533
Views
0
CrossRef citations to date
0
Altmetric
Research Article

On Global Existence and Regularity of Solutions for a Transport Problem Related to Charged Particles

Abstract

The paper considers a class of linear Boltzmann transport equations which models charged particle transport, for example in dose calculation of radiation therapy. The equation is an approximation of the exact transport equation containing hyper-singular integrals in its collision terms. The paper confines to the global case where the spatial domain G is the whole space R3. Existence results of solutions for the due initial value problem are formulated by applying variational methods. In addition, some regularity results of solutions are verified in scales of relevant anisotropic mixed-norm Sobolev spaces.

AMS CLASSIFICATION:

1. Introduction

This paper deals with the existence and regularity of solutions of the linear Boltzmann transport equation (BTE) (1) a(x,E)ψE+c(x,E)ΔSψ+d(x,ω,E)·Sψ+ω·xψ+Σ(x,ω,E)ψKrψ=f(1) in R3×S×I. Here S is the unit sphere (velocity direction domain) and I=[E0,Em] is the energy interval. x is the gradient with respect x-variable and S and ΔS are the gradient and Laplacian on S, respectively. The solution satisfies the initial condition (2) ψ(.,.,Em)=0(2) where Em is the so-called cutoff energy. This condition guarantees that the overall problem is well-posed. The restricted collision operator Kr is a partial integral type operator (see section 3.1 below) (3) Krψ=SIσ1(x,ω,ω,E,E)ψ(x,ω,E)dEdω+Sσ2(x,ω,ω,E)ψ(x,ω,E)dω+I02πσ3(x,E,E)ψ(x,γ(E,E,ω)(s),E)dsdE.(3)

The term ψω·xψ on the left in Equation(1) is called a convection (or advection) operator, the term ψΣψ is a scattering operator and the term ψKrψ is a restricted collision operator. The term ψd·Sψ is needed when external forces (electro-magnetic fields, for example) are present. The term ψaψE represents the energy attenuation and cΔSψ is linked to the diffusion of the angular variable on the sphere. On the right, the function f=f(x,ω,E) represent the internal source. The solution ψ of the problem (1, 2) describes the fluence of the considered particle. We remark that in the case where the spatial domain G=R3 one must impose an additional inflow boundary condition (4) ψ|Γ=g(4) where Γ is the “inflow boundary” {(y,ω,E)(G)×S×I|ω·ν(y)<0}. Generally speaking the existence and regularity analysis in this case is more sophisticated than in the global case G=R3.

In Tervo et al. (Citation2018), Tervo (Citation2019, section 6) and Tervo and Herty (Citation2020) one has given reasons to use the equations like Equation(1) for charged particle transport, for example for electrons and positrons in radiation therapy dose calculation. The starting point is that differential cross sections for charged particles may contain hyper-singularities with respect to energy variable and hence the corresponding exact (original) collision operator is a partial hyper-singular integral operator. This operator can be reasonably approximated which leads to an approximative equation of the form Equation(1).

We find that the operator in Equation(1) is the sum of the second order partial differential operator and a partial integral operator. Let (5) A(x,ω,E,D)ψ:=1a(ΔSψ+d·Sψ+ω·xψ+Σψ)(5)

Then the equation in Equation(1) is for a > 0 equivalent to (6) ψE+A(x,ω,E,D)ψ+1aKrψ=1af.(6)

In the case where Kr = 0 the problem (Equation1, 2) is an initial value problem for the partial differential equation (7) ψE+A(x,ω,E,D)ψ=1af,ψ(.,.,Em)=0.(7)

The existence and regularity results of this reduced problem mirror those of the complete problem (Equation1, 2).

The operator E+A(x,ω,E,D) is hyperbolic in nature in its wide sense (e.g. Rauch Citation2012, 47, Pazy Citation1983, 134). The literature contains numerous contributions for existence and regularity analysis of general partial differential initial boundary value problems which are hyperbolic in nature or which are formally dissipative beginning from Lax and Phillips (Citation1960, Theorem 3.2), Friedrichs (Citation1958) and Phillips and Sarason (Citation1966). More recent results can be found e.g. in Hörmander (Citation1985), Rauch and Massey (Citation1974), Rauch (Citation1985, Chapter XXIII), Morando et al. (Citation2009), Nishitani and Takayama (Citation1998), Nishitani and Takayama (Citation2000).

Some specific results concerning for existence and regularity of solutions of transport problems can also be found in the literature. We mention some of them. In the case where a=c=d=0 existence of solutions for transport problems like (1, 4) has been studied for single equations e.g. in Agoshkov (Citation1998), Dautray and Lions (Citation1999), Egger and Schlottbom (Citation2014) and for coupled systems in (Tervo and Kokkonen Citation2017). In Agoshkov (Citation1998, Chapter 4) a systematic study of the regularity of solutions is exposed. Therein a single monokinetic BTE is considered and the spatial domain G is a bounded subset of Rn,n=1,2,3 with sufficiently regular boundary. In the above references it is assumed that the restricted collision operator Kr satisfies a (partial) Schur criterion for boundedness (Halmos and Sunder Citation1978, 22). In Tervo et al. (Citation2018a, Citation2018b), we studied existence of solutions for the case c=d=0. In Frank et al. (Citation2010) additionally it has been assumed that a=a(E). In Tervo and Herty (Citation2020) we proved existence results for the problem Equation(1), Equation(4), Equation(2) when the spatial domain G was bounded.

Related results for (deterministic and linear) Fokker-Planck type equations can be found in Degond (Citation1986, Appendix A) and Tian (Citation2014). Methods in Degond (Citation1986) are closely related to our techniques in section 3 below. In Le Bris and Lions (Citation2008), existence results of solutions are shown for a class of time-dependent Fokker-Planck equations with irregular (having only Sobolev regularity) coefficients. Results in Chupin (Citation2010) consider existence results for a special form of stationary Fokker-Planck equation in weighted Sobolev spaces. In Herty et al. (Citation2012) and Herty and Sandjo (Citation2011) existence results are obtained in the context of dose calculation for optimal radiation treatment planning.

In Mokhtar-Kharroubi (Citation1991), Sobolev regularity results up to order 1 are proved when a=c=d=0 by applying singular integral methods (see also Zeghal Citation2012). In Bouchut (Citation2002) one has shown that a time-dependent transport problem satisfies a kind of “gain in x-regularity with the help of ω-regularity.” Note that the time-dependent equation is of the form Equation(1) when we replace time t with energy E. In Alonso and Sun (Citation2014, especially Theorem 5.3) and in some of its references one has considered regularity results for the mono-kinetic equation (here Sn1 denotes the unit sphere in Rn) ut+ω·xuKu=0inRn×Sn1,u(.,.,0)=u0 where the collision operator is of the form (Ku)(x,ω):=Sn1(u(x,ω)u(x,ω)σ(ω,ω)dω.

Chen and He (Citation2012) have investigated the regularity of solutions of time-dependent nonlinear equation (for general foundations of these equations see e.g. Ukai and Yang Citation2010) (8) ut+ω·xuQ(u,u)=0,u(.,.,.,0)=u0.(8)

The paper confines to the case n = 3 and to the periodic solutions in spatial variable. Q(.,.) is an appropriate bilinear form.

In this paper we restrict ourselves to the case where the spatial dimension n = 3 and the Lebesgue index p = 2. The regularity results are formulated utilizing the anisotropic Sobolev spaces H(m1,m2,m3)(R3×S×I°). For the first instance, we in section 2 introduce these spaces for integer indexes (m1,m2,m3) and bring up some of their properties. These spaces are subspaces of L2(R3×S×I).

In Section 3 we consider the existence and uniqueness of solutions for the problem (1, 2) in global case G=R3. The solution spaces are certain subspaces of L2(R3×S×I). The proofs are modifications of proofs given in Tervo and Herty (Citation2020) for the case of a bounded spatial domain G. The results are based on the Lions-Lax-Milgram Theorem. In Tervo et al. (Citation2018), section 6.2 we applied same kind of techniques in the case where c=d=0. In Tervo et al. (Citation2018a, Citation2018b), we considered related results rested on the m-dissipativity of the smallest closed extension of the partial differential part of the transport operator, the methods of which offer an alternative approach for existence analysis.

The focus of this paper is in Sections 4 and 5 where we study the regularity of solutions in the global case G=R3. We restrict ourselves in the main to the spatial regularity (x-regularity) but some outlines for the regularity with respect to all variables are stated as well. We shall find that the main principle operating in the global case is roughly speaking that the “regularity of solutions increases according to the regularity of data.” We proceed in the increasing order of complexity. For the first instance in section 4 we deal with merely the slowing down-convection-scattering equation (that is, Kr=0,c=d=0) and therein the proofs are founded on the known explicit formulae of solutions. These treatments suggest relevant anisotropic Sobolev spaces within which regularity results can be formulated. After that the total Equationequation (1) is handled in section 5 Therein the proofs are based on the use of partial differences and on the obtained a priori estimates.

We finally expose some notes for the case where G=R3. In this case we must impose additionally an inflow boundary condition Equation(4). This inflow boundary condition causes some problems because the corresponding initial inflow boundary value problem (Equation1, 2, 4) has the so called variable multiplicity (Morando et al. Citation2009; Nishitani and Takayama Citation1998; Nishitani and Takayama Citation2000). We also remark that to retrieve smoothness properties of solutions (with respect to evolution variable) in the case of hyperbolic problems, one always must assume that the relevant compatibility conditions are valid for the data. Actually, due to the needed inflow boundary condition Equation(4) the Sobolev regularity of solutions is strongly dependent on the properties of the so called escape time mapping t(x,ω) contrary to the global case G=Rn. This can be seen transparently e.g. from explicit solution formulas (cf. e.g. Tervo and Kokkonen Citation2017, section 4.2 and Tervo et al. Citation2018a, section 10.2). The inflow boundary condition causes that regularity of solutions in the case G=R3 is not necessarily increasing along the data with respect to (x,ω)-variable. In fact, it is known that the regularity of the solution of the transport equation is limited (with respect to x, for example) up to the fractional space H(s1,0,0)(G×S×I°),s1<3/2 regardless of the smoothness of the data (see counterexample given in Tervo and Kokkonen Citation2017, section 7.1).

2. Preliminaries

2.1. Basic notations and concepts

We restrict ourselves to the spatial dimension n = 3 and we consider the transport problem where the spatial (position) domain G=R3. In this case the related problems are often called global ones. Let S = S2 be the unit sphere in R3 (S is the velocity direction domain). Furthermore, let I=[E0,Em] (I is the energy interval) where 0E0<Em<. We shall denote by I° the interior of I. Define the (Sobolev) space W2(R3×S×I) by W2(R3×S×I):={ψL2(R3×S×I)|ω·xψL2(R3×S×I)} equipped with the inner product ψ,vW2(R3×S×I):=ψ,vL2(R3×S×I)+ω·xψ,ω·xvL2(R3×S×I).

The space W2(R3×S×I) is a Hilbert space and the space C01(R3,C1(S×I)) is a dense subspace of W2(R3×S×I) (e.g. Friedrichs Citation1944).

We recall that the following Green’s formulas are valid (9) R3×S×I(ω·xψ)vdxdωdE=R3×S×Iψ(ω·xv)dxdωdE(9) for every ψ,vW2(R3×S×I) and (10) R3×S×IΔSψ,vdxdωdE=R3×S×ISψ,SvdxdωdE(10) for every ψ,vL2(R3×I,H1(S)) for which ΔSψL2(3×S×I). Here S is the gradient on S and ΔS is the Laplace-Beltrami operator (so called Laplacian) on S. H1(S) is the standard Sobolev space on the compact manifold S.

2.2. Function spaces needed in regularity analysis

Let m=(m1,m2,m3)N03 be a multi-index. Define anisotropic Sobolev spaces Hm(R3×S×I°) by (11) Hm(R3×S×I°):={ψL2(R3×S×I°)|xαωβElψL2(R3×S×I°),for all|α|m1,|β|m2,lm3},(11) where the derivatives are taken in the distributional sense. Above {ωj|j=1,2} is a local basis of the tangent space T(S). We recall that for sufficiently smooth functions f:SR fωj|ω=uj(f°h)|u=h1(ω),j=1,2.

Here h:USS0 is a parametrization of SS0 where S0 has the surface measure zero. Moreover, recall that fL2(S) if and only if f°hL2(U,|Ah|du) where |Ah|:=1h2h. In L2(S) we use the inner product (12) f1,f2L2(S)=Sf1f2dω:=f1°h,f2°hL2(U,|Ah|du).(12)

The space Hm(R3×S×I°) is a Hilbert space when equipped with the inner product (13) ψ,vHm(R3×S×I°):=|α|m1,|β|m2,lm3xαωβElψ,xαωβElvL2(R3×S×I°).(13)

The corresponding norm is ψHm(R3×S×I°)=(|α|m1|β|m2lm3xαωβElψL2(R3×S×I°)2)12.

Note that for mm (that is, mjmj,j=1,2,3) Hm(R3×S×I°)Hm(R3×S×I°).

The tensor product Hm1(R3)Hm2(S)Hm3(I°) is a dense subspace of Hm(R3×S×I°) but generally the spaces Hm1(R3)Hm2(S)Hm3(I°) and Hm(R3×S×I°) are not equal (principles of these kind of results are found e.g. in Aubin Citation1979, Chapter 12). Moreover, we have

Theorem 2.1.

The space (14) D(R3×S×I):={ψ|R3×S×I°|ψC0(R3×S×R)}(14) is dense in Hm(R3×S×I°).

Proof.

The proof follows by using the standard cutting and Friedrichs mollifier smoothing techniques. □

Remark 2.2.

For multi-indexes of the form m=(m1,0,m3) the spaces Hm(R3×S×I°) can be characterized by using Fourier transforms. The (partial) Fourier transform with respect to (x, E) of ψL2(R3×S×R) (in the sense of tempered distributions) is given by

(F(x,E)ψ)(ξ,ω,η):=R3SRψ(x,ω,E)ei(ξ,η),(x,E)dxdE,(ξ,η)R3×R,ωS.

The inverse partial Fourier transform is then ψ(x,ω,E)=(F(x,E)1ψ)(x,ω,E):=(2π)4R3R(F(x,E)ψ)(ξ,ω,η)ei(ξ,η),(x,E)dξdη.

For multi-indexes m=(m1,0,m3) we define (15) Ĥm(R3×S×R):={ψL2(R3×S×R)|ψĤm(Rn×S×R)<}(15) where ψĤm(R3×S×R)2:=R3SR(1+|ξ|2)m1(1+|η|2)m3|(F(x,E)ψ)(ξ,ω,η)|2dξdωdη.

The space Ĥm(R3×S×R) is a Hilbert space when equipped with the inner product (16) ψ,vĤm(R3×S×R):=R3SR(F(x,E)ψ)(ξ,ω,η)(F(x,E)v)¯(ξ,ω,η)·(1+|ξ|2)m1(1+|η|2)m3dξdωdη.(16)

Due to Plancherel’s formula (17) Hm(R3×S×R)=Ĥm(R3×S×R)(17) and the inner products Equation(13) and Equation(16) are equivalent. The spaces Hm(R3×S×I°) are isomorphic to the factor spaces Hm(R3×S×R)/H0m(R3×S×(R\I)) where H0m(R3×S×(R\I)) is the completion of C0(R3×S×(R\I)) with respect to the inner product Equation(13).

We finally define the space (18) W,m(R3×S×I°):={fL(R3×S×I)|xαωβElfL(R3×S×I)<for|α|m1,|β|m2,lm3}(18) which is equipped with the norm fW,m(R3×S×I°):=max|α|m1,|β|m2,lm3xαω˜βElfL(R3×S×I).

3. Existence of solutions when the spatial domain is R3

We assume that the transport operator T is of the form (19) Tψ=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ+ΣψKrψ(19) where b(x,ω,E,ω)ψ is given by (20) b(x,ω,E,ω)ψ=c(x,E)ΔSψ+d(x,ω,E)·Sψ.(20)

Here ΔS is the Laplace-Beltrami operator on sphere and d=(d1,d2)d1ω1+d2ω2. d·S is the chosen Riemannian inner product on the tangent space (bundle) T(S). Denote (21) d(x,ω,E,ω)ψ:=d(x,ω,E)·Sψ.(21)

We consider an initial value transport problem (22) Tψ=f,ψ(.,.,Em)=0(22) in the global case G=R3 where fL2(R3×S×I). The existence of solutions for the problem Equation(22) can be studied, by applying e.g. the generalized Lax-Milgram Theorem, the so called Lions-Lax-Milgram Theorem. We will use the following statement that can be found e.g. in Treves (Citation1975, 403) or Grisvard (Citation1985, 234).

Theorem 3.1.

Let X and Y be Hilbert spaces, with Y continuously embedded into X. Assume that B(·,·):X×YR is a bilinear form satisfying the following properties with M0,c>0, (23) |B(u,v)|MuXvYuX,vY(boundedness)(23) and (24) B(v,v)cvX2vY(coercivity).(24)

Suppose that F:XR is a bounded linear form. Then there exists uX (possibly non-unique) such that (25) B(u,v)=F(v)vY.(25)

3.1. Assumptions for the restricted collision operator

We assume that the restricted collision operator is the sum (for some more details see Tervo et al. Citation2018, section 5) (26) Kr=Kr1+Kr2+Kr3.(26)

Here Kr1 is of the form (Kr1ψ)(x,ω,E)=S×Iσ1(x,ω,ω,E,E)ψ(x,ω,E)dωdE, where σ1:R3×S2×I2R is a non-negative measurable function such that (27) S×Iσ1(x,ω,ω,E,E)dωdEM1,S×Iσ1(x,ω,ω,E,E)dωdEM2,(27) for a.e. (x,ω,E)R3×S×I.

The operator Kr2 is of the form (Kr2ψ)(x,ω,E)=Sσ2(x,ω,ω,E)ψ(x,ω,E)dω, where σ2:R3×S2×IR is a non-negative measurable function such that (28) Sσ2(x,ω,ω,E)dωM1,Sσ2(x,ω,ω,E)dωM2,(28) for a.e. (x,ω,E)R3×S×I.

Finally, Kr3 is of the form (Kr3ψ)(x,ω,E)=I02πσ3(x,E,E)ψ(x,γ(E,E,ω)(s),E)dsdE where γ=γ(E,E,ω):[0,2π]S is a parametrization of the curve (an example for the choice of γ=γ(E,E,ω) is given in Tervo et al. Citation2018) Γ(E,E,ω)={ωS|ω·ωμ(E,E)=0}.

Moreover, σ3:R3×I2R is a non-negative measurable function such that (29) Iσ3(x,E,E)dEM1,Iσ3(x,E,E)dEM2,(29) for a.e. (x,E)R3×I. The following result is shown analogously to Theorem 5.13 in Tervo et al. (Citation2018) (in the reference in question we have assumed that the spatial domain G is bounded).

Theorem 3.2.

The operators Krj,j=1,2,3 are bounded operators L2(R3×S×I)L2(R3×S×I) and (30) Kr1M1M2,(30) (31) Kr2M1M2,(31) (32) Kr32πM1M2,.(32)

In order to render the operator ΣKr coercive (accretive), we shall assume that (33) Σ(x,ω,E)S×Iσ1(x,ω,ω,E,E)dωdESσ2(x,ω,ω,E)dω2πIσ3(x,E,E)dEc,(33) and (34) Σ(x,ω,E)S×Iσ1(x,ω,ω,E,E)dωdESσ2(x,ω,ω,E)dω2πIσ3(x,E,E)dEc,(34) for a.e. (x,ω,E)R3×S×I. In the sequel we assume that c>0.

Next result addresses coercivity (accretivity) of KrΣ under the above assumptions. The result is proven analogously to Theorem 5.14 in Tervo et al. (Citation2018a) (or Tervo et al. Citation2018a, section 4).

Theorem 3.3.

Suppose that the assumptions Equation(27), (Equation28, 29, 33) and Equation(34) are valid. Then (35) (ΣKr)ψ,ψL2(R3×S×I)cψL2(R3×S×I)2ψL2(R3×S×I).(35)

3.2. Existence of weak solutions

At first we verify (formally) the corresponding variational equation. Assume that ψ is a solution of Equation(22) in the classical sense and let vC01(R3,C1(I,C2(S))). By the Green’s formula Equation(10) we have S(ΔSψ)(x,ω,E)v(x,ω,E)dω=S(Sψ)(x,ω,E),(Sv)(x,ω,E)dω where Sψ,Sv is the chosen Riemannian inner product on S. Hence, (36) c(x,E)ΔSψ,vL2(3×S×I)=Sψ,c(x,E)SvL2(3×S×I):=3×S×ISψ,c(x,E)SvdxdωdE.(36)

Moreover, d(x,ω,E,ω)ψ,vL2(R3×S×I)=ψ,d*(x,ω,E,ω)vL2(R3×S×I) where d*(x,ω,E,ω) is the formal adjoint of d(x,ω,E,ω) that is, (37) d*(x,ω,E,ω)v=d(x,ω,E,ω)v(divSd)v=-d(x,ω,E)·Sv(divSd)v(37) where divS is the divergence on S. By integration by parts over I (38) aψE,vL2(R3×S×I)=R3Sa(x,Em)ψ(x,ω,Em)v(x,ω,Em)dωdxR3Sa(x,E0)ψ(x,ω,E0)v(x,ω,E0)dωdxψ,(av)EL2(R3×S×I)=ψ,avEL2(R3×S×I)ψ,a)EvL2(R3×S×I)+a(.,Em)ψ(.,.,Em),v(.,.,Em)L2(R3×S)a(.,E0)ψ(.,.,E0),v(.,.,E0)L2(R3×S).(38)

Using Green’s formula Equation(9) we obtain (39) ω·xψ,vL2(R3×S×I)=ψ,ω·xvL2(R3×S×I).(39)

Finally, (40) Krψ,vL2(R3×S×I)=ψ,Kr*vL2(R3×S×I)Σψ,vL2(R3×S×I)=ψ,Σ*vL2(R3×S×I)(40) where Σ*=Σ and Kr* is the adjoint of Kr which can be computed (but we omit computations here).

As a conclusion we see that if ψ is a classical solution of problem Equation(22) then the following weak formulation is fulfilled (41) B(ψ,v):=ψ,avEL2(R3×S×I)ψ,a)EvL2(R3×S×I)a(.,E0)ψ(.,.,E0),v(.,.,E0)L2(R3×S)Sψ,c(x,E)SvL2(R3×S×I)+ψ,d*(x,ω,E,ω)vL2(R3×S×I)ψ,ω·xvL2(R3×S×I)+ψ,Σ*vKr*vL2(R3×S×I)=f,vL2(R3×S×I).(41)

Define in C01(R3,C1(I,C2(S))) inner products (42) ψ,vH:=ψ,vL2(R3×S×I)+ψ(.,.,E0),v(.,.,E0)L2(R3×S)+ψ(.,.,Em),v(.,.,Em)L2(R3×S)+ψ,vL2(R3×I,H1(S))(42) and (43) ψ,vĤ=ψ,vH+ω·xψ,ω·xvL2(R3×S×I)+ψE,vEL2(R3×S×I).(43)

Let H and Ĥ be the completions of C01(R3,C1(I,C2(S))) with respect to the inner products .,.H and .,.Ĥ, respectively.

We assume for the coefficients: (44) aL(R3,W,1(I)),cL(R3×I),djL(R3×I,W,1(S))(44) (45) (a)E(x,E)+(divSd)(x,ω,E))q1>0,a.e.,(45) (46) c(x,E)q2>0,a.e.,(46) (47) a(x,E0)q3>0,a(x,Em)q3>0,a.e..(47)

We proceed analogously to Tervo and Herty (Citation2020). At first, we show that the bilinear form B:C01(R3,C1(I,C2(S)))×C01(R3,C1(I,C2(S)))R obeys the following boundedness and coercivity conditions:

Theorem 3.4.

Suppose that the assumptions Equation(27), (Equation28, 29, 33, 34, 44–47) are valid. Then there exists a constant M > 0 such that (48) |B(ψ,v)|MψHvĤψ,vC01(R3,C1(I,C2(S)))(48) and (49) B(v,v)cvH2vC01(R3,C1(I,C2(S)))(49) where (50) c:=min{q12,q32,q2,12,c}.(50)

Proof.

A. The boundedness can be seen as in Tervo et al. (Citation2018a, Theorem 6.4), Tervo and Herty (Citation2020, Theorem 6.6) and so we omit details.

B. Secondly, we verify the coercivity Equation(49). By partial integration we have for vC01(R3,C1(I,C2(S))) (51) v,av)EL2(R3×S×I)=v,(av)EL2(R3×S×I)v(·,·,Em),a(·,Em)v(·,·,Em)L2(R3×S)+v(·,·,E0),a(·,E0)v(·,·,E0)L2(R3×S)(51) and then (52) v,av)EL2(R3×S×I)=12(v,aEvL2(R3×S×I)v(·,·,Em),a(·,Em)v(·,·,Em)L2(R3×S)+v(·,·,E0),a(·,E0)v(·,·,E0)L2(R3×S)).(52)

Using the Green’s formula we have (53) v,ω·xvL2(R3×S×I)=ω·xv,vL2(R3×S×I)(53) which implies (54) ω·xv,vL2(R3×S×I)=0.(54)

Furthermore, we have d(x,ω,E,ω)v,vL2(R3×S×I)=v,d*(x,ω,E,ω)vL2(R3×S×I) which implies (by recalling Equation(37) that (55) v,d*(x,ω,E,ω)vL2(R3×S×I)=12(divSd)v,vL2(R3×S×I).(55)

Finally, we have (56) Sv,c(x,E)SvL2(R3×S×I)=R3×S×Ic(x,E)(Sv)(x,ω,E),(Sv)(x,ω,E)dxdωdE.(56)

Inserting (Equation52, 54, 55) and Equation(56) in the expression of B(.,.,) (given in Equation(41)) with ψ=v we get in virtue of the assumptions Equation(45), (46, 47) and by Theorem 3.3 the required estimate Equation(49) as in Tervo and Herty (Citation2020, Theorem 6.6).

This completes the proof. □

Because C01(R3,C1(I,C2(S)))×C01(R3,C1(I,C2(S))) is dense in H×Ĥ and since Equation(48) holds, the bilinear form B(·,·):C01(R3,C1(I,C2(S)))×C01(R3,C1(I,C2(S)))R has a unique extension B˜(·,·):H×ĤR which satisfies (57) |B˜(ψ,v)|MψHvH^ψH,vH^(57) and (58) B˜(v,v)cvH2vĤ.(58)

Furthermore, define a bounded linear form (59) F:HR;F(ψ):=f,ψL2(R3×S×I).(59)

Note also that the embedding ĤH is continuous.

Let P(x,ω,E,D)ψ:=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ be the differential part of T. The space (60) HP(R3×S×I°):={ψL2(R3×S×I)|P(x,ω,E,D)ψL2(R3×S×I)in the weak sense}(60) is a Hilbert space when equipped with the inner product ϕ,vHP(G×S×I°)=ψ,vL2(R3×S×I)+P(x,ω,E,D)ψ,P(x,ω,E,D)vL2(R3×S×I).

Using this notation, the Equationequation (22) can be written shortly as P(x,ω,E,D)ψ+ΣψKrψ=f.

We formulate the existence of weak solutions without the initial condition ψ(.,.,Em)=0.

Theorem 3.5.

Suppose that the assumptions of Theorem 3.4 that is, (Equation27–29, 33, 34, 44–47) are valid. Let fL2(R3×S×I). Then the variational equation (61) B˜(ψ,v)=F(v)vĤ(61) has a solution ψH. Furthermore, ψHP(R3×S×I°) and it is a weak (distributional) solution of the equation (62) Tψ:=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ+ΣψKrψ=f.(62)

Proof.

The proof is similar as in Tervo and Herty (Citation2020, Theorem 6.8).□

Remark 3.6.

For fL2(R3×I,H1(S),vL2(R3×I,H1(S) we define f,v:=R3×I(f(x,.,E),v(x,.,E))dxdE where (f(x,.,E),v(x,.,E)) is the canonical duality between H1(S) and H1(S) Since (63) f,vfL2(R3×I,H1(S))vL2(R3×I,H1(S))(63) we find that the above Theorem 3.5 is valid more generally for fL2(R3×I,H1(S)).

3.3. Existence of solutions for the initial value problem

Let Q(x,ω,E,D)ψ:=a(x,E)ψE+ω·xψ.

Define the space HQ(R3×S×I°) HQ(R3×S×I°):={ψL2(R3×I,H1(S))|Q(x,ω,E,D)ψL2(R3×I,H1(S))} equipped with the inner product ψ,vHQ(R3×S×I°)=ψ,vL2(R3×I,H1(S))+Q(x,ω,E,D)ψ,Q(x,ω,E,D)vL2(R3×I,H1(S)).

Furthermore, define the traces γm(ψ):=ψ(·,·,Em),γ0(ψ):=ψ(·,·,E0). We verify the following trace theorem

Theorem 3.7.

Suppose that (64) aW,1(R3×I°)such that|a(x,E)|q4>0a.e.inR3×I.(64)

Then the trace operators γm:HQ(R3×S×I°)Lloc2(R3×S),γ0:HQ(R3×S×I°)Lloc2(R3×S), are well-defined and continuous.

Proof.

In virtue of density results like Friedrichs (Friedrichs Citation1944; Rauch Citation1985) the space C01(R3,C1(S×I)) is dense in HQ(R3×S×I°) and so it suffices to show the below boundedness estimate only for ψC01(R3,C1(S×I)).

By the Green’s formula for ψC01(R3,C1(S×I)) (65) 2R3×S×I1a(ω·xψ)ψdxdωdE=R3×S×Iω·x(1a)ψ2dxdωdE(65) where we used that ω·x(1aψ)=1aω·xψ+ω·x(1a)ψ.

Note that ψE+1aω·xψ=1aQ(x,ω,E)ψ.

Consider the operator γm. Let ηC0(R3×S×R) such that η(.,.,E0)=0. Then we get by Equation(65) and Equation(63) (66) (ηψ)(.,.,Em)L2(R3×S)2=R3×S(ηψ)2(x,ω,Em)dωdx=R3×SE0EmE((ηψ)2(x,ω,E))dEdωdx=R3×S×I2(ηψ)(x,ω,E)(ηψ)E(x,ω,E)dEdωdx=2R3×S×I(ηψ)(x,ω,E)((ηψ)E(x,ω,E)+1aω·x(ηψ))(x,ω,E)dEdωdx+R3×S×Iω·x(1a)(ηψ)2(x,ω,E)dxdωdE=2R3×S×I(ηψ)(x,ω,E)1a(Q(x,ω,E)(ηψ))(x,ω,E)dEdωdx+R3×S×Iω·x(1a)(ηψ)2(x,ω,E)dxdωdE(ηψL2(R3×I,H1(S)2+1aQ(x,ω,E,D)(ηψ)L2(R3×I,H1(S))2)+ω·x(1a)L(R3×S×I)ηψL2(R3×S×I)2(66)

Since Q(x,ω,E)(ηψ)=ηQ(x,ω,E)ψ+(Q(x,ω,E)η)ψ and for qL(R3×I,W,1(S)) qψL2(R3×I,H1(S))qL(R3×I,W,1(S))ψL2(R3×I,H1(S)) and qUL2(R3×I,H1(S))qL(R3×I,W,1(S))UL2(R3×I,H1(S)) we conclude by Equation(66) (67) (ηψ)(.,.,Em)L2(R3×S)2C(ψL2(R3×I,H1(S))2+Q(x,ω,E,D)ψL2(R3×I,H1(S))2)(67) and so the assertion holds for γm (by choosing e.g. η=θ1θ2 where θ1C0(R3×S) and θ2C0() such that θ2(E0)=0,θ2(Em)=1). The assertion for γ0 is similarly proved which completes the proof. □

Let P* be the formal adjoint operator of P(x,ω,E,D) that is, P*(x,ω,E,D)v=avEaEv+b*(x,ω,E,ω)vω·xv where the formal adjoint of b(x, ω, E, ∂ω) is b*(x,ω,E,ω)v=c(x,E)ΔSv+d*(x,ω,E,ω)v.

We have the next generalized Green’s formula

Lemma 3.8.

Suppose that (44, 70) hold and that ψHP(R3×S×I°)L2(R3×I,H1(S)) and vĤ for which (supp(v))(R3×S×I) is a compact subset of (R3×S×I)=(R3×S×{Em})(R3×S×{E0}). Then (here .,. is the duality defined in Remark 3.6 above) (68) P(x,ω,E,D)ψ,vP*(x,ω,E,D)v,ψ=R3×S(a(·,Em)ψ(·,·,Em)v(·,·,Em)a(·,E0)ψ(·,·,E0)v(·,·,E0))dxdω.(68)

The proof follows by applying the standard Green’s formula and density arguments.

Remark 3.9.

The Green formula has some additional generalizations. Especially, Equation(68) holds for ψ=v in the case when ψHP(R3×S×I°)L2(R3×I,H1(S)) such that γm(ψ),γ0(ψ)L2(R3×S) (cf. Dautray and Lions Citation1999, 225).

Under the assumption Equation(66) the weak solution ψ of the Equationequation (62) obtained in Theorem 3.5 can be shown to be a solution of the initial value problem. We have

Theorem 3.10.

Suppose that the assumptions of Theorem 3.5 and Equation(64) are valid. Let fL2(R3×S×I). Then the initial value transport problem (69) a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ+ΣψKrψ=fψ(.,.,Em)=0(69) has a unique solution ψHHP(R3×S×I°). In addition, the solution ψ obeys the apriori estimate (70) ψHCfL2(R3×S×I).(70)

Proof.

The proof is analogous to the proof of Theorem 5.7 (Tervo et al. Citation2018a) (items (ii)-(iii) of the proof) and we omit the detailed treatments. □

The estimate Equation(70) implies that (71) ψHCTψL2(R3×S×I)(71) for all ψD:={ψHHP(R3×S×I°)|ψ(.,.,Em)=0} where, as above (72) Tψ=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ+ΣψKrψ.(72)

Actually, instead of Equation(71) more can be said

Theorem 3.11.

Suppose that the assumptions of Theorem 3.10 are valid. Then the apriori estimate (73) cψH2Tψ,ψL2(R3×S×I)for allψD(73) holds.

Proof.

The proof follows by the generalized Green’s formula using the same kind of estimates as utilized in the proof of the above Theorem 3.4. We omit the detailed proof. □

4. Regularity results of solutions emerging from explicit formulas of solutions

In some cases the transport equation can be solved explicitly. The obtained solution formulas will imply regularity of solutions. The results contain typical (anisotropic) regularity properties of solutions that can be said in the global case G=R3. Moreover, the below case studies expose relevant scales within which the regularity results can be formulated. Hence it is reasonable to consider the subsequent computational methods. To keep the computations limited we focus on the spatial regularity (x-regularity) but some outlooks for regularity with respect to all variables are exposed as well.

4.1. On regularity with respect to x-variable

4.1.1. Convection-scattering equation

We begin by considering a convection-scattering equation (74) Tψ:=ω·xψ+Σψ=f.(74) where Σ obeys (75) Σ=Σ(x,ω,E)c>0(75) and where fL2(R3×S×I). The solution ψ is obtained explicitly (Tervo and Kokkonen Citation2017, sections 4 and 5 or Dautray and Lions Citation1999, 244) and it is given by (76) ψ=0e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt.(76)

Moreover, the solution obeys an estimate (77) ψL2(R3×S×I)1cfL2(R3×S×I).(77)

We utilize repeatedly (without any mention) the following result from analysis.

Theorem 4.1.

Let GR3. be open and let (Y,μ) be a measure space. Suppose that f:G×YR is a measurable mapping such that the partial derivatives αfxα(x,y) exist for |α|r and that there exist gαL1(Y) such that |αfxα(x,y)|gα(y),xG,yY.

Then (78) αxα(Yf(x,y)dy)=Yαfxα(x,y)dy(78) for |α|r.

Proof.

For the proof we refer e.g. to Folland (Citation1999, 56). □

We begin with

Theorem 4.2.

Suppose that ΣW,(m1,0,0)(R3×S×I°) such that Equation(75) holds and that fH(m1,0,0)(R3×S×I°). Then the solution ψL2(R3×S×I) of the Equationequation (74) belongs to H(m1,0,0)(R3×S×I°).

Proof.

A. Constant Σ. To illustrate actual computations we, at first, consider the assertion in the case where Σ=Σ0 is constant. In this case the solution is (by Equation(76)) (79) ψ=0eΣ0tf(xtω,ω,E)dt(79) and by Equation(77) (80) ψL2(R3×S×I)1Σ0fL2(R3×S×I).(80)

Assume that fC0(R3×S×R). For a general fH(m1,0,0)(R3×S×I°) the claim is obtained by a limiting process as exposed below. For all .. (by Lemma 4.1) (81) (αψxα)(x,ω,E)=0eΣ0t(αfxα)(xtω,ω,E)dt.(81)

Furthermore, we find by the Cauchy-Schwartz inequality that (82) |(αψxα)(x,ω,E)|(0eΣ0tdt)1/2(0eΣ0t|(αfxα)(xtω,ω,E)|2dt)1/2(82) and then for |α|m1 (83) αψxαL2(R3×S×I)2=R3SI|(αψxα)(x,ω,E)|2dxdωdE1Σ0R3SI(0eΣ0t|(αfxα)(xtω,ω,E)|2dt)dxdωdE=1Σ00eΣ0t(R3SI|(αfxα)(xtω,ω,E)|2dxdωdE)dt=1Σ02αfxαL2(R3×S×I)2(83) where we noticed by using the change of variables x=xtω that R3|(αfxα)(xtω,ω,E)|2dx=R3|(αfxα)(x,ω,E)|2dx.

Due to Equation(83) (84) ψH(m1,0,0)(R3×S×I°)1Σ0fH(m1,0,0)(R3×S×I°)for allfC0(R3×S×R).(84)

Suppose, more generally that fH(m1,0,0)(R3×S×I°). Then there exists a sequence {fn}C0(R3×S×R) such that fnfH(m1,0,0)(R3×S×I°)0 for n (recall Theorem 2.1). Let (85) ψn(x,ω,E):=0eΣ0tfn(xtω,ω,E)dt.(85)

By the inequality Equation(86) ψnψmH(m1,0,0)(R3×S×I°)1Σ0fnfmH(m1,0,0)(R3×S×I°) and so {ψn} is a Cauchy sequence in H(m1,0,0)(R3×S×I°). Let ψH(m1,0,0)(R3×S×I°) such that ψnψH(m1,0,0)(R3×S×I°)0 for n. Since on the other hand, by Equation(80) ψnψL2(R3×S×I°)1Σ0fnfL2(R3×S×I°)0 we conclude that ψ=ψH(m1,0,0)(R3×S×I°), as desired.

B. Variable Σ. Secondly, we consider the claim more generally for ΣW,(1,0,0)(R3×S×I°).

B.1. At first, suppose that m1=1. Let fC0(R3×S×I°). By routine computations we get (by Lemma 4.1) (86) (ψ)xj(x,ω,E)=0(0tΣxj(xsω,ω,E)ds)e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt+0e0tΣ(xsω,ω,E)dsfxj(xtω,ω,E)dt,(86)

Furthermore, we find that by Equation(75) (87) e0tΣ(xsω,ω,E)dsect.(87)

Moreover, (88) |0tΣxj(xsω,ω,E)ds|ΣxjL(R3×S×I)t.(88)

Using same kind of techniques as in Part A (of the present proof), these estimates imply that there exists C > 0 such that (89) ψH(1,0,0)(R3×S×I°)CfH(1,0,0)(R3×S×I°)for allfC0(R3×S×I°).(89)

Hence by similar limiting arguments which we applied in Part A show that for fH(1,0,0)(R3×S×I°) the solution ψH(1,0,0)(R3×S×I°).

B.2. Let more generally m1N and fC0(R3×S×I°) The analogous computations as in Parts A and B.1. for higher derivatives show that (90) ψH(m1,0,0)(R3×S×I°)CfH(m1,0,0)(R3×S×I°)for allfC0(R3×S×I°).(90)

For example, (91) 2ψxkxj(x,ω,E)=0(0t2Σxkxj(xsω,ω,E)ds)e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt+0(0tΣxj(xsω,ω,E)ds)(0tΣxk(xsω,ω,E)ds)·e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt+0(0tΣxj(xsω,ω,E)ds)e0tΣ(xsω,ω,E)dsfxk(xtω,ω,E)dt+0(0tΣxk(xsω,ω,E)ds)e0tΣ(xsω,ω,E)dsfxj(xtω,ω,E)dt+0e0tΣ(xsω,ω,E)ds2fxkxj(xtω,ω,E)dt(91) and so by using similar kind of arguments as above we get Equation(90). We omit further details of this technically more complex part but notice that the Leibniz’s rule αψxα(x,ω,E)=βα(αβ)0αβxαβ(e0tΣ(xsω,ω,E)ds)βfxβ(xtω,ω,E)dt. is useful in computations. That is why, we are able to conclude (by utilizing limiting methods as above) that for fH(m1,0,0)(R3×S×I°) the solution ψH(m1,0,0)(R3×S×I°) which completes the proof. □

4.1.2. A Continuous slowing down convection-scattering equation

We deal with the following special case of a slowing down convection-scattering equation. Suppose that a=a(E) is independent of x and Σ=Σ(x,ω) is independent of E. We assume that a:IR is continuous and that (92) infEIa(E)=:κ>0,aEL(I).(92)

Σ is assumed to belong to L(R3×S) and (93) Σ(x,ω)c>0a.e.(x,ω)R3×S.(93)

In addition, for simplicity, we assume that E0=0. Consider the problem of the form (94) Tψ:=(aψ)E+ω·xψ+Σ(x,ω)ψ=f(x,ω,E),ψ(.,.,Em)=0.(94)

One can show that the solution of the problem (cf. Example 10.1 of Tervo et al. Citation2018) (95) Tψ=f,ψ(.,.,Em)=0(95) is (96) ψ(x,ω,E)=1a(E)(0R(Em)R(E)e0sΣ(xτω,ω)dτf˜(xsω,ω,R(E)+s)ds)(96) where R(E)=0E1a(τ)dτ and f˜(x,ω,η)=a(R1(η))f(x,ω,R1(η)).

The solution Equation(96) possesses the following x-regularity

Theorem 4.3.

Suppose that aW,m1(I°) and ΣW,(m1,0)(R3×S) such that Equation(92), Equation(93) hold. Furthermore, suppose that fH(m1,0,0)(R3×S×I°). Then the solution ψL2(R3×S×I) of the problem Equation(94) belongs to H(m1,0,0)(R3×S×I°).

Proof.

We omit the proof but the below computations in Example 4.4 for a more simple case will shed light to the assertion. □

To illustrate the claim of the above theorem we elaborate the following example.

Example 4.4.

Let a = 1 and let Σ=Σ0>0 be constant. Then by Equation(96) the solution of the transport problem (97) ψE+ω·xψ+Σ0ψ=f,ψ(.,.,Em)=0(97) is (98) ψ(x,ω,E)=0EmEeΣ0sf(xsω,ω,E+s)ds.(98)

Let m1N For |α|m1 (99) (xαψ)(x,ω,E)=0EmEeΣ0s(xαf)(xsω,ω,E+s)ds.(99)

Assuming that fH(m10,,0)(R3×S×I°) we see by applying similar type of computations as above in section 4.1.1 that by Equation(99) ψH(m1,0,0)(R3×S×I°).

4.2. Some outlines to regularity results with respect to all variables

We give some depictions for regularity with respect to all variables (x,ω,E) emerged from explicit formulas. Consider the equation (100) ω·xψ+Σψ=f(x,ω,E)(100) where (101) Σ=Σ(x,ω,E)c>0a.e..(101)

Recall that (102) ψ=0e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt.(102)

We have

Theorem 4.5.

Suppose that ΣW,(m1+m2,m2,m3)(R3×S×I°) such that Equation(101) holds and that fH(m1+m2,m2,m3)(R3×S×I°). Then the solution ψL2(R3×S×I) of the Equationequation (100) belongs to H(m1,m2,m3)(R3×S×I°).

Proof.

We omit the proof but the next Example 4.6 illustrates the assertion; especially the fact that (to guarantee the stated ω-regularity) we need the regularity up to order m1+m2 for f and Σ with respect to x-variable. □

Example 4.6.

In this example we compute some special cases.

A. Let m1=0,m2=1,m3=0. Assume that ΣW,(1,1,0)(R3×S×I°) and fC0(R3×S×I°). Let IS be the identity mapping of S (that is, IS(ω)=ω). We notice that IS is a C-mapping since S is a C-manifold. By the chain rule we find that (103) ωj(f(xtω,ω,E))=txf(xtω,ω,E),ISωj(ω)+fωj(xtω,ω,E)(103) and similarly for ωj(Σ(xsω,ω,E)). Moreover, ISωj|ω=Ω¯j(ω),j=1,2 where Ω¯j(ω) are tangent vectors on S at ω that is, Ω¯1(ω)=(ω2,ω1,0),Ω¯2(ω)=(ω1ω3ω12+ω22,ω2ω3ω12+ω22,ω12+ω22).

Hence by Equation(102) we have for j = 1, 2 (104) ψωj(x,ω,E)=0(0t((sΩ¯j(ω),x+ωj)Σ)(xsω,ω,E)ds)·e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt+0e0tΣ(xsω,ω,E)ds((tΩ¯j(ω),x+ωj)f)(xtω,ω,E)dt.(104)

Furthermore, (105) 0t|((sΩ¯j(ω),x+ωj)Σ)(xsω,ω,E)|0t(Ω¯j(ω)L(S)xΣL(R3×S×I)s+ΣωjL(R3×S×I))=:C1t2+C2t(105) and (106) |((tΩ¯j(ω),x+ωj)f)(xtω,ω,E)|Ω¯j(ω)L(S)t|xf(xtω,ω,E)|+|fωj(xtω,ω,E)|.(106)

Hence by Equation(104), Equation(87) (107) |ψωj(x,ω,E)|0ect(C1t2+C2t)|f(xtω,ω,E)|dt+0ect(tΩ¯j(ω)L(R3×S×I)|xf(xtω,ω,E)|+|fωj(xtω,ω,E)|)(107) which implies as in section 4.1.1 that (108) ψH(0,1,0)(R3×S×I°)CfH(1,1,0)(R3×S×I°)for allfC0(R3×S×I°).(108)

From the estimate Equation(108) it follows as above that for fH(1,1,0)(R3×S×I°) the solution ψH(0,1,0)(R3×S×I°).

B. Let m1=0,m2=0,m3=1. Suppose that ΣW,(0,0,1)(R3×S×I°) and fC0(R3×S×I°). Then we have (109) ψE(x,ω,E)=0(0tΣE(xsω,ω,E)ds)e0tΣ(xsω,ω,E)dsf(xtω,ω,E)dt+0e0tΣ(xsω,ω,E)dsfE(xtω,ω,E)dt(109) where |0tΣE(xtω,ω,E)ds|ΣEL(R3×S×I)t.

Thus again similar computations as in section 4.1.1 show that (110) ψH(0,0,1)(R3×S×I°)CfH(0,0,1)(R3×S×I°)for allfC0(R3×S×I°)(110) from which it follows that for fH(0,0,1)(R3×S×I°) the solution ψH(0,0,1)(R3×S×I°).

For a special case of continuous slowing down equation brought up in section 4.1.2 we have

Theorem 4.7.

Suppose that aW,m3(I°) and ΣW,(m1+m2,m2)(R3×S) such that Equation(92), Equation(93) hold. Furthermore, suppose that fH(m1+m2+m3,m2,m3)(R3×S×I°). Then the solution ψL2(R3×S×I) of the problem (111) (aψ)E+ω·xψ+Σ(x,ω)ψ=f(x,ω,E),ψ(.,.,Em)=0.(111) belongs to H(m1,m2,m3)(R3×S×I°).

Proof.

We omit the proof. The Example 4.8 below gives some insight for the claim. □

Example 4.8.

Consider the case of Example 4.4 for m1=m2=0,m3=2. Recall that ψ is given by Equation(98). Suppose that (112) fH(2,0,2)(R3×S×I°).(112)

Then we see that (113) (Eψ)(x,ω,E)=eΣ0(EmE)f(x(EmE)ω,ω,Em)+0EmEeΣ0s(Ef)(xsω,ω,E+s)ds(113) and furthermore (114) (E2ψ)(x,ω,E)=Σ0eΣ0(EmE)f(x(EmE)ω,ω,Em)eΣ0(EmE)(ω·xf)(x(EmE)ω,ω,Em)+eΣ0(EmE)(Ef)(x(EmE)ω,ω,Em)+0EmEeΣ0s(E2f)(xsω,ω,E+s)ds.(114)

Due to the Sobolev imbedding Theorem (e.g. Friedman Citation1976, 22) (115) uW,k(I°)CuHk+1(I°),uHk+1(I°),k=0,1(115) and so by Equation(114) EfL2(R3×S×I)CfH(1,0,2)(R3×S×I) which implies the claim for m=(0,0,2).

Remark 4.9.

The assumptions in Theorems 4.5 and 4.7 are not strict ones. In fact, the above computations show that in Theorem 4.5 it suffices only to assume that

fk=0m2H(m1+k,m2k,m3)(R3×S×I°),Σk=0m2W,(m1+k,m2k,m3)(R3×S×I°).

Similarly, in Theorem 4.7 it suffices only to assume that

fk=0m2j=0m3H(m1+k+j,m2k,m3j)(R3×S×I°)

Σk=0m2j=0m3W,(m1+k+j,m2k,m3j)(R3×S×I°).

The claim of Theorem 4.7 runs along general regularity results for evolution type equations. The above computations show that the use of explicit formulas in retrieving regularity results for transport equations is very limited.

5. On spatial regularity of solutions for the total transport equation by applying partial differences

In this section we consider the regularity of solutions of the complete transport equation (116) Tψ:=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ+ΣψKrψ=f(116) where b(x,ω,E,ω)ψ=c(x,E)ΔSψ+d(x,ω,E)·Sψ. We assume that the assumptions of Theorem 3.10 are valid. Then for any fL2(R3×S×I) the solution ψHHP(R3×S×I°) of the initial value problem (117) Tψ=f,ψ(.,.,Em)=0(117) exists. In addition, there exists a constant C0 such that (118) ψHCTψL2(R3×S×I).(118)

Recall that Pψ=P(x,ω,E,D)ψ:=a(x,E)ψE+b(x,ω,E,ω)ψ+ω·xψ

Our basic tools to prove regularity are the application of pertinent a priori estimates and partial differences. For simplicity, we restrict ourselves to the case Kr=Kr1 that is, Kr is of the form (119) (Krψ)(x,ω,E)=S×Iσ(x,ω,ω,E,E)ψ(x,ω,E)dωdE,(119) where σ:R3×S2×I2R is a non-negative measurable function such that Equation(27) holds.

In the following we, for simplicity, assume that a, c and d are independent of x. In more general cases it seems that the techniques based on the use of partial convolutions are more appropriate than the partial differences (see the Discussion section below). We recall the assumptions of Theorem 3.10 for this case: (120) ΣL(R3×S×I),Σ0a.e.inR3×S×I,(120) (121) aW,1(I),cL(I),djL(I,W,1(S))(121) (122) (aE(E)+(divSd)(ω,E))q1>0,a.e.,(122) (123) c(E)q2>0,a.e.,(123) (124) a(E0)<0,a(Em)<0,(124) (125) |a(E)|q4>0a,e,.(125) (126) S×Iσ(x,ω,ω,E,E)dωdEM1,S×Iσ(x,ω,ω,E,E)dωdEM2,(126) (127) Σ(x,ω,E)S×Iσ(x,ω,ω,E,E)dωdEc,(127) (128) Σ(x,ω,E)S×Iσ(x,ω,ω,E,E)dωdEc(128) where c is a strictly positive constant.

Key techniques based on difference approaches are contained in the proof of the next theorem.

Theorem 5.1.

Let m1N. Assume that a, c and d are independent of x and that the assumptions Equation(120), Equation(121), Equation(122), Equation(123), Equation(124), Equation(125), Equation(126), Equation(127), and Equation(128) are valid. Furthermore, suppose that (129) ΣL(S×I,W,1(R3),(129) and that (130) esssup(ω,E)S×ISIσ(.,ω,ω,E,E)W,m1(R3)dωdE=:Mm1<,esssup(ω,E)S×ISIσ(.,ω,ω,E,E)W,m1(R3)dωdE=:Mm1<.(130)

Let fH(m1,0,0)(R3×S×I°) and let ψHHP(R3×S×I°) be a solution of the problem (131) Tψ=f,ψ(.,.,Em)=0.(131)

Then ψH(m1,0,0)(R3×S×I°). Furthermore, there exists a constant C0 such that (132) ψH(m1,0,0)(R3×S×I°)CTψH(m1,0,0)(R3×S×I°).(132)

Proof.

A. At first, we assume that m1=1. Let ψHHP(R3×S×I°) and define (the difference quotient with respect to xj variable) (133) (δh,jψ)(x,ω,E):=ψ(x+hej,ω,E)ψ(x,ω,E)h.(133)

In the sequel we denote shortly δh=δh,j (for a fixed j{1,2,3}). We find that (134) (Σ(δhψ))(x,ω,E)=1h[Σ(x,ω,E)(ψ(x+hej,ω,E)ψ(x,ω,E))]=1h[Σ(x,ω,E)ψ(x+hej,ω,E)+Σ(x+hej,ω,E)ψ(x+hej,ω,E)Σ(x+hej,ω,E)ψ(x+hej,ω,E)Σ(x,ω,E)ψ(x,ω,E)]=δh(Σψ)(x,ω,E)(δhΣ)(x,ω,E)ψ(x+hej,ω,E)(134) and similarly (135) (Kr(δhψ))(x,ω,E)=S×Iσ(x,ω,ω,E,E)(δhψ)(x,ω,E)dωdE=δh(Krψ)(x,ω,E)S×I(δhσ)(x,ω,ω,E,E)ψ(x+hej,ω,E)dωdE.(135)

Denote ((δhKr)ψ)(x,ω,E):=S×I(δhσ)(x,ω,ω,E,E)ψ(x,ω,E)dωdE.

We see that S×I(δhσ)(x,ω,ω,E,E)ψ(x+hej,ω,E)dωdE=((δhKr)(τh,jψ))(x,ω,E) where (τh,jψ)(x,ω,E):=ψ(x+hej,ω,E) (the translation with respect to xj variable). We denote shortly τh,j=τh. Hence (136) (Kr(δhψ))(x,ω,E)=δh(Krψ)(x,ω,E)((δhKr)(τh,jψ))(x,ω,E).(136)

Since a, c and d are independent of x we see that δhψHP(R3×S×I) and (137) P(δhψ)=δh(Pψ).(137)

In fact, denote (δ¯hv)(x,ω,E):=v(xhej,ω,E)v(x,ω,E). Recalling that the formal adjoint of P=P(x,ω,E,D) is P*v=avEaEv+cΔSvd·SvdivS(d)vω·xv,vC0(R3×S×I°) we find that (138) δ¯h(P*v)=P*(δ¯hv).(138)

Furthermore, we have for all f,vL2(R3×S×I) (139) δhf,vL2(R3×S×I)=f,δ¯hvL2(R3×S×I)(139) and so we obtain by Equation(138) for vC0(R3×S×I°) (140) δh(Pψ),vL2(R3×S×I)=Pψ,δ¯hvL2(R3×S×I)=ψ,P*(δ¯hv)L2(R3×S×I)=ψ,δ¯h(P*v)L2(R3×S×I)=δhψ,P*vL2(R3×S×I)(140) that is, Equation(137) holds. Moreover, (141) δhψH(141) since if {ψn}C2(S,C1(G¯×I)) is a sequence such that ψnψ in H with n we see that δhψnδhψ in H and so Equation(141) holds (we omit details).

Applying Equation(137), Equation(134) and Equation(135) we get (140) T(δhψ)=P(δhψ)+Σ(δhψ)Kr(δhψ)=δh(Pψ)+δh(Σψ)(δhΣ)(τhψ)δh(Krψ)+(δhKr)(τhψ)=δh(Tψ)(δhΣ)(τhψ)+(δhKr)(τhψ).(140)

Since Tψ=f,δhψHHP(R3×S×I) (by Equation(137), Equation(141)) and δhψ(.,.,Em)=1h(ψ(.+hej,.,Em)ψ(.,.,Em))=0 we obtain in virtue of Equation(118), Equation(142) (143) δhψHCT(δhψ)L2(R3×S×I)=Cδhf(δhΣ)(τhψ)+(δhKr)(τhψ)CδhfL2(R3×S×I)+CδhΣL(R3×S×I)τhψL2(R3×S×I)+CδhKrτhψL2(R3×S×I)=CδhfL2(R3×S×I)+CδhΣL(R3×S×I)ψL2(R3×S×I)+CδhKrψL2(R3×S×I).(143)

Since by Equation(129) Σ(.,ω,E)W,1(R3) we obtain by the Morrey’s inequality (in Sobolev spaces) that a.e. (x,ω,E) |Σ(x+hej,ω,E)Σ(x,ω,E)|C1hxjΣL(R3×S×I) and so (144) δhΣL(R3×S×I)C1xjΣL(R3×S×I).(144)

In the similar way we see by Equation(136) that a.e. S×I|σ(x+hej,ω,ω,E,E)σ(x,ω,ω,E,E)|dωdEC2hS×I(xjσ)(.,ω,ω,E,E)L(R3) and S×I|σ(x+hej,ω,ω,E,E)σ(x,ω,ω,E,E)|dωdEC2hS×I(xjσ)(.,ω,ω,E,E)L(R3) and and so by Theorem 3.2 (145) δhKrC2M1M1(145) where M1 and M1 are as in Equation(130) (with m1=1).

B. For vC0(R3×S×I°) v(x+hej,ω,E)v(x,ω,E)h=01vxj(x+t(hej),ω,E)dt and so (146) δhvL2(R3×S×I)2=R3×S×I|01vxj(x+t(hej),ω,E)dt|2dxdωdE01R3×S×I|vxj(x+t(hej),ω,E)|2dxdωdEdt=R3×S×I|vxj(x,ω,E)|2dxdωdEvH(1,0,0)(R3×S×I°)2.(146)

Since C0(R3×S×I°) is dense in H(1,0,0)(R3×S×I°) we obtain by Equation(146) (147) δhfL2(R3×S×I)fH(1,0,0)(R3×S×I°).(147)

Let {hk} be a sequence such that hk0 with k. By virtue of Equation(143), Equation(144), Equation(145), Equation(147) the sequence {δhkψ} is bounded in H and hence in L2(R3×S×I). That is why there exists a subsequence {δhkiψ}iN which convergences weakly to an element UL2(R3×S×I) (e.g. Yosida Citation1980, 126) and so for any vC0(R3×S×I°) (148) δhkiψ,vL2(R3×S×I)U,vL2(R3×S×I).(148)

On the other hand, (149) δhkiψ,vL2(R3×S×I)=1hki(R3×S×Iψ(x+hkiej,ω,E)v(x,ω,E)dxdωdER3×S×Iψ(x,ω,E)v(x,ω,E)dxdωdE)=1hki(R3×S×Iψ(z,ω,E)v(zhkiej,ω,E)dzdωdER3×S×Iψ(z,ω,E)v(z,ω,E)dzdωdE)ψ,vxjL2(R3×S×I)=((ψ)xj)(v),i.(149)

That is why (ψ)xj=UL2(R3×S×I),j=1,2,3. As a conclusion we get that ψH(1,0,0)(R3×S×I°).

C. We show the a priori estimate Equation(132) for m1=1. Let for |α|=1 ((xαKr)ψ)(x,ω,E):=S×I(xασ)(x,ω,ω,E,E)ψ(x,ω,E)dωdE.

The assumption (130) implies that for any ψH(1,0,0)(R3×S×I°) the partial derivative xα(Krψ),|α|=1 exists (in the weak sense) and (150) (xα(Krψ)(x,ω,E)=S×Ixα(σ(x,ω,ω,E,E)ψ(x,ω,E))dωdE=S×I(xασ)(x,ω,ω,E,E)ψ(x,ω,E)dωdE+S×Iσ(x,ω,ω,E,E)(xαψ)(x,ω,E)dωdE=((xαKr)ψ)(x,ω,E)+(Kr(xαψ))(x,ω,E).(150)

Since a, c and d do not depend on x we have (similarly to Equation(137)) (151) P(xjψ)=xj(Pψ)=xj(fΣψ+Krψ)=xjf(xjΣ)ψΣ(xjψ)+(xjKr)ψ+Kr(xjψ)(151) and then (by the assumptions Equation(129), Equation(130) (with m1=1) P(xjψ)L2(R3×S×I) because ψH(1,0,0)(R3×S×I°) and fH(1,0,0)(R3×S×I°). Hence xjψHP(R3×S×I°).

In addition, (152) xjψH.(152)

This can be shown as follows. Due to Part B the sequence {δhkψ} therein is bounded in H. Hence by the Banach-Saks’s Theorem there exists a subsequence {δhkiψ} such that the running average sequence 1Ni=1Nδhkiψ convergences, say to an element U in H. On the other hand, by Equation(149) 1Ni=1Nδhkiψ convergences weakly to xjψ in L2(R3×S×I). Hence U=xjψ and so Equation(152) holds. Similarly, 1Ni=1N(δhkiψ)(.,.,Em)(xjψ)(.,.,Em) in L2(R3×S) and since 1Ni=1N(δhkiψ)(.,.,Em)=0 we have (xjψ)(.,.,Em)=0.

We find that (153) xj(Tψ)T(xjψ)=xj(Pψ)+xj(Σψ)xj(Krψ)P(xjψ)Σ(xjψ)+Kr(xjψ)=(xjΣ)ψ(xjKr)ψ.(153)

Hence, due to Equation(118) (154) 1CxjψH(T(xjψ)L2(R3×S×I)=xj(Tψ)(xjΣ)ψ+(xjKr)ψL2(R3×S×I)xj(Tψ)L2(R3×S×I)+xjΣL(R3×S×I)ψL2(R3×S×I)+xjKrψL2(R3×S×I)(154) and then again by Equation(118) (155) 1CxjψHxj(Tψ)L2(R3×S×I)+CxjΣL(R3×S×I)TψL2(R3×S×I)+CxjKrTψL2(R3×S×I)(155) which implies that (156) ψH(1,0,0)(R3×S×I°)CTψH(1,0,0)(R3×S×I°),(156) with C:=C\3maxj=1,2,3{1+CxjΣL(3×S×I)+CxjKr} as desired.

D. Next we verify that the claim holds for m1=2. Let Uj:=xjψ. By virtue of Part C, UjHP(R3×S×I°)H and by Equation(159) (recall that f=Tψ) (157) TUj=xjf(xjΣ)ψ+(xjKr)ψ=:FjL2(R3×S×I).(157)

In addition, Uj(.,.,Em)=0. Hence we are able to apply the results of Parts A-C of the present proof by replacing ψ with Uj which gives that UjH(1,0,0)(R3×S×I°) and by Equation(156) (158) UjH(1,0,0)(R3×S×I°)CTUjH(1,0,0)(R3×S×I°).(158)

Hence (159) xjψH(1,0,0)(R3×S×I°)2C2xjf+(xjΣ)ψ+(xjKr)ψH(1,0,0)(R3×S×I°)2=C2xjf+(xjΣ)ψ+(xjKr)ψL2(R3×S×I°)2+C2k=13xk(xjf+(xjΣ)ψ+(xjKr)ψ)L2(R3×S×I)2C2xjf+(xjΣ)ψ+(xjKr)ψL2(R3×S×I°)2+3C2k=13((xkxj)fL2(R3×S×I)2+xk((xjΣ)ψ)L2(R3×S×I)2+xk((xjKr)ψ)L2(R3×S×I)2).(159)

Furthermore, we have (160) xk((xjΣ)ψ)=(xkxjΣ)ψ+(xjΣ)xkψ(160) (161) xk((xjKr)ψ)=(xkxjKr)ψ+(xjKr)xkψ(161) where ((xkxjKr)ψ)(x,ω,E):=S×I(xkxjσ)(x,ω,ω,E,E)ψ(x,ω,E)dωdE.

In virtue of the assumption (130) (with m1=2) and Theorem 3.2 the operators (xkxjKr) are bounded operators L2(R3×S×I)L2(R3×S×I). Hence we finally get by Equation(159) (162) xjψH(1,0,0)(R3×S×I°)23C2(xjfL2(R3×S×I°)2+xjΣW,1(R3×S×I°)2ψL2(R3×S×I°)2+xjKr2ψL2(R3×S×I°)2)+3C2k=13(xkxjfL2(R3×S×I)2+xkxjΣL(R3×S×I)2ψL2(R3×S×I)2+xjΣL(R3×S×I)2xkψL2(R3×S×I)2+xkxjKr2ψL2(R3×S×I)2+xjKr2xkψ)L2(R3×S×I)2).(162)

Noting that f=Tψ we conclude (as above) from Equation(162) and Equation(118), Equation(156) that ψH(2,0,0)(R3×S×I°) and that there exists C0 such that (163) ψH(2,0,0)(R3×S×I°CTψH(2,0,0)(R3×S×I°)(163) which is the claim for m1=2.

E. For the general m1N0 the assertion is obtained by the induction. The induction step is similar to Part D and we neglect these technicalities. However, we notice that this step needs the generalization of derivation rules Equation(160), Equation(161) which is obtained by the Leibniz’s rules (164) xα(Σψ)=βα(αβ)(xαβΣ)xβψ=Σ(xαψ)+R1,αψ,|α|m1(164) where R1,αψ:=β<α(αβ)(xαβΣ)xβψ and (165) (xα(Krψ))(x,ω,E)=S×Ixα[σ(x,ω,ω,E,E)ψ(x,ω,E)]dωdE=S×Iβα(αβ)(xαβσ)(x,ω,ω,E,E)xβψ(x,ω,E)dωdE=:(Kr(xαψ))(x,ω,E)+(R2,αψ)(x,ω,E),|α|m1(165) where (R2,αψ)(x,ω,E):=SIβ<α(αβ)(xαβσ)(x,ω,ω,E,E)xβψ(x,ω,E)dωdE.

This finishes the proof. □

Remark 5.2.

A. Note that the assumptions (130) are equivalent to (166) σ=σ(x,ω,ω,E,E)L(S×I,L1(S×I,W,m1(R3)))L(S×I,L1(S×I,W,m1(R3))).(166)

B. Suppose that the assumptions of Theorem 5.1 are valid for every m1N0 and let fW(,0,0)(R3×S×I°). Then the solution ψ of the problem Equation(131) belongs to W(,0,0)(R3×S×I°)C(R3,L2(S×I)).

6. Discussion

We propose that the above regularity result of Theorem 5.1 can be generalized for x-dependent a,c and d and as in section 4 results with respect to all variables x,ω,E can be achieved, at least when Kr=Kr1. We conjecture the following result:

Let m=(m1,m2,m3)N03. Suppose that the assumptions of Theorem 3.10 are valid for Kr=Kr1. Furthermore, suppose that (167) Σ,a,c,dW,(m1+m2+m3,m2,m3)(R3×S×I°)),(167)

Finally, we suppose that (168) σW,(m1+m2+m3,m2,m3)(R3×S×I°,L1(S×I)),(168) (169) σW,(m1+m2+m3,m2,m3)(R3×S×I°,L1(S×I)).(169)

Let fH(m1+m2+m3,m2,m3)(R3×S×I°) and let ψHP(R3×S×I°) be a solution of the problem (170) Tψ=a(x,E)ψE+c(x,E)ΔSψ+d(x,ω,E)·Sψ+ω·xψ+ΣψKrψ=f(x,ω,E),ψ(.,.,Em)=0.(170)

Then ψH(m1,m2,m3)(R3×S×I°).

The assumptions Equation(167), Equation(168), Equation(169) can be weakened along the above Remark 4.9.

In addition to the difference techniques applied here we propose that the proofs can be based on the application of the Friedrich’s mollifier smoothing Jϵψ for ψL2(R3×S×R). The convolution (in mollifier) can be taken simultaneously with respect to all variables. We notice that the mollifier can also be defined on S (see e.g. Fukuoka Citation2006). For the mollifier it holds that JϵψH(,,)(3×S×). The relevant a priori estimates together with the (consequences of) well-known Friedrich’s Lemma enable us to deduce the same kind of conclusions as in the proof of the above Theorem 5.1. One additional method would be the application of the known regularity theory of evolution equations.

We finally notice that in the case where G=R3 the relevant problem is an initial inflow boundary value problem of the form (171) Tψ=f,ψ|Γ=g,ψ(.,.,Em)=0.(171)

The regularity of the solution ψ is limited, for example in x-variable, to the scale H(s,0,0)(G×S×I),s<3/2 (Tervo and Kokkonen Citation2017, Example 7.4). This is due to the fact that the initial inflow boundary value problem has the so called variable multiplicity (e.g. Nishitani and Takayama Citation1996; Nishitani and Takayama Citation2000), which causes irregularity on the inflow boundary.

Acknowledgments

The author thanks an anonymous referee for his/her improvements of the manuscript.

References

  • Agoshkov, V. 1998. Boundary value problems for transport equations. Berlin, Germany: Springer Science + Business Media.
  • Alonso, R., and W. Sun. 2014. The radiative transfer equation in the forward-peaked regime. arXiv:1411.0163v1 [math.AP].
  • Aubin, J.-P. 1979. Applied functional analysis. Hoboken, NJ: John Wiley and Sons.
  • Bouchut, F. 2002. Hypoelliptic regularity in kinetic equations. J. Math. Pures Appl. 81 (11):1135–59. doi:10.1016/S0021-7824(02)01264-3
  • Cessenat, M. 1984. Théorémes de trace Lp pour des espaces de fonctions de la neutronique. C.R. Acad. Sc. Paris, t. 299, Serie 1, number 16.
  • Chen, Y., and L. He. 2012. Smoothing estimates for Boltzmann equation with full-range interactions. Spatially inhomogeneous case. Arch. Rational Mech. Anal. 203 (2):343–77. doi:10.1007/s00205-011-0482-3
  • Chupin, L. 2010. Fokker-Planck equation in bounded domain. Ann. Inst. Fourier 60 (1):217–55. doi:10.5802/aif.2521
  • Dautray, R., and J.-L. Lions. 1999. Mathematical analysis and numerical methods for science and technology. Vol. 6. Evolution Problems II. Berlin, Germany: Springer.
  • Degond, P. 1986. Global existence of smooth solutions for the Vlasov-Fokker-Planck equation in 1 and 2 space dimensions. Ann. Sci. École Norm. Sup. 4e serie 19 (4):519–42. doi:10.24033/asens.1516
  • Egger, H., and M. Schlottbom. 2014. An Lp-theory for stationary radiative transfer. Appl. Anal. 93 (6):1283–96. doi:10.1080/00036811.2013.826798
  • Folland, G. B. 1999. Real analysis: Modern techniques and their applications. Hoboken, NJ: Wiley and Sons.
  • Frank, M., M. Herty, and A. N. Sandjo. 2010. Optimal radiotherapy treatment planning governed by kinetic equations. Math. Models Methods Appl. Sci. 20 (04):661–78. doi:10.1142/S0218202510004386
  • Friedman, A. 1976. Partial differential equations. Malabar, FL: Robert E. Krieger Publishing Co.
  • Friedrichs, K. O. 1944. The identity of weak and strong extensions of differential operators. Trans. Amer. Math. Soc. 55:132–51. doi:10.1090/S0002-9947-1944-0009701-0
  • Friedrichs, K. O. 1958. Symmetric positive linear differential equations. Comm. Pure Appl. Math. 11:333–418.
  • Fukuoka, R. 2006. Mollifier smoothing of tensor fields on differentiable manifolds and applications to Riemannian geometry. arXiv:math/06088230v1 [math.DG].
  • Grisvard, P. 1985. Elliptic problems in nonsmooth domains, vol. 24 of Monographs and Studies (Mathematics). London, UK: Pitman.
  • Halmos, P. R., and V. S. Sunder. 1978. Bounded Integral Operators on L2 Spaces. Berlin, Germany: Springer.
  • Herty, M., C. Jörres, and A. N. Sandjo. 2012. Optimization of a model Fokker-Planck equation. Kinetic Relat. Models 5 (3):485–503. doi:10.3934/krm.2012.5.485
  • Herty, M., and A. N. Sandjo. 2011. On optimal treatment planning in radiotherapy governed by transport equations. Math. Models Methods Appl. Sci. 21 (2):345–59. doi:10.1142/S0218202511005076
  • Hörmander, L. 1985. The analysis of linear partial differential operators III. Berlin, Germany: Springer.
  • Lax, P. D., and R. S. Phillips. 1960. Local boundary conditions for dissipative symmetric linear differential operators. Comm. Pure Appl. Math. 13 (3):427–55. doi:10.1002/cpa.3160130307
  • Le Bris, C., and P.-L. Lions. 2008. Existence and uniqueness of solutions to Fokker-Planck type equations with irregular coefficients. Comm. Partial Diff. Equations 33 (7):1272–317. doi:10.1080/03605300801970952
  • Mokhtar-Kharroubi, M. 1991. W1,p regularity in transport theory. Math. Models. Methods. Appl. Sci. 1 (4):477–99.
  • Morando, A., P. Secchi, and P. Trebeschi. 2009. Characteristic initial boundary value problems for symmetrizable systems. Rend. Sem. Mat. Univ. Torino 67:229–45.
  • Nishitani, T., and M. Takayama. 1996. A characteristic initial boundary value problem for a symmetric positive system. Hokkaido Math. J. 25 (1):167–82. doi:10.14492/hokmj/1351516716
  • Nishitani, T., and M. Takayama. 1998. Characteristic initial boundary value problems for symmetric hyperbolic systems. Osaka J. Math. 35:629–57.
  • Nishitani, T., and M. Takayama. 2000. Regularity of solutions to non-uniformly characteristic boundary value problems for symmetric systems. Commun. Part. Diff. Eq. 25 (5–6):987–1018.
  • Pazy, A. 1983. Semigroups of linear operators and applications to partial differential equations. Berlin, Germany: Springer.
  • Phillips, R. S., and L. Sarason. 1966. Singular symmetric positive first order differential operators. J. Math. Mech. 15 (2):235–71.
  • Rauch, J. 1985. Symmetric positive systems with boundary characteristic of constant multiplicity. Trans. Amer. Math. Soc. 291 (1):167–87. doi:10.1090/S0002-9947-1985-0797053-4
  • Rauch, J. 2012. Hyperbolic partial differential equations and geometric optics. Providence, RI: American Mathematical Society.
  • Rauch, J., and F. Massey. 1974. Differentiability of solutions to hyperbolic initial-boundary value problems. Trans. Amer. Math. Soc. 189:303–18. doi:10.2307/1996861
  • Tervo, J. 2019. On linear hypersingular Boltzmann transport equation and its variational formulation. arXiv:1808.09631v3 [math.AP].
  • Tervo, J., and M. Herty. 2020. On approximation of a hyper-singular transport operator and existence of solutions. Methods Appl. Anal. 27 (2):125–52. doi:10.4310/MAA.2020.v27.n2.a2
  • Tervo, J., and P. Kokkonen. 2017. On existence of L1-solutions for coupled Boltzmann transport equation and radiation therapy treatment planning. arXiv:1406.3228v2 [math.OC].
  • Tervo, J., P. Kokkonen, M. Frank, and M. Herty. 2018. On approximative linear Boltzmann transport equation for charged particle transport. Math. Models Methods Appl. Sci. 28 (14):2905–39. doi:10.1142/S0218202518500641
  • Tervo, J., P. Kokkonen, M. Frank, and M. Herty. 2018a. On existence of L2-solutions of coupled Boltzmann continuous slowing down transport equation system. arXiv:1603.05534v3 [math.OC].
  • Tervo, J., P. Kokkonen, M. Frank, and M. Herty. 2018b. On existence of solutions for Boltzmann continuous slowing down transport Equation. J. Math. Anal. Appl. 460 (1):271–301. doi:10.1016/j.jmaa.2017.11.052
  • Tian, R. 2014. Existence, uniqueness and regularity property of solutions to Fokker-Planck type equations. J. Math. Anal. Appl. 2 (1):53–63.
  • Treves, F. 1975. Basic linear partial differential equations. Cambridge, MA: Academic Press.
  • Ukai, S., and T. Yang. 2010. Mathematical theory of Boltzmann equation. Kowloon, Hong Kong: City University of Hong Kong, Reprint.
  • Yosida, K. 1980. Functional analysis. 6th ed. Berlin, Germany: Springer.
  • Zeghal, A. 2012. Sobolev regularity in neutron transport theory. ISRH Math. Anal. 2012:1–22. doi:10.5402/2012/379491