Loading [MathJax]/jax/output/SVG/jax.js

A geometric-analytic study of linear differential equations of order two

  • Received: 01 February 2020 Revised: 01 August 2020 Published: 19 October 2020
  • Primary:34A05, 34A25;Secondary:34A30, 34A26

  • We study second order linear differential equations with analytic coefficients. One important case is when the equation admits a so called regular singular point. In this case we address some untouched and some new aspects of Frobenius methods. For instance, we address the problem of finding formal solutions and studying their convergence. A characterization of regular singularities is given in terms of the space of solutions. An analytic-geometric classification of such linear polynomial homogeneous ODEs is obtained by the use of techniques from geometric theory of foliations means. This is done by associating to such an ODE a rational Riccati differential equation and therefore a global holonomy group. This group is a computable group of Moebius maps. These techniques apply to classical equations as Bessel and Legendre equations. We also address the problem of deciding which such polynomial equations admit a Liouvillian solution. A normal form for such a solution is then obtained. Our results are concrete and (computationally) constructive and are aimed to shed a new light in this important subject.

    Citation: Víctor León, Bruno Scárdua. A geometric-analytic study of linear differential equations of order two[J]. Electronic Research Archive, 2021, 29(2): 2101-2127. doi: 10.3934/era.2020107

    Related Papers:

    [1] Víctor León, Bruno Scárdua . A geometric-analytic study of linear differential equations of order two. Electronic Research Archive, 2021, 29(2): 2101-2127. doi: 10.3934/era.2020107
    [2] Huifu Xia, Yunfei Peng . Pointwise Jacobson type necessary conditions for optimal control problems governed by impulsive differential systems. Electronic Research Archive, 2024, 32(3): 2075-2098. doi: 10.3934/era.2024094
    [3] Shuang Wang, FanFan Chen, Chunlian Liu . The existence of periodic solutions for nonconservative superlinear second order ODEs: a rotation number and spiral analysis approach. Electronic Research Archive, 2025, 33(1): 50-67. doi: 10.3934/era.2025003
    [4] Li Yang, Chunlai Mu, Shouming Zhou, Xinyu Tu . The global conservative solutions for the generalized camassa-holm equation. Electronic Research Archive, 2019, 27(0): 37-67. doi: 10.3934/era.2019009
    [5] Eteri Gordadze, Alexander Meskhi, Maria Alessandra Ragusa . On some extrapolation in generalized grand Morrey spaces with applications to PDEs. Electronic Research Archive, 2024, 32(1): 551-564. doi: 10.3934/era.2024027
    [6] Noelia Bazarra, José R. Fernández, Ramón Quintanilla . Numerical analysis of a problem in micropolar thermoviscoelasticity. Electronic Research Archive, 2022, 30(2): 683-700. doi: 10.3934/era.2022036
    [7] Mufit San, Seyma Ramazan . A study for a higher order Riemann-Liouville fractional differential equation with weakly singularity. Electronic Research Archive, 2024, 32(5): 3092-3112. doi: 10.3934/era.2024141
    [8] Yi Wei . The Riccati-Bernoulli subsidiary ordinary differential equation method to the coupled Higgs field equation. Electronic Research Archive, 2023, 31(11): 6790-6802. doi: 10.3934/era.2023342
    [9] Huanhuan Li, Lei Kang, Meng Li, Xianbing Luo, Shuwen Xiang . Hamiltonian conserved Crank-Nicolson schemes for a semi-linear wave equation based on the exponential scalar auxiliary variables approach. Electronic Research Archive, 2024, 32(7): 4433-4453. doi: 10.3934/era.2024200
    [10] Yin Yang, Sujuan Kang, Vasiliy I. Vasil'ev . The Jacobi spectral collocation method for fractional integro-differential equations with non-smooth solutions. Electronic Research Archive, 2020, 28(3): 1161-1189. doi: 10.3934/era.2020064
  • We study second order linear differential equations with analytic coefficients. One important case is when the equation admits a so called regular singular point. In this case we address some untouched and some new aspects of Frobenius methods. For instance, we address the problem of finding formal solutions and studying their convergence. A characterization of regular singularities is given in terms of the space of solutions. An analytic-geometric classification of such linear polynomial homogeneous ODEs is obtained by the use of techniques from geometric theory of foliations means. This is done by associating to such an ODE a rational Riccati differential equation and therefore a global holonomy group. This group is a computable group of Moebius maps. These techniques apply to classical equations as Bessel and Legendre equations. We also address the problem of deciding which such polynomial equations admit a Liouvillian solution. A normal form for such a solution is then obtained. Our results are concrete and (computationally) constructive and are aimed to shed a new light in this important subject.



    Since the first appearance of Newton's laws of motion (see Axioms or Laws of Motion in [17] page 13), the study of ordinary differential equations has been associated with fundamental problems in physics and science in general. Many are the applications as universal gravitation and planetary dynamics, dynamics of particles under the action of a force field as the electromagnetic field, thermodynamics, meteorology and weather forecast, study of climate phenomena as typhoons and hurricanes, aerodynamics and hydrodynamics, atomic models, etc. Thanks to the nature of Newton's laws and other laws as Maxwell's equations or Faraday's and Kepler's laws [12], most of the pioneering work is first or second order ordinary differential equations (ODEs). Of special interest are the laws of the oscillatory movement (pendulum equation and Hill lunar movement equation [11]) and Hooke's law (spring extension or compression). A number of classical equations are, or have nice approximations by, linear equations with analytic coefficients. Among the linear equations the homogeneous case is a first step and quite meaningful. To be able to solve classical ordinary linear homogeneous differential equations is an important and active subject in mathematics. The arrival of features like scientific computing brings back the problem of finding solutions via power series. In this direction, a classical and powerful method is due to Frobenius. The main point is that Frobenius method works pretty well in a suitable class of second order linear ODEs, so called regular singular ODEs.

    In this paper we study second order linear differential equations with analytic coefficients under the viewpoint of finding solutions and studying their convergence. In very few words, we study forgotten as well as new aspects of Frobenius method. We start with the convergence of formal solutions. We also discuss the characterization of the so called regular singularities in terms of the space of solutions. An analytic-geometric classification of these polynomial ODEs is obtained via associating to such an ODE a Riccati differential equation and therefore a global holonomy group. This group is a computable group of Moebius maps. Next we apply these techniques and results to classical equations as Bessel and Legendre equations. Finally, we study the existence and form of Liouvillian solutions for polynomial ODEs.

    Next we give a more detailed description of our results.

    In Section 3 we discuss the problem of convergence of formal solutions for linear homogeneous ODEs of order two of the form

    a(x)y+b(x)y+c(x)y=0 (1)

    where a,b,c are analytic functions at x0R. We recall that there are examples of ODEs admitting a formal solution that is nowhere convergent (cf. Example 3.13). Our next result may be seen as a version of a theorem due to Malgrange (see [14, 15]) and also to Mattei-Moussu (see [16]) for holomorphic integrable systems of differential forms. By a formal solution centered at x0R of an ODE we shall mean a formal power series ˆy(x)=n=0an(xx0)n with complex coefficients anC. We prove:

    Theorem A. Consider a second order ordinary differential equation given by (1). Suppose also that there exist two linearly independent formal solutions ˆy1(x) and ˆy2(x) centered at x0 of equation (1). Then x0 is an ordinary point or a regular singular point of (1). Moreover, ˆy1(x) and ˆy2(x) are convergent.

    A formal solution associated to a regular singularity is always convergent:

    Theorem B. Consider a second order ordinary differential equation given by (1). Suppose (1) has at x0 an ordinary point or a regular singular point. Given a formal solution ˆy(x) of (1) then this solution is convergent. Indeed, this solution converges in the same disc type neighborhood where the coefficients a(x),b(x),c(x) are analytic.

    We shall say that a function u(x) for x in a disc |x|<R centered at the origin 0R,C is an analytic combination of log and power (anclop for short) if it can be written as u(x)=α(x)+log(x)β(x)+γ(x)xr for some analytic functions α(x),β(x),γ(x) defined in the disc |x|<R and rR or rC. In the real case we assume that x>0 in case we have β0 or γ0 and a power xr with rRQ.

    Definition 1.1 A one-variable complex function u(z) considered in a domain UC will be called analytic up to log type singularities (autlos for short) if:

    1. u(z) is holomorphic in Uσ where σU is a discrete set of points, called singularities.

    2. Given a singularity pσ either p is a removable singularity of u(z) or there is a germ of real analytic curve γçolon[0,ϵ)U such that γ(0)=p and u(z) is holomorphic in Dγ[0,ϵ) for some disc DU centered at p.

    A one variable real function u(x) defined in an interval JR will be called analytic up to log type singularities (autlos for short) if, after complexification, the corresponding function uC(z), which is defined in some neighborhood J×{0}UC is analytic up to log type singularities, as defined above.

    With such notions we obtain the following characterization of regular singularities:

    Theorem C (characterization of regular points). Consider a second order ordinary differential equation given by (1). Then the following conditions are equivalent:

    (i) The equation admits two linearly independent solutions y1(x),y2(x) which are anclop (analytic combinations of log and power). Then x0 is an ordinary point or a regular singular point for the ODE.

    (ii) The equation admits two solutions y1(x),y2(x) which are autlos (analytic up to logarithmic singularities).

    (iii) The equation has an ordinary point or a regular singular point at x0.

    We start with a polynomial second order linear equation of the form

    a(z)u+b(z)u+c(z)u=0 (2)

    with a,b,c are complex polynomials of a variable z. By introducing the change of coordinates t=u/u we obtain a first order Riccati equation which writes as

    dtdz=a(z)t2+b(z)t+c(z)a(z).

    Definition 1.2. The Riccati differential equation above is called Riccati model of the ODE (2).

    By its turn, since the work of Paul Painlevé (see [18]), a polynomial Riccati equation is studied from the point of view of its transversality with respect to the vertical fibers z=constant, even at the points at the infinity u=. With the advent of the theory of foliations, due to Ehresmann, the notion of holonomy was introduced as well as the notion of global holonomy of a foliation transverse to the fibers of a fibration. This is the case of a polynomial Riccati foliation once placed in the ruled surface P1(C)×P1(C), where P1(C)=Cçup{} is the Riemann sphere. This allows us to introduce the notion of global holonomy of a second order linear equation as above. Then we proceed the study of the equation from this group theoretical point of view, since the global holonomy will be a group of Moebius maps of the form tαt+βγt+δ. We do calculate this group in some special cases and reach some interesting consequences for the original ODE.

    Theorem D. Consider a second order ODE given by u+b(z)u+c(z)u=0 where b,c are complex polynomials of a variable z. Then the equation above admits a general solution of the form

    u,k(z)=kexp(z0D(ξ)B(ξ)A(ξ)C(ξ)dξ),k,C

    where A,B,C,D are entire functions satisfying ADBC0.

    We apply these techniques for studying the classical Bessel and Legendre equations (cf. Examples 5.10 and 5.13).

    One important class of solutions for ODEs is the class of Liouvillian solutions, following the work of Liouville, Rosenlicht and Ross among other authors. The question whether a polynomial first order ODE admits a Liouvillian solution or first integral has been addressed by M. Singer in [24] and others. We refer to [24] for the notion of Liouvillian function in n complex variables. Such a function has (holomorphic) analytic branches in some Zariski dense open subset of Cn. In particular we can ask whether an ODE admits such a solution.

    Question 1.3. What are the polynomial ODEs of the form (2) admitting a Liouvillian solution u(z)? What are the possible Liouvillian solutions?

    Our contribution to the above problem is:

    Theorem E. Consider a complex ODE of the form L(u):=a(z)u+b(z)u+c(z)u=0 where the coefficients a(z),b(z),c(z) are complex polynomials. Then we have the following:

    (i) If L[u]=0 admits a solution satisfying a Liouvillian relation then it has a Liouvillian first integral (cf. Corollary pages 674, 675 [24]).

    (ii) If L[u]=0 admits a Liouvillian solution then it has a Liouvillian first integral (cf. Corollary page 674, 675 [24]).

    (iii) If L[u]=0 admits a Liouvillian first integral then its solutions are Liouvillian and given by one of the forms below:

    (a) u(z)=exp(zγ(η)dη)[kzexp(η2γ(ξ)b(ξ)a(ξ)dξ)dη+] for constants k,C and γ(z) a rational solution for the Riccati equation.

    (b) u(z)=k1+k2zexp(ηb(ξ)a(ξ)dξ)dη, for constants k1,k2C.

    In this section we recall briefly the classical Frobenius method. We consider equations that write in the form a(x)y+b(x)y+c(x)y=0 for some real analytic functions a(x),b(x),c(x) at some point x0R. We say that x0 is an ordinary point if a(x0)0. Nevertheless, most of the relevant equations are connected to the singular (non-ordinary) case. We can mention the Bessel equation x2y+xy+(x2ν2)y=0, whose range of applications goes from heat conduction, to the model of the hydrogen atom (see [1]). This equation has the origin x=0 as a singular point. Another remarkable equation is the Laguerre equation xy+(ν+1x)y+λy=0 where λ,νR are parameters. This equation is quite relevant in quantum mechanics, since it appears in the modern quantum mechanical description of the hydrogen atom. According to Frobenius a singular point x=x0 of the ODE a(x)y+b(x)y+c(x)y=0 is regular if the following limits limxx0(xx0)b(x)a(x) and limxx0(xx0)2c(x)a(x) exist and are finite.

    The Frobenius method consists in associating to the original ODE an Euler equation, i.e., an equation of the form A(xx0)2y+B(xx0)y+Cy=0 and looking for solutions (to this equation) of the form y0(x)=(xx0)r. This gives an algebraic equation of degree two Ar(r1)+Br+C=0, so called indicial equation, whose zeroes r give solutions y0(x)=(xx0)r of the Euler equation. The Euler equation associated to the original ODE with a regular singular point at x=x0 is given by (xx0)2y+p0(xx0)y+q0y=0 where p0=limxx0(xx0)b(x)a(x) and q0=limxx0(xx0)2c(x)a(x).

    In this case we have the following classical theorem of Frobenius:

    Theorem 2.1 (Frobenius theorem, Theorem 5.6.1 in [2] pages 293, 294, [8], Theorem 3 and 4 in [6] pages 158, 175). Assume that the ODE

    (xx0)2y+(xx0)b(x)y+c(x)y=0

    has a regular singularity at x=x0, where the functions b(x),c(x) are analytic with convergent power series in |xx0|<R. Then there is at least one solution of the form

    y(x)=|xx0|rn=0dn(xx0)n

    where r is a root of the indicial equation, d0=1 where the series converges for |xx0|<R.

    Frobenius method actually consists in looking for solutions of the form

    y1(x)=|xx0|rn=0dn(xx0)n (3)

    where r is the zero of the indicial equation having greater real part. Whether there is a second linearly independent solution is related to the roots of the indicial equation. Indeed, there is some zoology and in general the second solution is of the form

    y2(x)=|xx0|˜rn=0˜dn(xx0)n (4)

    in case there is a second root ˜r of the indicial equation and this root is such that r˜rZ. If ˜r=r then there is a solution of the form

    y2(x)=y1(x)log|xx0|+|xx0|r+1n=0ˆdn(xx0)n. (5)

    Finally, if 0r˜rN then we have a second solution of the form

    y2(x)=ky1(x)log|xx0|+|xx0|˜rn=0çheckdn(xx0)n. (6)

    Each of the series in equations (3), (4), (5) and (6) converges for |xx0|<R and defines a function that is analytic in some neighborhood of x=x0.

    As referred above, Frobenius theorem statement is found in the book of Boyce-DiPrima and in the book of E. Coddington. Nevertheless, we are afraid that a more detailed and complete proof of the convergence of the formal part of the solutions in the disc |xx0|<R (ie., in the common disc where the coefficients of the ODE are analytic) can only be found in Coddington's book.

    Consider a second order ordinary differential equation given by a(z)u+b(z)u+c(z)u=0 where a,b,c are holomorphic in a neighborhood of the origin 0C. We shall mainly address two questions:

    Question 3.1

    (i) Under what conditions can we assure that the origin is an ordinary point or a regular singular point of the equation?

    (ii) Is it that a formal solution of the ODE is always convergent?

    Let us fix some notations that will be used from now on:

    On=C{z1,,zn} is the ring of holomorphic functions at 0Cn.

    ˆOn=C{{z1,,zn}} is the ring of formal series in n indeterminates over C.

    Mn is the field of meromorphic functions.

    ˆMn is the field of fractions of ˆOn.

    Recall that the field of fractions of an integral domain D is the smallest field in which the field embeds. It consists of fractions having elements of D as numerator and nonzero elements of D as denominator in the standard way. Thus ˆMn is the field consisting of fractions of the form ˆfˆg where ˆf,ˆgˆOn and ˆg0.

    Let us give a first proof of the convergence in Theorem A:

    Proof of the convergence in Theorem A. In order to simplify our notation we shall assume x0=0. We consider equation

    a(z)u+b(z)u+c(z)u=0 (7)

    where a,b,c are complex analytic (holomorphic) functions in neighborhood |z|<R. According to [19] there is an integrable complex analytic one-form Ω in C3 defined as follows

    Ω=a(z)ydx+a(z)xdy+[a(z)y2+b(z)xy+c(z)x2]dz.

    Indeed,

    dΩ=[2xc(z)+yb(z)+ya(z)]dxdz+[2ya(z)+xb(z)xa(z)]dydz+2a(z)dxdy
    ΩdΩ=[2y2a(z)2xya(z)b(z)+xya(z)a(z)]dxdydz+[2x2a(z)c(z)xya(z)b(z)xya(z)a(z)]dxdydz+[2y2a(z)2+2xya(z)b(z)+2x2a(z)c(x)]dxdydz=0.

    This one-form is tangent to the vector field in C3 associated to the reduction of order of the ODE. Indeed, as a first step we perform the classical order reduction process where equation (7) is rewritten after the following 'change of coordinates': x=u,y=u,z=z. We then obtain

    x=u=y,y=u=b(z)a(z)uc(z)a(z)u=b(z)a(z)yc(z)a(z)x,z=1.

    Therefore, a natural vector field X associated to equation (7) is given by

    X(x,y,z)=yx(b(z)a(z)y+c(z)a(z)x)y+z.

    Note that

    Ω(X)=y2a(z)xa(z)(b(z)a(z)y+c(z)a(z)x)+[y2a(z)+xyb(z)+x2c(z)]=0.

    Moreover, given two linearly independent solutions u1(z) and u2(z) of the ODE the function

    H(x,y,z)=xu1(z)yu1(z)xu2(z)yu2(z)

    is a first integral for the form Ω, ie., dHΩ=0. Indeed,

    dH=1(xu2yu2)2[y(u1u2u1u2)dxx(u1u2u1u2)dy[y2(u1u2u1u2)+xy(u1u2u1u2)+x2(u1u2u1u2)]dz]

    note that

    u1u2u1u2=ba(u1u2u1u2) and u1u2u1u2=ca(u1u2u1u2)

    then

    dH=1(xu2yu2)2[y(u1u2u1u2)dxx(u1u2u1u2)dy[y2(u1u2u1u2)+xyba(u1u2u1u2)+x2ca(u1u2u1u2)]dz]=u1u2u1u2a(xu2yu2)2(aydx+axdy+[ay2+bxy+cx2]dz)=u1u2u1u2a(xu2yu2)2Ω.

    By hypothesis there exist two linearly independent formal solutions ˆu1 and ˆu2 of equation (7). Each solution writes as a formal complex power series

    ˆuj(z)=n=0ajnznC{{z}}.

    According to the above, there exists a formal first integral

    H(x,y,z)=xˆu1(z)yˆu1(z)xˆu2(z)yˆu2(z)

    of the integrable one-form Ω above. Now note that H,1/HˆO3. Indeed, if HˆO3 then there exist ai,j,kC such that

    xˆu1(z)yˆu1(z)xˆu2(z)yˆu2(z)=i+j+k=0ai,j,kxiyjzk

    then

    xˆu1(z)yˆu1(z)formal series containingonly one term inxandy=(xˆu2(z)yˆu2(z))formal series containingonly one term inxandy(i+j+k=0ai,j,kxiyjzk) only depends onz

    so there is fˆO1 such that

    xˆu1(z)yˆu1(z)=(xˆu2(z)yˆu2(z))f(z)

    equivalently

    x[ˆu1(z)f(z)ˆu2(z)]=y[ˆu1(z)f(z)ˆu2(z)]

    so we have

    ˆu1(z)=f(z)ˆu2(z) and ˆu1(z)=f(z)ˆu2(z)

    therefore ˆu1 and ˆu2 are linearly dependent which contradicts the hypothesis. Similarly 1/HˆO3.

    Now we recall the following convergence theorem:

    Theorem 3.2 (Cerveau-Mattei, [5], Theorem 1.1 page 106). Let Ω be a germ at 0Cn of an integrable holomorphic 1-form and H=fgˆMn a purely formal meromorphic first integral of Ω, i.e. ΩdH=0 and H,1/HˆOn. Then H converges, i.e., HMn.

    From the above theorem H=fg where f,gO3 are relatively prime. Hence we have

    [xˆu1(z)yˆu1(z)]g(x,y,z)=[xˆu2(z)yˆu2(z)]f(x,y,z)

    as f and g are relatively prime then f divides xˆu1(z)yˆu1(z) e g divides xˆu2(z)yˆu2(z) from there exists α,β,ξ,ηO1 and kˆO1 such that

    xˆu1(z)yˆu1(z)=k(z)[xα(z)yβ(z)]andxˆu2(z)yˆu2(z)=k(z)[xξ(z)yη(z)]

    equivalently we have

    x[ˆu1(z)k(z)α(z)]=y[ˆu1(z)k(z)β(z)]andx[ˆu2(z)k(z)ξ(z)]=y[ˆu2(z)k(z)η(z)]

    so we have

    ˆu1(z)=k(z)α(z),ˆu1(z)=k(z)β(z),ˆu2(z)=k(z)ξ(z)andˆu2(z)=k(z)η(z)

    therefore

    ˆu1(z)ˆu1(z)=α(z)β(z)andˆu2(z)ˆu2(z)=ξ(z)η(z)

    thus

    ˆu1(z)=Aexp(zα(w)β(w)dw)andˆu2(z)=Bexp(zξ(w)η(w)dw),

    for some A,B constants, are convergent.

    We stress the fact that we are not assuming the ODE to be regular at x0.

    Consider the linear homogeneous second order ODE

    a(x)y+b(x)y+c(x)y=0 (8)

    where a(x),b(x),c(x) are differentiable real or complex functions defined in some open subset UR,C. We may assume that U is an open disc centered at the origin 0R,C. We make no hypothesis on the nature of the point x=0 as a singular or ordinary point of (8). Given two solutions y1 and y2 of (8) their wronskian is defined by W(y1,y2)(x)=y1(x)y2(x)y2(x)y1(x).

    Claim 3.3 The wronskian W(y1,y2) satisfies the following first order ODE

    a(x)w+b(x)w=0. (9)

    This is a well-known fact and we shall not present a proof, which can be done by straightforward computation. Most important, the above fact allows us to introduce the notion of wronskian of a general second order linear homogeneous ODE as (8) as follows:

    Definition 3.4 The wronskian of (8) is defined as the general solution of (9).

    Hence, in general the wronskian is of the form

    W(x)=Kexp(xb(η)a(η)dη) (10)

    where K is a constant.

    A well-known consequence of the above formula is the following:

    Lemma 3.5 Given solutions y1(x),y2(x) the following conditions are equivalent:

    (i) W(y1,y2)(x) is identically zero.

    (ii) W(y1,y2)(x) vanishes at some point x=x0.

    (iii) y1(x),y2(x) are linearly dependent.

    Let us analyze the consequences of this form. We shall consider the origin as the center of our disc domain. In what follows the coefficients are analytic in a neighborhood of the origin.

    Case (1). If ba has poles of order r>1 at the origin: In this case we can write

    b(x)a(x)=Arxr++A2x2+A1x+d(x)

    where A1,A2,,Ar are constant, Ar0 and d is analytic at the origin. Hence

    W(x)=K|x|A1exp(Ar(r1)xr1++A2x)exp(˜d(x))

    where ˜d is analytic. Now observe that exp(Ar(r1)xr1++A2x) is neither analytic nor formal. Therefore, in this case, W is neither analytic nor formal.

    Case (2): ba has poles of order 1 at the origin. In this case b(x)a(x)=A1x+d(x) and W(x)=K|x|A1exp(˜d(x)). If W is analytic or formal then we must have A1{0,1,2,3,}.

    Summarizing we have:

    Lemma 3.6 Assume that the wronskian W of the ODE a(x)y+b(x)y+c(x)y=0, with analytic coefficients, is analytic or formal. Then ba has a pole of order r1 at the origin. Moreover, we must have W(x)=K|x|Aexp(f(x)), where A{0,1,2,3,} and f is analytic.

    Now we are able to prove the remaining part of Theorem A:

    End of the proof of Theorem A. We have already proved the first part. Let us now prove that the origin is an ordinary point or a regular singularity of the ODE. This is done by means of the two following claims:

    Claim 3.7 The quotient ba has poles of order 1 at the origin.

    Proof. Indeed, since by hypothesis there are two formal linearly independent functions, the wronskian is formal. Thus, from the above discussion we conclude.

    The last part is done below. For simplicity we shall assume that x=zC and that the coefficients are complex analytic (holomorphic) functions.

    Claim 3.8 We have limz0z2c(z)a(z)C.

    Proof. Write a(z)u+b(z)u+c(z)u=0 and a(z)=zk according to the local form of holomorphic functions. Since limz0zb(z)a(z)C we must have b(z)a(z)=˜b(z)z for some holomorphic function ˜b(z) at 0. Assume that the Claim is not true, then c(z)a(z) must have a pole of order 3 at 0. Thus we may write c(z)a(z)=˜c(z)z3+ν for some holomorphic function ˜c(z) at 0 and some νN. We may choose ν such that ˜c(0)0. We have for the ODE above z3+νu+z2+ν˜b(z)u+˜c(z)u=0. For sake of simplicity we will assume that ˜c(0)=1 and ν=0. This does not affect the argumentation below. We write ˜b(z)=b0+b1z+b2z2+ and ˜c(z)=1+c1z+c2z2+ in power series. Substituting this in the ODE we obtain z3u+z2(b0+b1z+b2z2+)u+(1+c1z+c2z2+)u=0. Now we write u(z)=n=0anzn in power series. Thus we have

    n=3[(n1)(n2)an1+nk=2(k1)ak1bnk+nk=0akcnk]zn+[a2c0+a1(b0+c1)+a1b0]z2+(a0c1+a1c0)z+a0c0=0

    then a0=a1=a2=0 and for n3 we have

    (n1)(n2)an1+nk=2(k1)ak1bnk+nk=0akcnk=0.

    Hence an=0 for all n0, ie., u=0 is the only possible formal solution. This proves the claim by contradiction.

    The two claims above end the proof of Theorem A.

    Next we present a proof that also implies Theorem A.

    Proof of Theorem B. First of all we are assuming that the origin is an ordinary point or a regular singularity of the ODE. If it is an ordinary point, then by the classical existence theorem for ODEs there are two linearly independent analytic solutions and any solution, formal or convergent, will be a linear combination of these two solutions. Such a solution is therefore convergent.

    Thus we may write the ODE as x2y+xb(x)y+c(x)y=0 where the new coefficients b(x) and c(x), obtained after renaming xb(x)/a(x) and x2c(x)/a(x) conveniently, are analytic.

    Let us consider a formal solution ˆy(x)=n=0dnxn. We can write ˆy(x)=xr1(1+φ(x)) for some r10 and φ(x) a formal function with φ(0)=0. In other words, r1{0,1,2,} is the order of ˆy(x) at the origin. Then we have ˆy(x)=r1xr11(1+φ(x))+xr1φ(x) and ˆy(x)=r1(r11)xr12(1+φ(x))+2r1xr11φ(x)+xr1φ(x). Substituting this in the ODE x2ˆy(x)+xb(x)ˆy(x)+c(x)ˆy(x)=0 and dividing by xr1 we obtain

    x2φ(x)+x(2r1+b(x))φ(x)+(r1(r11)+r1b(x)+c(x))(1+φ(x))=0.

    For x=0, since φ(0)=0, we then obtain the equation r1(r11)+r1b(0)+c(0)=0. The above is exactly the indicial equation associated to the original ODE. We then conclude that the original ODE has an indicial equation with a root r1 that belongs to the set of non-negative integers. Let now rZ be the other root of the indicial equation. There are two possibilities:

    (i) rr1. In this case, then according to Frobenius classical theorem we conclude that there is at least one solution yr(x)=xrn=0enxn which is convergent. There are two possibilities:

    (i.1) yr(x) and ˆy(x) are linearly dependent: in this case, yr(x)=çdotˆy(x) for some constant R,C. Then r=r1 and therefore yr(x) is analytic and the same holds for ˆy(x). More precisely, ˆy(x) is analytic in the same neighborhood |x|<R where b(x),c(x) are convergent.

    (i.2) yr(x) and ˆy(x) are linearly independent: Since yr(x) is analytic and seeing yr(x) as a formal solution, we have two linearly independent formal solutions. From what we have seen above in Theorem A both solutions are convergent in the common disc domain of analyticity of the functions b(x),c(x).

    (ii) r1r. In this case, then according to Frobenius classical theorem we conclude that there is at least one solution ˜yr1(x)=xr1n=0fnxn, where the power series is convergent. There are two possibilities:

    (ii.1) ˜yr1(x) and ˆy(x) are linearly dependent: in this case, ˜yr1(x)=˜çdotˆy(x) for some constant ˜R,C. Then r=r1 and therefore ˜yr1(x) is analytic and the same holds for ˆy(x). More precisely, ˆy(x) is analytic in the same neighborhood |x|<R where b(x),c(x) are convergent.

    (ii.2) ˜yr1(x) and ˆy(x) are linearly independent: in this case, ˜yr1(x) is analytic and seeing ˜yr1(x) as a formal solution, we have two linearly independent formal solutions. From what we have seen above in Theorem A both solutions are convergent in the common disc domain of analyticity of the functions b(x),c(x).

    The above proof still makes use of the convergence part in Theorem A, thus it cannot be used to give an alternative proof of Theorem A. Let us work on a totally independent proof of Theorem A based only on classical methods of Frobenius and ODEs. For this sake we shall need a few lemmas.

    We consider the ODE x2y+xb(x)y+c(x)y=0 with a regular singular point at the origin.

    Lemma 3.9. Let ˆy(x) be a formal solution of the ODE. Then we must have ˆy(x)=xr(1+n=1anxn) where r is a root of the indicial equation of the ODE.

    Remark 3.10 Let r{0,1,2,} be a root of the indicial equation and assume that we have two solutions ˆy1(x)=xr(1+φ1(x)) and ˆy2(x)=xr(1+φ2(x)) which are formal. Then we have two cases: (i) r1. In this case W(ˆy1,ˆy2)(0)=0. In this situation we must have W(ˆy1,ˆy2)(x)=0 and therefore ˆy1,ˆy2 are linearly dependent. (ii) r=0.

    Let us proceed. We are assuming now that we have two formal solutions ˆy1,ˆy2 for the ODE above. We write ˆyj(x)=xrj(1+φj(x)) for some formal series φj(x) that satisfies φj(0)=0. The exponents rj are non-negative integers and from what we have seen above, these are roots of the indicial equation r(r1)+rb(0)+c(0)=0 of the ODE. We may assume that r1r2.

    So we have the following possibilities:

    (i) r1=r2. If this is the case we cannot a priori assure that the indicial equation has only the root r=r1=r2. Anyway, if r0 then from what we have seen above the formal solutions ˆy1,ˆy2 are linearly dependent. This is a contradiction. Thus we must have r=0. If r=0 is the only root of the indicial equation then we have a basis of the solution space given by y1(x)=1+n=1enxn and y2(x)=y1(x)log|x|+n=1fnxn. If a linear combination ˆy(x)=c1y1(x)+c2y2(x) is a formal function then necessarily c2=0. Thus any two formal solutions are linearly dependent. Assume now that r=0 is not the only root of the indicial equation. Denote by ˜rZ the other root of the indicial equation. There are two possibilities:

    (a) ˜r>0 then there is a basis of solutions given by y1(x)=x˜r(1+n=1gnxn) and y2(x)=ay1(x)log|x|+|x|0(1+n=1hnxn). Let y(x)=c1y1(x)+c2y2(x) be a formal power series. Then y(x)=(c1+ac2log|x|)x˜r(1+n=1gnxn)+c2(1+n=1hnxn). If y(x) is a formal power series then we must have ac2=0 and therefore y(x)=c1x˜r(1+n=1gnxn)+c2(1+n=1hnxn). In particular, since ˜rN, y(x) is convergent. This shows that the formal solutions ˆy1(x),ˆy2(x) are convergent and this is the only possible case where they can be linearly independent.

    (b) ˜r<0 then there is a basis of solutions given by y1(x)=1+n=1pnxn and y2(x)=ay1(x)log|x|+x˜r(1+n=1qnxn). Write y(x)=c1y1(x)+c2y2(x) for a linear combination of y1(x) and y2(x). Then y(x)=(c1+ac2log|x|)(1+n=1pnxn)+c2(x˜r(1+n=1qnxn)). If y(x) is a formal series then necessarily ac2=0 (because of the term log|x|) and also c2=0 in this case because ˜r<0. Thus we get y(x)=c1y1(x) which is convergent. This shows that again we must have that ˆy1 and ˆy2 are multiple of y1 and therefore they are linearly dependent, contradiction again.

    (ii) 0<r1r2=NN. This case follows from facts already used above. Since r1>r2 and since each rj is a root of the indicial equation, we conclude that these are the roots of the indicial equation. By Frobenius theorem there is a basis of the solutions given by y1(x)=xr1(1+n=1snxn) and y2(x)=ay1(x)log|x|+|x|r2(1+n=1tnxn). If y(x)=c1y1(x)+c2y2(x) is a formal power series then we must have ac2=0 and y(x)=c1xr1(1+n=1snxn)+c2xr2(1+n=1tnxn) which is convergent. This shows that ˆy1,ˆy2 must be convergent.

    We are now in conditions of giving a second proof to Theorem A.

    Alternative proof of Theorem A. Indeed, from the second part of the proof (which is based only on classical methods of Frobenius and ODEs) we know that the origin is an ordinary point or a regular singular point of the ODE. Given the two linearly independent formal solutions ˆyj(x),j=1,2, from the above discussion, the solutions ˆy1(x),ˆy2(x) are analytic.

    The next couple of examples show that the information on the wronskian (whether it is convergent, formal, etc) is not enough to infer about the nature of the solutions.

    Example 3.11 (convergent wronskian but no formal solution)This is an example of an ODE with a convergent wronskian but admitting no formal solution. The ODE x3yx2yy=0 has a non-regular singular point at the origin. From what we have observed above the wronskian W of two linearly independent solutions of the ODE satisfies the following first order ODE x3wx2w=0 whose solution is of the form W(x)=Kexp(xη2η3dη)=Kx for some constant K. It is now easy to check that there are no formal solutions besides the trivial.

    We now give an example of an ODE with non-convergent wronskian and admitting no formal solution but the trivial one.

    Example 3.12 (non-convergent wronskian no formal solution) The ODE x3yxyy=0 has a non-regular singular point at the origin. Indeed, the wronskian is solution of the first order ODE x3wxw=0 which has solutions of the form W(x)=Kexp(xηη3dη)=Kexp(1x) where K is a constant. The ODE only admits the trivial formal solution.

    Next we give an example of an equation admitting a formal but not a convergent solution.

    Example 3.13 Consider the equation x2yy12y=0. The origin x0=0 is a singular point, but not is regular singular point, since the coefficient -1 of y does not have the form xb(x), where b is analytic for 0. Nevertheless, we can formally solve this equation by power series k=0akxk, where the coefficients ak satisfy the following recurrence formula (k+1)ak+1=[k2k12]ak, for every k=0,1,2,. If a00, applying the quotient test to this expression we have that

    |ak+1xk+1akxk|=|k2k12k+1|çdot|x|,

    when k, provided that |x|0. Hence, the series converges only for x=0, and therefore does not represent a function in a neighborhood of x=0.

    We shall now prove Theorem C.

    Proof of Theorem C. We shall first consider the complex analytic case. We start then with a complex analytic ODE of the form a(z)u+b(z)u+c(z)u=0. Let us assume that this equation admits two linearly independent solutions u1(z),u2(z), which are of autlos type in some neighborhood of the origin z=0C.

    The wronskian W(u1,u2)(z) satisfies the first order ODE a(z)w+b(z)w=0 and since it is given by W(u1,u2)(z)=u1(z)u2(z)u1(z)u2(z), it is also of autlos type in some neighborhood of the origin z=0C. Using then the above first order ODE and arguments similar to those in the proof of Lemma 3.6 we conclude that b(z)/a(z) must have a pole of order 1 at the origin, otherwise W(u1,u2)(z) would have an essential singularity at the origin. Following now a similar reasoning in the proof of Claim 3.8 in the second part of the proof of Theorem A we conclude that c(z)/a(z) must have a pole of order 2 at the origin. This shows that the singularity at the origin is regular, or the origin is an ordinary point. If we start with a real analytic ODE then we consider its complexification. The fact that there are two linearly independent solutions of autlos type for the original ODE implies that there are two linearly independent solutions for the corresponding complex ODE, by definition these solutions will be of autlos type. Once we have concluded that the complex ODE has a regular singularity or an ordinary point at the origin, the same holds for the original real analytic ODE. Thus (ii)(iii). The classical Frobenius theorem shows that (iii)(i). Finally, it is clear from the definitions that (i)(ii).

    The next examples show how sharp is the statement of Theorem C.

    Example 4.1. Consider the equation

    z3uzu+u=0. (11)

    The origin z0=0 is a singular point, but not is regular singular point. It is easy to see that u1=z is a solution of equation (11). Making use of the method of reduction of order we can construct a second solution u2 linearly independent with u1. Hence we have that

    u2(z)=zz(exp(wvv3dv)w2)dw=zexp(1z).

    Note that u2 is not holomorphic.

    Remark 4.2 Consider a second order differential equation of the form

    z3a(z)u+z2b(z)u+c(z)u=0 (12)

    where a,b,c are holomorphic at the origin with a(0)0 and c(0)0. We shall see that (12) admits no formal solution. Indeed we assume that u(z)=n=0dnzn is a formal solution of (12) then we have dn=0 for all n0. Observe that there exists the limit limx0x3b(x)x3a(x)=b0a0 and the limit below does not exist limx0x2c(x)x3a(x).

    Remark 4.3. Consider a second order differential equation of the form

    z2a(z)u+b(z)u+c(z)u=0 (13)

    where a,b,c are holomorphic at the origin with a(0)0, b(0)0 and c(0)0. We shall see that (13) always admits non trivial formal solution. Indeed we assume that u(z)=n=0dnzn is a formal solution of (13) then d1=c0d0b0, d2=(b1c0+c20+c1b0)d0b20 and for n2 we have

    dn+1=1b0(n+1)(cnd0nk=1[k(k1)ank+kbnk+1+cnk]dk).

    Observe that the coefficients of the series depend on d0, since we look for non trivial formal solutions it suffices to choose d00. Hence, there exist non trivial formal solution. Also note that there exists the limit limx0x2c(x)x2a(x)=c0a0 and the following limit is not finite limx0xb(x)x2a(x).

    Example 4.4. Consider a second order differential equation given by

    z2u+bu+cu=0 (14)

    where b and c are nonzero constants. Observe that the origin is a non regular singular point of (14). Next we shall see that there exist non trivial formal solutions for (14). Let us assume that

    u(z)=n=0anzn (15)

    is a non trivial formal solution of (14). We have a1=ca0b, a2=c2a02b2 and

    an+1=(n2n+c)anb(n+1), for all n=2,3,. (16)

    Observe that the coefficients of the series depend on a0, since we look for non trivial formal solutions it suffices to choose a00. Hence, there exist non trivial formal solution. Observe now that this formal solution is not convergent. Applying the ratio test to the expressions (15) and (16), we have that

    |an+1zn+1anzn|=|n2n+cb(n+1)|çdot|z|,

    when n, whenever |z|0. Hence, the series converges only for z=0.

    We shall now exhibit method of associating to a homogeneous linear second order ODE a Riccati differential equation. Consider a second order ODE given by

    a(z)u+b(z)u+c(z)u=0

    where a,b,c are analytic functions, real or complex, of a variable z real or complex, defined in a domain UR,C. According to [19] there is an integrable one-form

    Ω=a(z)ydx+a(z)xdy+[a(z)y2+b(z)xy+c(z)x2]dz

    that vanishes at the vector field corresponding to the reduction of order of the ODE, i.e., ω(X)=0 where

    X(x,y,z)=yx(b(z)a(z)y+c(z)a(z)x)y+z.

    As a consequence the orbits of X are tangent to the foliation FΩ given by the Pfaff equation Ω=0.

    First of all we remark that we can write Ω as follows

    Ωx2=a(z)d(yx)+[a(z)(yx)2+b(z)(yx)+c(z)]dz.

    Thus, by introducing the variable t=yx we see that the same foliation FΩ can be defined by the one-form ω below:

    ω=a(z)dt+[a(z)t2+b(z)t+c(z)]dz.

    By its turn ω=0 defines a Riccati foliation which writes as

    dtdz=a(z)t2+b(z)t+c(z)a(z).

    Definition 5.1. The Riccati differential equation above is called Riccati model of the ODE a(z)u+b(z)u+c(z)u=0.

    Remark 5.2. The Riccati model can be obtained in a less geometrically clear way by setting t=u/u as a new variable. Sometimes it is also useful to consider the change of variable w=u/u which leads to the Riccati equation dwdz=c(z)w2+b(z)w+a(z)a(z).

    Let η=(E,π,B,F) be a (locally trivial) fibration with total space E, fiber F, base B and projection π:EB. A foliation F on E is transverse to the fibre bundle η if: (a) for each pE, the leaf Lp of F with pLp is transverse to the fiber π1(q), q=π(p); (b) dim(F)+dim(F)=dim(E); (c) for each leaf L of F, the restriction π|L:LB is a covering map. According to a theorem of Ehresmann (see [4, Ch. V], [9, Ch. Ⅱ]) if the fiber F is compact, then conditions (a) and (b) together already imply (c).

    Let us state the exact notion of Riccati foliation we consider. We shall consider a complex manifold M admitting a locally trivial holomorphic fibration πçolonMB, onto a complex manifold B, with fiber F. A singular holomorphic foliation F on M will be called a Riccati foliation on M if (i) codimF=dimF and (ii) there is a ramification set σB, analytic of codimension 1, such that π1(σ) is invariant (a union of F-invariant fibers), and the restriction of F to Mπ1(σ) is a foliation transverse to the fibre bundle π|Mπ1(σ)çolonMπ1(σ)Bσ in the classical sense of Ehresmann (see for instance [4, Ch. V]). In particular we have sing(F)π1(σ). Most relevant is the fact, well-known for foliations transverse to fibre bundles, that (iii) in case the fundamental group of Bσ is finitely generated, F|Mπ1(σ) is conjugate to the suspension of a subgroup of holomorphic diffeomorphisms GDiff(F), given by a (so called holonomy or monodromy) homomorphism φçolonπ1(Bσ)Diff(F).

    In view of (iii) our notion of Riccati foliation is quite general. Indeed, in our framework, a Riccati foliation on a fibered space πçolonMB, with fiber F, may be seen as an extension, holomorphic with singularities, of a foliation given by a suspension of a group of complex diffeomorphisms GDiff(F), obtained as a representation π1(Bσ)Diff(F), where σ is a codimension 1 analytic subset of B. This idea is reinforced by the following. When M=P1(C)×P1(C), a holomorphic foliation with singularities, is of Riccati type dydx=a(x)y2+b(x)y+c(x)p(x), if and only if, it is transverse to a generic vertical line P1(C)x0={x0}×P1(C)P1(C)×P1(C). This fact, widely used by Paul Painlevé in his memoire (see [18]), has been extended in a natural way in [21]. Indeed, in [21] the notion of Riccati foliation adopted is pretty much the same we use here, by considering suitable (natural) fibrations.

    It is well-known that a complex rational Riccati differential equation dydx=a(x)y2+b(x)y+c(x)p(x) induces in the complex surface P1(C)×P1(C) a foliation F with singularities, having the following characteristics:

    (i) The foliation has a finite number of invariant vertical lines {x0}×P1(C). These lines are given by the zeroes of p(x) and possibly by the line {}×P1(C).

    (ii) For each non-invariant vertical line {x0}×P1(C) the foliation has its leaves transverse to this line.

    (iii) From Ehresmann we conclude that the restriction of F to (P1(C)σ)×P1(C), where σ×P1(C) is the set of invariant vertical lines, is a foliation transverse to the fibers of the fiber space (P1(C)σ)×P1(C)P1(C)σ with fiber P1(C) and projection given by π(x,y)=x.

    (iv) The restriction π|L of the projection to each leaf L of the Riccati foliation defines a covering map LP1(C)σ.

    In particular, there is a global holonomy map which is defined as follows: choose any point x0σ as base point and consider the lifting of the closed paths γπ1(P1(C)σ) to each leaf LF by the restriction π|L above. Denote the lift of γ starting at the point (x0,z){x0}×P1(C) by ˜γz. If the end point of ˜γz is denoted by (x0,hγ(z)) then the map zhγ(z) depends only on the homotopy class of γπ1(P1(C)σ). Moreover, this defines a complex analytic diffeomorphism h[γ]Diff(P1(C)) and the map π1(P1(C)σ)Diff(P1(C)),[γ]h[γ] is a group homomorphism (see Proposition 1.16 in [13] page 24). The image Hol(F)Diff(P1(C)) is called global holonomy of the Riccati equation. It is well-known from the theory of foliations transverse to fiber spaces that the global holonomy classifies the foliation up to fibered conjugacy (see Theorem 3 in [4] page 99). This will be useful to us in what follows. Recall that Diff(P1(C)) as meant above is the projectivization of the special linear group ie., Diff(P1(C))çongPSL(2,C) (see [10] page 64) meaning that every global holonomy map can be represented by a Moebius map T(z)=a1z+a2a3z+a4 where a1,a2,a3,a4C and a1a4a2a3=1. Thus the global holonomy group of a Riccati foliation identifies with a group of Moebius maps. Another important fact which comes from the general theory of suspension foliations is:

    (v) Given a non-vertical leaf L of a Riccati foliation F in P1(C)×P1(C), the leaf L is conformally equivalent, via the projection π|LçolonLB=P1(C)σ, to the holomorphic covering of B associate to the subgroup Fix(L)Hol(F) stabilizer of L in the global holonomy Hol(F).

    (vi) Fix a point p{x0}×P1(C), x0σ, and consider as transverse section the germ of disc induced by the vertical fiber {x0}×P1(C). The holonomy group of the non-vertical leaf L through this point p is conjugate to the subgroup of germs at p of elements of the global holonomy group Hol(F) that fix the point p.

    In particular, if the group Fix(L) is trivial then L is conformally equivalent to the basis B=P1(C)σ.

    Definition 5.3 (holonomy of a second order ODE) Given a linear homogeneous second order ODE with complex polynomial coefficients

    a(z)u+b(z)u+c(z)u=0

    we call the holonomy of the ODE the global holonomy group of the corresponding Riccati model.

    Remark 5.4. As we have seen above we can also obtain a Riccati model by any of the changes of variables t=u/u or w=u/u. From the viewpoint of ODEs these models may seem distinct. Nevertheless, they differ only up to the change of coordinates t=1/w. Moreover, both have the same global holonomy group, since the point at infinity is always considered in the definition of global holonomy group. Indeed, the ideal space for considering a Riccati equation from the geometrical viewpoint, is the space C×C.

    Let us investigate some interesting cases. First consider a Riccati foliation F assuming that σ is a single point. Thus we may assume that in affine coordinates (x,y) the ramification point is the point x=. Then we may write F as given by a polynomial differential equation dydx=a(x)y2+b(x)y+c(x). The global holonomy of F is given by an homomorphism ϕçolonπ1(P1(C)σ)Diff(P1(C)). Since σ is a single point we have P1(C)σ=C is simply-connected and therefore the global holonomy is trivial. By the classification of foliations transverse to fibrations (see [4, Ch. V]) there is a fibered biholomorphic map ΦçolonC×P1(C)C×P1(C) that takes the foliation F into the foliation H given by the horizontal fibers C×{y},yP1(C).

    Lemma 5.5. A holomorphic diffeomorphism ΦçolonC×P1(C)C×P1(C) preserving the vertical fibration writes in affine coordinates (x,y)C2C×P1(C) as Φ(x,y)=(Ax+B,a(x)y+b(x)c(x)y+d(x)) where a,b,c,d are entire functions satisfying adbc0, 0A,BC.

    Proof of Lemma 5.5. The fact that the vertical fibration is preserved means that in coordinates (x,y)C×CP1(C) the map Φ is of the form Φ(x,y)=(f(x),g(x,y)). Similarly, the inverse map Φ1 is also of the form Φ1(x,y)=(˜f(x),˜g(x,y)). The holomorphic map fçolonCC is then an entire automorphism (diffeomorphism of C). It is well-known that the automorphisms of C is the group of affine maps, so that f(x)=Ax+B for some A,BC with A0. Again, since the vertical fibration x=constant is preserved, for each fixed xC the map Φxçolon{x}×P1(C){Ax+B}×P1(C),(x,y)Φ(x,y)=(Ax+B,g(x,y)) is a diffeomorphism. In particular the map gx:yg(x,y) is a holomophic diffeomorphism of P1(C). It is well-known that the group of automorphisms (holomorphic diffeomorphisms) of the Riemann sphere is the group of Moebius maps of the form yay+bcy+d with a,b,c,dC,adbc=1. Hence, locally we must have g(x,y)=aj(x)y+bj(x)cj(x)y+dj(x) for some holomorphic functions aj(x),bj(x),cj(x),dj(x) satisfying satisfying aj(x)dj(x)bj(x)cj(x)0,xDj, defined in some open subset UjC such that the union jJUj is a cover of the complex plane C. These functions aj(x),bj(x),cj(x),dj(x) are not unique, but they are unique up to multiplication by a nonzero complex number, i.e., they correspond to a unique element of the projectivization PGL(2,C). Because C is a simply-connected domain, by standard analytic continuation we conclude that the map g induces a holomorphic map CGL(2,C),Cxgx(y)=g(x,y), ie, there are entire functions a(x),b(x),c(x),d(x) satisfying adbc0 such that g(x,y)=a(x)y+b(x)c(x)y+d(x).

    The map Φ(x,y)=(Ax+B,a(x)y+b(x)c(x)y+d(x)) takes leaves of F into leaves of the horizontal fibration H:y=constant. Since the second projection map σ2çolonC×P1(C)P1(C),(x,y)y is a first integral for H, it follows that the composition g=σ2çircΦçolonC×P1(C)P1(C) is a meromorphic first integral for F. Another point is that, as already mentioned in (iv) above, given any leaf L of F, the restriction π|L of the projection to each leaf L of the Riccati foliation defines a covering map LP1(C)σ. In our current situation we have σ={(,0)}, so that we have a covering map LC. In particular we conclude that the leaves of F are diffeomorphic with C (including the one contained in the invariant fiber {(0,)}×P1(C)).

    Let us now apply this to our framework of second order linear ODEs.

    Proof of Theorem D. Beginning with the ODE a(z)u+b(z)u+c(z)u=0 the Riccati model is

    dtdz=a(z)t2+b(z)t+c(z)a(z).

    Thus if we assume that a(z)=1 then we have for this Riccati equation that σ={} as considered in § 5.2 above. This implies that F admits a meromorphic first integral gçolonC×P1(C)P1(C) of the above form g(z,t)=A(z)t+B(z)C(z)t+D(z) for some entire functions A(t),B(t),C(t),D(t). Given a leaf L of the Riccati foliation there is a constant P1(C) such that g(z,t)= for all (t,z)L. Hence t=D(z)B(z)A(z)C(z) for all (t,z)L. This defines a meromorphic parametrization zt(z) of the leaf. Since we have t=yx=uu therefore u(z)=kexp(z0t(ξ)dξ) is a solution of the ODE with kC a constant. This gives

    u,k(z)=kexp(z0D(ξ)B(ξ)A(ξ)C(ξ)dξ),k,C;

    as general solution of the original ODE. Notice that u,k(z)u,k(z)=D(z)B(z)A(z)C(z) so that if 12 then the corresponding solutions u1,k1 and u2,k2 generate a nonzero wronskian, and therefore they are linearly independent solutions for all k10k2.

    Illustrating Theorem D we have:

    Example 5.6. Consider the equation given by

    uzuu=0. (17)

    From what we have observed above we known that for a(z)=1, b(z)=z and c(z)=1 there exists a Riccati equation given by

    dtdz=1+ztt2.

    It is not difficult to see that t=z is a solution of the differential Riccati model. Hence

    t=exp(z22)+zexp(η22)dη+z,

    where is constant, is a solution of the differential Riccati model. In the construction of the Riccati equation associated to (17) it is considered that t=uu where u is a solution of (17). Hence

    u(z)=kexp(z22)(+zexp(η22)dη)

    where k is constant. It is a straightforward computation to show that u is a solution of (17). Rewriting u we have

    u(z)=kexp(z[η+exp(η22)+ηexp(ξ22)dξ]dη)=kexp(zη[exp(η22)ηηexp(ξ22)dξ]ηexp(ξ22)dξ(1)dη).

    Note that A(η)=ηexp(ξ22)dξ, B(η)=exp(η22)ηηexp(ξ22)dξ, C(η)=1 and D(η)=η are entire functions satisfying

    A(η)D(η)B(η)C(η)=exp(η22)0.

    Next we investigate the case where the ODE generates a Ricatti foliation R having a ramification set σ that consists of two points. In this case the holonomy group of the ODE is cyclic generated by a single Moebius map. Let us make a general study of this case. We start by assuming that σ={0,} i.e., that the invariant vertical lines are {0}×P1(C) and {}×P1(C) in P1(C)×P1(C). The basis Bσ is the Riemann sphere P1(C) minus two points, the origin and the point at infinity. This corresponds to C minus the origin, which is kwnon as the cylinder C. Since the fundamental group of the cylinder C is isomorphic to Z, the global holonomy group of R is cyclic generated by a single Moebius map. About the conformal and topological type of the leaves of R we have:

    Lemma 5.7. Each non-vertical leaf of R is either conformally equivalent to C or to the cylinder C.

    Proof. By a vertical leaf of R we mean a leaf contained in a vertical line z=constant. Since R is a foliation transverse to a fibration in C×P1(C), a non-vertical leaf L of R is, by the second coordinate projection, a holomorphic covering of the basis C. This leaf is then either conformally equivalent to C or to the cylinder C. This is a simple consequence of the classical Riemann-Koebe uniformization theorem (see [7]).

    Indeed, we can state:

    Proposition 5.8. The non-vertical leaves of R are diffeomorphic to the cylinder C.

    Proof. As we already observed a non-vertical leaf L of R is conformally equivalent to C or to C. If L has trivial holonomy group then (as we have seen in the final remarks preceding Definition 5.3) L is diffeomorphic to the basis B=C. Assume now that the holonomy of L is not trivial. We claim that still in this case L is diffeomorphic to C. Indeed, otherwise L mus be diffeomorphic to C, in which case L is simply-connected and then its holonomy group must be trivial, a contradiction.

    Next we give a concrete example of a second order ODE having a cyclic global holonomy group. This is an example that fits into Frobenius approach, since the origin is a regular singularity.

    Example 5.9. Consider the equation given by

    z2u+u=0.

    From what we observed above we have that for a(z)=z2, b(z)=0 and c(z)=1 there exists a Riccati equation given by

    dtdz=z2+t2z2.

    This is a Riccati equation with a single vertical invariant line in C×P1(C), the projective line {0}×P1(C). The line at infinity {}×P1(C) is also invariant, as it follows from a simple change of coordinates ˜z=1z.

    Before finishing this example, we give a word about the computation of the global holonomy. Since it is a homogeneous equation, the solutions of the Riccati differential equation are of the form

    t=K(1+i3)z1i3(1i3)z2Kzi32

    where K is constant. It is enough to compute a simple loop holonomy map with base point at t=0 for instance, z(θ)=z0eiθ with 0θ2π. After the corresponding computations we get a holonomy map as

    h(t0)=t(z(2π))=z0(t0(1+i3)2z0)e2π3t0(1i3)+2z0(2t0(1i3)z0)e2π32t0+(1+i3)z0.

    A fairly classical regular singularity type example is given below.

    Example 5.10 (Bessel equation) Consider the complex Bessel equation given by

    z2u+zu+(z2ν2)u=0

    where z,νC. Since a(z)=z2, b(z)=z and c(z)=z2ν2 the corresponding Riccati model is

    dtdz=z2t2+zt+z2ν2z2

    A change of coordinates to w=1z shows that the ramification set is σ={0,}. We are then in the cyclic global holonomy case. At this moment we shall not compute the Moebius map that generates the global holonomy. We are more interested in the geometry of the leaves which is given by Proposition 5.8.

    A non-regular singularity example is given below. This example is specially interesting in view of our Proposition 5.8. Indeed, it cannot be studied by the classical Frobenius approach since the origin is not a regular singularity.

    Example 5.11. Let us consider the following polynomial ODE

    znu+b(z)u+c(z)u=0

    where b,c are complex polynomials of a variable z. If n2 and b(0)0 or if n3 and c(0)0 or b(0).b(0)0 then z=0 is a non-regular singular point. Let us assume that this is the case. The corresponding Riccati equation is

    dtdz=znt2+b(z)t+c(z)zn.

    Changing coordinates w=1/z we obtain

    dtdw=t2+wnb(1/w)t+wnc(1/w)w2=wkt2+˜b(w)t+˜c(w)w2+k

    for some polynomials ˜b(w),˜c(w) and some kN. This shows that the ramification set σP1(C) consists of the points z=0 and z=. The basis is then the Riemman sphere minus two points. This corresponds to the complex plane C minus one point, ie., to the cylinder C=C{0}. The fundamental group of the basis is then isomorphic to Z, being cyclic generated by a simple non-trivial loop homotopy class. The holonomy of the ODE is then generated by a single Moebius map.

    Now we consider the case where σ consists of three points. We may assume that these are the points z=±1 and z=, i.e., σ={1,1,}. The considered Riccati foliation R on P1(C)×P1(C) has three invariant vertical lines: the lines {1}×P1(C),{1}×P1(C) and the line {}×P1(C) where stands for the point at infinity in the horizontal coordinate. The basis B=P1(C)σ corresponds to the Riemann sphere minus three points. The fundamental group π1(B) is then free with two generators. The global holonomy group is therefore generated by two Moebius maps which we shall not compute. Instead we shall work with the conformal type of the leaves as follows:

    Proposition 5.12. The non-vertical leaves of R are parabolic (in the sense of potential theory) Riemman surfaces of the form D/G where D is the unit disc and GSL(2,R) is a properly discontinous group.

    Proof. Again this is a consequence of the uniformization theorem of Riemman-Koebe: a non-vertical leaf admits a holomorphic covering onto the basis P1(C){1,1,}. Thanks to Picard little theorem there is no nonconstant holomorphic map CP1(C){1,1,} and therefore the universal covering of L is conformally equivalent to the unit disc D. The fact that these leaves are parabolic is a consequence of the fact that the basis P1(C){1,1,} is clearly parabolic (it is a compact Riemann surface minus a finite number of points, therefore it does not admit a non-constant bounded from above subharmonic function).

    The following is an example with a holonomy group generated by two Moebius maps.

    Example 5.13 (Legendre equation) Consider the equation of Legendre given by

    (1z2)u2zu+α(α+1)u=0

    where αC. From what we observed above we have that for a(z)=1z2, b(z)=2z and c(z)=α(α+1) there exists a Riccati equation given by

    dtdz=(1z2)t22zt+α(α+1)1z2.

    Putting w=1z we have

    dtdw=(w21)t22wt+α(α+1)w2w2(w21).

    As above, we are not aiming to compute the global holonomy at this time. We want to hightlight the qualitative information we obtain about the nature of the leaves by applying Proposition 5.12.

    As a final application of our methods we have:

    Example 5.14 (an equation without solutions) We consider the ODE zu+u+zu=0. In the Euler form we have z2u+zu+z2u=0 which was obtained from Besssel equation (Example 5.10) for ν=0. This last has indicial equation r(r1)+r=0 which gives as only solution r=0. Then Frobenius theorem assures the existence of an analytic solution of the form u(z)=z0n=0anzn. Let us take t=u/u and consider the corresponding Riccati equation

    dtdz=zt2+t+zz.

    Rewriting this equation we have

    zt(z)=zt2(z)t(z)z (18)

    Claim 5.15. Equation (18) admits no non-trivial formal solution.

    Proof. Indeed, let us assume that t(z)=n=0anzn is a formal solution of (18). Thus we have

    n=2[(n+1)an+cn1]zn+(2a1+c0+1)z+a0=0

    then a0=0, 2a1+c0+1=0 and (n+1)an+cn1=0 for all n=2,3,. Hence an=0 for all n=0,1,2,.

    The conclusion is that the solution given by Frobenius method vanishes at the origin (easy to see already from Frobenius type computations of the solution) and (more interesting) the solutions of the original ODE are Riemann surfaces of the logarithmic, since each solution u is of the form u=Ket for some constant K.

    In this section we shall refer to the notion of Liouvillian function as introduced in [24]. We stress the fact that the generating basis field is the one of rational functions. Thus a Liouvillian function of n complex variables x1,,xn will be a function belonging to a Liouvillian tower of differential extensions k0k1çdotskr starting the field k0 of rational functions k0=C(x1,,xn) equipped with the partial derivatives xj.

    Recall that a Liouvillian function is always holomorphic in some Zariski open subset of the space Cn. Nevertheless, it may have several branches. Let us denote by Dom(F)Cn the domain of F as the biggest open subset where F has local holomorphic branches. This allows the following definition:

    Definition 6.1 (Liouvillian solution, Liouvillian first integral, Liouvillian relation) Given an equation a(z)u+b(z)u+c(z)u=0, a Liouvillian function u(z) of the variable z will be called a solution of the ODE if we have a(z)u+b(z)u+c(z)u=0 in some nonempty open subset where u(z) is holomorphic. A three variables Liouvillian function F(x1,x2,x3) will be called a first integral of the ODE if given any local solution u0(z) of the ODE, defined for z in a disc D(z0,r)C, we have that F(z,u0(z),u0(z)) is constant for |zz0|<r provided that (z,u0(z),u0(z))Dom(F), for all zD(z0,r). Similarly we shall say that a solution u0(z) of the ODE, defined for zDom(u0)C, satisfies a Liouvillian relation if there is a Liouvillian function F(x1,x2,x3) such that {(z,u(z),u(z))C3,zDom(u)}çapDom(F) and F(z,u0(z),u0(z))=0 in some dense open subset of Dom(u0).

    Let us recall a couple of classical results:

    Theorem 6.2 (Singer, [24]). Assume that the polynomial first order ODE dxdz=P(x,y), dydz=Q(x,y) admits a Liouvillian first integral. Then there are rational functions U(x,y), V(x,y) such that Uy=Vx and the differential form Q(x,y)dxP(x,y)dy admits the integrating factor R(x,y)=exp[(x,y)(x0,y0)U(x,y)dx+V(x,y)dy].

    Theorem 6.3 (Rosenlicht [20], Singer [24]). Let p(z),q(z) be Liouvillian functions and L(y)=y+p(z)y+q(z)y. If L(y)=0 has a Liouvillian first integral then all solutions are Liouvillian. If L(y)=0 has a nontrivial Liouvillian solution, then this equation has a Liouvillian first integral.

    Example 6.4 (Bernoulli ODEs) Recall that a Bernoulli differential equation of power 1 is one of the form dydx=a1(x)y+a2(x)y2p(x). If we perform a change of variables as (x,y)(x,y1/k) then we obtain an equation of the form dydx=yk+1a2(x)+ya1(x)kp(x) which will be called a Bernoulli equation of power k.

    We prove the existence of a first integral for Ω=0 of Liouvillian type. First we observe that Ω=0 can be given by

    Ωp(x)yk+1=kdyyk+1(a2(x)p(x)+a1(x)p(x)yk)dx=0.

    Let now f(x) be such that f(x)f(x)=a1(x)p(x) and let g(x) be such that g(x)=a2(x)p(x)f(x). Then Ω=0 can be given by

    kdyyk+1+f(x)g(x)dx+f(x)ykf(x)dx=0.

    Therefore F(x,y)=g(x)1f(x)yk defines a first integral for Ω=0 which is clearly of Liouvillian type.

    Before proving Theorem E we shall need a lemma:

    Lemma 6.5. Let dydx=c(x)y2+b(x)y+a(x)a(x) be a rational Riccati ODE, where a(x),b(x), c(x) are complex polynomials. Assume that there is a Liouvillian first integral. Then we have the following possibilities:

    1. The equation is linear of the form a(x)yb(x)y=a(x).

    2. Up to a rational change of coordinates of the form Y=yA(x)/B(x), the equation is a Bernoulli equation dYdx=˜c(x)Y2+˜b(x)Y˜a(x).

    Proof. Let Ω=[c(x)y2+b(x)y+a(x)]dxa(x)dy. The ODE is equivalent to Ω=0. According to Singer [24] (Theorem 6.2 above) there is a rational 1-form η=U(x,y)dx+V(x,y)dy such that dη=(UyVx)dydx=0 and exp(η) is an integrating factor for Ω. This means that d(Ω/exp(η))=0 and therefore dΩ=ηΩ.

    We shall split our argumentation in two cases according to the existence of an invariant algebraic curve other than one of the vertical lines:

    Case 1. Ω=0 admits some invariant algebraic curve CC2 which is not a vertical line x=cC. First we let us prove that this curve is a graph of a rational function y=R(x). We may choose an irreducible polynomial f(x,y) such that f(x,y)=0 describes this non-vertical algebraic solution. Now we observe that the leaves of the Riccati foliation defined by Ω=0 on P1(C)×P1(C) are, except for those contained in the invariant vertical fibers, all transverse to the vertical fibers {x}×P1(C)P1(C)×P1(C). Thus we conclude that the partial derivative fy(x,y)=fy(x,y) never vanishes for each x such that a(x)0. Since f(x,y) is a polynomial the partial derivative fy(x,y) is also a polynomial. Take a value of x such that yfy(x,y) has no zero. Since this map is a polynomial, it must be constant say fy(x)=B(x). In particular the second partial derivative fyy(x,y)=2fy2(x,y) is identically zero for all (x,y) except maybe for a finite number of values of x. Then the polynomial fyy is identically zero and because f is a polynomial we must have f(x,y)=A(x)+B(x)y for some polynomial A(x). This shows that the non-vertical solution is a graph of the form y(x)=A(x)B(x).

    Now, as a classical procedure we may perform a change of variables as follows: write Y=yy(x) to obtain:

    dYdx=c(x)Y2+(2c(x)y(x)+b(x))Ya(x)=B(x)c(x)Y2+(b(x)B(x)+2c(x)A(x))Ya(x)B(x).

    This is a Bernoulli type equation (see [23] Example 4.2 page 771).

    Case 2. There is no invariant algebraic curve other than the vertical lines. Denote by (η) the polar divisor of η in C2. Let us prove that this polar divisor is a finite union of invariant vertical lines: Firstly notice that the polar set of η is invariant by Ω=0 (see [3] or [23] Lemma 5.5 page 772). By the hypothesis we then conclude that (η){xC,a(x)=0}×C is a union of vertical invariant lines.

    Now, from the Integration lemma in [22, Page 174;5, Page 5] we have

    η=rj=1λjdxxxj+d(g(x,y)rj=1(xxj)nj1)

    where g(x,y) is a polynomial function, xjC,λjC and njN is the order of the poles of Ω in the component (x=xj) of the polar set for every j=1,...,r. Now we use the equation dΩ=ηΩ to obtain

    [a(x)+2yc(x)+b(x)]dxdy=a(x)rj=1λjxxjdxdy+d(g(x,y)rj=1(xxj)nj1)Ω

    where

    d(g(x,y)rj=1(xxj)nj1)Ω=dgrj=1(xxj)nj1Ωga(x)d(1rj=1(xxj)nj1)dy.

    Notice that

    dgrj=1(xxj)nj1Ω=gy[c(x)y2+b(x)y+a(x)]rj=1(xxj)nj1dxdygxa(x)rj=1(xxj)nj1dxdy

    Notice that the left side is [a(x)+2yc(x)+b(x)] has no term in y2 so that we must have gyc(x)=0 in the right side. If gy=0 then g=g(x) and from the left side we must have c(x)=0. This shows that we must always have c(x)=0. This implies that the original equation is a linear homogeneous equation of the form dydx=b(x)y+a(x)a(x) which can be written as a(x)yb(x)y=a(x).

    Proof of Theorem E. Let us prove the second part, ie., the equivalence. We assume that L[u](z)=a(z)u+b(z)u+c(z)u=0 admits a Liouvillian first integral. We consider the change of coordinates t=uu which gives the following Riccati model R:dtdz=a(z)t2+b(z)t+c(z)a(z). We claim:

    Claim 6.6. The Riccati model R also admits a Liouvillian first integral.

    Proof. By hypothesis the ODE L[u]=a(z)u+b(z)u+c(z)u=0 has a Liouvillian first integral. By the Corollary in [24] page 674 all solutions of L[u]=0 are Liouvillian. This implies, by Theorem 1 in [24] page 674, that R admits a Liouvillian first integral.

    From the Lemma 6.5 have then two possibilities:

    Case 1. There is a solution γ(z)=A(z)/B(z) for the Riccati equation, where A,B are polynomials. In this case there is a rational change of coordinates of the form T=tA(z)/B(z) that takes the Riccati model R into a Bernoulli foliation B: dTdz=T2˜b(z)T where ˜b(z)=b(z)a(z)+2γ(z). In this case the original ODE L[u](z)=0 becomes ˜L[U](z)=U+˜b(z)U=0 after a rational change of coordinates

    U=exp(zTdη)=uexp(zγ(η)dη)

    where γ(z)=A(z)B(z). This shows that we have Liouvillian solutions to the ODE which are given by

    u(z)=exp(zγ(η)dη)[+kzexp(ηb(ξ)a(ξ)dξ)çdotexp(η2γ(ξ)dξ)dη]

    for constants k,C.

    Case 2. We have c(z)=0 and therefore the original ODE is of the form L[u]=a(z)u+b(z)u=0. Thus the solutions are Liouvillian given by u(z)=k1+k2zexp(ηb(ξ)a(ξ)dξ)dη for constants k1,k2C.



    [1] H. A. Bethe and E. E. Salpeter, Quantum Mechanics of One- and Two-Electron Atoms, Springer-Verlag, Berlin-Göttingen-Heidelberg; Academic Press Inc., New York 1957.
    [2] W. E. Boyce and R. C. DiPrima, Elementary Differential Equations and Boundary Value Problems, 10th edition, John Wiley & Sons, New York. 2012.
    [3] Holomorphic foliations with Liouvillian first integrals. Ergod. Theory Dyn. Syst. (2001) 21: 717-756.
    [4] C. Camacho and A. Lins Neto, Geometric Theory of Foliations, Progress in Math. Birkhäusser, Boston, Basel and Stuttgart, 1985. doi: 10.1007/978-1-4612-5292-4
    [5] D. Cerveau and J.-F. Mattei, Formes intégrables holomorphes singulières, Astérisque, vol. 97, Société Mathématique de France, Paris, 1982.
    [6] E. A. Coddington, An Introduction to Ordinary Differential Equations, Dover Publications, New York, 1989.
    [7] H. M. Farkas and I. Kra, Riemann Surfaces, Graduate Texts in Mathematics, 71. Springer-Verlag, New York-Berlin, 1980.
    [8] Ueber die Integration der linearen Differentialgleichungen durch Reihen. J. Reine Angew. Math. (1873) 76: 214-235.
    [9] C. Godbillon, Feuilletages. Études géométriques, Progress in Mathematics, vol. 98, Birkhäuser Verlag, Basel, 1991.
    [10] P. Griffiths and J. Harris, Principles of Algebraic Geometry, Wiley Classics Library, John Wiley & Sons Inc., New York, 1994. doi: 10.1002/9781118032527
    [11] On the part of the motion of the lunar perigee which is a function of the mean motions of the sun and moon. Acta Math. (1886) 8: 1-36.
    [12] J. Kepler, Astronomia Nova, 1609, edited in J. Kepler, Gesammelte Werke, vol Ⅲ, 2nd edition, C.H. Beck, München, 1990.
    [13] A. Lins Neto and B. Scárdua, Complex Algebraic Foliations, Expositions in Mathematics, vol. 67, Walter de Gruyter GmbH, Berlin/Boston, 2020.
    [14] Frobenius avec singularités. I. Codimension un. Inst. Hautes Études Sci. Publ. Math. (1976) 46: 163-173.
    [15] Frobenius avec singularités. Ⅱ. Le cas général. Invent. Math. (1977) 39: 67-89.
    [16] Holonomie et intégrales premières. Ann. Sci. École Norm. Sup. (1980) 13: 469-523.
    [17] I. Newton, The Mathematical Principles of Natural Philosophy, Bernard Cohen Dawsons of Pall Mall, London, 1968.
    [18] P. Painlevé, Leçcons sur la Théorie Analytique des Équations Différentielles, Librairie Scientifique A. Hermann, Paris, 1897.
    [19] F. Reis, Methods from Holomorphic Foliations in Differential Equations, Ph. D thesis, IM-UFRJ in Rio de Janeiro, 2019.
    [20] On Liouville's theory of elementary functions. Pacific J. Math. (1976) 65: 485-492.
    [21] Construction of vector fields and Ricatti foliations associated to groups of projective automorphism. Conform. Geom. Dyn. (2010) 14: 154-166.
    [22] Transversely affine and transversely projective holomorphic foliations. Ann. Sci. École Norm. Sup. (1997) 30: 169-204.
    [23] Differential algebra and Liouvillian first integrals of foliations. J. Pure Appl. Algebra (2011) 215: 764-788.
    [24] Liouvillian first integrals of differential equations. Trans. Amer. Math. Soc. (1992) 333: 673-688.
  • Reader Comments
  • © 2021 the Author(s), licensee AIMS Press. This is an open access article distributed under the terms of the Creative Commons Attribution License (http://creativecommons.org/licenses/by/4.0)
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

Metrics

Article views(2699) PDF downloads(281) Cited by(0)

Other Articles By Authors

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog