }(t, d x)- \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (x) \Phi (t, d x) + \frac{2}{1-\gamma } \int _{\mathbb R_*} \varphi '(x) x \Phi (t, d x) \\&\qquad - \frac{1}{\int _{\mathbb R_*}x^{\gamma +\lambda }\Phi (t,d x)} \int _{\mathbb R_*} \varphi '(x) x^{\gamma + \lambda }\Phi (t, d x) \nonumber \\&\quad =\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x,y) \left[ \varphi (x+y)- \varphi (x) - \varphi (y) \right] \Phi (t, d x) \Phi (t, d y), \nonumber \end{aligned}$$
(5.4)

for every test function \(\varphi \in C^1_c(\mathbb R_*).\)

Given the positive numbers \(k_1, k_2, \rho ^*\), we define the subset \(S(k_1, k_2, \rho ^*) \) of \({\mathcal {M}}_+(\mathbb R_*) \) as

$$\begin{aligned} S(k_1, k_2, \rho ^*):=\left\{ H \in {\mathcal {M}}_+(\mathbb R_*): \quad \begin{aligned}&H((0, \rho ^*))=0, \int _{\mathbb R_*} xH(d x ) =1, \\&k_1 \le \int _{\mathbb R_*} x^{\gamma +\lambda } H(d x ) \le k_2 \end{aligned} \right\} .\nonumber \\ \end{aligned}$$
(5.5)

Theorem 5.3

Assume \(K_{R}\) to be a truncated kernel defined as in (5.3) as a function of a homogeneous symmetric kernel K satisfying (1.5), (1.6), with parameters \(\gamma , \lambda \in \mathbb R\) such that (1.7) holds and \(\gamma + 2\lambda \ge 1 \) and \(\gamma > -1.\) Then for all \(k_1, k_2>0\) and sufficiently large \(R_0>0\) such that \(\Phi _{0} \in S\left( 2 k_1, \frac{k_2}{2}, \rho (k_2)\right) \cap \left\{ H \in {\mathcal {M}}_+(\mathbb R_*): H((4R, \infty ) )=0 \right\} \) for \(R >R_0\) and for \(\rho (k_2)\) given by (4.2), there exists a unique solution \(\Phi \in C^{1}([0,T];{\mathcal {M}}_{+}(\mathbb R_*))\) of Eq. (5.1) in the sense of Definition 5.2, for \(T>0\) sufficiently small. For every \( t \in [0, T]\), we have that \(\Phi (t, \cdot ) \in S\left( k_1, k_2, \rho ( k_2 )\right) \cap \left\{ H \in {\mathcal {M}}_+(\mathbb R_*): H((4R, \infty ) )=0 \right\} .\)

Lemma 5.4

Let \(T>0\) and let \(R>0\). Assume \(\gamma < 1 \) and \(\gamma + \lambda <1.\) Assume \(\alpha \in C([0, T]) \) satisfying

$$\begin{aligned} 0< k_1 \le \alpha (t) \le k_2 < \infty , \ \text { for all } \ t \in [0, T], \end{aligned}$$
(5.6)

for some positive constants \(k_1\) and \(k_2\). Consider the ODE

$$\begin{aligned} \frac{d x(t) }{dt}= V(x(t), \alpha (t)), \quad x(0)=x_0 >0, \end{aligned}$$
(5.7)

with \(V(x, \alpha ):=\frac{x^{\gamma +\lambda } }{\alpha }- \frac{2}{1-\gamma } x \). If \(x_0\ge \rho (k_2)\) with \(\rho (k_2)\) given by (4.2), then (5.7) has a unique solution \(X(t,x_0, \alpha ) \ge \rho (k_2)\). Let \(X(t, x_{0,1}, \alpha _1)\) and \(X(t, x_{0,2}, \alpha _2)\) be the solutions of Eq. (5.7) with respect to the functions \(\alpha _1\) and \(\alpha _2\) satisfying (5.6) and \(x_{0,1},x_{0,2} \in [\rho (k_2),4R]\). Then

$$\begin{aligned} |X(t,x_{0,1}, \alpha _1) - X(t,x_{0,2}, \alpha _2) |\le L_1 |x_{0,1}-x_{0,2}| + T L_2 \Vert \alpha _1-\alpha _2 \Vert _{[0,T]},\nonumber \\ \end{aligned}$$
(5.8)

for every \(t \in [0, T]\), where \(L_1=L_1(T, \gamma , \lambda ,k_1, k_2, R)>0,\) and \( L_2=L_2(T, \gamma , \lambda ,k_1, k_2, R)>0\) and where we denote by \(\Vert \cdot \Vert _{[0,T]}\) the norm \(\Vert f \Vert _{[0,T]}:= \sup _{ t \in [0, T]} | f(t) |\), for \(f\in C([0,T])\).

Proof

Since \(V(x, \alpha )>0\), for every \(x \le \rho (k_2)\), we deduce that the set \(\{ x \ge \rho (k_2)\} \) is an invariant region of (5.7). This also implies that a unique solution exists.

Equation (5.7) is a Bernoulli equation that can be reduced to the linear ODE

$$\begin{aligned} \frac{ dy }{ dt} =\frac{ 1-(\gamma +\lambda ) }{\alpha (t)} - \frac{2(1-(\gamma +\lambda ))}{1-\gamma } y, \quad y_0:= x_0^{1-(\gamma +\lambda )} \end{aligned}$$

via the change of variable \(y=x^{1-(\lambda + \gamma )}\). We deduce that

$$\begin{aligned} Y(t,x_0^{1 -(\gamma +\lambda ) }, \alpha )= x_0^{1 -(\gamma +\lambda ) } e^{- \frac{2(1-(\gamma +\lambda ))}{1-\gamma } t } + \int _0^t \frac{(1-\gamma -\lambda )}{\alpha (s) } e^{- \frac{2(1-(\gamma +\lambda ))}{1-\gamma } (t-s) } d s. \end{aligned}$$

Since \(\alpha _1, \alpha _2 \) satisfy (5.6) and since \(x_{0,1}, x_{0,2} \in [\rho (k_2), 4R]\), the above formula implies that

$$\begin{aligned} \left| Y(t,x_{0,1}^{1 -(\gamma +\lambda ) }, \alpha _1)-Y(t,x_{0,2}^{1 -(\gamma +\lambda ) }, \alpha _2)\right| \le&\left| x_{0,1}^{1 -(\gamma +\lambda ) } - x_{0,2}^{1 -(\gamma +\lambda ) }\right| \\&+ C(T,\gamma ,\lambda , k_1) \Vert \alpha _1 -\alpha _2 \Vert _{[0, T]}. \end{aligned}$$

Since the set \( \{x \ge \rho (k_2)\} \) is invariant for (5.7), the above inequality, together with the definition of y as a function of x, implies (5.8). \(\square \)

We will solve Eq. (5.1) using Lagrangian coordinates. To this end we introduce the notation

$$\begin{aligned} \chi ^\alpha _{\varphi }(x,y,t):=\varphi (\ell _{\alpha }(x,y,t))- \varphi (x) - \varphi (y), \end{aligned}$$

for \(\varphi \in C_c(\mathbb R_*)\), where \(\ell _\alpha \) is defined for \(x >0\), \(y>0\) and \(t \ge 0\) as

$$\begin{aligned} X(t, \ell _\alpha (x,y,t), \alpha )=X(t,x, \alpha ) + X(t,y, \alpha ). \end{aligned}$$
(5.9)

The function \(\ell _\alpha \) is well defined because for every fixed time \(t\ge 0\) the function \(x \mapsto X(t,x, \alpha )\) is increasing.

Lemma 5.5

Let \(K_R\) be a truncated kernel defined as in (5.3) as a function of a homogeneous symmetric kernel K satisfying (1.5), (1.6), with \(\gamma , \lambda \in \mathbb R\) satisfying (1.7). Let \(k_1, k_2, R\) be three positive constants. Assume that the initial condition \(\Phi _0\) is such that \( 2k_1 \le \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi _0(d x ) \le \frac{k_2}{2}\) and that \(\Phi _0((0,\rho (k_2)) \cup (4R, \infty ))=0\) for \(\rho (k_2)\) given by (4.2).

For a sufficiently small time \(T>0\), depending on R, there exists a function \(F \in C([0, T], {\mathcal {M}}_{+}(\mathbb R_*))\), with \(F (0, \cdot )=\Phi _0\), that satisfies

$$\begin{aligned}&\int _{\mathbb R_*} \varphi (x) F(t, d x) = \int _{\mathbb R_*} \varphi (x) \Phi _0( d x) \\&\quad + \!\frac{1}{2} \int _0^t \int _{\mathbb R_*} \!\int _{\mathbb R_*}\! K_R(X(s,x, \alpha ),X(s,y,\alpha ))e^{\frac{1+\gamma }{1-\gamma }s}\chi ^\alpha _{\varphi }(x,y,s)F(s, d x) F(s, d y) d s, \nonumber \end{aligned}$$
(5.10)

for every \(\varphi \in C_c(\mathbb R_*)\), with

$$\begin{aligned} \alpha (t):=e^{\frac{1+\gamma }{1-\gamma } t}\int _{\mathbb R_*} \left( X(t,x, \alpha )\right) ^{\gamma +\lambda } F(t, d x ), \quad \forall t \in [0,T]. \end{aligned}$$
(5.11)

The function F is such that \( k_1 \le \alpha (t) \le k_2\) and such that

$$\begin{aligned} F(t,(0,\rho (k_2) )\cup (4R, \infty ))=0, \quad \forall t \in [0,T]. \end{aligned}$$
(5.12)

Proof

We define the set

$$\begin{aligned} \tilde{X}:= \left\{ H \in {\mathcal {M}}_{+}(\mathbb R_*): H((0,\rho (k_2)) \cup (4R, \infty ))=0\right\} \end{aligned}$$

and the set

$$\begin{aligned} X_T:=\left\{ (F,\alpha ) \in C([0,T]; {\mathcal {M}}_{+}(\mathbb R_*)) \times C([0,T]): \begin{aligned}&F(t,\cdot ) \in \tilde{X}, k_1 \le \alpha (t) \le k_2, \forall t \in [0, T],\\&\sup _{t\in [0,T]}\int _{\mathbb {R}_{*}}F(t,d x)\le 1+\int _{\mathbb {R}_{*}}\Phi _{0}(d x) \end{aligned} \right\} . \end{aligned}$$

We endow the set \({\mathcal {M}}_{+}(\mathbb R_*) \) with the Wasserstein metric \(W_{1}\). The reason for this choice will become clear in the proof of the contractivity of the evolution operator corresponding to (5.10), (5.11). We endow the set \(C([0,T], {\mathcal {M}}_+(\mathbb R_*))\) with the metric induced by the distance

$$\begin{aligned} d_T(\mu , \nu ):=\sup _{t \in [0,T] } W_1(\mu (t, \cdot ), \nu (t, \cdot )). \end{aligned}$$

Similarly, we endow C([0, T]) with the norm \(\Vert f \Vert _{[0,T]}:= \sup _{ t \in [0, T]} | f(t) |\).

For every \((F, \alpha ) \in X_T\), we define the operator \({\mathcal {T}}[F,\alpha ](t): C_c(\mathbb R_*) \rightarrow \mathbb R_{+}\) as \({\mathcal {T}} [F,\alpha ](t):= {\mathcal {T}}_1[F,\alpha ](t)+{\mathcal {T}}_2[F,\alpha ](t)\) with

$$\begin{aligned} \langle {\mathcal {T}}_1[F,\alpha ](t), \varphi \rangle =&\, \int _{\mathbb R_*} \varphi (x) e^{-\int _{0}^{t}a[F,\alpha ](s,x)d s} \Phi _0( d x),\\ a[F,\alpha ](t,x):=&\,e^{\frac{\gamma +1 }{1-\gamma }t} \int _0^\infty K_R(X(t,x,\alpha ), X(t,y,\alpha )) F(t, d y),\\ \langle {\mathcal {T}}_2[F,\alpha ](t), \varphi \rangle =&\,\frac{1}{2} \int _0^t \int _{\mathbb R_*} \int _{\mathbb R_*} e^{- \int _s^t a[F](v,x) dv} K_R(X(s,x,\alpha ),X(s,y, \alpha ))\\ {}&\times e^{\frac{1+\gamma }{1-\gamma }s}\varphi (\ell _\alpha (x,y,s)) \cdot F(s, d x) F(s, d y)d s. \end{aligned}$$

Moreover, given \((F, \alpha ) \in X_T \), we define the operator \({\mathcal {A}}[F,\alpha ]: [0, T] \rightarrow \mathbb R_*\) as

$$\begin{aligned} {\mathcal {A}}[F,\alpha ](t):= e^{\frac{1+\gamma }{1-\gamma } t}\int _{\mathbb R_*} \left( X(t,x, \alpha )\right) ^{\gamma +\lambda } F(t, d x ), \quad \forall t \in [0,T]. \end{aligned}$$

We can now rewrite (5.10), (5.11) in a fixed point form, namely as \((F,\alpha )={\mathcal {F}}[F, \alpha ]\) where \({\mathcal {F}}[F,\alpha ]:= ({\mathcal {T}}[F,\alpha ], {\mathcal {A}}[{\mathcal {T}}[F,\alpha ], \alpha ])\).

We prove now that \({\mathcal {F}}: X_T \rightarrow X_T\). Since for every \((F,\alpha ) \in X_T\), the operator \({\mathcal {T}}[F,\alpha ](t)\) is linear and continuous, we deduce that it can be identified with an element of \({\mathcal {M}}_{+, b} (\mathbb R_*).\) Moreover, the operator \(t \mapsto {\mathcal {T}}[F,\alpha ](t)\) is a continuous map from \(\mathbb R_{+}\) to \({\mathcal {M}}_{+, b} (\mathbb R_*)\), hence \({\mathcal {T}} [F,\alpha ] \in C([0, T], {\mathcal {M}}_{+,b}(\mathbb R_*))\). Similarly, for every \((F,\alpha ) \in X_T \), we have that \({\mathcal {A}}[F, \alpha ] \in C([0,T])\).

Consider a test function \(\varphi \) such that \(\varphi (x)=0\) for every \( x \in [\rho (k_2), \infty )\). Then, since \(\Phi _0((0,\rho (k_2)))=0\), we deduce that

$$\begin{aligned} \langle {\mathcal {T}}_1[F,\alpha ](t), \varphi \rangle = \int _{[\rho (k_2), \infty )}\varphi (x) e^{-\int _{0}^{t}a[F,\alpha ](s,x)d s} \Phi _0(d x) =0. \end{aligned}$$

Similarly, since \(\ell _\alpha (s,x,y) \ge x \) and since \((F,\alpha ) \in X_T\), we deduce that

$$\begin{aligned} \langle {\mathcal {T}}_2[F,\alpha ](t), \varphi \rangle&\le \frac{C(R)}{2} \int _0^t \int _{[\rho (k_2), \infty )} \int _{[\rho (k_2), \infty )} e^{\frac{1+\gamma }{1-\gamma }s}\varphi (\ell _\alpha (x,y,s))\\ {}&\quad \times F(s, d x) F(s, d y)d s=0. \end{aligned}$$

Therefore, \(({\mathcal {T}}[F,\alpha ])(t,(0,\rho (k_2)))=0\), for every \(t \in [0, T].\)

Consider any test function \(\varphi \) such that \(\varphi (x)=0\) in [0, 4R]. Since \(\Phi _0((4R, \infty ))=0\), we have that

$$\begin{aligned} \langle {\mathcal {T}}_1[F,\alpha ](t), \varphi \rangle \le \int _{(0, 4 R]} \varphi (x) e^{-\int _{0}^{t}a[F,\alpha ](s,x)d s} \Phi _0(d x) =0. \end{aligned}$$

Moreover, using the notation \(S_1(s):=\{ (x,y) \in \mathbb R_*^2: X(s,x,\alpha ) \le R, X(s,y,\alpha ) \le R \} \), \(S_2(s):=\{ (x,y) \in \mathbb R_*^2: X(s,x,\alpha ) > R, X(s,y,\alpha ) \le R \} \) and \(S_3(s):=\{ (x,y) \in \mathbb R_*^2: X(s,x,\alpha )> R, X(s,y,\alpha ) > R \} \), we have

$$\begin{aligned}&\langle {\mathcal {T}}_2[F,\alpha ](t), \varphi \rangle \\&\quad \le \frac{1}{2} \int _0^t \iint _{S_1(s)} e^{\frac{1+\gamma }{1-\gamma }s} K_R(X(s,x,\alpha ),X(s,y,\alpha ))\varphi (\ell _\alpha (x,y,s))F(s, d x) F(s,d y)d s\\&\qquad + \frac{1}{2} \int _0^t \iint _{S_2(s)} e^{\frac{1+\gamma }{1-\gamma }s} K_R(X(s,x,\alpha ),X(s,y,\alpha ))\varphi (\ell _\alpha (x,y,s))F(s, d x) F(s,d y)d s \\&\qquad + \frac{1}{2} \int _0^t \iint _{S_3(s)} e^{\frac{1+\gamma }{1-\gamma }s} K_R(X(s,x,\alpha ),X(s,y,\alpha ))\varphi (\ell _\alpha (x,y,s))F(s, d x) F(s,d y)d s. \end{aligned}$$

For xy such that \(X(s,x,\alpha ) \ge R\) or \(X(s,y,\alpha ) \ge R\), we have that \(K_R(X(s,x,\alpha ),X(s,y,\alpha ))=0\). Thus, the second and the third terms above are equal to zero.

To see that the first term is equal to zero, notice that \(X(t, x, \alpha ) \ge x e^{- \frac{2}{1-\gamma } t}\). We can thus select T sufficiently small so that \(X(t, x, \alpha ) \ge \frac{x}{2}\). This implies that \( \frac{1}{2} \ell _\alpha (x,y,s) \le X(s, \ell _\alpha (x,y,s), \alpha )= X(s, x, \alpha ) + X(s, y, \alpha ) \le 2 R \). Hence \(\ell _\alpha (x,y,s) \le 4R\). Since \(\varphi (x)=0\) for every \(x \le 4 R\), the desired conclusion follows.

We now prove that \(k_1 \le {\mathcal {A}}[ {\mathcal {T}}[F,\alpha ],\alpha ](t) \le k_2\). To this end notice that

$$\begin{aligned}&\left| {\mathcal {A}}[{\mathcal {T}}[F,\alpha ],\alpha ](t) - \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi _0(d x) \right| \\&\quad \le \int _{\mathbb R_*} \left| e^{\frac{1+\gamma }{1-\gamma } t } (X(t,x,\alpha ))^{\gamma +\lambda }-x^{\gamma +\lambda } \right| {\mathcal {T}}[F,\alpha ](t, d x) \\&\qquad + \int _{\mathbb R_*} x^{\gamma +\lambda } \left| {\mathcal {T}}[F,\alpha ](t, d x) -\Phi _0(d x) \right| . \end{aligned}$$

Notice that

$$\begin{aligned} \left| e^{\frac{1+\gamma }{1-\gamma } t } (X(t,x,\alpha ))^{\gamma +\lambda }-x^{\gamma +\lambda } \right|&\le e^{\frac{1+\gamma }{1-\gamma } t } \left| (X(t,x,\alpha ))^{\gamma +\lambda }-x^{\gamma +\lambda } \right| \\ {}&\quad + \frac{1+\gamma }{1-\gamma } T x^{\gamma +\lambda }\\ {}&\le T C(T, R, k_2) . \end{aligned}$$

Finally, we have that the term

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma +\lambda } \left| {\mathcal {T}}[F,\alpha ](t, d x) - \Phi _{0}( d x) \right| \end{aligned}$$

can be arbitrarily small by taking T small. This is due to the definition of the map \(T[F,\alpha ]\) and the upper bound for F in the definition of the space \(X_{T}\).

We deduce that, for small time T, we have that \( \left| {\mathcal {A}}[F,\alpha ](t) - \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi _0(d x ) \right| \le C(T),\) for every \(t \in [0, T] \), where the constant C(T) can be made arbitrarily small.

We now check that the map \({\mathcal {F}}\) is a contraction. To this end, we use the fact that \(|e^{-x_{1}}-e^{-x_{2}}|\le |x_{1}-x_{2}|,\) for \(x_{1},x_{2}\ge 0.\) Consider \((F,\alpha ), (G,\beta ) \in X_{T}\). Then, for every Lipschitz function \(\varphi \) with \(||\varphi ||_{Lip }\le 1\), we have that

$$\begin{aligned}&\bigg |\int _{\mathbb R_*} \varphi (x) [{\mathcal {T}}_1 [F,\alpha ]-{\mathcal {T}}_{1}[G,\beta ]](t,d x)\bigg |\\&\quad \le \int _{\mathbb R_*} \int _0^t \left| \varphi (x) \left[ a[F,\alpha ](s,x)- a[G,\beta ](s,x) \right] \right| d s \Phi _0(d x) \\&\quad \le \int _{\mathbb R_*} \int _0^t \left| \varphi (x) \left[ \int _{\mathbb R_*} K_R(X(s,x,\alpha ),X(s,y,\alpha )) F(s, d y) \right. \right. \\&\qquad -\left. \left. \int _{\mathbb R_*} K_R(X(s,x,\beta ),X(s,y,\beta )) G(s, d y) \right] d s \Phi _0(d x) \right| \\&\quad \le I_1[\varphi ] + I_2[\varphi ], \end{aligned}$$

where

$$\begin{aligned} I_1[\varphi ]&:=\int _{\mathbb R_*} \int _0^t \bigg | \varphi (x) \int _{\mathbb R_*} [ K_R(X(s,x,\alpha ),X(s,y,\alpha )) \\&\quad - K_R(X(s,x,\beta ),X(s,y,\beta )) ] F(s, d y) \bigg | d s \Phi _0(d x) \end{aligned}$$

and

$$\begin{aligned} I_2[\varphi ]:= \int _{\mathbb R_*} \int _0^t \left| \varphi (x) \int _{\mathbb R_*} K_R(X(s,x,\beta ),X(s,y,\beta )) \left( F(s, d y)- G(s, d y) \right) d s \Phi _0(d x ) \right| . \end{aligned}$$

Then, the differentiability of the kernel \(K_R\), together with inequality (5.8), implies that

$$\begin{aligned}&\sup _{||\varphi ||_{Lip \le 1}} I_1[\varphi ] \le C(R) T \Vert \Phi _0\Vert _{TV} (1+ \Vert \Phi _0\Vert _{TV}) \Vert \alpha -\beta \Vert _{[0,T]}. \end{aligned}$$

Similarly,

$$\begin{aligned}&\sup _{||\varphi ||_{Lip \le 1}} I_2[\varphi ] \le C(R) T \Vert \Phi _0\Vert _{TV} d_T(F,G). \end{aligned}$$

Using again the differentiability of the kernel \(K_R\), inequality (5.8) and the definition of \(\ell _\alpha \), we obtain that

$$\begin{aligned}&\sup _{||\varphi ||_{Lip \le 1}} \bigg |\int _{\mathbb R_*} \varphi (x)[ {\mathcal {T}}_2 [F,\alpha ](t,d x)-{\mathcal {T}}_{2}[G,\beta ](t,d x)]\bigg | \\&\quad \le C(T,R)(1+||\Phi _{0}||_{TV})^{2} d_T(F,G) \\&\qquad +C(T,R)(1+||\Phi _{0}||_{TV})^{2} \Vert \alpha -\beta \Vert _{[0,T]} , \end{aligned}$$

where T can be selected so that \(C(T,R)(1+||\Phi _{0}||_{TV})^{2} \le \frac{1}{4} \).

We remark that, in order to obtain this bound, we used the properties of the Wasserstain distance. Similar estimates cannot be obtained by endowing \({\mathcal {M}}_+(\mathbb R_*)\) with the total variation norm.

To understand this better, let \(\alpha ,\beta \in C([0,T])\), with \(k_{1}\le \alpha ,\beta \le k_{2}\). Let \(s\in [0,T]\). Using the fact that \(||\varphi ||_{Lip }\le 1\) and that we can assume we work on a sufficiently nice compact set due to the support of the measures, we can derive the estimate

$$\begin{aligned} |\varphi (\ell _\alpha (x,y,s))-\varphi (\ell _\beta (x,y,s))|&\le |\ell _\alpha (x,y,s)-\ell _\beta (x,y,s)|\\&\lesssim |X(s,\ell _\alpha (x,y,s),\alpha )-X(s,\ell _\beta (x,y,s),\alpha )|\\&\le |X(s,\ell _\alpha (x,y,s),\alpha )-X(s,\ell _\beta (x,y,s),\beta )|\\&\quad +|X(s,\ell _\beta (x,y,s),\beta )-X(s,\ell _\beta (x,y,s),\alpha )|. \end{aligned}$$

Using (5.8) and (5.9), we have

$$\begin{aligned} |X(s,\ell _\alpha (x,y,s),\alpha )-X(s,\ell _\beta (x,y,s),\beta )|&\le |X(s,x,\alpha )-X(s,x,\beta )|\\&\quad + |X(s,y,\alpha )-X(s,y,\beta )|\\&\le 2TL_{2}||\alpha -\beta ||_{[0,T]} \end{aligned}$$

and thus

$$\begin{aligned} |\varphi (\ell _\alpha (x,y,s))-\varphi (\ell _\beta (x,y,s))|\lesssim T ||\alpha -\beta ||_{[0,T]}. \end{aligned}$$
(5.13)

More precisely, inequality (5.13) was needed to prove the contractivity of the map \(\mathcal {T}_{2}\).

In order to prove the upper bound for the map \(T[F,\alpha ],\) we first observe that

$$\begin{aligned} \langle {\mathcal {T}}_1[F,\alpha ](t), 1 \rangle = \int _{[\rho (k_2), \infty )} e^{-\int _{0}^{t}a[F,\alpha ](s,x)d s} \Phi _0(d x) \le ||\Phi _{0}||_{TV} \end{aligned}$$

and then that

$$\begin{aligned} \langle {\mathcal {T}}_2[F,\alpha ](t), 1 \rangle&\le \frac{C(R)}{2}e^{\frac{1+\gamma }{1-\gamma }t} \int _0^t \int _{[\rho (k_2), \infty )} \int _{[\rho (k_2), \infty )} F(s, d x) F(s, d y)d s\\&\le \frac{C(R)}{2}e^{\frac{1+\gamma }{1-\gamma }T}T(1+||\Phi _{0}||_{TV})^{2}\le 1, \end{aligned}$$

for T sufficiently small. Thus,

$$\begin{aligned} \sup _{t\in [0,T]}\langle {\mathcal {T}}[F,\alpha ](t), 1 \rangle \le 1+||\Phi _{0}||_{TV}. \end{aligned}$$

Finally,

$$\begin{aligned}&\Vert {\mathcal {A}} [{\mathcal {T}}[F,\alpha ], \alpha ]- {\mathcal {A}} [{\mathcal {T}}[G,\beta ], \beta ] \Vert _{[0,T]} \\&\quad \le C(R) d_T([{\mathcal {T}}[F,\alpha ],{\mathcal {T}}[G,\beta ])\\&\qquad + C_1(T,R)T(1+||\Phi _{0}||_{TV}) \Vert \alpha - \beta \Vert _{[0,T]} \\&\quad \le C(T,R)T(1+||\Phi _{0}||_{TV})^{2} [d_T(F,G)+\Vert \alpha - \beta \Vert _{[0,T]}], \end{aligned}$$

where the constant \(TC(T,R)(1+||\Phi _{0}||_{TV})^{2}\) can be made smaller than \(\frac{1}{4}\) by selecting T small. The operator \({\mathcal {F}}\) is therefore a contraction and, by Banach fixed point theorem, we conclude that it has a unique fixed point \((F, \alpha ) \in X_T\). \(\square \)

Proof of Theorem 5.3

Assume \(T>0\) to be as in Lemma 5.5. Let \((F, \alpha ) \in C([0,T], {\mathcal {M}}_+(\mathbb R_*)) \) \(\times C([0,T])\) be the solution of (5.10), (5.11). Let \(X(t,y, \alpha )\) be the solution of Eq. (5.7). Let \( \Phi \in C^1([0,T], {\mathcal {M}}_+(\mathbb R_*))\) be the function defined by duality as

$$\begin{aligned} \int _{\mathbb R_*} \varphi (x) \Phi (t, d x)=\int _{\mathbb R_*} \varphi (X(t,y, \alpha )) e^{\frac{1+\gamma }{1-\gamma }t} F(t, d y), \end{aligned}$$
(5.14)

for every \(\varphi \in C_c(\mathbb R_*)\). We prove that the function \(\Phi \) is such that

$$\begin{aligned} \Phi (t,(0, \rho (k_2)) \cup (4R, \infty ))=0, \quad \forall t \in [0,T], \end{aligned}$$
(5.15)

for \(\rho (k_2)\) defined as in (4.2). To this end notice that, since \(F(t,(0,\rho (k_{2}))=0\), for every \(t\in [0,T]\), and since \( X(t, y, \alpha ) \ge \rho (k_2)\), for every \(y \ge \rho (k_2)\), we have that

$$\begin{aligned} \int _{\mathbb R_* } \varphi (x) \Phi (t, d x) =\int _{[\rho (k_2),\infty )} \varphi (X(t,y, \alpha )) e^{\frac{1+\gamma }{1-\gamma }t} F(t, d y)=0, \end{aligned}$$

for every test function \(\varphi \) such that \(\varphi (x)=0\) if \(x \ge \rho (k_2)\). Similarly, since \(t \mapsto X(t, x, \alpha ) \) is a decreasing function for large values of x, we deduce that \( X(t, x, \alpha )\le 4R\), for every \(x \le 4R \). Hence,

$$\begin{aligned} \int _{\mathbb R_* } \varphi (x) \Phi (t, d x) =\int _{[\rho (k_2), 4 R]} \varphi (X(t,y, \alpha )) e^{\frac{1+\gamma }{1-\gamma }t} F(t, d y)=0, \end{aligned}$$

for every test function \(\varphi \) such that \(\varphi (x)=0\) if \(x \le 4 R\).

We now prove that \(\Phi \) satisfies (5.4). By its definition as the fixed point of the operator \({\mathcal {F}}\), we deduce that \(F \in C^1([0, T], {\mathcal {M}}_+(\mathbb R_*))\). Differentiating both sides of the equality (5.14) in time, we deduce that

$$\begin{aligned} \frac{d }{d t} \int _{\mathbb R_*} \varphi (x) \Phi (t, d x)&= \int _{\mathbb R_*} \varphi (X(t, y, \alpha )) e^{\frac{1+\gamma }{1-\gamma } t } \partial _t F(t,d y) \\&\quad +\int _{\mathbb R_*} \varphi '(X(t, y,\alpha )) \bigg [ \frac{(X(t,y,\alpha ))^{\gamma +\lambda }}{\alpha (t)} \\&\quad -\frac{2}{1-\gamma } X(t,y,\alpha ) \bigg ] e^{\frac{1+\gamma }{1-\gamma } t }F(t, d y) \\&\quad + \frac{1+\gamma }{1-\gamma }\int _{\mathbb R_*} \varphi (X(t, y, \alpha )) e^{\frac{1+\gamma }{1-\gamma } t } F(t,d y). \end{aligned}$$

Using the fact that F satisfies Eq. (5.10), we deduce that

$$\begin{aligned}{} & {} \int _{\mathbb R_*} \varphi (X(t, y, \alpha )) e^{\frac{1+\gamma }{1-\gamma } t } \partial _t F(t,d y) \\{} & {} \qquad = \frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*}e^{ \frac{2(1+\gamma )}{1-\gamma }t} F(t,d x) F(t,d y)K_R(X(t,x,\alpha ),X(t,y,\alpha ))\\{} & {} \quad \qquad \times [ \varphi (X(t, x,\alpha )+X(t, y,\alpha ))- \varphi (X(t,y,\alpha ))- \varphi (X(t, x,\alpha ))]. \end{aligned}$$

Hence, using again the equality (5.14), we have that

$$\begin{aligned} \frac{d }{d t} \int _{\mathbb R_*} \varphi (x) \Phi (t,d x){} & {} = \int _{\mathbb R_*} \varphi '(X(t, y,\alpha )) \left[ \frac{(X(t,y,\alpha ))^{\gamma +\lambda }}{\alpha (t)}\right. \\{} & {} \quad \left. -\frac{2}{1-\gamma } X(t,y,\alpha )\right] e^{\frac{1+\gamma }{1-\gamma } t } F(t,d y) \\{} & {} \quad + \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (X(t, y,\alpha )) e^{\frac{1+\gamma }{1-\gamma } t } F(t,d y) \\{} & {} \quad + \frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(X(t,x,\alpha ),X(t,y,\alpha )) e^{ \frac{2(1+\gamma )}{1-\gamma }t}F(t,d x) F(t,d y) \\{} & {} \cdot \left[ \varphi (X(t, x,\alpha )+X(t, y,\alpha ))- \varphi (X(t,y,\alpha ))- \varphi (X(t, x,\alpha )) \right] \\{} & {} = \int _{\mathbb R_*} \varphi '(y) \left[ \frac{y^{\gamma +\lambda }}{M_{\gamma +\lambda }(\Phi (t))} -\frac{2}{1-\gamma } y\right] \Phi (t,d y) \\{} & {} \quad + \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (y) \Phi (t,d y) \\{} & {} \quad + \frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x ,y)\left[ \varphi ( x+y)- \varphi (y) \right. \\{} & {} \quad \left. - \varphi (x) \right] \Phi (t,d x)\Phi (t,d y). \end{aligned}$$

We conclude that \(\Phi \) satisfies (5.4). \(\qquad \square \)

5.2 Existence of a stationary solution for the truncated equation

Theorem 5.6

Assume \(K_R \) to be a truncated kernel as in (5.3) defined as a function of the homogeneous symmetric kernel K that satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R\) such that (1.7) holds and such that \(\gamma + 2\lambda \ge 1\), \( \gamma >-1 \). There exists \({\overline{R}}>0\) such that, for every truncation parameter \(R> {\overline{R}}\), there exists a \(\Phi \in {\mathcal {M}}_{+}(\mathbb R_*)\) satisfying the equation

$$\begin{aligned}&\int _{\mathbb R_*} \varphi '(x) \left[ \frac{2}{1-\gamma }x - \frac{x^{\gamma + \lambda }}{\int _{\mathbb R_*}z^{\gamma +\lambda }\Phi (d z)} \right] \Phi (d x) - \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (x) \Phi (d x)\nonumber \\&\quad =\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x,y) \left[ \varphi (x+y)- \varphi (x) - \varphi (y) \right] \Phi ( d x) \Phi (d y), \end{aligned}$$
(5.16)

for every \(\varphi \in C^{1}_c(\mathbb R_*)\). The solution is such that

$$\begin{aligned} \int _{\mathbb R_*} x \Phi (d x) = 1, \quad \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (d x) \le C_{1} \quad and \quad \int _{\mathbb R_*} x^{2-\gamma -\lambda } \Phi (d x) \le C_{2},\nonumber \\ \end{aligned}$$
(5.17)

for some constants \(C_{1}, C_{2}>0\) that do not depend on the truncation R. Additionally, \(\Phi \) is such that \(\Phi ((0, \rho (C_1)))=0\) where \(\rho (C_1)\) is given by (4.2).

To prove the theorem we introduce the semigroup \(\{S(t)\}_{\{ t \ge 0\}}\) defined as \(S(0)\Phi _0 =\Phi _0\) and \(S(t) \Phi _0 = \Phi (t, \cdot )\), where \(\Phi \) is the solution of Eq. (5.4) constructed in the previous section. We want to apply Tychonoff fixed point theorem to prove the existence of a stationary solution for Eq. (5.1). To this end we need to find an invariant region for S(t) (Proposition 5.7) and prove the continuity in the weak-\(^{*}\) topology of the map \(\Phi \mapsto S(t)\Phi \) for every time t (Proposition 5.12). In this way we prove that for every time t there exists a fixed point \(\hat{\Phi }_t\) such that \(S(t)\hat{\Phi }_t = \hat{\Phi }_t.\) We will then prove that S is continuous (Proposition 5.8) in time and conclude that the limit as t tends to zero of \(\hat{\Phi }_t \) is a solution of Eq. (5.16). This is a standard method used in the study of coagulation equations to prove existence of self-similar profiles. For more details, see [8, Theorem 1.2].

Proposition 5.7

(Invariant region) Assume \(K_R \) to be a truncated kernel as in (5.3) defined as a function of the homogeneous symmetric kernel K that satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R\) such that (1.7) holds and such that \(\gamma + 2\lambda \ge 1\), \(\gamma >-1 \). Then there exist some positive constants \(C_1, C_2\) that do not depend on the truncation R such that the set

$$\begin{aligned} P:=\left\{ H \in {\mathcal {M}}_{+}(\mathbb R_*) : \quad \begin{aligned}&M_1(H )=1,\ H((0, \rho (C_1)) \cup (4R,\infty ) )=0, \\&\frac{1}{C_{2}} \le M_{\gamma + \lambda }(H) \le C_1 \end{aligned} \right\} \end{aligned}$$
(5.18)

is invariant under the evolution operator S(t),  where \(\rho (C_1)\) was defined in (4.2).

Proof

Let \(\Phi _0\) and \(K_R\) be as in Lemma 5.5. Let T be as in Lemma 5.5. We prove that there exist two constants \(C_{1},C_{2}>0\) that do not depend on the truncation parameter R such that the solution \(\Phi \) obtained in Theorem 5.3 is such that for every time \(t \in [0, T] \)

  1. (1)

    \(\int _{\mathbb R_* } x \Phi (t,d x) = 1\);

  2. (2)

    \(\int _{\mathbb R_* } x^{\gamma + \lambda } \Phi (t, d x) \le \max \{ C_1,\int _{\mathbb R_* } x^{\gamma + \lambda } \Phi _{0}( d x)\}\);

  3. (3)

    \(\int _{\mathbb R_* } x^{2- \gamma - \lambda } \Phi (t, d x) \le \max \{C_{2},\int _{\mathbb R_* } x^{2-\gamma -\lambda } \Phi _{0}( d x)\}\).

Additionally, we prove that we can conclude from (2) that \(\Phi (t,(0,\rho (C_{1}))=0\), for every \(t\in [0,T]\).

We first notice that, since \(\Phi (t, \cdot )\) has compact support, we can consider in Eq. (5.4) test functions \(\varphi \) such that \(\varphi (x)=x^{k} \), for \(k\in \mathbb {R}.\)

Let us prove (1). We consider in Eq. (5.4) a test function \(\varphi \) such that \(\varphi (x)=x \). We deduce that

$$\begin{aligned} \frac{d}{dt} \int _0^\infty \xi \Phi (t, d \xi )=1- \int _0^\infty \xi \Phi (t,d \xi ). \end{aligned}$$

Hence

$$\begin{aligned} \int _0^\infty \xi \Phi (t,\xi ) d \xi = 1+ \left( \int _{\mathbb R_*} \xi \Phi _0(d \xi ) -1 \right) e^{-t} \end{aligned}$$

which since \(\int _{\mathbb R_*} \xi \Phi _0(d \xi )=1 \) implies

$$\begin{aligned} \int _{\mathbb R_*}\xi \Phi (t, d \xi ) = 1. \end{aligned}$$
(5.19)

We prove (2). We start by proving this for \(\gamma + 2 \lambda >1. \) Consider a test function \(\varphi \) such that \(\varphi (x)=x^{\gamma +\lambda }\) in Eq. (5.4). Then, we can see that the moment \(M_{\gamma + \lambda }\) satisfies the following

$$\begin{aligned}&\partial _{t}M_{\gamma +\lambda } \le - \frac{2\lambda + \gamma -1 }{1-\gamma } M_{\gamma +\lambda } + \frac{(\gamma +\lambda )}{M_{\gamma +\lambda }} M_{2(\gamma +\lambda )-1} \\ {}&+ c_{3}\int _{[\rho (2C_{1}), \frac{R}{4}] } \int _{[\rho (2C_{1}), \frac{R}{4} ] } (x^{\gamma + \lambda } y^{-\lambda } + x^{-\lambda } y^{\gamma + \lambda } )\\ {}&\quad \times [ (x+y)^{\gamma + \lambda } - x^{\gamma + \lambda } - y^{\gamma + \lambda }] \Phi (t, d x ) \Phi (t, d y ) \end{aligned}$$

since, due to the definition of \(K_{R}\) in (5.3), we have that on the set \([\rho (2C_{1}),\frac{R}{4}]^{2}\), there exists a constant \(C>0\) such that \(K_{R}\ge \frac{K}{C}\).

We denote by \(z:=\frac{y}{x}\). Notice that since \(\gamma + \lambda >0\) when \(y \le x \) we have that

$$\begin{aligned}&\left[ x^{\gamma + \lambda } y^{-\lambda } + x^{-\lambda } y^{\gamma + \lambda } \right] \left[ (x+y)^{\gamma + \lambda } - x^{\gamma + \lambda } - y^{\gamma + \lambda } \right] \\&\quad \le 2 x^{2(\gamma + \lambda )} y^{-\lambda } [ (1+z)^{\gamma + \lambda } - 1- z^{\gamma + \lambda } ]\\&\quad \le 2 x^{2(\gamma + \lambda )} y^{-\lambda } \left( \left( \gamma + \lambda \right) z- z^{\gamma + \lambda } \right) \\&\quad \le 2 (\gamma + \lambda -1 ) y^{\gamma } x^{\gamma + \lambda }. \end{aligned}$$

As a consequence, by symmetry, we deduce that

$$\begin{aligned}&\partial _{t}M_{\gamma +\lambda } \lesssim - \frac{2\lambda + \gamma -1 }{1-\gamma } M_{\gamma +\lambda } + \frac{(\gamma +\lambda )}{M_{\gamma +\lambda }} M_{2(\gamma +\lambda )-1} - 2c_{3} \left( 1-\gamma -\lambda \right) M_{\gamma } M_{\gamma + \lambda }\nonumber \\&\qquad +2c_{3}(1-\gamma -\lambda )\int _{(\frac{R}{4},\infty )}x^{\gamma }\Phi (d x)M_{\gamma +\lambda }+2c_{3}(1-\gamma -\lambda )\int _{(\frac{R}{4},\infty )}x^{\gamma +\lambda }\Phi (d x)M_{\gamma }\nonumber \\&\quad \le - \frac{2\lambda + \gamma -1 }{1-\gamma } M_{\gamma +\lambda } + \frac{(\gamma +\lambda )}{M_{\gamma +\lambda }} M_{2(\gamma +\lambda )-1} -2c_{3} \left( 1-\gamma -\lambda \right) M_{\gamma } M_{\gamma + \lambda }\nonumber \\&\qquad +4^{1-\gamma }R^{\gamma -1}2c_{3}(1-\gamma -\lambda )M_{\gamma +\lambda }+c_{4}M_{\gamma }\nonumber \\&\quad = - \left[ \frac{2\lambda + \gamma -1 }{1-\gamma }-\tilde{\varepsilon }\right] M_{\gamma +\lambda } + \frac{1}{M_{\gamma +\lambda }} [ (\gamma +\lambda ) M_{2(\gamma +\lambda )-1} - 2c_{3} \left( 1-\gamma -\lambda \right) M_{\gamma } M_{\gamma + \lambda }^2\nonumber \\&\qquad +c_{4}M_{\gamma +\lambda }M_{\gamma }], \end{aligned}$$
(5.20)

with \(\tilde{\varepsilon }:=4^{1-\gamma }R^{\gamma -1}2c_{3}(1-\gamma -\lambda )\). Notice we can take \(\tilde{R}\) sufficiently large so that \(- \frac{2\lambda + \gamma -1 }{1-\gamma }+\tilde{\varepsilon }<0\), for every \(R\ge \tilde{R}\).

Since \(\gamma < 2(\gamma +\lambda ) -1 \le 1 \) we deduce, by interpolation, using that \(M_{1}(\Phi (t))=1\), for all \(t\in [0,T]\), that there exists two positive constants \(c_5\) and \(c_6\) such that

$$\begin{aligned} M_{ 2(\gamma +\lambda ) -1 } \le c_5 M_\gamma + c_6. \end{aligned}$$

Hence, since \(\gamma + \lambda >0\) then

$$\begin{aligned} \partial _{t} M_{\gamma +\lambda } \le&\,- \left[ \frac{2\lambda + \gamma -1 }{1-\gamma }-\tilde{\varepsilon }\right] M_{\gamma +\lambda } + \frac{1}{M_{\gamma +\lambda }} [ (\gamma +\lambda ) \left[ c_5 M_\gamma + c_6 \right] \\&- 2c_3 \left( 1-\gamma -\lambda \right) M_{\gamma } M_{\gamma + \lambda }^2 +c_{4}M_{\gamma +\lambda }M_{\gamma }]. \end{aligned}$$

Multiplying by \(M_{\gamma +\lambda }\) the inequality we deduce that

$$\begin{aligned} M_{\gamma +\lambda } \partial _{t} M_{\gamma +\lambda } \le&- \left[ \frac{2\lambda + \gamma -1 }{1-\gamma }-\tilde{\varepsilon }\right] M_{\gamma +\lambda }^2 + [ (\gamma +\lambda ) \left[ c_5 M_\gamma + c_6 \right] \\&- 2c_3 \left( 1-\gamma -\lambda \right) M_{\gamma } M_{\gamma + \lambda }^2+c_{4}M_{\gamma +\lambda }M_{\gamma }] \end{aligned}$$

which readjusting the constants implies

$$\begin{aligned}&\partial _{t} M_{\gamma +\lambda }^2 \le - c_3 M_{\gamma +\lambda }^2 + M_\gamma \left( c_4 +c_{5}M_{\gamma +\lambda } - c_6 M_{\gamma + \lambda }^2 \right) + c_7 \end{aligned}$$
(5.21)

for \(c_3, c_4, c_5, c_6,c_{7}>0.\) This implies that the set \(\left\{ \Phi \in {\mathcal {M}}_{+}(\mathbb R_*): M_{\gamma +\lambda } \le \frac{ c_5+\sqrt{c_5^{2}+4c_{4}c_{6}}}{2 c_{6} } \right\} \) is invariant when \(\gamma + 2 \lambda >1.\)

We now consider the case \(\gamma + 2 \lambda =1.\) First of all, notice that, for every \(M < R \), we have

$$\begin{aligned} \int _{[M, \infty ) } x^{\gamma + \lambda } \Phi (t, d x) \le M^{\gamma +\lambda -1 } \int _{[M, \infty )} x \Phi (t, d x ) \le M^{\gamma + \lambda -1}. \end{aligned}$$

Notice that an upper bound is obvious for the points \(t\in [0,T]\) for which \(M_{\gamma +\lambda }(\Phi (t))\le 1\). If there exist \(t_{1},t_{2}\in [0,T]\), \(t_{1}<t_{2}\), such that \(M_{\gamma +\lambda }(\Phi (t_{1}))<1\) and \(M_{\gamma +\lambda }(\Phi (t_{2}))>1\), by the continuity in time of \(\Phi ,\) we have that there exists \(\overline{t}\in [t_{1},t_{2}]\) such that \(M_{\gamma +\lambda }(\Phi (\overline{t}))=1\) and \(\varepsilon _{1}\in (0,1)\) such that \(M_{\gamma +\lambda }(\Phi (s))\ge 1\) on \([\overline{t},\overline{t}+\varepsilon _{1}]\).

On the interval \([\overline{t},\overline{t}+\varepsilon _{1}]\), we can apply the following logic.

We select R large enough so that we can select M to be such that \((1-\delta )^{\frac{1}{1-\gamma -\lambda }}< M < R\), but independent on R, and we deduce that there exists a \(\delta >0\) such that

$$\begin{aligned} \int _{(0, M ) } x^{\gamma + \lambda } \Phi (t,d x)&= \int _{\mathbb R_* } x^{\gamma + \lambda }\Phi (t, d x) - \int _{[M, \infty ) } x^{\gamma + \lambda } \Phi (t, d x) \\&\ge \int _{\mathbb R_* } x^{\gamma + \lambda }\Phi (t, d x)- M^{\gamma + \lambda -1 }\ge 1- M^{\gamma + \lambda -1 } \ge \delta . \end{aligned}$$

As a consequence we deduce that

$$\begin{aligned} M_\gamma \ge \int _{(0, M)} x^{\gamma } \Phi (t, d x ) \ge M^{-\lambda } \int _{(0,M)} x^{\gamma +\lambda } \Phi (t, d x) \ge \delta M^{- \lambda }. \end{aligned}$$
(5.22)

Substituting the test function \(\varphi (x)=x^{\gamma + \lambda } \) in Eq. (5.4), we deduce that there exists a constant \(c>0\) such that

$$\begin{aligned}&\partial _t M_{\gamma + \lambda } \le \frac{\gamma + \lambda }{M_{\gamma +\lambda } } M_{2(\gamma + \lambda ) -1 } - c (1-\gamma -\lambda ) M_{\gamma +\lambda } M_\gamma + \tilde{\epsilon } M_{\gamma +\lambda } + c M_\gamma , \end{aligned}$$

for \(\tilde{\varepsilon }\in (0,1)\), which can be made sufficiently small as before. Hence, since for suitable constants \(c_3, c_4 >0\) we have that \(M_{2(\gamma +\lambda )-1} \le c_3 M_\gamma + c_4\), similarly to (5.20), we deduce that

$$\begin{aligned}&\frac{1}{2}\partial _t M^2_{\gamma + \lambda } \le c_5 M_{\gamma } (\gamma + \lambda ) +(\gamma + \lambda ) c_6 - c_7 M^2_{\gamma + \lambda } M_\gamma +\tilde{\varepsilon } M_{\gamma +\lambda }^{2}+c_{8} M_\gamma M_{\gamma +\lambda }. \end{aligned}$$

Using (5.22) we deduce that there exists a constant \(c_{9}>0\) such that

$$\begin{aligned}&\frac{1}{2}\partial _t M^2_{\gamma + \lambda } \le M_{\gamma } \left[ c_5 (\gamma + \lambda ) +(\gamma + \lambda ) c_{9} +c_{8}M_{\gamma +\lambda }-(c_7-\tilde{\varepsilon }\delta ^{-1} M^{\lambda }) M^2_{\gamma + \lambda } \right] . \end{aligned}$$

Choosing now \(\tilde{\varepsilon }\) sufficiently small such that \(-c_7+\tilde{\varepsilon }\delta ^{-1} M^{\lambda }<0\), we have that there exists a constant \(C_{1}>0\) such that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma + \lambda } \Phi (t, d x ) \le C_{1}. \end{aligned}$$
(5.23)

We now prove that \(\left( S(t) \Phi _0 \right) ((0, \rho (C_1)))=0\). Notice that, by Lemma 5.5, we know that \(\left( S(t) \Phi _0 \right) ((0, \rho (2C_{1})))=0\), where we recall that \(\rho ( 2 C_1) \le \rho (C_1)\). Consider now a sequence of test functions \( \{\varphi _n \}\) such that, for every \(n\in \mathbb {N}\), \(\varphi _{n}\) is decreasing and supported in \((0, \rho (C_1))\). Since \(\varphi _n \) is decreasing, we deduce that the solution F of Eq. (5.10) satisfies

$$\begin{aligned} \int _{\mathbb R_*} \varphi _n(x) F(t, d x )&\le \int _{\mathbb R_*} \varphi _n(x) \Phi _{0}(d x )+ C(R) \int _{0}^{t} \int _{\mathbb R_*} \varphi _n(x) F(s, d x ) d s, \\&\quad \text { for every } n \in \mathbb N, \end{aligned}$$

where we used in addition that \(M_{1}(F)\) is bounded. Using Grönwall’s inequality, we deduce that

$$\begin{aligned} \int _{\mathbb R_*} \varphi _n(x) F(t, d x ) =0, \end{aligned}$$

for every \(t\in [0,T]\) and \(n\in \mathbb {N}\).

We then let \(\varphi _{n}\) converge pontwise to \(\chi _{(0, \rho (C_1))}\) and obtain that

$$\begin{aligned} F(t, (0, \rho (C_1))) =0, \end{aligned}$$

for every \(t\in [0,T]\).

Consider now a test function \(\varphi \) such that \(\varphi (x)=0\) for every \(x \ge \rho (C_1)\) in equality (5.14). Then

$$\begin{aligned} \int _{\mathbb R_* }\varphi (x) \Phi (t, d x )= e^{\frac{1+\gamma }{1-\gamma }t} \int _{[\rho (C_1), \infty ) }\varphi (X(t, y, \alpha ) ) F(t, d y). \end{aligned}$$

Using the fact that \(X(t, y, \alpha ) \ge \rho (C_1)\) for every \(y \ge \rho (C_1)\), we deduce that \(\Phi (t, (0, \rho (C_1)))=0\), for every \(t \in [0, T]\).

We conclude by proving (3). We consider a test function \(\varphi \) equal to \(x^{2-\gamma - \lambda } \) in Eq. (5.4) and deduce that

$$\begin{aligned}&\partial _t \int _{\mathbb {R}_{*}} \xi ^{2-\gamma - \lambda }\Phi (t,d \xi ) - \frac{3-3\gamma - 2 \lambda }{1-\gamma }\int _{\mathbb {R}_{*}}\xi ^{2-\gamma - \lambda }\Phi (t,d \xi ) - \frac{ 2-\gamma -\lambda }{M_{\gamma +\lambda }(\Phi (t)) } M_{1}(\Phi (t)) \nonumber \\&\quad = \frac{1}{2}\int _{[ \rho (C_1), \infty )}\int _{[ \rho (C_1), \infty )} K_R(\xi ,z) \left[ (z+\xi )^{2-\gamma -\lambda }- \xi ^{2-\gamma -\lambda }-z^{2-\gamma -\lambda } \right] \Phi (t,d \xi )\Phi (t,d z). \end{aligned}$$
(5.24)

By Cauchy–Schwarz inequality it follows that

$$\begin{aligned} \frac{1}{\int _{[ \rho (C_1), \infty )} \xi ^{\gamma +\lambda } \Phi (t,d \xi )} \le \frac{\int _{[ \rho (C_1), \infty )}\xi ^{2-\gamma - \lambda } \Phi (t,d \xi )}{ \left( \int _{[ \rho (C_1), \infty )}\xi \Phi (t,d \xi ) \right) ^2} = \int _{[ \rho (C_1), \infty )} \xi ^{2-\gamma - \lambda } \Phi (t,d \xi ).\nonumber \\ \end{aligned}$$
(5.25)

Plugging (5.25) into Eq. (5.24) and using the fact that the total mass is equal to 1 proved in (5.19), we deduce that

$$\begin{aligned}&\partial _t \int _{[ \rho (C_1), \infty )} \xi ^{2-\gamma - \lambda }\Phi (t,d \xi ) \le \left( \frac{3\gamma -3+2\lambda }{1-\gamma } + 2-\gamma -\lambda \right) \int _{[ \rho (C_1), \infty )} \xi ^{2-\gamma - \lambda }\Phi (t,d \xi ) \\&+ \frac{1}{2}\int _{[ \rho (C_1), \infty )} \int _{[ \rho (C_1), \infty )}K_R(\xi ,z) \left[ (z+\xi )^{2-\gamma -\lambda }- \xi ^{2-\gamma -\lambda }-z^{2-\gamma -\lambda } \right] \Phi (t,d \xi )\Phi (t,d z). \end{aligned}$$

Hence, since \(-1< \gamma <1\) and \(0< \gamma + \lambda <1 \), we have

$$\begin{aligned} \frac{3\gamma -3+2\lambda }{1-\gamma } + 2-\gamma -\lambda =- \frac{(\gamma +1)(1-\gamma -\lambda )}{1-\gamma } <0. \end{aligned}$$

By symmetry,

$$\begin{aligned}&\frac{1}{2}\int _{[ \rho (C_1), \infty )} \int _{[ \rho (C_1), \infty )}K_R(\xi ,z) \left[ (z+\xi )^{2-\gamma -\lambda }- \xi ^{2-\gamma -\lambda }-z^{2-\gamma -\lambda } \right] \Phi (t, d \xi )\\ {}&\quad \qquad \times \Phi (t,d z) \\ {}&\quad \le \int _{[ \rho (C_1), \infty )} \int _{[\rho (C_1), z]} K_R(\xi ,z) \left[ (z+\xi )^{2-\gamma -\lambda }- \xi ^{2-\gamma -\lambda }-z^{2-\gamma -\lambda } \right] \\ {}&\qquad \times \Phi (t,d \xi )\Phi (t,d z). \end{aligned}$$

Assume \(\xi \le z\). Denote \(\eta :=\frac{\xi }{z}\in (0,1]\) and observe

$$\begin{aligned}&K_{R}(\xi ,z)[(\xi +z)^{2-\gamma -\lambda }-\xi ^{2-\gamma -\lambda }-z^{2-\gamma -\lambda }]\nonumber \\&\quad \le K(\xi ,z) z^{2-\gamma -\lambda }[(1+\eta )^{2-\gamma -\lambda }-1-\eta ^{2-\gamma -\lambda }]\nonumber \\&\quad \le C K(\xi ,z)z^{2-\gamma -\lambda }\eta \nonumber \\&\quad \le C z^{\gamma }(\eta ^{\gamma +\lambda }+\eta ^{-\lambda }) z^{2-\gamma -\lambda }\eta \le 2C z^{2-\lambda }\eta ^{1-\lambda }\nonumber \\&\quad \le 4C (z\xi ^{1-\lambda }+\xi z^{1-\lambda }). \end{aligned}$$
(5.26)

Since \(\rho (C_1)\le \xi \) and \(\rho (C_1)\le z\), then \(z^{1-\lambda }\le \rho (C_1)^{-\lambda }z\). Hence

$$\begin{aligned} \frac{d }{dt } \int _{\mathbb R_*} \xi ^{2-\gamma -\lambda } \Phi (t, d \xi ) \le - c_3 \int _{\mathbb R_*} \xi ^{2-\gamma -\lambda } \Phi (t, d \xi ) + c(\rho (C_1)), \end{aligned}$$

for suitable constants \(c_3, c(\rho (C_1))>0\). Then, (3) follows. \(\square \)

Proposition 5.8

(Time-continuity of the semigroup) Assume \(K_R \) to be a truncated kernel as in (5.3) defined as a function of the homogeneous symmetric kernel K that satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R\) such that (1.7) holds and such that \(\gamma > -1\) and \(\gamma + 2 \lambda \ge 1 \). Let \(\Phi _0\in P\). Let \(T>0\) be as in Theorem 5.3. The map \(S(\cdot )\Phi _0:[0,T]\rightarrow P\) is continuous in time, for every fixed \(\Phi _{0}\), where P was defined in (5.18).

Proof

Let \(T>0\). We want to estimate the value of \(|S(t) \Phi _0- S(s) \Phi _0|\), for \(s,t\in [0,T]\). Assume without loss of generality that \(s\le t\). By the definition of the operator S we know that

$$\begin{aligned}&\int _{\mathbb R_*} \varphi (x) [{\Phi }(t, d x)-{\Phi }(s, d x)]- \frac{1+\gamma }{1-\gamma } \int _{s}^{t}\int _{\mathbb R_*} \varphi (x) \Phi (z, d x)d z \\&\qquad + \frac{2}{1-\gamma } \int _{s}^{t}\int _{\mathbb R_*} \varphi '(x) x \Phi (z, d x)d z\\&\qquad - \int _{s}^{t} \frac{1}{\int _{\mathbb R_*}x^{\gamma +\lambda }\Phi (z,d x)} \int _{\mathbb R_*} \varphi '(x) x^{\gamma + \lambda }\Phi (z, d x)d z \\&\quad =\frac{1}{2}\int _{s}^{t}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x,y) \left[ \varphi (x+y)- \varphi (x) - \varphi (y) \right] \Phi (z, d x) \Phi (z, d y)d z. \end{aligned}$$

We have that there exists a constant \(C_{2}>0\) such that \(M_{\gamma +\lambda }(\Phi (r))\ge \frac{1}{C_{2}}\), for every \(r\in [0,T]\).

As \(\Phi \) has compact support, we can conclude that the exists a constant \(C>0\), which may depend on \(\rho (C_{1}),R,T, \gamma ,\lambda \), but independent on st, such that

$$\begin{aligned} \bigg |&\int _{\mathbb R_*} \varphi (x) [{\Phi }(t, d x)-{\Phi }(s, d x)]\bigg |\le C|t-s|, \end{aligned}$$

thus giving us the desired continuity of the semigroup. \(\square \)

Lemma 5.9

(Dual equation) Assume \(K_R\) to be a truncated kernel as in (5.3) defined as a function of a homogeneous symmetric kernel K that satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R \) satisfying (1.7)0 and with \(\gamma > -1\) and \(\gamma + 2 \lambda \ge 1 \). Let \(\Phi _{1},\Phi _{2}\in P\) be two solutions of (5.4) with initial conditions \(\Phi _{in ,1},\Phi _{in ,2}\in P\), respectively, and \(T>0\) be as in Theorem 5.3. Then there exists a unique solution \(\varphi \in C ^{1}([0,T],C ^{1}_{c}([\rho (C_{1}),\infty ))\), with \(\varphi (T,\cdot )=\psi (\cdot )\), where \(\psi \) is an arbitrary function in \(C ^{1}_{c}([\rho (C_{1}),\infty ))\), which solves the following equation:

$$\begin{aligned}&\partial _{t}\varphi (t, \xi )+ \frac{1+\gamma }{1-\gamma } \varphi (t, \xi ) - \frac{2\xi }{1-\gamma } \partial _\xi \varphi (t, \xi ) + \frac{ \xi ^{\gamma + \lambda }}{M_{\gamma +\lambda }(\Phi _1)}\partial _\xi \varphi (t, \xi ) \nonumber \\&+\mathbb {L}(\varphi )(t,\xi )=0, \end{aligned}$$
(5.27)

where

$$\begin{aligned} \mathbb {L}(\varphi )(t,\xi ):=&- \frac{\xi ^{\gamma +\lambda } }{M_{\gamma +\lambda }(\Phi _1) M_{\gamma +\lambda }(\Phi _2) } \int _{\mathbb R_*} \partial _z \varphi (t, z) z^{\gamma +\lambda } \Phi _2(t, d z) \\&+ \frac{1}{2}\int _{\mathbb R_*} K_{R}(\xi ,\eta )[\varphi (t,\xi +\eta )-\varphi (t,\xi )-\varphi (t,\eta )](\Phi _{1}(t, d \eta )\\&+\Phi _{2}(t, d \eta )). \end{aligned}$$

Remark 5.10

We prove the statement of the lemma for a modified operator \(\overline{\mathbb L}(\varphi )\) which preserves the continuity in the variable \(\xi \), is equal to zero when \(\xi \ge 8R\) and \(\overline{\mathbb L}(\varphi )={\mathbb L}(\varphi )\) if \(\xi \in [\rho (C_{1}),4R]\). Due to the support of \(\Phi _{1}\) and \(\Phi _{2}\), when proving the continuity of the map S(t) in the weak-\(^*\) topology, it suffices to analyse the operator \(\overline{\mathbb {L}}(\varphi )\). Thus, it is enough to prove the statement of the lemma only for the operator \(\overline{\mathbb {L}}(\varphi )\). Notice that, for example, the operator \(\mathbb {L}(\varphi )\) does not preserve compactness because of the presence of the term \(\frac{\xi ^{\gamma +\lambda } }{M_{\gamma +\lambda }(\Phi _1) M_{\gamma +\lambda }(\Phi _2) }\int _{\mathbb R_*} \partial _z \varphi (t, z) z^{\gamma +\lambda } \Phi _2(t, d z)\). We keep the notation \(\mathbb {L}(\varphi )\) for simplicity. We omit further details as the proof consists of standard methods used in the study of coagulation equations, see, for example, [10].

Proof of Lemma 5.9

First, we use the method of characteristics. We define \(X(t, \xi )\) to be the solution of the ODE

$$\begin{aligned} x'(t)&= - \frac{2}{1-\gamma }x + \frac{1}{\int _{\mathbb R_*}z^{\gamma +\lambda }\Phi _1(t,d z)} x^{\gamma + \lambda } \end{aligned}$$

with initial condition \(x(0)=\xi . \)

In this way, Eq. (5.27) can be rewritten in the fixed point form

$$\begin{aligned} \varphi (t, X(t, \xi )) = {\mathcal {L}}[\varphi ] (t, \xi ), \end{aligned}$$
(5.28)

where

$$\begin{aligned} {\mathcal {L}} [\varphi ] (t, \xi ) :=&\,\varphi (T, X(T, \xi )) + \frac{1+\gamma }{1-\gamma } \int _t^T \varphi (s, X(s, \xi )) d s \\&+\int _t^T \mathbb {L}(\varphi )(s, X(s , \xi )) d s. \end{aligned}$$

Our strategy for proving the statement of the lemma is to apply Banach fixed point theorem. The operator \({\mathcal {L}}\) maps \({C}^{1}([0,T],C ^{1}_{c}([\rho (C_{1}), \infty ) ))\) in itself.

We prove that the operator \({\mathcal {L}}\) is a contraction if we endow \(Y:= C([0,T], C^1_c([\rho (C_{1}), \infty ))\) with the norm \(\Vert \varphi \Vert _Y:= \sup _{ t \in [0, T] } \left( \sup _{ x \in \mathbb R_*} |\varphi (t, x) | \right. \) \( + \left. \sup _{x \in \mathbb R_*} | \partial _x \varphi (t, x) | \right) \).

To this end, we notice that

$$\begin{aligned} {\mathcal {L}} [\varphi _1] (t, \xi ) - {\mathcal {L}} [\varphi _2 ] (t, \xi ) =&\, \frac{1+\gamma }{1-\gamma } \int _t^T \left( \varphi _1 (s, X(s, \xi )) - \varphi _2 (s, X(s, \xi )) \right) d s \\&+\int _t^T \left[ \mathbb {L}(\varphi _1)(s, X(s , \xi )) - \mathbb {L}(\varphi _2)(s, X(s , \xi )) \right] d s. \end{aligned}$$

We notice that

$$\begin{aligned}&\int _t^T \left[ \mathbb {L}(\varphi _1)(s, X(s , \xi )) - \mathbb {L}(\varphi _2)(s, X(s , \xi )) \right] d s \\&\quad = -\int _t^T \frac{1}{M_{\gamma +\lambda }(\Phi _1) M_{\gamma +\lambda }(\Phi _2) } \int _{\mathbb R_*} \left( \partial _z \varphi _1 (s, z) - \partial _z \varphi _2 (s, z) \right) z^{\gamma +\lambda } \Phi _2(s, d z) X(s, \xi )^{\gamma +\lambda } d s \\&\qquad + \frac{1}{2}\int _t^T\int _{\mathbb R_*} K_{R}(X(s , \xi ),\eta )[\varphi _1(s,X(s , \xi )+\eta )- \varphi _2(s,X(s , \xi )+\eta ) -\varphi _1(s,X(s , \xi ))\\&\qquad +\varphi _2(s,X(s , \xi )) -\varphi _1(s,\eta ) + \varphi _2(s,\eta ) ] (\Phi _{1}(s, d \eta )+\Phi _{2}(s, d \eta ))d s. \end{aligned}$$

From this we deduce that

$$\begin{aligned} \Vert {\mathcal {L}} [ \varphi _1] - {\mathcal {L}} [\varphi _2] \Vert _Y \le T c(\rho (C_{1}), R, \Phi _1, \Phi _2) \Vert \varphi _1 - \varphi _2 \Vert _Y \end{aligned}$$

and hence \({\mathcal {L}} \) is a contraction for sufficiently small times T. We can extend the solution to all possible times noting that the contraction constant \(c(\rho (C_{1}), \Phi _1, \Phi _2, R)\) does not depend on the final condition \(\psi \). We thus deduce that there exists a solution \(\varphi \) of the fixed point \(\varphi = {\mathcal {L}} [\varphi ].\) \(\qquad \square \)

We now prove that the found solution is Lipschitz continuous.

Proposition 5.11

Assume \(K_R\) to be a truncated kernel as in (5.3) defined as a function of a homogeneous symmetric kernel K such that it satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R \) satisfying (1.7) and such that \(\gamma > -1\) and \(\gamma + 2 \lambda \ge 1 \). Let \(T>0\) be as in Theorem 5.3. Let \(\varphi \in C ^{1}([0,T],C ^{1}([\rho (C_1),8R]))\) with initial datum \(\varphi (T,\cdot )\) be the function found in Lemma 5.9. Assume, in addition, that \(\sup _{ \xi \in [\rho (C_1),8R]}|\varphi (T,\xi )|\le 1\) and that \(\varphi (T,\xi )\) is Lipschitz. Then \(\varphi \) is Lipschitz continuous, in the sense that, for every \(t\in [0,T],\) there exists \(C(t)>0\) such that

$$\begin{aligned} \sup _{s\in [0,t]}|\varphi (s,\xi )-\varphi (s,\tilde{\xi })|\le C(t)|\xi -\tilde{\xi }|, \end{aligned}$$

for every \(\xi ,\tilde{\xi }\in [\rho (C_1),8R]\). Moreover, C(t) may depend on the norm of \(\Phi _{1}\) and \(\Phi _{2}\), but is otherwise independent of the choice of \(\Phi _{1}\) and \(\Phi _{2}\).

Proof

Notice first that, since \(\sup _{ \xi \in [\rho (C_1),8R]}|\varphi (T,\xi )|\le 1\), there exists a constant \(C>0\), which depends on the norm of \(\Phi _{1}\) and \(\Phi _{2}\) and the parameters \(\rho (C_{1})\), R and t, such that \(\sup _{s\in [0,t],\xi \in [\rho (C_1),8R]}\) \(|\varphi (s,\xi )|\le C\). This can be seen by looking at (5.28). We will use Grönwall’s inequality in (5.28) in order to prove that \(\varphi \) is Lipschitz:

$$\begin{aligned} |\varphi (t,X(t,\xi ))-\varphi (t,X(t,\tilde{\xi }))|&\le |\varphi (T,X(T,\xi ))-\varphi (T,X(T,\tilde{\xi }))|\\&+\frac{1+\gamma }{1-\gamma } \int _t^T |\varphi (s, X(s, \xi ))-\varphi (s,X(s,\tilde{\xi }))| d s\\&+\int _{t}^{T}|\mathbb L [\varphi ] (s, X(s,\xi ))-\mathbb L [\varphi ] (s, X(s,\tilde{\xi }))|d s. \end{aligned}$$

In order to bound the term

$$\begin{aligned} \int _{t}^{T}|\mathbb L [\varphi ] (s, X(s,\xi ))-\mathbb L [\varphi ] (s, X(s,\tilde{\xi }))|d s, \end{aligned}$$

we need to estimate

$$\begin{aligned} I_{1}&= \bigg |\frac{1}{2}\int _{\mathbb R_*} K_{R}(X(t,\xi ),\eta )[\varphi (t,X(t,\xi )+\eta ) \\&\quad -\varphi (t,X(t,\xi ))-\varphi (t,\eta )](\Phi _{1}(t, d \eta )+\Phi _{2}(t, d \eta ))\\&\quad -\frac{1}{2}\int _{\mathbb R_*} K_{R}(X(t,\tilde{\xi }),\eta )[\varphi (t,X(t,\tilde{\xi })+\eta )\\&\quad -\varphi (t,X(t,\tilde{\xi }))-\varphi (t,\eta )](\Phi _{1}(t, d \eta )+\Phi _{2}(t, d \eta ))\bigg | \end{aligned}$$

and

$$\begin{aligned} I_{2}= \frac{|X(t,\xi )^{\gamma +\lambda }-X(t,\tilde{\xi })^{\gamma +\lambda }|}{M_{\gamma +\lambda }(\Phi _1) M_{\gamma +\lambda }(\Phi _2) } \int _{\mathbb R_*} |\partial _z \varphi (t, z)| z^{\gamma +\lambda } \Phi _2(t, d z). \end{aligned}$$

For \(I_{1},\) by the definition of \(K_{R}\) in (5.3), we have \(K_{R}\) is \(C^{1}\) and we can assume it has compact support as we are only interested in the region \([\rho (C_{1}),8R]^{2}\). Thus, we have that the first derivative of \(K_{R}\) is bounded from above. Moreover, there exist constants \(L_{1}(t), L_{2}(t)>0\) such that \(L_{1}(t)|\xi -\tilde{\xi }|\le |X(t,\xi )-X(t,\tilde{\xi })|\le L_{2}(t)|\xi -\tilde{\xi }|\).

For \(I_{2},\) we use

$$\begin{aligned}&\frac{1}{M_{\gamma +\lambda }(\Phi _1) M_{\gamma +\lambda }(\Phi _2) }\int _{\mathbb R_*} |\partial _z \varphi (t, z)| z^{\gamma +\lambda } \Phi _2(t, d z)\nonumber \\ {}&\le \frac{C_{1}}{(C_{2})^{2}}\sup _{z\in [\rho (C_{1}),8R]}|\partial _z \varphi (t, z)|. \end{aligned}$$

If we prove that there exists a constant \(C>0,\) which can depend on \(\rho (C_{1}),R,\) and the norms of \(\Phi _{1},\) and \(\Phi _{2},\) but does not vary depending on the choice of \(\Phi _{1},\Phi _{2}\) such that

$$\begin{aligned} \sup _{z\in [\rho (C_{1}),8R]}|\partial _z \varphi (t, z)|\le C, \end{aligned}$$
(5.29)

then there exists a constant \(C>0,\) which can depend on \(\rho (C_{1}),\) R and the norms of \(\Phi _{1}\) and \(\Phi _{2}\) such that

$$\begin{aligned}&|\varphi (t,X(t,\xi ))-\varphi (t,X(t,\tilde{\xi }))|\nonumber \\&\qquad \le C |X(t,\xi )-X(t,\tilde{\xi })|+ C\int _{t}^{T} |\varphi (s,X(s,\xi ))-\varphi (s,X(s,\tilde{\xi }))|. \end{aligned}$$
(5.30)

Let us now prove (5.29). From Eq. (5.28) and the fact that \(K_{R}\) is in \(C^{1}\) with compact support, we obtain an upper bound for \(|\partial _{2}\varphi (t,X(t,\xi ))|\).

We then use that there exist some constants \(\overline{c}(t), \overline{\overline{c}}(t)>0\) such that \(\overline{c}(t)\xi \le X(t,\xi )\le \overline{\overline{c}}(t)\xi \), with \(\xi \in [\rho (C_{1}),8R]\) in order to obtain the desired bound for \(\sup _{z\in [\rho (C_{1}),8R]}|\partial _z \varphi (t, z)|\).

We use Grönwall’s inequality in (5.30) and obtain that \(|\varphi (t,X(t,\xi )) -\varphi (t,X(t,\tilde{\xi }))|\le C(t) |X(t,\xi )-X(t,\tilde{\xi })|.\) Thus, we can conclude that also \(|\varphi (t,\xi )-\varphi (t,\tilde{\xi })|\le C(t) |\xi -\tilde{\xi }|\) since \(X(t,\xi )\) is Lipschitz continuous in the \(\xi \) variable. \(\square \)

Proposition 5.12

(Continuity of the semigroup in the weak topology) Assume \(K_R\) to be a truncated kernel as in (5.3) defined as a function of a homogeneous symmetric kernel K such that it satisfies (1.5), (1.6), for parameters \(\gamma , \lambda \in \mathbb R \) satisfying (1.7) and such that \(\gamma > -1\) and \(\gamma + 2 \lambda \ge 1 \). For every time \(t>0\) the map

$$\begin{aligned} S(t):P\rightarrow P \end{aligned}$$

is continuous in the weak-\(^{*}\) topology, where P was defined in (5.18).

Proof

Let \(\delta >0\). In order to prove continuity in the weak-\(^{*}\) topology of the semigroup and because of the support of \(\Phi _{1}\) and \(\Phi _{2}\), it is enough to prove that, if for every \(\psi \in C_c([\rho (C_{1}),\infty ))\), with \(||\psi ||_{\infty }\le 1\), we have that

$$\begin{aligned} \int _{\mathbb R_*} \psi (x) \left( \Phi _{in , 1 } (d x) - \Phi _{in , 2} ( d x ) \right) \text { is sufficiently small,} \end{aligned}$$
(5.31)

then we have that for every \(\psi \in C_c([\rho (C_{1}),\infty ))\), with \(||\psi ||_{\infty }\le 1,\)

$$\begin{aligned} \int _{\mathbb R_*} \psi (x) \left( \Phi _1 (t, d x) - \Phi _2(t, d x ) \right) \le \delta , \end{aligned}$$

where \(\Phi _1 \) and \(\Phi _2 \) are the solutions of Eq. (5.4) with initial conditions \(\Phi _{in , 1} \) and \(\Phi _{in , 2} \), respectively. To this end we notice that since \(\Phi _1\) and \(\Phi _2 \) satisfy Eq. (5.4) then

$$\begin{aligned}{} & {} \int _{\mathbb R_*} \varphi (t,x) [{\Phi _{1}}(t, d x)-{\Phi _{2}}(t, d x)]-\int _{\mathbb R_*} \varphi (0,x) [{\Phi _{in ,1}}( d x)-{\Phi _{in ,2}}( d x)] \\{} & {} \qquad -\int _{0}^{t}\int _{\mathbb R_*} \partial _{s}\varphi (s,x) [{\Phi _{1}}(s, d x)-{\Phi _{2}}(s, d x)]d s\\{} & {} \qquad - \frac{1+\gamma }{1-\gamma } \int _{0}^{t}\int _{\mathbb R_*} \varphi (s,x) [{\Phi _{1}}(s, d x)-{\Phi _{2}}(s, d x)]d s \\{} & {} \qquad +\frac{2}{1-\gamma }\int _{0}^{t} \int _{\mathbb R_*} \partial _x \varphi (s,x) x [{\Phi _{1}}(s, d x)-{\Phi _{2}}(s, d x)]d s \\{} & {} \qquad - \int _{0}^{t}\frac{1}{\int _{\mathbb {R}_{*}}z^{\gamma +\lambda }\Phi _{1}(s,d z)} \int _{\mathbb R_*} \partial _x \varphi (s, x) x^{\gamma + \lambda }\Phi _{1}(s, d x)d s\\{} & {} \qquad +\int _{0}^{t}\frac{1}{\int _{\mathbb {R}_{*}}z^{\gamma +\lambda }\Phi _{2}(s,d z)} \int _{\mathbb R_*} \partial _x \varphi (s, x)x^{\gamma + \lambda }\Phi _{2}(s, d x)d s \\{} & {} \quad =\frac{1}{2}\int _{0}^{t}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x,y) \chi _{\varphi }(s,x,y) [{\Phi _{1}}(s, d y)\\{} & {} \qquad +{\Phi _{2}}(s, d y)][{\Phi _{1}}(s, d x)-{\Phi _{2}}(s, d x)]d s, \end{aligned}$$

where \(\chi _{\varphi }(s,x,y):=\varphi (s,x+y)-\varphi (s,x)-\varphi (s,y)\), for every \(\varphi \in C([0, t], C_c^1(\mathbb R_*)).\) To simplify the computations, we adopt the notation \(\tilde{\Phi }:= \Phi _1 - \Phi _2 \) and we notice that \(\tilde{\Phi } \) satisfies the following equation for every \(\varphi \in C^1([0, t], C_c^1(\mathbb R_*))\)

$$\begin{aligned}&\int _{\mathbb R_*} \varphi (t,x) \tilde{\Phi }(t, d x)-\int _{\mathbb R_*} \varphi (0,x) [{\Phi _{in ,1}}( d x)-{\Phi _{in ,2}}( d x)] \\&\quad \quad -\int _{0}^{t}\int _{\mathbb R_*} \partial _{s}\varphi (s,x) \tilde{\Phi }(s, d x)d s - \frac{1+\gamma }{1-\gamma } \int _{0}^{t}\int _{\mathbb R_*} \varphi (s,x) \tilde{\Phi }(s, d x)d s\\&\qquad +\frac{2}{1-\gamma } \int _{0}^{t}\int _{\mathbb R_*} \partial _x \varphi (s,x) x\tilde{\Phi }(s, d x)d s \\&\qquad - \int _{0}^{t}\frac{1}{\int _{\mathbb R_*}z^{\gamma +\lambda }\Phi _{1}(s,d z)} \int _{\mathbb R_*} \partial _x \varphi (s, x) x^{\gamma + \lambda }\tilde{\Phi }(s, d x)d s \\&\qquad +\int _{0}^{t}\int _{\mathbb R_*} \mathbb L[\varphi ](s, x) \tilde{\Phi } (s, d x)d s. \end{aligned}$$

First, we make the following notation:

$$\begin{aligned} C_{norm}:=\sup _{s\in [0,t]}\int _{\mathbb {R}_{*}}[\Phi _{1}+\Phi _{2}](s,d x)<\infty . \end{aligned}$$

Notice that \(C_{norm}\) can be bounded from above by

$$\begin{aligned} C_{norm}=\sup _{s\in [0,t]}\int _{\mathbb {R}_{*}}[\Phi _{1}+\Phi _{2}](s,d x)\le C(R,\rho (C_{1}),t) \int _{\mathbb {R}_{*}}[\Phi _{in ,1}+\Phi _{in ,2}](d x). \end{aligned}$$

Let now \(\varphi \) be the solution found in Lemma 5.9 with coagulation kernel \(K_{R}\). By Proposition 5.11, we know that there exists a constant \(C(t,R,\Phi _{1},\Phi _{2})\), that depends only on the norm of \(\Phi _{1}\) and \(\Phi _{2}\), but is otherwise independent of the choice of \(\Phi _{1}\) and \(\Phi _{2}\), such that \(|\varphi (s,\xi )-\varphi (s,\tilde{\xi })|\le C(t,R,\Phi _{1},\Phi _{2})|\xi -\tilde{\xi }|,\) for every \(\xi ,\tilde{\xi }\in [\rho (C_{1}),8R]\) and \(s\in [0,t]\). We thus look at the set:

$$\begin{aligned} \mathcal {K}_{\rho (C_{1}),R}:=&\{\chi \in C ([\rho (C_{1}),8R])| |\chi (\xi )-\chi (\tilde{\xi })|\le C(t,R,\Phi _{1},\Phi _{2})|\xi -\tilde{\xi }|,\\&for all \xi , \tilde{\xi }\in [\rho (C_{1}),8R]\}\subset C ([\rho (C_{1}),8R]). \end{aligned}$$

The set \( \mathcal {K}_{\rho (C_{1}),R}\) is totally bounded, thus there exist \(N\in \mathbb {N}\) and \(\chi _{1},\ldots ,\chi _{N}\in \mathcal {K}_{\rho (C_{1}),R}\) such that \( \mathcal {K}_{\rho (C_{1}),R}\subseteq \cup _{i=1}^{N}B(\chi _{i}, \frac{\delta }{4N C_{norm}})\).

Then we obtain that

$$\begin{aligned} \int _{\mathbb R_*} \varphi (t,x) \tilde{\Phi }(t, d x)&=\int _{\mathbb R_*} \varphi (0,x) [{\Phi _{in ,1}}(d x)-{\Phi _{in ,2}}( d x)] =:T_{1}. \end{aligned}$$
(5.32)

We can bound \(T_{1}\) by

$$\begin{aligned} T_{1}&\le \min _{i=1}^{N}\int _{\mathbb R_*} |[\varphi (0,x)-\chi _{i}(x) ][{\Phi _{in ,1}}(d x)-{\Phi _{in ,2}}( d x)] | \\&\qquad +\max _{i=1}^{N}\big (\int _{\mathbb R_*} \chi _{i}(x) [{\Phi _{in ,1}}(d x)-{\Phi _{in ,2}}( d x)]\big )\\&\le \frac{\delta }{4C_{norm}}\int _{\mathbb R_*} [{\Phi _{in ,1}}(d x)+{\Phi _{in ,2}}( d x)] + \max _{i=1}^{N}\int _{\mathbb R_*} \chi _{i}(x) [{\Phi _{in ,1}}(d x)\\&\qquad -{\Phi _{in ,2}}( d x)]\le \delta , \end{aligned}$$

where in the last step we used (5.31). \(\square \)

Proof of Theorem 5.6

By Proposition 5.12 we know that the operator \(\Phi \mapsto S(t) \Phi \) is continuous in the weak-\(^{*}\) topology. Additionally, thanks to Proposition 5.7, we know that P is an invariant region for S(t). Since P is also convex and compact and since we have proven that the map \(t \mapsto S(t)\) is continuous, we apply Theorem 1.2 in [8] to deduce that there exists a \(\Phi \) such that \(S(t)\Phi =\Phi \). \(\square \)

5.3 Existence of the self-similar profile

To keep the notation lighter, in the previous sections we denoted with \(\Phi \) the solution of the truncated problem. In this section, since we pass to the limit as R tends to infinity, we add the label R to \(\Phi \) and we will denote with \(\Phi \) the self-similar profile. For simplicity, we denote by \(\rho :=\rho (C_{1}),\) where \(\rho (C_{1})\) was defined in (4.2).

Proof of Theorem 5.1

For every \(R>1\), sufficiently large, there exist \(T>0\) and \(\Phi _{R}\in C^1([0,T];{\mathcal {M}}_{+}(\mathbb R_*))\) such that

$$\begin{aligned}&\int _{\mathbb R_*} \varphi (x) \dot{\Phi }_{R}(t, d x)- \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \varphi (x) \Phi _{R}(t, d x) + \frac{2}{1-\gamma } \int _{\mathbb R_*} \varphi '(x) x \Phi _{R}(t, d x) \\&- \frac{1}{M_{\gamma +\lambda }(\Phi _{R}(t))} \int _{\mathbb R_*} \varphi '(x) x^{\gamma + \lambda }\Phi _{R}(t, d x) \\&=\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K_R(x,y) \left[ \varphi (x+y)- \varphi (x) - \varphi (y) \right] \Phi _{R}(t, d x) \Phi _{R}(t, d y), \end{aligned}$$

where \(K_{R}\) is the coagulation kernel defined in (5.3). Notice that the bounds in Proposition 5.7 are independent of \(R>1\).

Thanks to Theorem 5.6 we know that a measure \(\Phi _R \in {\mathcal {M}}_+(\mathbb R_*) \) satisfying Eq. (5.16) and satisfying the bounds (5.17) exists. We notice also that, since \(\gamma + \lambda >0\),

$$\begin{aligned} \int _{(0,\infty )}\Phi _{R}(d x)=\int _{[\rho ,\infty )}\Phi _{R}(d x)\le \rho ^{-\gamma -\lambda }\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)\le \rho ^{-\gamma -\lambda }C_{1}. \end{aligned}$$
(5.33)

Hence, Banach-Alaoglu Theorem implies that there exists \(\Phi \) such that

$$\begin{aligned} \Phi _R \rightharpoonup \Phi \text { as } R \rightarrow \infty \end{aligned}$$
(5.34)

in the weak-\(^{*}\) topology.

We now prove that the measure \(\Phi \) in (5.34) satisfies Eq. (3.3) by taking the limit as \(R \rightarrow \infty \). Fix \(\varphi \in C_{c}(\mathbb {R}_{*})\). We start with passing to the limit in the coagulation term.

Let \(\varepsilon \in (0,1)\). We first show that

$$\begin{aligned}&\bigg | \int _{[\rho ,\infty )}\int _{[\rho ,\infty )}\left( K (\xi ,z)- K_{R}(\xi ,z) \right) \Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \bigg |\nonumber \\&\le \varepsilon , \end{aligned}$$
(5.35)

for sufficiently large R. We then prove that

$$\begin{aligned} \int _{[\rho ,\infty )}\int _{[\rho ,\infty )} K(\xi ,z) \Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \end{aligned}$$
(5.36)

converges to

$$\begin{aligned} \int _{[\rho ,\infty )}\int _{[\rho ,\infty )}K (\xi ,z)\Phi (d \xi )\Phi (d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \end{aligned}$$
(5.37)

as \(R \rightarrow \infty .\)

For (5.35), we have that:

$$\begin{aligned}&\bigg | \int _{[\rho ,\infty )}\int _{[\rho ,\infty )}\left( K (\xi ,z)- K_{R}(\xi ,z) \right) \Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \bigg | \\&\quad \le \int _{[\rho ,\frac{R}{4}]}\int _{[\rho ,\frac{R}{4}]}\bigg |\left( K (\xi ,z)- K_{R}(\xi ,z) \right) \Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \bigg | \\&\qquad + 2\int _{(\frac{R}{4},\infty )}\int _{[\rho ,\infty )}\bigg |\left( K (\xi ,z)- K_{R}(\xi ,z) \right) \Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \bigg |\\&\quad =T_{1}+T_{2}. \end{aligned}$$

For the first term, we have

$$\begin{aligned} T_{1}\le e^{-R}3||\varphi ||_{\infty }\rho ^{-2}, \end{aligned}$$

where we used the definition of \(K_{R}\) in (5.3) and the fact that the total mass of the measures is equal to 1.

For the region \((\frac{R}{4}, \infty )\times [\rho , \infty )\), we use \(K_R \le K\) to prove that

$$\begin{aligned} T_{2}&\le 12||\varphi ||_{\infty } \int _{(\frac{R}{4},\infty )}\int _{[\rho ,\infty )} K(\xi ,z) \Phi _{R}(d z)\Phi _{R}(d \xi )\\&\le c \left( R^{-\gamma -2\lambda }M_{\gamma +\lambda }^{2}(\Phi _{R}) +R^{\gamma +\lambda -1}\rho ^{-\gamma -2\lambda } M_{\gamma +\lambda }(\Phi _{R}) \right) , \end{aligned}$$

which gives a small contribution due to the uniform estimates for \(\Phi _{R}\) if we make R sufficiently large.

We now analyse (5.36). We consider a continuous function \(g: \mathbb R_+ \rightarrow [0, 1] \) such that \(g(x)=1\), when \(x\le 1\), and \(g(x)=0\), when \(x \ge 2\). We define the function \(p_M \) as

$$\begin{aligned} p_{M}(x,y)=g\left( \frac{x}{M}\right) g\left( \frac{y}{M}\right) , \end{aligned}$$
(5.38)

where M is a positive constant. By the construction of \(p_{M}\) and given the fact that \(\Phi _{R}\) is supported in the region \([\rho ,\infty )\), for every \(R>1\), we have that, given any function \(\varphi \in C_c(\mathbb R_*)\),

$$\begin{aligned}&\int _{[\rho ,\infty )}\int _{[\rho ,\infty )}K(\xi ,z)p_{M}(\xi ,z)\Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]\\&\qquad =\int _{[\rho ,2M]}\int _{[\rho ,2M]}K(\xi ,z)p_{M}(\xi ,z)\Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]. \end{aligned}$$

Therefore, since \(\Phi _R \rightharpoonup \Phi \) in the weak-\(^{*}\) topology, we can conclude that \(\Phi _R \Phi _{R}\rightharpoonup \Phi \Phi \) in the weak-\(^{*}\) topology as \(R\rightarrow \infty \) if we work on the set \([\rho ,2M]^{2}\) and we deduce that

$$\begin{aligned} \int _{[\rho ,\infty )}\int _{[\rho ,\infty )}K(\xi ,z)p_{M}(\xi ,z)\Phi _{R}(d \xi )\Phi _{R}(d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \end{aligned}$$

converges to

$$\begin{aligned} \int _{[\rho ,\infty )}\int _{[\rho ,\infty )}K(\xi ,z)p_{M}(\xi ,z)\Phi (d \xi )\Phi (d z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )] \end{aligned}$$

as R tends to infinity.

To conclude that, for every test function \(\varphi \in C_c(\mathbb R_*), \) we have that (5.36) converges to (5.37) as \(R \rightarrow \infty \), we have to prove that the reminder terms, namely

$$\begin{aligned}{} & {} \int _{[\rho ,\infty )}\int _{[M,\infty )}K(\xi ,z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]\Phi _{R}(d \xi )\Phi _{R}(d z), \end{aligned}$$
(5.39)
$$\begin{aligned}{} & {} \int _{[M,\infty )}\int _{[\rho ,\infty )}K(\xi , z) [\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]\Phi _{R}(d \xi )\Phi _{R}(d z), \end{aligned}$$
(5.40)
$$\begin{aligned}{} & {} \int _{[\rho ,\infty )}\int _{[M,\infty )}K(\xi ,z)[\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]\Phi (d \xi )\Phi (d z), \end{aligned}$$
(5.41)
$$\begin{aligned}{} & {} \int _{[M,\infty )}\int _{[\rho ,\infty )}K(\xi , z) [\varphi (z+\xi )-\varphi (z)-\varphi (\xi )]\Phi (d \xi )\Phi (d z), \end{aligned}$$
(5.42)

tend to zero as \(M\rightarrow \infty \).

Let us look at the term (5.39). In this case we are in the region where \(\{\xi \ge M\}\), hence \(\xi ^{\gamma +\lambda -1}\le M^{\gamma +\lambda -1}\) as \(\gamma +\lambda <1\). Therefore, for every \(\varphi \in C _{c}(\mathbb {R}_{>0})\)

$$\begin{aligned}&\int _{[\rho ,\infty )}\int _{[M,\infty )}\xi ^{\gamma +\lambda }z^{-\lambda }\Phi _{R}(d \xi )\Phi _{R}(d z)|\varphi (z+\xi )-\varphi (z)-\varphi (\xi )|\\&\quad \lesssim M^{\gamma +\lambda -1}\rho ^{-\gamma -2\lambda } M_{\gamma +\lambda }(\Phi _{R}), \end{aligned}$$

where we remind that \(M_{\gamma +\lambda }(\Phi _{R})\) is bounded uniformly in \(R>1\) and that the mass of \(\Phi _R\) is equal to one. Similarly, the fact that \(\xi \ge M\) implies that \(\xi ^{-\gamma -2\lambda }\le M^{-\gamma -2\lambda }\) since \(\gamma +2\lambda > 0\). We then obtain that

$$\begin{aligned}&\int _{[\rho ,\infty )}\int _{[M,\infty )}z^{\gamma +\lambda }\xi ^{-\lambda }\Phi _{R}(d \xi )\Phi _{R}(d z)|\varphi (z+\xi )-\varphi (z)-\varphi (\xi )|\\&\quad \le C M^{-\gamma -2\lambda }M_{\gamma +\lambda }^{2}(\Phi _{R}). \end{aligned}$$

From these two inequalities and the fact that \(\gamma + \lambda <1 \) and \(\gamma + 2 \lambda >0\) we deduce that the term (5.39) tends to zero as \(M \rightarrow \infty .\) By a symmetric argument we prove that the term (5.40) tends to zero as \(M\rightarrow \infty \). The fact that the two terms (5.41) and (5.42) tend to zero as \(M \rightarrow \infty \) follows similarly by the fact that the \(\gamma + \lambda \) moment of \(\Phi \) is bounded.

The linear terms for the self-similar profiles of the truncated problem will converge as \(R\rightarrow \infty \) to the desired terms thanks to (5.34).

To conclude, we need to prove the convergence of the remainder term. We thus have to prove that

$$\begin{aligned} \frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)}\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x) \end{aligned}$$

converges to

$$\begin{aligned} \frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi (d x)}\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x) \end{aligned}$$

as \(R\rightarrow \infty \). We have that

$$\begin{aligned}&\bigg | \frac{1}{M_{\gamma +\lambda }(\Phi _{R})}\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\frac{1}{M_{\gamma +\lambda }(\Phi )}\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |\\&\quad \le \bigg | \frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)}\bigg |\bigg |\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |\\&\qquad + \bigg | \frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)}-\frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi (d x)}\bigg |\bigg |\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |. \end{aligned}$$

We can analyse the two terms separately. For the first term, we know by (5.25) that the \(M_{\gamma +\lambda }(\Phi _{R})\) moment is uniformly bounded from below and thus

$$\begin{aligned} \bigg |&\frac{1}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)}\bigg |\bigg |\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg | \\&\le M_{2-\gamma -\lambda }(\Phi _{R})\bigg |\int _{[\rho ,\infty )} \partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |\\&\le C_{2}\bigg |\int _{[\rho ,\infty )} \partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg | \end{aligned}$$

and then we use that \(\bigg |\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi _{R}(d x)-\int _{[\rho ,\infty )} \partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |\rightarrow 0\) by (5.34).

For the second term, we split the integral into a region with compact support and the remainder regions. More precisely,

$$\begin{aligned} \bigg |&\frac{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi (d x)-\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)}{\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi (d x)}\bigg |\bigg |\int _{[\rho ,\infty )}\partial _{x}\varphi (x)x^{\gamma +\lambda }\Phi (d x)\bigg |\\&\le C\bigg |\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi (d x)-\int _{[\rho ,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)\bigg |M_{2-\gamma -\lambda }(\Phi _{R})M_{2-\gamma -\lambda }(\Phi )M_{\gamma +\lambda }(\Phi )\\&\lesssim C\bigg |\int _{[\rho ,\infty )}g\left( \frac{x}{M}\right) \left[ x^{\gamma +\lambda }\Phi (d x)-x^{\gamma +\lambda }\Phi _{R}(d x)\right] \bigg |+2C\bigg |\int _{[M, \infty )}x^{\gamma +\lambda }\Phi (d x)\bigg |\\&\qquad +2C\bigg |\int _{[M, \infty )}x^{\gamma +\lambda }\Phi _{R}(d x)\bigg |. \end{aligned}$$

We have that \(\bigg |\int _{[\rho , \infty )}g(\frac{x}{M})[x^{\gamma +\lambda }\Phi (d x)-x^{\gamma +\lambda }\Phi _{R}(d x)]\bigg |\rightarrow 0\) as \(R\rightarrow \infty \) by (5.34). For the regions close to infinity, we use

$$\begin{aligned} \bigg |\int _{[M,\infty )}x^{\gamma +\lambda }\Phi _{R}(d x)\bigg |\le M^{\gamma +\lambda -1}, \end{aligned}$$

as the total mass is uniformly bounded. Since \(\gamma +\lambda <1,\) the remaining term goes to zero as \(M\rightarrow \infty \).

As a last step we prove that, if \(\Phi ((0, \delta ))=0\) for some \(\delta >0\), then \(\Phi ((0, \rho (M_{\gamma +\lambda })))=0\) for \(\rho (M_{\gamma +\lambda })\) given by (3.4). Since the proof is standard we only sketch it here. Consider \(x \in [\frac{\delta }{2},\frac{1}{2}\rho (M_{\gamma +\lambda }))\), then, since \(\Phi \) satisfies (3.1), we have

$$\begin{aligned} c \Phi (x)&\le \left( \frac{x^{\gamma +\lambda } }{M_{\gamma +\lambda }} - \frac{2}{1-\gamma } x \right) \Phi (x)\\&\le \frac{|1+\gamma |}{1-\gamma } \int _\delta ^x \Phi (\eta ) d \eta + \int _{\delta }^x \int _\delta ^{y-\delta } K(y-\eta ,\eta ) \Phi (y-\eta ) \Phi (\eta ) d \eta d y \\&\le \tilde{c} \int _\delta ^x \Phi (\eta ) d \eta , \end{aligned}$$

for some positive constants \(\tilde{c}, c >0\). Using Grönwall’s lemma, this implies that \(\Phi (x) \le e^{\frac{\tilde{c}}{c} x }\Phi (\frac{\delta }{2}) =0\). Hence \(\Phi ([\delta , \frac{1}{2}\rho (M_{\gamma +\lambda })))=0\). Iterating this argument, we deduce that \(\Phi ((0, \rho (M_{\gamma +\lambda } )))=0\). The argument can be made rigorous by working with the weak formulation of Eq. (3.1). \(\square \)

5.4 Properties of the constructed self-similar profile

In this section we aim at proving Theorem 3.3. In particular we prove that the self-similar profile, whose existence has been proven in Theorem 5.1, satisfies the properties stated in Theorem 3.3. In other words we have to prove the following theorem.

Theorem 5.13

Assume K to be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that

$$\begin{aligned} -1 < \gamma , \quad \gamma +2\lambda \ge 1. \end{aligned}$$

Let \(\Phi \) be the self-similar profile constructed in Theorem 5.1. Then

$$\begin{aligned} \int _{\mathbb R_*} e^{Lx } \Phi (d x) < \infty \end{aligned}$$

for some \(L>0\) and \(\Phi \) is absolutely continuous with respect to the Lebesgue measure. Then \(\Phi (d x )=\phi (x) d x \) and \(\phi \) is such that

$$\begin{aligned} \limsup _{x \rightarrow \infty } e^{ \overline{L} x} \phi (x) < \infty \end{aligned}$$

for a constant \(\overline{L}>0\).

To prove this theorem we start by proving that every self-similar profile \(\Phi \) in the sense of Definition 3.1 and such that \(\Phi ((0, \delta ))=0\), for a positive \(\delta \), is absolutely continuous with respect to the Lebesgue measure and satisfies some moment bounds.

Proposition 5.14

(Regularity) Assume K to be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that \( \gamma +2\lambda \ge 1.\) Let \(\Phi \in {\mathcal {M}}_+(\mathbb R_*)\) be a self-similar profile as in Definition 3.1. Assume additionally that \(\Phi ((0, \delta ))=0\) for \(\delta >0\). Then \(\Phi \) is absolutely continuous with respect to the Lebesgue measure. Its density \(\phi \in L^1([\delta , \infty ))\) satisfies the following equation

$$\begin{aligned}&\int _0^z x \phi (x) dx - \frac{2}{1-\gamma } z^2 \phi (z) + \frac{1}{M_{\gamma +\lambda }(\phi )} z^{\gamma +\lambda +1 } \phi (z) \nonumber \\&\quad - \frac{1}{M_{\gamma +\lambda }(\phi ) } \int _0^z x^{\gamma +\lambda }\phi (x) d x \\&= - J_\phi (z), \quad a.e.\ z > \delta \nonumber \end{aligned}$$
(5.43)

where

$$\begin{aligned} J_\phi (z):= \int _0^z \int _{z-x}^\infty x K(x,y) \phi (x) \phi (y) d y d x. \end{aligned}$$
(5.44)

Proof

We just sketch the proof as similar arguments have been repeatedly used in the analysis of coagulation equations, see for instance the proof of Lemma 4.9 in [9]. An analogous proof will be presented in the proof of Proposition 6.2. Let \(\Phi \) be a solution of (3.1), then

$$\begin{aligned} \partial _\xi \left[ \left( \frac{\xi ^{\gamma + \lambda }}{M_{\gamma +\lambda } } - \frac{2}{1-\gamma } \xi \right) \Phi (\xi ) \right] = \frac{1+\gamma }{1-\gamma } \Phi (\xi ) + \mathbb K [\Phi ](\xi ). \end{aligned}$$

Using the fact that

$$\begin{aligned} \mathbb K [\Phi ] (\xi ) \le \int _{\delta }^{\xi -\delta } K(\xi -\eta ,\eta ) \Phi (\eta ) \Phi (\xi -\eta ) d \eta , \quad \xi > \delta \end{aligned}$$

and that \(\int _{\delta }^{\xi -\delta } K(\xi -\eta ,\eta ) \Phi (\eta ) \Phi (\xi -\eta ) d \eta \) is a bounded measure in every compact set, we deduce that the measure

$$\begin{aligned} \left( \frac{\xi ^{\gamma + \lambda }}{M_{\gamma +\lambda } } - \frac{2}{1-\gamma } \xi \right) \Phi (\xi ) \end{aligned}$$

is absolutely continuous with respect to the Lebesgue measure on the set \((\delta , \infty )\) if \(\delta > \rho (M_{\gamma +\lambda })\), while it is absolutely continuous with respect to the Lebesgue measure in the set \((\delta ,\rho (M_{\gamma +\lambda })) \cup (\rho (M_{\gamma +\lambda }), \infty )\) if \(\delta \le \rho (M_{\gamma +\lambda })\). In the latter case, we therefore have that \(\Phi (d x )=\phi (x) d x + c_0 \delta _{\rho (M_{\gamma +\lambda })}(x) \) with density \(\phi \in L^1(\mathbb R_*) \) and with \(c_0 \in \mathbb R\). To exclude that the self-similar solution \(\Phi \) has a Dirac in \(\rho (M_{\gamma +\lambda })\), we notice that

$$\begin{aligned}&\int _\delta ^{\xi -\delta } K(\xi - \eta , \eta ) c_0\delta _{\rho (M_{\gamma +\lambda })}(\eta ) c_0 \delta _{\rho (M_{\gamma +\lambda })}(\xi -\eta ) d \eta \\&= c_0^2 \int _\delta ^{\xi -\delta } K(\xi -\eta , \eta ) \delta _{\rho (M_{\gamma +\lambda })} (\eta ) \delta _{\rho (M_{\gamma +\lambda })}(\xi -\eta ) d \eta \end{aligned}$$

is non-zero only if \(\xi = 2 \rho (M_{\gamma +\lambda }) \). Hence the coagulation operator applied to measure \(c_0 \delta _{\rho (M_{\gamma +\lambda })}\) produces a Dirac in \(2 \rho (M_{\gamma +\lambda })\), contradicting the fact that \(\Phi _{|(\rho (M_{\gamma +\lambda }), \infty )}\) is absolutely continuous with respect to the Lebesgue measure. As a consequence, we deduce that \(c_0=0\). \(\square \)

Proposition 5.15

(Moment bounds) Assume K to be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that \( \gamma +2\lambda \ge 1.\) Let \(\Phi \in {\mathcal {M}}_+(\mathbb R_*)\) be a self-similar profile as in Definition 3.1 and assume additionally that \(\Phi ((0,\delta ))=0\), where \(\delta >0\). Then

$$\begin{aligned} \int _{\mathbb R_*} x^\mu \Phi (d x) = \int _{[\delta , \infty )} x^\mu \Phi ( d x) < \infty , \quad \forall \mu \in \mathbb R. \end{aligned}$$

Proof

For \(\mu \le \gamma + \lambda \) the statement follows by the fact that \(\Phi ((0, \delta ))=0\) and that by definition of self-similar profile we know that

$$\begin{aligned} \int _{[\delta , \infty )} x^{\gamma + \lambda } \Phi (d x) < \infty . \end{aligned}$$

We want to prove the statement for \(\mu > \gamma + \lambda .\) As a first step, we prove that there exists a \(\overline{\delta }>0\) such that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma + \lambda +\overline{\delta }} \phi (x) d x < \infty , \end{aligned}$$
(5.45)

where \(\phi \) is the density of \(\Phi \). As a second step we will prove that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma + \lambda +n \overline{\delta }} \phi (x) d x < \infty \end{aligned}$$
(5.46)

for \(n \ge 1\) implies that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma + \lambda +(n+1)\overline{\delta }} \phi (x) d x < \infty . \end{aligned}$$
(5.47)

The desired conclusion then follows by induction.

Step 1

Consider an integrable function \(\varphi \) and integrate both sides of Eq. (5.43) against the function \(\psi (x)= \int _x^\infty \varphi (y) d y\) to deduce that

$$\begin{aligned}&\int _0^\infty \psi (z) z \phi (z) d z - \int _0^\infty \phi (z) \psi '(z) \left( \frac{z^{\gamma +\lambda +1}}{M_{\gamma + \lambda }} - \frac{2}{1-\gamma } z^2 \right) d z \nonumber \\&\quad - \frac{1}{M_{\gamma +\lambda } } \int _0^\infty \psi (z) z^{\gamma +\lambda } \phi (z) d z \nonumber \\&= \int _0^\infty \int _0^\infty x K(x,y) \left( \psi (x+y)-\psi (x)\right) \phi (x) \phi (y) d y d x . \end{aligned}$$
(5.48)

We remark that the above formulation requires that \(\psi \) has to decay fast enough so that all the integrals in (5.48) are finite.

Select \(\overline{\delta }>0\) such that \(\max \{0, 1- 2 (\gamma +\lambda )\}< \overline{\delta }< 1-\gamma - \lambda \) and consider \(\psi _R(x)=x^{\gamma +\lambda + \overline{\delta }-1 }\) if \(x \le R\) while \(\psi _R(x)=R^{\overline{\delta }} x^{\gamma +\lambda -1 }\) if \(x > R\). Using (5.48) and the fact that \(\gamma + 2\lambda \ge 1 \) and that \(\psi _R(x+y)-\psi _R(x) \le 0\) we deduce that

$$\begin{aligned} 0< \frac{\gamma + 2\lambda -1 +2\overline{\delta }}{1-\gamma } \int _0^R z^{\gamma + \lambda + \overline{\delta }} \phi (z) d z \le \frac{1}{M_{\gamma +\lambda } } \int _0^\infty z^{2(\gamma +\lambda )+ \overline{\delta }-1 } \phi (z) d z < \infty , \end{aligned}$$

where the moment \(M_{2(\gamma +\lambda )+ \overline{\delta }-1 }\) is bounded because, due to the choice of \(\overline{\delta }\), we have that \(2(\gamma +\lambda ) + \overline{\delta }-1 < \gamma +\lambda \). Passing to the limit as R tends to infinity we deduce that (5.45) holds.

Step 2

We assume that inequality (5.46) holds and we want to prove (5.47). Taking \(\psi _R(x)= x^{\gamma + \lambda + (n+1) \overline{\delta }-1 }\) when \(x \le R\) and \(\psi _R(x)= R^{\overline{\delta }} x^{\gamma + \lambda + n \overline{\delta }-1 }\) when \(x > R\) in (5.48) we deduce that

$$\begin{aligned}&0 <\frac{\gamma +2\lambda -1 +2\overline{\delta }(n+1 )}{1-\gamma } \int _0^R z^{\gamma + \lambda + (n+1)\overline{\delta }} \phi (z) d z \\&\quad \le \frac{1}{M_{\gamma +\lambda } } \int _0^\infty z^{2(\gamma +\lambda )+(n+1) \overline{\delta }-1 } \phi (z) d z \\&\qquad + \int _\delta ^\infty \int _\delta ^\infty x K(x,y) \left[ \psi _R(x+y)-\psi _R(x) \right] \phi (x) \phi (y) d y d x \\&\qquad + \int _0^\infty \phi (z) \psi _R'(z) z^{\gamma +\lambda +1} \phi (z) d z. \end{aligned}$$

Thanks to the choice of \(\overline{\delta }\), we have that \(2(\gamma +\lambda )+(n+1) \overline{\delta }-1 \le \gamma +\lambda + n \overline{\delta }\) and the desired conclusion follows for n such that \(\gamma +\lambda +(n+1)\overline{\delta }-1 < 1\) by the fact that in that case \(\psi _R\) is decreasing and hence the coagulation term is negative as well as the term \(\int _0^\infty \phi (z) \psi _R'(z) z^{\gamma +\lambda +1} \phi (z)\). We examine now the case in which \(\gamma +\lambda +(n+1)\overline{\delta }-1 \ge 1\). In this case we have that

$$\begin{aligned}&\int _\delta ^\infty \int _\delta ^\infty x K(x,y) \left[ \psi _R(x+y)-\psi _R(x) \right] \phi (x) \phi (y) d y d x \\&\quad \le c \int _\delta ^\infty \int _\delta ^\infty x K(x,y) \left[ (x+y)^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right. \\&\qquad \left. -x^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right] \phi (x) \phi (y) d y d x \\&\quad \le c \iint _{\{(x,y) \in [\delta , \infty )^{2} : x \le y\}} x^{1-\lambda } y^{\gamma +\lambda } \left[ (x+y)^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right. \\&\qquad \left. -x^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right] \phi (x) \phi (y) d y d x\\&\qquad + c \iint _{\{(x,y) \in [\delta , \infty )^{2} : x \ge y\}} x^{1+\gamma +\lambda } y^{-\lambda } \left[ (x+y)^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right. \\&\qquad \left. -x^{\gamma +\lambda +(n+1) \overline{\delta }-1} \right] \phi (x) \phi (y) d y d x \\&\quad \le c \int _\delta ^\infty x^{1-\lambda } \phi (x) d x \int _\delta ^\infty y^{2(\gamma +\lambda ) +(n+1) \overline{\delta }-1 } \phi (y) d y < \infty . \end{aligned}$$

Since we also have that

$$\begin{aligned} \int _0^\infty \phi (z) \psi _R'(z) z^{\gamma +\lambda +1} \phi (z) d z \le c(n) \int _0^\infty \phi (z) z^{2(\gamma +\lambda ) + (n+1)\delta -1} \phi (z) d z < \infty \end{aligned}$$

the desired conclusion follows. \(\square \)

Lemma 5.16

(Exponential bound) Assume K to be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that \( \gamma +2\lambda \ge 1.\) Let \(\Phi \in {\mathcal {M}}_+(\mathbb R) \) be as in Proposition 5.14 and let \(\phi \) be its density. Then there exist two positive constants L and M such that

$$\begin{aligned} \int _{M}^\infty \phi (z) e^{L z } d z < \infty . \end{aligned}$$
(5.49)

Proof

Adapting the approach used in [11] we define the function \(\Psi _a \) as

$$\begin{aligned} \Psi _a(L):= \int _M^\infty \frac{e^{L \min \{ x, a \}}}{\min \{ x, a \}} x^2 \phi (x) d x \end{aligned}$$

where \(M > \delta \). Hence, by its definition

$$\begin{aligned} \Psi _a'(L)= \int _M^\infty e^{L \min \{ x, a \}}x^2 \phi (x) d x. \end{aligned}$$

We consider a function \(\psi \) in (5.48) with \(\psi '(x):= e^{L \min \{ x, a \}} \) to deduce that

$$\begin{aligned} \frac{2}{1-\gamma } \Psi _a' (L)&\le \frac{M^{\gamma +\lambda -1}}{M_{\gamma +\lambda }} \int _M^\infty e^{L \min \{ x,a \}} x^{2} \phi (x) d x \\&\quad +\frac{1}{M_{\gamma +\lambda }} \int _\delta ^M e^{L \min \{ x,a \}} x^{\gamma + \lambda +1 } \phi (x) d x \\&\quad +\int _{\mathbb R_*} \int _{\mathbb R_*} x K(x,y) \int _{y}^{x+y} e^{L \min \{ w, a \}} d w \phi (x) \phi (y) d x d y \\&\quad + \int _\delta ^{{\overline{z}} } \left( \frac{1}{M_{\gamma +\lambda }} x^{\gamma + \lambda } -x \right) \psi (x) \phi (x) d x, \end{aligned}$$

for \( {\overline{z}}:= \left( \frac{1}{ M_{\gamma +\lambda }}\right) ^{\frac{1}{1-\gamma - \lambda }}\), where we are assuming without loss of generality that \({\overline{z}}> \delta \) and since for every \(z \ge {\overline{z}}\) we have

$$\begin{aligned} z- \frac{z^{\gamma + \lambda }}{M_{\gamma + \lambda }} > 0. \end{aligned}$$
(5.50)

Thus, there exists a \(c \ge 0\) such that

$$\begin{aligned}&\frac{2}{1-\gamma } \Psi _a' (L) \le \Psi _a' (L) \frac{M^{\gamma +\lambda -1}}{M_{\gamma +\lambda }}+ c \\ {}&\quad +\int _{\mathbb R_*}\int _{\mathbb R_*} x K(x,y) \int _{y}^{x+y} e^{L \min \{ w, a \}} d w \phi (x) \phi (y) d x d y . \end{aligned}$$

As in [9, 11] we can deduce, using Jensen’s inequality together with the fact that \(\gamma + \lambda < 1 \) and that \(- \lambda < 1 \), that

$$\begin{aligned} \int _{\mathbb R_*} \int _{\mathbb R_*} x K(x,y) \int _{y}^{x+y} e^{L \min \{ w, a \}} d w \phi (x) \phi (y) d x d y \le \Psi _a(L)^{1-\gamma -\lambda } \Psi _a'(L)^{\gamma +\lambda }. \end{aligned}$$

This, together with the fact that we can take M arbitrary large, implies that

$$\begin{aligned} c(\gamma ,\lambda ) \Psi _a' (L) \le \Psi _a (L)^{1-\gamma -\lambda } \Psi _a' (L)^{\gamma +\lambda } + c, \end{aligned}$$
(5.51)

for a positive constant c and for \( c(\gamma , \lambda )= \frac{2}{1-\gamma }-\frac{M^{\gamma +\lambda -1}}{M_{\gamma + \lambda }}>0.\)

By the definition of \(\Psi _a\) we have that

$$\begin{aligned} \Psi _a(0) \le \frac{2}{1-\gamma } \left( 1+ \frac{c_1}{a} \right) \rightarrow \frac{2}{1-\gamma } \text { as } a \rightarrow \infty . \end{aligned}$$

If we prove that \( \limsup _{a \rightarrow \infty } \Psi _a(L) < \infty \) for some L, then we can conclude. Indeed we would have

$$\begin{aligned} M \int _M^\infty e^{Lx } \phi (x) d x \le \int _M^\infty e^{Lx } x \phi (x) d x < \infty . \end{aligned}$$

Let us prove that \( \limsup _{a \rightarrow \infty } \Psi _a(L) < \infty \) for some L. First of all notice that this is true if \(\Psi _a'(L) \le 1 \) as in this case \(\Psi _a (L) \le L + \lim _{a \rightarrow \infty } \Psi _a(0) \) and the desired conclusion follows. If instead \(\Psi _a' >1 \), then

$$\begin{aligned} c(\gamma , \lambda ) \Psi _a'(L) \le \Psi _a (L)^{1-\gamma -\lambda } \Psi _a' (L)^{\gamma +\lambda }+c \le \alpha \Psi _a (L)^{1-\gamma -\lambda } \Psi _a' (L)+c. \end{aligned}$$

Without loss of generality we can assume \(\alpha >0\) such that \( \lim _{a \rightarrow \infty } \Psi _a(0) \ne \alpha \). This implies that a solution of the ODE

$$\begin{aligned} \Psi _a' \left( 1- \alpha \Psi _a^{1-\gamma -\lambda } \right) =c \end{aligned}$$

exists for small intervals around 0 and is such that \(\lim _{a \rightarrow \infty } \Psi _a (L) < \infty \) for L in that interval. \(\qquad \square \)

Proposition 5.17

(Exponential decay) Assume K to be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that \(\gamma +2\lambda \ge 1.\) Let \(\Phi \in {\mathcal {M}}_+(\mathbb R) \) be as in Proposition 5.14 and let \(\phi \) be its density. Then there exists a positive constant \(\tilde{M}\) such that

$$\begin{aligned} \limsup _{z \rightarrow \infty } \phi (z) e^{\tilde{M} z } < \infty . \end{aligned}$$
(5.52)

Proof

From Eq. (5.43) we deduce that

$$\begin{aligned}&\left( \frac{2}{1-\gamma } z^2 - \frac{z^{\gamma +\lambda +1 }}{M_{\gamma +\lambda }(\phi )} \right) \phi (z) \\&\qquad \le \int _0^z x \phi (x) d x - \frac{1}{M_{\gamma +\lambda }(\phi )}\int _0^z x^{\gamma +\lambda }\phi (x) d x + J_\phi (z). \end{aligned}$$

We now show that \(J_\phi \) decays exponentially and that the term

$$\begin{aligned} \int _0^z x \phi (x) d x - \frac{1}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y } \int _0^z x^{\gamma +\lambda }\phi (x) d x \end{aligned}$$

decays exponentially too.

First of all, we prove that there exists a constant \(M_1>0\) such that \( J_\phi (z) \le e^{- M_1 z}\) for large z. To this end we notice that the bound (5.49) implies that

$$\begin{aligned} J_\phi (z)&= \int _0^z \int _{z-x}^\infty e^{- L(x+y) }e^{L(x+y) } x K(x,y) \phi (x) \phi (y) d y d x \\&\le e^{- Lz } \int _0^z \int _{z-x}^\infty e^{L(x+y) } x K(x,y) \phi (x) \phi (y) d y d x \le e^{-M_1 z } . \end{aligned}$$

On the other side, for large values of z we have that

$$\begin{aligned}&\int _0^z x \phi (x) d x - \frac{1}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y } \int _0^z x^{\gamma +\lambda }\phi (x) d x \\&\quad = \frac{1}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y } \left( \int _0^z x \phi (x) d x \int _0^\infty y^{\gamma + \lambda } \phi (y) d y - \int _0^z x^{\gamma +\lambda }\phi (x) d x\right) \\&\quad \le \frac{1}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y } \left( \int _0^\infty y^{\gamma + \lambda } \phi (y) d y - \int _0^z x^{\gamma +\lambda }\phi (x) d x\right) \\&\quad \le \frac{1}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y } \int _z^\infty y^{\gamma + \lambda } \phi (y) d y \le C(\rho (M_{\gamma + \lambda })) e^{- M_2 z }, \end{aligned}$$

where \(M_2 >0\).

We deduce that for large values of z

$$\begin{aligned} \left( \frac{2z^2 }{1-\gamma } - \frac{z^{\gamma +\lambda +1 }}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y} \right) \phi (z) \le c \max \{e^{-z M_1}, e^{-z M_2} \}. \end{aligned}$$

The conclusion follows since, for every \(z > \left( \frac{1-\gamma }{2 \int _0^\infty x^{\gamma +\lambda } \Phi (d x)} \right) ^{\frac{1}{1-\gamma -\lambda }}\), we have

$$\begin{aligned} \frac{2z^2 }{1-\gamma } - \frac{z^{\gamma +\lambda +1 }}{\int _0^\infty y^{\gamma + \lambda } \phi (y) d y}>0. \end{aligned}$$

\(\square \)

6 Non-existence results and properties of the self-similar profiles

To study the non-existence of the self-similar solutions we proceed by contradiction and start by assuming that a self-similar solution exists. To find a contradiction we need to analyse the properties of each self-similar profile. In the case \(\gamma + 2 \lambda >1 \) the fundamental properties that we prove to be true are that \(\Phi ((0, \delta ))=0\) for some positive \(\delta >0\) and that \(\Phi \) decays sufficiently fast for large sizes. When \(\gamma +2\lambda =1\), it is possible to prove that \(\Phi \) decays fast for large sizes, but we have not been able to prove that \(\Phi ((0,\delta ))=0\) for some \(\delta >0\). This is the reason why we require the additional condition (3.5) in Theorem 3.5.

6.1 Properties of the self-similar profiles

In this section we study the properties of each self-similar solution that do not rely on the existence of the self-similar solution and hence do not rely on the assumption \(\gamma >-1\).

Theorem 6.1

(Properties of the self-similar profiles) Let K be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds.

  1. 1.

    If \(\gamma +2\lambda > 1\), then every self-similar profile \(\Phi \) as in Definition 3.1 is such that \(\Phi ((0, \rho (M_{\gamma +\lambda }) ))=0\) where \(\rho (M_{\gamma +\lambda }) \) is given by (3.4). Additionally, \(\Phi \) is such that

    $$\begin{aligned} \int _{\mathbb R_*} e^{Lx } \Phi (d x) < \infty \end{aligned}$$
    (6.1)

    for \(L>0\) and it is absolutely continuous with respect to the Lebesgue measure. Its density \(\phi \) is such that

    $$\begin{aligned} \limsup _{x \in \mathbb R_*} \phi (x) e^{ \overline{L} x} \le c \end{aligned}$$
    (6.2)

    for constants \(\overline{L}, c>0\).

  2. 2.

    If \(\gamma + 2\lambda =1 \) and if, in addition, \(\Phi ((0, \delta ))=0\) for some \(\delta >0\), then \(\Phi ([\delta , \rho (M_{\gamma +\lambda })))=0\) where \( \rho (M_{\gamma +\lambda })\) is given by (3.4). In addition, (6.1) and (6.2) hold.

First of all we prove, see Proposition 6.2, that when \(\gamma + 2 \lambda >1\), each solution of Eq. (1.19) in the sense of Definition 3.1 is zero near the origin. The statement of Theorem 6.1 then follows by Proposition 5.14 and Proposition 5.17 and by the fact that when \(\gamma +\lambda =1 \) we are assuming that there exists a \(\delta >0\) such that \(\Phi ((0, \delta ))=0\).

Proposition 6.2

(Support of the self-similar solution) Let K be a homogeneous symmetric coagulation kernel, of homogeneity \(\gamma \), satisfying (1.5), (1.6), with \(\gamma , \lambda \) such that (1.7) holds and such that \(\gamma +2\lambda > 1.\) Let \(\Phi \in {\mathcal {M}}_+(\mathbb R_*)\) be a self-similar profile as in Definition 3.1. Then \( \Phi ((0, \overline{\xi }))=0\) for

$$\begin{aligned} \overline{\xi }:= \min \left\{ \left( \frac{1}{M_{\gamma +\lambda }(\Phi )} \right) ^{\frac{1}{1-\gamma -\lambda }}, \left( \frac{1-\gamma }{ 2M_{\gamma +\lambda }(\Phi )} \right) ^{\frac{1}{1-\gamma -\lambda }} \right\} . \end{aligned}$$

Proof

We start explaining the proof in a heuristic manner without entering in the technical details that will be explained later. Equation (1.19) can be rewritten in the following flux form

$$\begin{aligned} \partial _x \left( J_\Phi (x) + \left( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma }\right) \Phi (x ) \right) =\left( \frac{x^{\gamma +\lambda }}{M_{\gamma +\lambda }} -x \right) \Phi (x ), \end{aligned}$$
(6.3)

where \(J_\phi \) is given by (5.44). For x smaller than \(\overline{\xi }\) we have that \( \frac{x^{\gamma +\lambda }}{M_{\gamma +\lambda }} -x \ge 0\) and hence the function

$$\begin{aligned} x \mapsto J_\Phi (x) + \left( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma }\right) \Phi (x ) \end{aligned}$$

is increasing and right-continuous at \(x=0\). This implies that

$$\begin{aligned} \lim _{ x \rightarrow 0^+ } \left( J_\Phi (x) + \left( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma }\right) \Phi (x ) \right) =L. \end{aligned}$$

Since for every \(x \le \overline{\xi }\) we also have that \( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma } \ge 0 \) we deduce that \(L \ge 0.\)

We make a scaling argument to identify the value of L. To this end we notice that the units of the flux are given by \([J_\Phi (x)] =[\Phi ]^2 [x]^{3+\gamma }\), where from now on we indicate with \([\cdot ]\) the dimensional properties of a quantity, hence

$$\begin{aligned}{}[\Phi ] =\left[ x\right] ^{-\frac{(3+ \gamma ) }{2}}. \end{aligned}$$

We deduce that

$$\begin{aligned}{}[x^{\gamma + \lambda +1} \Phi ] = [x^{\gamma + \lambda +1} ] [\Phi ] =[x^{\frac{\gamma + 2 \lambda -1}{2}} ]. \end{aligned}$$

Since \(\gamma + 2 \lambda > 1 \) this implies that as \(x \rightarrow 0 \) the dominant term in Eq. (6.3) is \(J_\Phi \), hence \(J_\Phi (x)\sim L\) as \(x \rightarrow 0\). Finally, we prove that \(L=0\), in agreement with the statement proven in [10] that, when \(\gamma + 2 \lambda \ge 1 \), there are no solutions to the constant flux equation. Integrating (6.3), we deduce that

$$\begin{aligned} J_\Phi (x) + \left( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma }\right) \Phi (x ) = \int _0^x \left( \frac{z^{\gamma +\lambda }}{M_{\gamma +\lambda }} -z \right) \Phi (z) d z. \end{aligned}$$

A detailed analysis of this ODE for \(\Phi \) implies that \(\Phi =0\) on the interval \(\left( 0, \left( \frac{1-\gamma }{2 M_{\gamma +\lambda } } \right) ^{\frac{1}{1-\gamma -\lambda } } \right) \), see Step 4.

We now explain the proof in detail. Testing (3.3) with a function of the form \(\psi (\xi )=\xi \varphi (\xi ),\) where \(\varphi \in C ^{1}_{c}(\mathbb R_*),\) we get to the following equation:

$$\begin{aligned} \int _{(0,\infty )}&\left( \xi -\frac{1}{M_{\gamma +\lambda }} \xi ^{\gamma +\lambda } \right) \varphi (\xi )\Phi (d \xi ) =\int _{(0,\infty )}\varphi '(\xi )W[\Phi ](d \xi ) , \end{aligned}$$
(6.4)

where

$$\begin{aligned} W[\Phi ](d x)= J_\Phi (x) d x + \left( \frac{x^{\gamma +\lambda +1 } }{ M_{\gamma + \lambda } } - \frac{2 x^2 }{1- \gamma }\right) \Phi (d x ). \end{aligned}$$
(6.5)

Step 1: Regularity of \(W[\Phi ]\) and an integral representation formula

First of all we prove that, for every \(\overline{\delta }>0\), the restriction of the measure \(W[\Phi ] \) on the interval \([\overline{\delta }, 1 ]\) is absolutely continuous with respect to the Lebesgue measure. This follows from the fact that for every function \(\varphi \in C^1_c([\overline{\delta }, \overline{\xi }])\) with \( 0< \overline{\delta }< 1\) we have that

$$\begin{aligned} \int _{\mathbb R_*} \varphi '(\xi )W[\Phi ](d \xi ) \le c(\Phi , \delta , \gamma , \lambda ) \Vert \varphi \Vert _\infty ; \end{aligned}$$

hence \(W[\Phi ]\), on the interval \([\overline{\delta },\overline{\xi }]\), has a density \(W_\phi \in BV([\overline{\delta }, \overline{\xi }] )\), see [1]. Since \(\overline{\delta }\) is arbitrary we deduce that \(W[\Phi ]\) has a density \(W_\phi \in L^{1}_{loc}((0, \overline{\xi }] )\). This implies that the measure \(\Phi _{| (0, \overline{\xi }]}\) is absolutely continuous with respect to the Lebesgue measure, its density is \(\phi \).

We now prove that \( W_\phi \) is increasing on \((0, \overline{\xi }]\). To this end we consider a sequence of functions \( \{\varphi _n \}_n \subset C^1_c(\mathbb R_* ) \) such that \(\varphi _n \rightarrow \chi _{[\xi _1, \xi _2]} \) pointwise, with \(0<\xi _1, \xi _2 < \overline{\xi }\), \(\varphi _n '(\xi ) \ge 0 \) for \(\xi \in \left( 0, \frac{\xi _1+\xi _2}{2}\right) \) and such that \(\varphi _n '(\xi ) \le 0 \) for \(\xi \in \left( \frac{\xi _1+\xi _2}{2}, \infty \right) \). Substituting \(\varphi _n\) in (6.4) we deduce that for every \(n\ge 1\)

$$\begin{aligned} \int _0^\infty&\xi \varphi _n(\xi )\phi (\xi ) d \xi -\frac{1}{M_{\gamma +\lambda }} \int _0^\infty \varphi _n(\xi ) \xi ^{\gamma +\lambda }\phi (\xi ) d \xi =\int _0^\infty \varphi _n'(\xi )W_\phi ( \xi ) d \xi . \end{aligned}$$

Notice that

$$\begin{aligned} \int _0^\infty \xi \varphi _n(\xi )\phi (\xi ) d \xi \rightarrow \int _{[\xi _1, \xi _2 ] } \xi \phi (\xi ) d \xi \quad \text { as } \quad n \rightarrow \infty \end{aligned}$$

and that

$$\begin{aligned} \int _0^\infty \xi ^{\gamma +\lambda } \varphi _n(\xi ) \phi (\xi ) d \xi \rightarrow \int _{[\xi _1, \xi _2 ] } \xi ^{\gamma +\lambda } \phi (\xi ) d \xi \quad \text { as } \quad n \rightarrow \infty . \end{aligned}$$

Additionally, since \(W_\phi \) is in \(L^1_{loc}((0, \overline{\xi }))\) and satisfies (6.4) we deduce that \(W_\phi \in W^{1}_{loc}(0, \overline{\xi })\) and hence is continuous in \((0, \overline{\xi })\). Hence we have that

$$\begin{aligned} \int _{(0,\infty )}\varphi _n'(\xi )W_\phi ( \xi ) d \xi \rightarrow W_\phi (\xi _1) - W_\phi (\xi _2) \quad \text { as } \quad n \rightarrow \infty . \end{aligned}$$

As a consequence, we deduce that

$$\begin{aligned} W_\phi (\xi _2) - W_\phi (\xi _1) = \int _{[\xi _1, \xi _2] } \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi . \end{aligned}$$
(6.6)

Since if \(\xi \le \overline{\xi }\), then \(\frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \ge 0\), we deduce that there exists a \(\overline{\xi }\) such that \(W_\phi \) is increasing in \((0, \overline{\xi }]\). Hence, since \(W_\phi \) is bounded, we have that the following limit exists

$$\begin{aligned} 0 \le L:= \lim _{\xi \rightarrow 0^+} W_\phi (\xi ) < \infty . \end{aligned}$$

Due to (3.2) we can take the limit as \(\xi _1 \rightarrow 0\) in (6.6) and deduce that

$$\begin{aligned} W_\phi (\xi _2) - L = \int _{(0, \xi _2] } \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi . \end{aligned}$$
(6.7)

Step 2: we prove that \(\int _{\mathbb R_*} J_\phi (\varepsilon x) \varphi (x) d x \rightarrow L \int _{\mathbb R_*}\varphi (x) d x\)

Since \(W_\phi \) is bounded in \((0, \overline{\xi }] \) and \(\frac{1}{M_{\gamma +\lambda }} \xi ^{\gamma +\lambda } -\frac{2}{1-\gamma } \xi \ge 0\) for every \(\xi \le \overline{\xi }\) we have that \(J_\phi \) is also bounded in \((0, \overline{\xi }] \). Since

$$\begin{aligned}{}[2/3 z, z ] \times [2/3 z, z ] \subset \left\{ x \in \mathbb R_*^2: 0< x \le z, z-x \le y < \infty \right\} =:\Omega _z \end{aligned}$$
(6.8)

we deduce that for every \(z \in (0, \overline{\xi }] \) we have that

$$\begin{aligned} z^{\gamma + 1 } \left( \int _{2z/3 }^z \phi (x) d x \right) ^2 \le J_\phi (z) \le C, \end{aligned}$$

hence

$$\begin{aligned} \frac{1}{z} \int _{2z/3}^z \phi (x) d x \le \frac{C}{z^{\frac{\gamma + 3}{2} }} \quad z \in (0, \overline{\xi }]. \end{aligned}$$
(6.9)

Equation (6.7) implies that

$$\begin{aligned} \int _{\mathbb R_*} \varphi \left( x\right) J_{\phi }(\varepsilon x) d x&=\frac{1}{\varepsilon }\int _{\mathbb R_*} \varphi \left( \frac{x}{\varepsilon }\right) J_{\phi }( x) d x \nonumber \\&= - \frac{1}{\varepsilon } \int _{\mathbb R_*} \varphi \left( \frac{x}{\varepsilon }\right) \left[ \frac{x^{\gamma +\lambda +1}}{M_{\gamma +\lambda }}- \frac{2}{1-\gamma }x^2\right] \phi (x) d x \\&+ \frac{1}{\varepsilon } \int _{\mathbb R_*} \varphi \left( \frac{x}{\varepsilon }\right) \int _0^x \left[ \frac{z^{\gamma +\lambda }}{M_{\gamma +\lambda }}- z\right] \phi (z) d z d x + L \int _{\mathbb R_*} \varphi (x) d x \nonumber \end{aligned}$$
(6.10)

where \(\varphi \in C_c((0, \overline{\xi }))\) and where \(\varepsilon >0\).

Thanks to the bounds (3.2) for \(\phi \) and to the fact that \(\varphi \) is compactly supported we can pass to the limit as \(\varepsilon \rightarrow 0\) in Eq. (6.10) and deduce that

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0 } \int _{\mathbb R_*} J_{\phi }(\varepsilon \xi ) \varphi (\xi ) d \xi = L \int _{\mathbb R_*} \varphi (x) d x \end{aligned}$$

for every \(\varphi \in C^1_c((0, \overline{\xi }))\).

Step 3: we prove that \(L=0\)

We want to prove that for every \(\varphi \in C^1_c((0, \overline{\xi }))\)

$$\begin{aligned} \lim _{\varepsilon \rightarrow 0 } \int _{\mathbb R_*} J_{\phi }(\varepsilon \xi ) \varphi (\xi ) d \xi =0. \end{aligned}$$

Using the formula for the fluxes we obtain that

$$\begin{aligned} \frac{1}{\varepsilon } \int _{\mathbb R_*} J_{\phi }( \xi ) \varphi \left( \frac{\xi }{\varepsilon }\right) d \xi&= \int _{\mathbb R_*} \varphi (\xi ) J_{\phi }(\varepsilon \xi ) d \xi \\&= \int _{\mathbb R_*} \int _{\mathbb R_*} x K(x,y) \left( \psi \left( \frac{x+y}{\varepsilon }\right) - \psi \left( \frac{x}{\varepsilon }\right) \right) \Phi (d x ) \Phi ( d y )\\&=I_1 + I_2 \end{aligned}$$

where \(\psi (x)= \int _0^x \varphi (y) d y\) and where

$$\begin{aligned} I_1 = \int _{\mathbb R_*} \int _{(y,\infty )} x K(x,y) \left( \psi \left( \frac{x+y}{\varepsilon }\right) - \psi \left( \frac{x}{\varepsilon }\right) \right) \Phi (d x ) \Phi ( d y ) \end{aligned}$$

and

$$\begin{aligned} I_2 = \int _{\mathbb R_*} \int _{(0, y)} x K(x,y) \left( \psi \left( \frac{x+y}{\varepsilon }\right) - \psi \left( \frac{x}{\varepsilon }\right) \right) \Phi (d x ) \Phi ( d y ). \end{aligned}$$

Since \(\varphi \) is compactly supported we have, by definition, that \(\text{ supp }\left[ \psi \left( x+y\right) - \right. \) \(\left. \psi \left( x\right) \right] \subset \left\{ (x,y) \in \mathbb R_*^2: x \le R, \ x+y \ge \overline{\delta }>0 \right\} \) for suitable \(\overline{\delta }>0\) and \(R>0.\) As a consequence, using (3.2) we deduce that

$$\begin{aligned} |I_1 |&\le c\int _{(0, \varepsilon R)} \int _{(\overline{\delta }\varepsilon /2, \varepsilon R)} y^{-\lambda } x^{1+\gamma +\lambda } \left| \psi \left( \frac{x+y}{\varepsilon }\right) - \psi \left( \frac{x}{\varepsilon }\right) \right| \Phi (d x) \Phi (d y) \\&\le C(\psi ) \frac{1}{\varepsilon } \int _{(0, \varepsilon R)} y^{1-\lambda } \Phi (d y) \int _{(\overline{\delta }\varepsilon /2, \varepsilon R)}x^{1+\gamma +\lambda } \Phi (d x) \rightarrow 0 \text { as } \varepsilon \rightarrow 0. \end{aligned}$$

Using again (3.2) we have

$$\begin{aligned} | I_2 |&= \int _{\mathbb R_*} \int _{(0,y )} x K(x,y) \left| \psi \left( \frac{x+y}{\varepsilon }\right) - \psi \left( \frac{x}{\varepsilon }\right) \right| \Phi (d x ) \Phi ( d y ) \\&\le C(\psi ) \int _{(\varepsilon \delta /2, \infty )} y^{\gamma +\lambda } \Phi ( d y ) \int _{(0, \varepsilon R)} x^{1-\lambda } \Phi (d x ) \rightarrow 0 \text { as } \varepsilon \rightarrow 0. \end{aligned}$$

Hence \( \lim _{\varepsilon \rightarrow 0 } \int _{\mathbb R_*} J_{\phi }(\varepsilon \xi ) \varphi (\xi ) d \xi =0\) and therefore \(L=0.\)

Step 4: \(\Phi \) is zero near zero

Since \(L=0\) we deduce by (6.7) that

$$\begin{aligned} W_\phi (z) = \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi \quad \text { for } z \in (0, \overline{\xi }). \end{aligned}$$
(6.11)

Or, equivalently, since \(J_\phi \in BV(\mathbb R_*)\)

$$\begin{aligned}&J_\phi (z) + \left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \phi (z) = \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi \nonumber \\&\quad \text { for } a.e. z \in (0, \overline{\xi }). \end{aligned}$$
(6.12)

We now rewrite (6.12) in the following way

$$\begin{aligned}&\int _0^z \int _{z-x}^z x K(x,y) \Phi (d y) \Phi ( d x) + \int _0^z \int _z^\infty x K(x,y) \Phi (d y) \Phi (d x) \\&+ \left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \phi (z) = \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi . \end{aligned}$$

If \(\Phi \) is different to zero near the origin, then

$$\begin{aligned} \int _0^z \int _z^\infty x K(x,y) \Phi ( d x) \Phi (d y )&\ge c_1 \int _0^z x^{1-\lambda } \Phi ( d x) \int _z^\infty y^{\gamma +\lambda } \Phi ( d y) \\&\quad + c_1 \int _0^z x^{1+\gamma +\lambda } \Phi (d x) \int _z^\infty y^{-\lambda } \Phi (d y ) \\&\ge C_+ \int _0^z x^{1-\lambda } \Phi ( d x) \end{aligned}$$

where we have used that for sufficiently small z we have that

$$\begin{aligned} \int _z^\infty y^{\gamma +\lambda } \Phi (d y) \ge C_+ \end{aligned}$$

for a strictly positive constant \(C_+\).

This implies that for z small we have that

$$\begin{aligned} 0&=J_\phi (z) + \left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \phi (z) - \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi \\&\ge C_+ \int _0^z x^{1-\lambda } \phi ( x) d x+ \left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \phi (z) \\&\quad - \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi \\&\ge \left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \phi (z) + \int _0^z \left( C_+ \xi ^{1-\lambda } - \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } \right) \phi (\xi ) d \xi . \end{aligned}$$

Since we are assuming that \(\Phi \) is different from zero near the origin and since \(\gamma +2\lambda >1\), we deduce that \(\int _0^z \left( C_+ \xi ^{1-\lambda } - \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } \right) \phi (\xi ) d \xi >0\). This, together with the fact that for \(z \le \overline{\xi }\) we have \(\left( \frac{z^{\gamma + \lambda +1} }{M_{\gamma + \lambda } }- \frac{2 z^2 }{1-\gamma }\right) \ge 0\) leads to a contradiction. Hence there exists a \(\delta >0\) such that \(\phi (x)=0\) for \(x \in (0, \delta )\). We now prove that \(\phi (x)=0\) for every \(x < \overline{\xi }\). To this end we define the function H as

$$\begin{aligned} H(z):= \int _0^z \left( \frac{1}{M_{\gamma +\lambda } } \xi ^{\gamma + \lambda } - \xi \right) \phi (\xi ) d \xi . \end{aligned}$$

Due to (6.12) we have that H satisfies

$$\begin{aligned} \frac{d}{dz} H(z)= H(z)Q(z)-J_\phi (z)Q(z), \quad H(\delta )=0 \quad \text { for } \ a.e. \ z\in (\delta , \overline{\xi })\nonumber \\ \end{aligned}$$
(6.13)

where

$$\begin{aligned} Q(z):=\frac{(1-\gamma ) ( z^{\gamma + \lambda } - M_{\gamma + \lambda }z) }{(1-\gamma ) z^{\gamma + \lambda +1} - 2 M_{\gamma + \lambda } z^2} > 0, \quad z \in (\delta ,\overline{\xi }). \end{aligned}$$
(6.14)

The solution of Eq. (6.13) is the function

$$\begin{aligned} H(z)= - \int _\delta ^z \frac{J_\phi (s) }{Q(s)} e^{ \int _s^z Q(x) d x } d s. \end{aligned}$$
(6.15)

Expression (6.15) for H implies that H is negative for \(z \in (\delta , \overline{\xi })\). On the other hand, if we assume that \(\phi \) is different from zero in \((0, \overline{\xi })\) we deduce that H is positive on \(z \in (\delta , \overline{\xi })\), by definition. This contradiction implies that \(H(z)=0\) for \(z \in (\delta ,\overline{\xi }).\)

\(\square \)

Remark 6.3

We remark that Eq. (6.11) is valid both for \(\gamma + 2 \lambda >1\) and for \(\gamma + 2 \lambda =1\), but the same argument we used to prove that \(\Phi ((0, \delta ))=0\) does not work in the case \(\gamma + 2 \lambda =1\).

Proof of Theorem 6.1

As in the proof of Theorem 5.1 we prove that if \(\Phi ((0, \delta ))=0\) for some \(\delta >0\) then \(\Phi ((0, \rho (M_{\gamma +\lambda }) ))=0\) for \(\rho (M_{\gamma +\lambda }) \) given by (3.4). To conclude the proof combine this with Proposition 5.14 and Proposition 5.17 with Proposition 6.2. \(\qquad \square \)

6.2 Non-existence results

Proof of Theorem 3.5

We consider now the case \(\gamma + 2\lambda >1\) and \(\gamma \le -1 \) in which \(\gamma + \lambda \) could be bigger than zero or smaller than zero, see Fig. 1. We will provide a proof that relies on the technical proof of Proposition 6.2. We proceed by contradiction. Assume that a self-similar solution exists. Then thanks to Proposition 6.2 we know that \(\Phi ((0, \delta ))=0\) for some \(\delta >0\). Additionally, from Proposition 5.17 we know that \(\lim _{x \rightarrow \infty } \Phi (x)=0\) exponentially.

We can therefore consider the test function \(\varphi \equiv 1\) in Eq. (3.3). This implies

$$\begin{aligned} 0\le - \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \Phi ( d x) =-\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K(x,y)\Phi ( d x) \Phi ( d y)<0, \end{aligned}$$

which is a contradiction.

Assume now that \(\gamma + 2 \lambda =1 \) and that \(\gamma \le -1\). Notice that in this case we necessarily have that \(\gamma +\lambda \le 0 \). Let us assume by contradiction that a self-similar solution exists and is such that

$$\begin{aligned} \int _{(0,1]} x^{-\lambda }\Phi (d x) < \infty . \end{aligned}$$

We show that

$$\begin{aligned} \int _{\mathbb R_*} \Phi (d x ) < \infty . \end{aligned}$$

Notice that this is true by assumption when \(\gamma +\lambda =0\), therefore we restrict the attention to the case \(\gamma +\lambda <0.\) To this end we adapt the argument used in the proof of Proposition 5.15, we refer there for the technical aspects, and prove that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma +\lambda } \Phi (d x)< \infty \quad \text { implies } \quad \int _{\mathbb R_*} x^{\gamma +\lambda +\overline{\delta }} \Phi (d x) < \infty \end{aligned}$$

when \( 0< -\gamma -\lambda<\overline{\delta }< 1-\gamma -\lambda \). To see this we consider \(\psi (x)=x^{\gamma +\lambda +\overline{\delta }-1 } \) in Eq. (5.48) to deduce that

$$\begin{aligned} \frac{2\overline{\delta }}{1-\gamma } M_{\gamma +\lambda +\overline{\delta }} \le \frac{M_{2(\gamma +\lambda )+\overline{\delta }-1}}{M_{\gamma +\lambda }}. \end{aligned}$$

Now we notice that our choice of \(\overline{\delta }\) implies that

$$\begin{aligned} - \lambda \le 2(\gamma +\lambda )+\overline{\delta }-1 \le \gamma + \lambda . \end{aligned}$$

Hence \(M_{2(\gamma +\lambda )+\overline{\delta }-1}< \infty \) and the desired conclusion follows.

Similarly, we can prove that

$$\begin{aligned} \int _{\mathbb R_*} x^{\gamma +\lambda + n \overline{\delta }} \Phi (d x)< \infty \quad \text { implies } \quad \int _{\mathbb R_*} x^{\gamma +\lambda +(n+1)\overline{\delta }} \Phi (d x) < \infty . \end{aligned}$$

As a consequence, taking n sufficiently large, we deduce that \(M_0(\Phi )< \infty \). We can therefore consider the test function \(\varphi \equiv 1\) in Eq. (3.3). This implies

$$\begin{aligned} 0\le - \frac{1+\gamma }{1-\gamma } \int _{\mathbb R_*} \Phi ( d x) =-\frac{1}{2}\int _{\mathbb R_*} \int _{\mathbb R_*} K(x,y)\Phi ( d x) \Phi ( d y)<0, \end{aligned}$$

which is a contradiction. \(\qquad \square \)