Self-similar source-type solutions to the three-dimensional Navier-Stokes equations
Abstract
We formalise a systematic method of constructing forward self-similar solutions to the Navier-Stokes equations in order to characterise the late stage of decaying process of turbulent flows. (i) In view of critical scale-invariance of type 2 we exploit the vorticity curl as the dependent variable to derive and analyse the dynamically-scaled Navier-Stokes equations. This formalism offers the viewpoint from which the problem takes the simplest possible form. (ii) Rewriting the scaled Navier-Stokes equations by Duhamel principle as integral equations, we regard the nonlinear term as a perturbation using the Fokker-Planck evolution semigroup. Systematic successive approximations are introduced and the leading-order solution is worked out explicitly as the Gaussian function with a solenoidal projection. (iii) By iterations the second-order approximation is estimated explicitly up to solenoidal projection and is evaluated numerically. (iv) A new characterisation of nonlinear term is introduced on this basis to estimate its strength quantitatively. We find that for the 3D Navier-Stokes equations. This should be contrasted with for the Burgers equations and for the 2D Navier-Stokes equations. (v) As an illustration we explicitly determine source-type solutions to the multi-dimensional the Burgers equations. Implications and applications of the current results are given.
Navier-Stokes equations,self-similarity, scale-invariance
1 Introduction
Self-similarity is a tool of fundamental importance in analysing partial differential equations, including the construction of solutions and the determination of their stability. It is also useful for numerical and asymptotic methods of studying partial differential equations. For general aspects of self-similarity and its applications we refer the readers to e.g. [1, 2].
In this paper we will study the so-called source-type self-similar solution to the Navier-Stokes equations. Our motivation for this is as follows. First of all, it will give us a particular self-similar solution to the Navier-Stokes equations, which characterises the decaying process in the late stage of evolution. Second, it is likely to give useful information as to how we may handle more general solutions.
It may be in order to have a look at previous works which are related to this paper. It has been shown under mild conditions that no nontrivial smooth backward self-similar solution exists to the Navier-Stokes equations. On the other hand it is known that nontrivial forward self-similar solutions do exist, but their explicit functional forms are not known, except for some asymptotic results. It is of interest to see how they actually behave because such solutions contain important information regarding more general solutions. This is particularly the case when the governing equations are exactly linearisable, e.g. the Burgers equations. While it is not expected that the Navier-Stokes equations are exactly soluble in general, we might still obtain insights into the nature of their solutions.
In [3] the existence of forward self-similar solutions for small data was proven by using a fixed-point theorem in a Besov space (see below). There, initial data for the self-similar solution (3D) are assumed to be homogeneous of degree in velocity
and the existence of a self-similar solution of the form
has been established under the assumption that initial data is small in some Besov space. Moreover, it has been proved that the self-similar profile satisfies (in their notations)
where denotes a heat operator at time 1 and is small. In [4] using a locally Hölder class in the smallness assumption has been removed and it is furthermore shown that
where and denotes a constant with some norm of . Those studies indicate that the self-similar solution is close to the heat flow in the late stage. However, studies on the determination of a specific functional form of self-similar solutions are few and far between, except for an attempt in[5]. We also note that the existence of generalised self-similar solutions (in the sense that the scaling holds only at a set of discrete values of ) was studied subsequently, e.g. [7, 6].
Our basic strategy in practice is as follows. After recasting the dynamically-scaled Navier-Stokes equations as integral equations via the Duhamel principle, we regard the nonlinear term as a perturbation using the Fokker-Planck evolution semigroup. Systematic successive approximations are then introduced and the first-order solution is worked out explicitly as the Gaussian function with a solenoidal projection. The second-order approximation is also evaluated numerically to assess the strength of the nonlinear term.
We will construct an approximate solution to the 3D Navier-Stokes equations valid in the long-time limit, using the vorticity curl . In Sections 4 we will see why this is the most convenient variable from the reaction of the Navier-Stokes equations under dynamic scaling.
Here we appreciate the suitability of such a choice of the unknown by comparing the source-type solutions to the Navier-Stokes equations and their self-similar solutions at criticality. By definition the source-type solution for nonlinear parabolic PDEs is a solution in a scaled space, which starts from the Dirac mass in some dependent variable and ends up like a near-identity of the Gaussian function in the long-time limit. It serves as an analogue of the fundamental solution to nonlinear PDEs.
The unknown whose -norm is marginally divergent is suitable for describing the late-stage evolution. This is because this self-similar solution satisfies the same scaling as the Dirac mass and both of them belong to a Besov space near . In one dimension, in the limit of , we have roughly
which suggests that the velocity is convenient in this case.
In two dimensions it is the vorticity which is the most convenient, as can be seen from
where . Recall that those scaling properties of velocity in 1D or vorticity in 2D, are the same as that of the Dirac mass; in -dimensions.
Now consider Besov spaces whose norms are given by
where and represents band-filtered velocity at frequency . It is known that in the Dirac delta mass is embedded as
(1) |
see e.g. [8]. In particular we have for any
While the velocity we have
and correspondingly for the vorticity curl,
Hence in three dimensions this and the Dirac mass belong to the same function class , with in (1).
The rest of this paper is organised as follows. In Section 2 after reviewing critical scale-invariance of type 2 (to be defined below) using the Burgers equation we introduce successive approximations of determining the self-similar profile. On this basis we introduce and quantify the strength of nonlinearity. Higher-dimensional Burgers equations are also discussed. In Section 3 we have a brief look at the 2D Navier-Stokes equations. In the main Section 4 we study the self-similar solutions of the 3D Navier-Stokes equations utilising the vorticity curl as the unknown to achieve critical scale-invariance of type 2. We carry out successive approximations and determine the strength of nonlinearity as introduced above. Section 5 will be devoted to the Summary and outlook. Some further details and derivations are given in Appendices.
2 Burgers equations
We review the source-type solution of the Burgers equation with an emphasis on the critical scale-invariance. Our approach is novel in the introduction of its successive approximations, in preparation for handling the 3D Navier-Stokes equations, and in employing a new method of estimating the strength of nonlinearity on this basis. Also described are the source-type solutions of the Burgers equations in -dimensions.
2.1 Critical scale-invariance
We consider the Burgers equation [9]
(2) |
which satisfies static scale-invariance under
This means that if is a solution, so is for any . It is readily checked that
which shows that the -norm is scale-invariant.
Let us clarify the two kinds of critical scale-invariance. Type 1 scale-invariance is achieved when we use a dependent variable whose physical dimension is the same as . Type 1 is deterministic in nature where the additional term arising in the governing equations under dynamic scaling is minimised in number. Type 2 instead is statistical in nature where the additional terms under dynamic scaling are maximised in number so that a divergence form is completed and the dynamically-scaled equations have the Fokker-Planck operator as the linearisation. In the former the dependent variable has the same physical dimension as kinematic viscosity, whereas in the latter the argument of the Hopf characteristic functional (the independent variable) has the same physical dimension as the reciprocal of kinematic viscosity [15]. This approach provides a viewpoint from which the problem appears in the simplest possible form.
Critical scale-invariance of type 1 is achieved with the velocity potential which is defined by . If is a solution, so is Under dynamic scaling for the velocity potential we have
(3) |
whose linearisation has the Ornstein-Uhlenbeck operator. This is called type 1 (deterministic) scale-invariance where the number of additional terms is minimised, that is, only the drift term remains. Under dynamic scaling for velocity , we find
(4) |
whose linearisation is the Fokker-Planck equation. Here the zooming-in parameter has the same physical dimension as and is on the same order of it. With this type 2 (statistical) scale-invariance where the number of additional terms is maximised meaning that a divergence form is completed with the addition of term. As it is a second-order equation it has two independent solutions, of which we will focus on the Gaussian one. See Appendix C for the other non-Gaussian kinds of solutions.
Equation (4) is exactly soluble and its steady solution is called the source-type solution [11, 12]:
(5) |
The name has come from the time zero asymptotics
where and Observe that (5) is a near-identity transformation of the Gaussian function. See [10, 11, 12].
It is also known that for we have
where
The simplest method for solving (4) without linearisation is as follows. Rewrite the equation
By changing variables to we find
which is readily integrable. Alternatively we may solve the equation (2.1) by regarding it as a Bernoulli equation.
It may be in order to comment on the significance of source-type solution. When we recast (5) as
(6) |
which is reminiscent of the celebrated Cole-Hopf transform. In other words, the source-type solution encodes the vital information of the nonlinear term in the case of the Burgers equation. Note that the error-function itself in (5, 6) is a self-similar solution to the heat equation. This suggests that studying source-type solution of the Navier-Stokes equations may give a hint on how to characterise their long-time evolution by a heat flow.
2.2 Successive approximations
The operator is not self-adjoint. It is possible to find a function 111With a slight abuse of notation this should be distinguished from the Gaussian function used in Section 3. such that holds, where is the adjoint of . In fact where denotes the Dawson’s integral, defined by [18]. However, because decays slowly at large distances as , it cannot be used as a Green’s function, at least, in the usual manner.
The inversion formula for can be obtained by an alternative method. Recall that based on the fundamental solution to the Poisson equation in 1D is given by
where * denotes convolution. Likewise for the fundamental solution to the Fokker-Planck equation in 1D we write
where
and f.p. denotes the finite part of Hadamard, e.g. [13, 14]. It can be verified by changing the variable from to using the solution of the Fokker-Planck equation
As we will consider the 3D Navier-Stokes equations, for which methods of exact solutions are unavailable, we treat (4) by approximate methods as an illustration. Because the inversion of is unwieldy, we will seek a workaround by which we can dispense with it.
First we convert it to an integral equation by the Duhamel principle for the Fokker-Planck operator
The long-time limit is given by
We may consider a number of different iteration schemes. For example, the following option (1), also known as the Picard iteration, requires the inversion :
Note that is a steady function at each step.
Alternatively we first consider the steady equation
and then introduce iteration schemes:
or
Note that iteration schemes (1) and (2a) coincide with each other at .
2.3 Estimation of the strength of nonlinearity
For the Burgers equation we can work out the two kinds of approximations to the second-order analytically. After straightforward algebra they are
where On this basis we estimate the strength of the nonlinear term The source-type solution is a near identity transform of the Gaussian function. In its series expansion in the Reynolds number after non-dimensionalisation, the nonlinear correction term has as its prefactor. Consider the scheme (1), or (2a) equivalently, taking without loss of generality. We have
Separating out the -dependence, or equivalently assuming that we define by
so that holds. The strength of nonlinearity is given by
The same goes for (2b) from the above expressions for the 1D Burgers equation. Altogether we find
We conclude that the typical strength of nonlinearity is irrespective of the choice of schemes.
2.4 Burgers equations in several dimensions
The source-type solution is basically a near-identity function of the Gaussian form. It has been seen how the source-type solutions show up in the long-time limit in one and two spatial dimensions in [15]. Here we will take a look at cases in three and higher dimensions. From the Cole-Hopf transform we have
As we are going for the type 2 scale-invariance, differentiating it twice we find
The denominator then tends to as where Hence
where the function is to be determined such that We can thus take
Therefore after collecting other terms of derivatives we find in three dimensions, say, with
(7) |
where
denotes the Reynolds number. Because is small the expression (7) is near-Gaussian. It can be verified that where We can also write
which reflects the Cole-Hopf transform more directly, that is, See Appendix A for the general form in -dimensions.
3 2D Navier-Stokes equation
We briefly recall the self-similar solution of the 2D Navier-Stokes equations, where the so-called Burgers vortex appears after dynamic scaling.
3.1 Critical scale-invariance
The Burgers vortex was originally introduced to represent the reaction of a vortex under the influence of the collective effect of surrounding vortices in the ambient medium. When we write the steady solution in velocity and vorticity using cylindrical coordinates
the solution takes the following forms
where denotes the velocity circulation.
In two dimensions dynamic scaling transforms take the following form
In the 2-dimensional case, in order to achieve the critical scale-invariance of type 2, we must choose the second spatial derivative of the stream function, which is the vorticity, as the unknown.
The scaled form of the vorticity equation in two dimensions reads
where satisfies the type 2 scale-invariance. It is known that the self-similar solution under scaling has a mathematically identical form as the Burgers vortex above. Indeed in the scaled variables the above expression can be written
Note that is an exact self-similar decaying solution222When an exact decaying solution is obtained by formally replacing which is known as the Lamb-Oseen vortex. with the following property
It also satisfies the following asymptotic property, for
where see e.g. [16].
3.2 Interpretation
Because the source-type solution coincides with the linearised solution, the inhomogeneous terms on the right-hand side of the approximations at each order vanish identically. Hence there is no way to set up successive approximations that can capture non-zero nonlinear corrections. The strength of nonlinearity is identically zero; .
4 3D Navier-Stokes equations
We will describe two approaches for handling the scaled 3D Navier-Stokes equations perturbatively. First we describe a general framework based on the Green’s function. Second we describe an iterative approach which is specifically suited for calculations associated with the 3D Navier-Stokes problem.
4.1 Critical scale-invariance
We consider the 3D Navier-Stokes equations written in four different dependent variables. Starting from the vector potential and taking a curl successively we have
(8) |
where and p.v. denotes a principal-value integral. We also have because of the incompressibility condition. The derivation of (8)1 can be found in [17]. The final fourth equation (8)4 is obtained by taking the Laplacian of the velocity equations (8)2. For the equations we may alternatively take a curl on the vorticity equations to obtain a form of equation different from the final line in (8), which is useful for handling inviscid fluids (details to be found in Appendix B). Under dynamic scaling
the 3D Navier-Stokes equations in four different unknowns are transformed respectively to
(9) |
where
It is to be noted that the coefficient of the linear term increases in number with the increasing order of derivatives and for the variable a divergence is completed in the convective term. Observe that type 1 scale-invariance is achieved with and type 2 scale-invariance with .
4.2 Successive approximations
Using the Duhamel principle we convert the scaled Navier-Stokes equations (9)4 into integral equations
Here and the action of whose exponential operator is given by
(10) |
for any function as can be verified by combining the heat kernel and dynamic scaling transforms.
The inverse operator associated with the fundamental solution to the Fokker-Planck equation in 3D is defined by
where the Green’s function is given by
This can be verified by changing the variable from to in the solution (10) to the Fokker-Planck equation.
We consider the steady solution in the long-time limit of
where denotes the leading-order approximation (to be made more explicit in next subsection). This is one form of integral equations we are supposed to handle.
On the other hand, steady equations are obtained by assuming in (9)4
or
This is yet another form of the steady Navier-Stokes equations after dynamic scaling. It is noted that one of the potential problems associated with the nonlinear term has been eliminated without having a recourse to the Green’s function. It is this virtually trivial fact that allows us to set up a simple successive approximation.
To summarise, the steady Navier-Stokes equations after dynamic scaling can be written as
(11) |
or, by we can express it solely in terms of as
(12) |
This is the set of equations that we need to solve.
In passing we note the following facts before proceeding to the specific results. By the definition of scaled variables it is easily seen that for
This means that if as we have
That is about the long-time asymptotics. On the other hand as time-zero asymptotics we have
where is the Dirac mass.
4.3 Leading-order approximations
Before discussing the second-order approximation we derive expressions of the first-order solutions in several different variables.
We will derive the basic formulas by solving the heat equation
The first-order approximation is given by
After applying the dynamic scaling
the linearised equations for the vector potential read
After dynamic scaling their solution is given by
where denotes the first-order approximation and the initial data. (The same convention applies to and in the following.) Taking a curl with respect to three times we find the expressions for the vorticity curl
For well-localised initial data we make use of the formula where Noting that we have333If the initial condition satisfies the similarity condition we have In this case is singular like .
where and This is the leading-order approximation for the scaled 3D Navier-Stokes equations.
The first-order (that is, the leading-order) approximation obtained above can be calculated explicitly because the Gaussian function is a radial function. Care should be taken that the leading-order approximations themselves are not radial because of the incompressibility condition. Indeed, in terms of the vorticity curl the first-order approximation is given by
where and summation is implied on repeated indices. In the second line we computed for the Gaussian function using and the final line by direct computations. Clearly the final expression is not radial.
Hereafter in this subsection we take for simplicity. Note that for can be evaluated by quadratures and their explicit form are as follows, which can be obtained most conveniently with the assistance of computer algebra. The results are
(14) | |||||
(15) | |||||
(16) | |||||
With these at hand the expressions for the several different unknowns, including the vorticity curl above, are ()
(17) | |||||
(18) | |||||
(19) | |||||
(20) | |||||
(21) |
where and denotes the error function. Here for convenience we have introduced two functions
both of which are continuous at (See Figure 1 below). Because all the fields are incompressible we also have
Regarding the derivations of (17)-(21), applying to (21) repeatedly we get (19) and (17), respectively. Then, taking a curl of (17) we get (18) and taking a curl of (19) we get (20). Finally simpler forms of (19) and (21) are obtained by taking a curl of (18) and (20), respectively. The final results in vectorial form read
(22) | |||||
(23) | |||||
(24) | |||||
(25) |
It is of interest to observe that . Using the above formula (20) with it is instructive to compare a component of the Gaussian function with that of the vorticity curl
(26) |
Figure 1 shows how is affected by the incompressible condition (solenoidality), in particular the peak value at is reduced by a factor of .

4.4 Estimation of the strength of nonlinearity
We first describe the numerical methods employed to obtain the approximate solutions. We use a centered finite-difference scheme in a box of size with discretised coordinates where and .
An important step in the determination of the second-order approximation is the inversion of the Laplacian operator. This was done by using a Poisson solver in the Intel MKL library. Parameters used are hence . The box size was chosen large enough to capture the decay of near the boundaries and was chosen large enough to resolve spatial structure in the center of the domain. As a code validation we calculated (4.6) numerically and compared against the analytical expression (26) and confirmed their agreement (figure omitted).
To evaluate the perturbation we make use of the iteration scheme (2b) illustrated in the previous section for simplicity. Consider a series expansion of the source-solution in . In comparison to the leading-order the nonlinear correction is proportional to .444After full non-dimensionalisation we have where As it is clear that the nonlinear term has a factor of in comparison to the leading-order approximation. Separating out or equivalently assuming that we define the strength of nonlinearity by the remaining factor. To be more specific the second-order solution in this case is given by
Separating out the -dependence we define by
(27) |
and put .


It turns out that the -component of the second-order correction along the line is identically equal to zero owing to the radial symmetry of the Gaussian function. For this reason we show in Figure 3 the -component of the first-order approximation as a function of along the line . It has a peak at the origin whose height is approximately 0.12. Accordingly we show in Figure 3 the -component of the second-order correction due to the nonlinearity as a function of . It has double peaks near the origin, but their value is small and is about 0.0015. Noting by (27) we can estimate the strength of nonlinearity in that cross-section as
Actually the maximum value of the nonlinear term in the above sense in is 0.0022, not much different from the above value. Thus the strength of nonlinearity is at most and we conclude
It should be noted that it is much smaller than the value of for the Burgers equations, whose solutions are known to remain regular all time. Because the difference between the Navier-Stokes and Burgers equations is the presence or absence of the incompressibility condition, it is the incompressibility that makes the value of smaller for the Navier-Stokes equations. On a practical side this also means that even if we add the second-order correction to the first-order term at low Reynolds number, say the superposed solution is virtually indistinguishable from the first-order approximation. In the final period of decay the Navier-Stokes flows are very close to (25), to within at which may be regarded as the building blocks for representing the late-stage of evolution, that is, as the counterpart of the Burgers vortex in two dimensions.
5 Summary and outlook
We have studied self-similar solutions to the fluid dynamical equations with particular focus on the so-called source-type solutions. As an illustration of successive approximation schemes we have discussed the 1D Burgers equation which is exactly soluble. In this case the velocity is the most convenient choice for its analysis. We have introduced a method of quantitatively assessing the strength of nonlinearity using the source-type solutions. Similar analyses have been carried out for higher-dimensional Burgers equations. For the Burgers equations we find in any dimensions.
We then move on to consider the 2D and 3D Navier-Stokes equations. In two dimensions we review the known results done using the vorticity. In three dimensions the most convenient choice of the unknown is the vorticity curl. We have formulated the dynamically-scaled equations using that variable and set up the successive approximation schemes. We have found that the second-order correction stemming from the nonlinear term gives rise to , an order of magnitude smaller than that for the Burgers equations. We are led to conclude that the incompressible condition makes smaller for the Navier-Stokes equations than for the Burgers equations.
The current approach relies on perturbative treatments. It may be challenging, but worthwhile to study the functional form of the solution by non-perturbative methods for further theoretical developments. It is also of interest to seek a fully non-linear solution by numerical methods. It is noted that this is at least one order of magnitude smaller than found for the Burgers equations whose solutions are known to remain regular all the time. As an application of the source-type solution, it is useful to characterise the late stage of statistical solutions of the Navier-Stokes equations [15].
Appendix A Source solution to the Burgers equations in dimensions
The source-type solution to the -dimensional Burgers equations takes the following form
where ’s are polynomials to be constructed below,
and
Construction
Consider a function separable in variables
where is a smooth function and . Define another function by
then we have
where is a sequence of polynomials in of degree not exceeding In fact, it is given by
or equivalently,
The first four of them are
Proof (Due to Yuji Okitani.)
Noting we have
while the penultimate expression of which may also be written
Likewise we have
Hence in general the recursion relationship is
where
Alternatively, by we compute
and
In general we find
and hence
Using the general expression we can estimate the nonlinearity in -dimensional cases. Write for small we have
so we find
As we deduce for all Hence the estimate obtained in the iteration (1) in Section 2(c) holds valid in any dimensions.
Appendix B Derivation of the vorticity curl equations
Recalling vector identities
and adding them and solving for we have
Taking we find
We then compute
Identifying the underlined part as the right-hand side, we obtain
that is,
Appendix C Steady Fokker-Planck equation
C.1 One-dimensional case
Because
(28) |
is a second-order equation, its general solution has two constants of integration. We solve it paying attention to the boundary conditions. First, upon integration we have
for some constant . If decays sufficiently rapidly as , then . Otherwise it’s possible to have . Let us proceed keeping and write
A further integration gives
with another constant . Here we have assumed that as However, if the second term survives with . The function
is the Dawson’s integral [18], which behaves as It also satisfies
where denotes the Hilbert transform
As we can write
with suitably redefined constants . This shows the general solution consists of two kinds of solutions, which we call a source-type solution and a kink-type one. The former converges to the Dirac mass and the latter to the Cauchy kernel in the limit of
It is of interest to note that if a solution is found, then its Hilbert transform gives the other solution. In fact applying to (28)
By e.g. [19], it follows that
which shows that also a solves the same equation. See the Fig.4 for a comparison of those fundamental solutions.

initial data | non-Gaussian steady solutions | |
---|---|---|
1D Burgers | , continuous | |
2D Navier-Stokes | , discontinuous | |
3D Navier-Stokes | , discontinuous |
C.2 Two-dimensional case
In the case of the 1D Burgers equation the singular self-similar initial condition becomes continuous after infinitesimal time evolution. The second steady solution, which is not Gaussian, is also continuous but not in . We should take it into account when we discuss long-time evolution.
For the 2D (and 3D) Navier-Stokes equations, the other steady solutions of non-Gaussian form, are in fact neither continuous nor in . They are discontinuous at the origin. Hence care should be taken in considering them when we discuss long-time evolution in a larger function space such as .


Under the assumption of radial symmetry the Fokker-Planck equation in two dimensions takes the following form
which is equivalent to
Indeed,
Setting we have that is,
Hence
where denotes the exponential integral and the Euler’s constant.
Derivation
Define
We have
Now, with
As
we find
Hence
that is,
Taking its f.p.,
Here
denotes the logarithmic integral and
C.3 Three-dimensional case
Under the assumption of radial symmetry, the Fokker-Planck equation in three dimensions takes the following form
where It is equivalent to
Indeed,
Setting we have Thus we can write
A further integration gives
where denotes the Dawson’s integral.
Derivation
As
we have
Noting
and dropping the term, we obtain the desired form.
Acknowledgement
This work has been supported by EPSRC grant EP/N022548/1.
Its major part was carried out when one of the authors (K.O.)
was affiliated with School of Mathematics and Statistics, the University of Sheffield.
References
- [1] Barenblatt GI. 2003 Scaling Cambridge: Cambridge University Press.
- [2] Dresner L. 1983 Similarity Solutions of Nonlinear Partial Differential Equations Boston: Pitman Publishing.
- [3] Cannone M and Planchon F. 1996. Self-similar solutions for Navier-Stokes equations in . Commun. in P.D.E. 21, 179–193.
- [4] Jia H and Šverák V. 2014. Local-in-space estimates near initial time for weak solutions of the Navier-Stokes equations and forward self-similar solutions. Invent. Math. 196, 233–265.
- [5] Brandolese L. 2009 Fine properties of self-similar solutions of the Navier–Stokes equations. Arch. Rat. Mech. Anal. 192, 375–401.
- [6] Bradshaw Z and Tsai T-P. 2019. Discretely self-similar solutions to the Navier–Stokes equations with data in Lloc2 satisfying the local energy inequality. Anal. PDE 12, 1943–1962.
- [7] Chae D and Wolf J. 2017. Removing discretely self-similar singularities for the 3D Navier–Stokes equations Commun. Partial Diff. Eqs. 42, 1359–1374.
- [8] Hairer M and Labbé C. 2017. The reconstruction theorem in Besov spaces. J. Funct. Anal. 273, 2578–2618.
- [9] Burgers JM. 1948. A mathematical model illustrating the theory of turbulence Adv. Appl. Mech. 1, 171–199.
- [10] Escobedo M and Zuazua E. 1991. Large time behavior for convection-diffusion equations in . J. Func. Anal. 100, 119–161.
- [11] Biler P, Karch G and Woyczyński WA. 1999. Asymptotics for multifractal conservation laws. Studia Mathematica 135, 231-252
- [12] Biler P, Karch G and Woyczyński WA. 2001. Critical nonlinearity exponent and self-similar asymptotics for Lévy conservation laws. Annal. IHP Anal. nonlin. 18, 613–637.
- [13] Bureau FJ. 1955. Divergent integrals and partial differential equations. Commun Pure Appl. Math. 8 143–202.
- [14] Yosida K. 1956 An operator-theoretical integration of the wave equation. J. Math. Soc. Jpn. 8 79–92.
- [15] Ohkitani K. 2020. Study of the Hopf functional equation for turbulence: Duhamel principle and dynamical scaling. Phys. Rev. E 101, 013104-1–15.
- [16] Gallay T and Wayne CE. 2005. Global stability of vortex solutions of the two-dimensional Navier-Stokes equation. Commun. Math. Phys. 255, 97–129.
- [17] Ohkitani K. 2015. Dynamical equations for the vector potential and the velocity potential in incompressible irrottational Euler flows: a refined Bernoulli theorem. Phys. Rev. E. 92, 033010-1–8.
- [18] Olver FW, Lozier DW, Boisvert RF and Clark CW. (eds.) 2010 NIST Handbook of Mathematical Functions Cambridge: Cambridge University Press.
- [19] King FW. 2009. Hilbert transforms Cambridge: Cambridge University Press.