Review Special Issues

Integrative cardiology-state of the art of mind body therapies for the treatment of cardiovascular disease and risk factors

  • Mind-body therapies are a heterogeneous group of interventions that seek to improve multiple aspects of somatic health by focusing on interactions between mental factors and physiological functions. There is a growing interest in modern Western culture in these forms of alternative medicine. Most of these therapies exert their effects via stress control. Induction of the relaxation response via neurohormonal, endocrine and immunological pathways may have beneficial effects in a variety of conditions, including oncological, neuropsychiatric and cardiovascular pathologies. Several randomized controlled trials have produced promising results, supporting a complementary role of mind-body therapies for both the prevention and treatment of the most prevalent cardiovascular problems.

    Citation: Bernardo Lopez Abel, María I Martínez-Soto, Maria Luz Couce. Integrative cardiology-state of the art of mind body therapies for the treatment of cardiovascular disease and risk factors[J]. AIMS Medical Science, 2018, 5(1): 80-89. doi: 10.3934/medsci.2018.1.80

    Related Papers:

    [1] Ariane Locat, Pascal Locat, Hubert Michaud, Kevin Hébert, Serge Leroueil, Denis Demers . Geotechnical characterization of the Saint-Jude clay, Quebec, Canada. AIMS Geosciences, 2019, 5(2): 273-302. doi: 10.3934/geosci.2019.2.273
    [2] Aleksander S. Gundersen, Ragnhild C. Hansen, Tom Lunne, Jean-Sebastien L’Heureux, Stein O. Strandvik . Characterization and engineering properties of the NGTS Onsøy soft clay site. AIMS Geosciences, 2019, 5(3): 665-703. doi: 10.3934/geosci.2019.3.665
    [3] Boris I. Birger . Internal friction in the Earth' crust and transverse seismic waves. AIMS Geosciences, 2022, 8(1): 84-97. doi: 10.3934/geosci.2022006
    [4] Michael Long, Kevin Flynn, Chris Menkiti, Ronan Travers, Kieran Simpson, David Gill, Miles Friedman . Characterisation and engineering properties of Dublin Boulder Clay—an update. AIMS Geosciences, 2025, 11(2): 387-441. doi: 10.3934/geosci.2025017
    [5] Vladimir Kompanichenko, Galina El-Registan . Advancement of the TI concept: defining the origin-of-life stages based on the succession of a bacterial cell exit from anabiosis. AIMS Geosciences, 2022, 8(3): 398-437. doi: 10.3934/geosci.2022023
    [6] Elvis Ishimwe, Richard A. Coffman, Kyle M. Rollins . Predicting blast-induced liquefaction within the New Madrid Seismic Zone. AIMS Geosciences, 2020, 6(1): 71-91. doi: 10.3934/geosci.2020006
    [7] Alan K Betts . The Earth's view of climate change. AIMS Geosciences, 2022, 8(2): 224-232. doi: 10.3934/geosci.2022013
    [8] Solve Hov, David Gaharia . Geotechnical characterization of index and deformation properties of Stockholm clays. AIMS Geosciences, 2023, 9(2): 258-284. doi: 10.3934/geosci.2023015
    [9] Annika Bihs, Mike Long, Steinar Nordal . Geotechnical characterization of Halsen-Stjørdal silt, Norway. AIMS Geosciences, 2020, 6(3): 355-377. doi: 10.3934/geosci.2020020
    [10] Dušan Berisavljević, Zoran Berisavljević . Engineering properties and characterisation of sediments at the Belgrade Waterfront test site. AIMS Geosciences, 2025, 11(3): 600-628. doi: 10.3934/geosci.20250226
  • Mind-body therapies are a heterogeneous group of interventions that seek to improve multiple aspects of somatic health by focusing on interactions between mental factors and physiological functions. There is a growing interest in modern Western culture in these forms of alternative medicine. Most of these therapies exert their effects via stress control. Induction of the relaxation response via neurohormonal, endocrine and immunological pathways may have beneficial effects in a variety of conditions, including oncological, neuropsychiatric and cardiovascular pathologies. Several randomized controlled trials have produced promising results, supporting a complementary role of mind-body therapies for both the prevention and treatment of the most prevalent cardiovascular problems.


    We focus in this work on families of planar real analytic vector fields $ \mathcal{X} = P(x, y; \lambda) \partial_x + Q(x, y; \lambda) \partial_y $ defined in a neighborhood of a monodromic singularity that can be placed at the origin without loss of generality. Here $ \lambda \in \mathbb{R}^p $ denotes the finite number of parameters in the family. We recall that a monodromic singular point of $ \mathcal{X} $ is a singularity of $ \mathcal{X} $ such that the associated flow rotates around it, and therefore a Poincaré map $ \Pi $ is well defined in a sufficiently small transversal section with an end point at the singularity. Centers and foci are examples of monodromic singularities. We will restrict the family to the monodromic parameter space $ \Lambda \subset \mathbb{R}^p $ defined as the parameter subset for which the origin is a monodromic singularity of $ \mathcal{X} $. The common characterization of $ \Lambda $ is via the blow-up procedure developed by Dumortier in [1]; see also Arnold [2]. It is worth emphasizing that Algaba and co-authors present in [3,4] an algorithmic procedure to determine the parameter restrictions that defines $ \Lambda $ in terms of the Newton diagram $ \mathbf{N}(\mathcal{X}) $ of the vector field $ \mathcal{X} $; see also [5].

    In [6], an explicit analytic first-order ordinary differential equation was obtained for the Poincaré map $ \Pi $ associated with a monodromic singularity (without local zero angular speed curves) of $ \mathcal{X} $ under the assumption of the existence of a Laurent inverse integrating factor $ V $ of $ \mathcal{X} $. Later on, in [7], an example of focus without Puiseux (a generalization of Laurent) inverse integrating factor is presente. This example shows that the differential equation obtained in [6] for $ \Pi $ is not universal.

    Initially, our first main objective in this work was to find an explicit analytic ordinary differential equation $ f(\rho, \Pi, \Pi', \ldots, \Pi^{k}) = 0 $ for the Poincaré map $ \Pi(\rho) $ associated to a monodromic singularity of $ \mathcal{X} $ where $ f $ only depends on objects that do not include the flow of $ \mathcal{X} $. If this objective were achieved, an alternative proof of the fact that $ \Pi $ admits an asymptotic Dulac series can be constructed via Bruno's theory [8,9] of asymptotic solutions of analytic differential equations at singularities. In this work we take the former universal object to be the analytic invariant curve $ F(x, y) = 0 $ through the monodromic singularity that there always exists, as it was proved in [10]. Using this idea, we find a first order ordinary differential equation for $ \Pi $ only depending on $ F $; see Theorem 1.1 in Section 1.1, although we have been able to prove its analyticity only under some assumptions; see, for example, Theorem 1.6. This equation generalizes the one obtained in [6] in the sense that both coincides in the particular case that $ F $ becomes an inverse integrating factor of $ \mathcal{X} $, as we show in Section 2.

    We denote by $ \mathbf{N}(\mathcal{X}) $ the Newton diagram of $ \mathcal{X} $ that is composed of edges joining both positive semi-axis; see [11] for a detailed account of this construction and also [3]. Each edge of $ \mathbf{N}(\mathcal{X}) $ has endpoints in $ \mathbb{N}^2 $ and slope $ -p/q $ with $ (p, q) \in \mathbb{N}^2 $ coprimes. From now on we denote by $ W(\mathbf{N}(\mathcal{X})) \subset \mathbb{N}^2 $ the set of all weights $ (p, q) $. We will use the weighted polar blow-up $ (x, y) \mapsto (\rho, \varphi) $ given by

    $ x=ρpcosφ,  y=ρqsinφ,
    $
    (1.1)

    with Jacobian

    $ J(\varphi, \rho) = \rho^{p+q-1}(p \cos^2\varphi + q \sin^2\varphi). $

    In these coordinates, the system associated to $ \mathcal{X} $ is written in the form $ \dot{\rho} = R(\varphi, \rho) $, $ \dot{\varphi} = \Theta(\varphi, \rho) $ with $ R(\varphi, 0) = 0 $ and $ \Theta(\varphi, \rho) = G_r(\varphi) + O(\rho) $. Therefore, the equations of the orbits of $ \mathcal{X} $ near the origin are governed by the equation

    $ dρdφ=F(φ,ρ):=R(φ,ρ)Θ(φ,ρ)
    $
    (1.2)

    well defined in $ C \backslash \Theta^{-1}(0) $ being the cylinder $ C \, = \, \left\{ (\theta, \rho) \in \mathbb{S}^1 \times \mathbb{R} \, : \, 0 \leq \rho \ll 1 \right\} $ with $ \mathbb{S}^1 = \mathbb{R}/ (2 \pi \mathbb{Z}) $. the local set of zero angular speed is defined as $ \Theta^{-1}(0) = \{ (\varphi, \rho) \in C : \Theta(\varphi, \rho) = 0, \rho \geq 0 \} $, and the set of characteristic directions $ \Omega_{pq} = \{ \varphi^* \in \mathbb{S}^1 : \Theta(\varphi^*, 0) = G_r(\varphi^*) = 0 \} $. Notice that the set $ \{\rho = 0 \} $ is invariant for the flow of (1.2), and it becomes either a periodic orbit or a polycycle according to whether $ \Omega_{pq} = \emptyset $ or $ \Omega_{pq} \neq \emptyset $, respectively.

    We also define the $ (p, q) $-critical parameters as the elements of the subset $ \Lambda_{pq} \subset \Lambda $ of the monodromic parameter space corresponding to vector fields with local curves of zero angular speed, that is,

    $ \Lambda_{pq} = \{ \lambda \in \Lambda : \Theta^{-1}(0) \backslash \{ \rho = 0\} \neq \emptyset \}. $

    We emphasize that $ \Lambda_{pq} = \emptyset $ when $ \Omega_{pq} = \emptyset $ but the converse is not true.

    Given $ (p, q) \in W(\mathbf{N}(\mathcal{X})) $, we take the $ (p, q) $-quasihomogeneous expansion

    $ X=jrXj,
    $
    (1.3)

    where $ r \geq 1 $ and $ \mathcal{X}_j $ are $ (p, q) $-quasihomogeneous vector fields of degree $ j $, and $ \mathcal{X}_r $ is called the leading vector field of $ \mathcal{X} $.

    Let $ F(x, y) = 0 $ be a real invariant analytic curve of $ \mathcal{X} $ with $ F(0, 0) = 0 $, and denote by $ K(x, y) $ the cofactor of $ F $, that is, $ \mathcal{X}(F) = K F $ holds. We recall that $ F $ always exists, as it was proved in [10] using the separatrix theorem of Camacho-Sad. In weighted polar coordinates this equation is transformed into

    $ ˆX(ˆF)=ˆKˆF,
    $
    (1.4)

    where $ \hat{\mathcal{X}} = \partial_\varphi + \mathcal{F}(\varphi, \rho) \partial_\rho $, $ \hat{F}(\varphi, \rho) = F(\rho^p \cos\varphi, \rho^q \sin\varphi) $ and $ \hat{K} $ is the cofactor of the invariant curve $ \hat{F} = 0 $ of $ \hat{\mathcal{X}} $ whose expression is

    $ ˆK(φ,ρ)=K(ρpcosφ,ρqsinφ)ρrΘ(φ,ρ).
    $
    (1.5)

    Let $ \Phi(\varphi; \rho_0) $ be the flow of $ \hat{\mathcal{X}} $ with initial condition $ \Phi(0; \rho_0) = \rho_0 $. In [10], it is proved that

    $ Ipq(ρ0)=PV2π0ˆK(φ,Φ(φ;ρ0))dφ,
    $
    (1.6)

    exists for any initial condition $ \rho_0 > 0 $ sufficiently small. Working in $ \Lambda \backslash \Lambda_{pq} $ we may also define the integral

    $ ζpq(ρ0)=2π0Fρ(φ,Φ(φ;ρ0))dφ,
    $
    (1.7)

    and the difference

    $ \alpha_{pq}(\rho_0) = \zeta_{pq}(\rho_0) - I_{pq}(\rho_0). $

    A fundamental result in this work is the following one.

    Theorem 1.1. The equation

    $ ˆF(0,Π(ρ0))=ˆF(0,ρ0)exp(Ipq(ρ0))
    $
    (1.8)

    holds. Moreover, in $ \Lambda \backslash \Lambda_{pq} $, the relations

    $ ζpq(ρ0)=log(Π(ρ0)),
    $
    (1.9)

    and

    $ ˆF(0,Π(ρ0))=ˆF(0,ρ0)exp(αpq(ρ0))Π(ρ0)
    $
    (1.10)

    are satisfied.

    Remark 1.2. We believe that the fundamental equation (1.10) holds in the whole monodromic space $ \Lambda $ and not only in $ \Lambda \backslash \Lambda_{pq} $, but we have no proof.

    Remark 1.3. If $ \mathcal{X} $ has $ \ell > 1 $ invariant analytic curves $ F_i = 0 $ with cofactors $ K_i $ for $ i = 1, \ldots, \ell $, then it also has the $ \ell $-parameter invariant analytic curve $ F = \prod_i F_i^{m_i} = 0 $ with arbitrary multiplicities $ m_i \in \mathbb{N} $ and cofactor $ K = \sum_i m_i K_i $. Sometimes we may find $ m_i $ that simplifies the expression of $ K $. As an example, the case when $ F $ becomes an inverse integrating factor of $ \mathcal{X} $, hence $ K = {\rm div}(\mathcal{X}) $, was studied in [6].

    Il'Yashenko in [12] proves that $ \Pi $ has a Dulac asymptotic expansion possessing a linear leading term. More specifically, one has $ \Pi(\rho_0) = \eta_1 \rho_0 + o(\rho_0) $. The computation of the first (generalized) Poincaré-Lyapunov quantity $ \eta_1 $ is quite involved and needs the use of cumbersome blow-ups computations; see for example [13]. Some sufficient conditions that guarantee the computation of $ \eta_1 $ are stated below. We will user the following expansions. The functions $ \Theta $ and $ R $ appearing in (1.2) have Taylor expansions at $ \rho = 0 $ given by $ \Theta(\varphi, \rho) = G_r(\varphi) + O(\rho) $ and $ R(\varphi, \rho) = F_r(\varphi) \rho + O(\rho^2) $. Moreover $ F(x,y) = F_{s}(x,y) + \cdots $ and $ K(x,y) = K_{\bar{r}}(x,y) + \cdots $ are the $ (p,q) $-quasihomogeneous expansions of the cofactor $ K $ associated to the analytic invariant curve $ F = 0 $. Here $ F_s $ and $ K_{\bar{r}} $ are the leading $ (p,q) $-quasihomogeneous polynomials of weighted degree $ s $ and $ \bar{r} $, respectively, and the dots denote higher $ (p,q) $-quasihomogeneous terms.

    Proposition 1.4. Assume that both $ I_{pq} $ and $ \zeta_{pq} $ can be extended with continuity to the origin. Then

    $ \log(\eta_1^s) = I_{pq}(0) = s \, \zeta_{pq}(0), $

    that is,

    $ \log(\eta_1^s) = PV \int_0^{2\pi} \frac{K_{\bar{r}}(\cos\varphi, \sin\varphi)}{G_r(\varphi)} d \varphi = s PV \int_0^{2\pi} \frac{F_r(\varphi)}{G_r(\varphi)} d \varphi, $

    provided that both principal values exist.

    Remark 1.5. It is worth emphasizing that, in general, $ I_{pq} $ cannot be extended by continuity at $ \rho_0 = 0 $. In [10], it is shown that expression (3.4) of $ \eta_1 $ is wrong in several examples.

    Clearly, if $ \Phi(\varphi; \rho_0) $ is analytic at $ \rho_0 = 0 $ for all $ \varphi \in \mathbb{S}^1 $ then $ \Pi $ is analytic at $ \rho_0 = 0 $. A sufficient condition for that to happen is that $ \Omega_{pq} = \emptyset $. Instead, a weaker condition to ensure the analyticity of $ \Pi $ is the following one.

    Theorem 1.6. If $ I_{pq} $ is analytic at the origin, then $ \Pi $ is too.

    The classical center-focus problem has been studied for decades. In [14,15], some particular degenerate systems with a monodromic singularity were studied. In [16,17], it is proved that some degenerate systems with a monodromic singularity are limit of differential systems with monodromic linear part. In [18,19,20] some sufficient conditions to have a center at a completely degenerate critical point are given. In [21], the relation between the reversivility and the center problem is studied. The textbook [22] is a good summary about the relations between the the center and cyclicity problems. In [23], a geometrical criteria to determine the existence of a center for certain differential systems is given. In [24], the Hopf-bifurcation formulas for some differential systems are established. Final, in [25], the authors solve the center problem for monodromic sigularities with characteristic directions and with inverse integrating factor and [26] the linear term of all the monodromic families known is obtained. However, no general characterization was known until the work [10], where it is proved the following theorem.

    Theorem 1.7 ([10]). Let $ \mathcal{X} $ be a family of analytic planar vector fields having a monodromic singular point at the origin and $ K $ the cofactor associated to a real analytic invariant curve through the origin. Then $ I_{pq}(\rho_0) $ exists for any initial condition $ \rho_0 > 0 $ that is sufficiently small, and the origin is a center if and only if $ I_{pq}(\rho_0) \equiv 0 $.

    We show other necessary center condition in $ \Lambda \backslash \Lambda_{pq} $.

    Proposition 1.8. If the origin is a center of $ \mathcal{X} $ and we restrict to the parameter space $ \Lambda \backslash \Lambda_{pq} $ then $ \alpha_{pq}(\rho_0) = \zeta_{pq}(\rho_0) \equiv 0 $.

    Notice that from Proposition 1.8 and relation $ \alpha_{pq}(\rho_0) = \zeta_{pq}(\rho_0) - I_{pq}(\rho_0) $, we obtain a new proof of the necessary part in Theorem 1.7.

    Example 1.9. We emphasize that there are focus with parameters in $ \Lambda \backslash \Lambda_{pq} \neq \emptyset $ and $ \alpha_{pq}(\rho_0) \not\equiv 0 $. As an example, we consider the family

    $ ˙x=λ1(x6+3y2)(y+μx)+λ2(x2+y2)(y+Ax3),˙y=λ1(x6+3y2)(x+μy)+λ2(x2+y2)(x5+3Ax2y).
    $
    (1.11)

    In [27], it is proved that

    $ \Lambda = \{ (\lambda_1, \lambda_2, \mu, A) \in \mathbb{R}^4 : 3 \lambda_1 - \lambda_2 > 0, \ \ \lambda_1 - \lambda_2 > 0\} $

    and the Poincaré map is the linear map $ \Pi(\rho_0) = \eta_1 \rho_0 $. Moreover, $ F(x, y) = (x^2+y^2)(x^6 + 3 y^2) $ is an inverse integrating factor of the full family (1.11).Using the weights $ (p, q) = (1, 1) \in W(\mathbf{N}(\mathcal{X})) $ we have $ \mathcal{X} = \mathcal{X}_2 + \cdots $, hence $ r = 2 $, and the forthcoming formula (2.3) yields $ \alpha_{11}(\rho_0) = -\log(\eta_1^7) $. The value $ \eta_1 = \exp(2 \pi \lambda_1 \mu + 2 \sqrt{3} \pi A \lambda_2 /3) $ is a consequence of the works [6,27]. Notice that in this example $ 0 \in \Omega_{11} $.

    We consider analytic degenerate infinity vector fields

    $ X=Xn+AXE
    $
    (1.12)

    with $ \mathcal{X}_n = P_{p+n}(x, y) \partial_x + Q_{q + n}(x, y) \partial_y $ a $ (p, q) $-quasihomogeneous polynomial vector field of degree $ n $, $ \mathcal{X}_E = p x \partial_x + q y \partial_y $ is the radial Euler field, and $ A(x, y) $ a real analytic function in $ \mathbb{R}^2 $ whose Taylor expansion at the origin starts with $ (p, q) $-quasihomogeneous terms of degree $ m-1 $ with $ m > n+1 $. In the work [28] the particular homogeneous case $ (p, q) = (1, 1) $ is analyzed, and here we generalize it.

    Proposition 1.10. Any degenerate infinity vector field (1.12) has the homogeneous algebraic invariant curve

    $ F(x,y) = p \, x \, Q_{q+n}(x,y) - q \, y \, P_{p+n}(x,y) = 0 $

    with cofactor

    $ K(x,y) = {\rm div}(\mathcal{X}_n) + (n + p + q) A(x,y). $

    Moreover, it also has the inverse integrating factor $ V = F H^{\frac{m-n-1}{d}} $ provided that $ A $ is $ (p, q) $-quasihomogeneous of degree $ m-1 $, where $ H $ is a $ (p, q) $-quasihomogeneous first integral of degree $ d $ of $ \mathcal{X}_n $.

    Proposition 1.11. Any degenerate infinity vector field (1.12) with a monodromic singularity at the origin becomes a focus provided that $ {\rm div}(\mathcal{X}_n) + (n + p + q) A(x, y) $ is a positive or negative defined function in a neighborhood of the origin. If additionally the origin is a center of $ \mathcal{X}_n $ and $ A $ is $ (p, q) $-quasihomogeneous then the Poincaré map associated with the origin is analytic and has the form $ \Pi(\rho) = \rho + \eta_{m-n} \rho^{m-n} (1+ O(\rho)) $ with

    $ \eta_{m-n} = \int_0^{2 \pi} \frac{(p \cos^2\varphi + q \sin^2\varphi) A(\cos\varphi, \sin\varphi)}{F(\cos\varphi, \sin\varphi) \left( H(\cos\varphi, \sin\varphi) \right)^{(m-n-1)/d}} \, d \varphi. $

    In particular, the origin is a center if and only if $ \eta_{m-n} = 0 $ and the cyclicity of the origin in family (1.12) is zero.

    Remark 1.12. We note that the $ (p,q) $-quasihomogeneous first integral $ H $ of the monodromic vector field $ \mathcal{X}_n $ appearing in Proposition 1.10 only exists in case the origin be a center of $ \mathcal{X}_n $. This statement can be inferred from statement (ⅱ) of Lemma 4 in [29] because the transformed first integral to $ (p,q) $-weighted polar coordinates becomes a Laurent first integral of $ \mathcal{X}_n $. In particular, $ \Pi(\rho) = \rho + o(\rho) $ in agreement with Proposition 1.11.

    In this section we restrict our attention to the monodromic subset $ \Lambda \backslash \Lambda_{pq} $ so that $ \Theta^{-1}(0) \backslash \{ \rho = 0\} = \emptyset $. In particular $ \Omega_{pq} $ can be empty or not.

    Let us assume the particular case that $ F(x,y) $ is an analytic inverse integrating factor of $ \mathcal{X} $; that is, $ {\rm div}(\mathcal{X}/F) \equiv 0 $ holds. Then the function

    $ V(φ,ρ)=ˆF(φ,ρ)ρrJ(φ,ρ)Θ(φ,ρ)
    $
    (2.1)

    is an inverse integrating factor of $ \hat{\mathcal{X}} $ in $ C \backslash \{\Theta^{-1}(0) \cup \{\rho = 0\} \} $, that is, $ V $ satisfies the equation $ \hat{\mathcal{X}}(V) = \partial_\rho(\mathcal{F}) \, V $. Applying the differential operator $ \hat{\mathcal{X}} $ on both sides of (2.1) and taking into account (1.4), we obtain the relation between $ \hat{K} $ and $ \partial_\rho \mathcal{F} $ given by

    $ ρF=ˆKˆX(log|ρrJΘ|).
    $
    (2.2)

    On the other hand, using (2.2) and recalling that $ \alpha_{pq}(\rho_0) = \zeta_{pq}(\rho_0) - I_{pq}(\rho_0) $, we obtain

    $ αpq(ρ0)=2π0ˆX(log|ρrJΘ|)(φ,Φ(φ;ρ0))dφ=2π0ddφ(log|ρrJΘ|(φ,Φ(φ;ρ0)))dφ,=log|Πr(ρ0)J(0,Π(ρ0))Θ(0,Π(ρ0))ρrJ(0,ρ0)Θ(0,ρ0)|,
    $
    (2.3)

    where we have used the $ 2 \pi $-periodicity of $ J $ and $ \Theta $ in the variable $ \varphi $. Using this expression of $ \alpha_{pq} $ and taking into account (2.1), Eq (1.10) in $ \Lambda \backslash \Lambda_{pq} $ is written in the simpler form $ V(0, \Pi(\rho_0)) = V(0, \rho_0) \, \Pi'(\rho_0) $. In this way we rediscover the formula obtained in [6] by other methods. This formula for the special case of degenerate differential systems without characteristic directions was given in [30]; see also [31].

    Proof. To prove the first statement, we evaluate (1.4) along the flow $ \Phi(\varphi; \rho_0) $, and we obtain

    $ ddφˆF(φ,Φ(φ;ρ0))=ˆK(φ,Φ(φ;ρ0))ˆF(φ,Φ(φ;ρ0)),
    $

    hence

    $ PV2π0ddφˆF(φ,Φ(φ;ρ0))ˆF(φ,Φ(φ;ρ0))dφ=PV2π0ˆK(φ,Φ(φ;ρ0))dφ,
    $
    (3.1)

    where the last principal value exist for any initial condition $ \rho_0 > 0 $ sufficiently small as it was proved in [10]. Therefore, (3.1) takes the form

    $ \hat{F}(2 \pi, \Phi(2 \pi; \rho_0)) = \hat{F}(0, \rho_0) \, \exp\left(I_{pq}(\rho_0) \right). $

    Using the $ 2 \pi $-periodicity of $ \hat{F} $ in the variable $ \varphi $ and the definition of $ \Pi $, the former equation becomes (1.8).

    To prove the second part, we use the definition of $ \Phi(\varphi; \rho_0) $, that is,

    $ Φφ(φ;ρ0)=F(φ,Φ(φ;ρ0)),  Φ(0;ρ0)=ρ0>0.
    $
    (3.2)

    Working in $ \Lambda \backslash \Lambda_{pq} $ we know that $ \mathcal{F} $ is analytic in $ C \backslash \{ \rho = 0 \} $; hence, $ \Phi $ is also analytic there, and differentiating both expressions in (3.2) with respect to $ \rho_0 $ yields

    $ φ(Φρ0(φ;ρ0))=Fρ(φ,Φ(φ;ρ0))Φρ0(φ;ρ0),   Φρ0(0;ρ0)=1.
    $
    (3.3)

    Clearly $ (\varphi, \Phi(\varphi; \rho_0)) \in C \backslash \{ \rho = 0 \} $ for all $ \varphi \in [0, 2 \pi] $, and therefore the function $ \partial_\rho \mathcal{F}(\varphi, \Phi(\varphi; \rho_0)) $ is continuous in $ \mathbb{S}^1 \times \{ 0 < \rho_0 \ll 1 \} $. Thus, we may integrate the first equality in (3.3), yielding

    $ \int_0^{2 \pi} \frac{\partial \mathcal{F}}{\partial \rho}(\varphi, \Phi(\varphi; \rho_0)) \, d \varphi = \int_0^{2 \pi} \frac{\frac{\partial}{\partial \varphi} \left( \frac{\partial \Phi}{\partial \rho_0}(\varphi; \rho_0) \right)}{\frac{\partial \Phi}{\partial \rho_0}(\varphi; \rho_0)} \, d \varphi. $

    So we obtain

    $ \int_0^{2 \pi} \frac{\partial \mathcal{F}}{\partial \rho}(\varphi, \Phi(\varphi; \rho_0)) \, d \varphi = \left[ \log\left( \frac{\partial \Phi}{\partial \rho_0}(\varphi; \rho_0) \right) \right]_{\varphi = 0}^{\varphi = 2 \pi} = \log\left( \frac{\partial \Phi}{\partial \rho_0}(2 \pi; \rho_0) \right) $

    that can be written as Eq (1.9). Finally, Eq (1.8) is rewritten as the fundamental ordinary differential equation (1.10).

    Proof. Using the definition (1.5) of $ \hat{K} $ together with the $ (p, q) $-quasihomogeneous expansion $ K(x, y) = K_{\bar{r}}(x, y) + \cdots $ and the Taylor expansion $ \Theta(\varphi, \rho) = G_r(\varphi) + O(\rho) $ with $ \bar{r} \geq r $ we obtain

    $ \hat{K}(\varphi, \rho) = \rho^{\bar{r}-r} \frac{K_{\bar{r}}(\cos\varphi, \sin\varphi)}{G_r(\varphi)} + O(\rho^{\bar{r}-r+1}). $

    From [10], we know that it is proved that the principal value $ I_{pq}(\rho_0) $ defined in (1.6) exists for any $ \rho_0 > 0 $ and small. If $ I_{pq} $ can be extended with continuity to the origin, then, using that $ \Phi(\varphi; 0) = 0 $, we would have

    $ I_{pq}(0) = \left\{ 0ifˉr>r,PV2π0Kˉr(cosφ,sinφ)Gr(φ)dφifˉr=r,
    \right. $

    in case that this principal value exists. Moreover, from the $ (p, q) $-quasihomogeneous expansion $ F(x, y) = F_{s}(x, y) + \cdots $, we could use Eq (1.8) to express

    $ Ipq(ρ0)=log|ˆF(0,Π(ρ0))ˆF(0,ρ0)|=log|Πs(ρ0)Fs(cosφ,sinφ)+O(Πs+1(ρ0))ρs0(Fs(cosφ,sinφ)+O(ρ0))|
    $

    whose extension to $ \rho_0 = 0 $ gives $ I_{pq}(0) = \log(\eta_1^s) $, taking into account that $ \Pi(\rho_0) = \eta_1 \rho_0 + o(\rho_0) $. Comparing both expressions of $ I_{pq}(0) $, we have that if $ \bar{r} > r $, then $ \eta_1 = 1 $, whereas

    $ log(ηs1)=PV2π0Kˉr(cosφ,sinφ)Gr(φ)dφ,
    $
    (3.4)

    when $ \bar{r} = r $ under the restriction that the former principal value exists.

    On the other hand, if $ \zeta_{pq} $ can be extended with continuity to the origin, then, by (1.9),

    $ \log(\eta_1) = \zeta_{pq}(0) = PV \int_0^{2\pi} \frac{F_r(\varphi)}{G_r(\varphi)} d \varphi, $

    assuming this last principal value exists. Then the proposition follows.

    Proof. We define

    $ f(\rho_0, \Pi) = \hat{F}(0, \Pi) - \hat{F}(0, \rho_0) \, \exp\left(I_{pq}(\rho_0) \right). $

    The analyticity of $ I_{pq} $ at the origin implies that (1.8) can be written as $ f(\rho_0, \Pi) = 0 $, where $ f $ is an analytic function in a neighborhood of $ (\rho_0, \Pi) = (0, 0) $. Therefore, the Poincaré map $ \Pi(\rho_0) = \eta_1 \rho_0 + o(\rho_0) $ must be a branch of $ f $ at the origin and, consequently, admits the convergent Puiseux expansion

    $ Π(ρ0)=i0ηi+1ρ1+in0
    $
    (3.5)

    with some index $ n \in \mathbb{N}^* $. Now we are going to compute $ n $.

    By Proposition 1.4 we know that $ I_{pq}(0) = \log(\eta_1^s) $. Then $ \exp(I_{pq}(\rho_0)) = \eta_1^s + O(\rho_0) $ with $ \eta_1 > 0 $. Using that $ \hat{F}(0, \rho_0) = \rho_0^s (F_s(1, 0) + O(\rho_0)) $, we obtain that $ f(\rho_0, \Pi) = F_s(1, 0) (\Pi^s - \eta_1^s \rho_0^s) + \cdots $ from where we deduce that the Newton diagram of $ f $ only contains the edge $ L $ joining the endpoints $ (s, 0) $ and $ (0, s) $ provided that $ F_s(1, 0) \neq 0 $. We can take $ F_s(1, 0) \neq 0 $ because, without loss of generality, we take $ 0 \not\in \Omega_{pq} $ after a rotation (if necessary), and moreover we claim that the eventual real roots of $ F_s(\cos\varphi, \sin\varphi) $ must belong to $ \Omega_{pq} $. To prove the claim, we observe that

    $ \hat{F}(\varphi, \rho) = F(\rho^p \cos\varphi, \rho^q \sin\varphi) = \rho^s \, [F_{s}(\cos\varphi, \sin\varphi) + o(\rho)], $

    hence $ \hat{F}(\varphi, \rho) / \rho^s = 0 $ is an invariant curve of the differential equation (1.2). Clearly that invariant curve can only intersect the monodromic polycycle $ \rho = 0 $ at its singularities $ (\varphi, \rho) = (\varphi^*, 0) $ with $ \varphi^* \in \Omega_{pq} $.

    Once we know that $ F_s(1, 0) \neq 0 $ so that the edge $ L $ has slope $ -1 $, then we compute

    $ f(\rho_0, \rho_0 \mu) = \rho_0^s((\mu^s-\eta_1^s) F_s(1,0) + O(\rho_0)). $

    From this expression we deduce that the determining polynomial $ \mathcal{P}(\mu) $ associated to $ L $ is $ \mathcal{P}(\mu) = F_s(1, 0) (\mu^s-\eta_1^s) $. Therefore $ \mu = \eta_1 $ is a simple root of $ \mathcal{P} $, and it follows that $ n = 1 $ in the Puiseux expansion (3.5); this is a classical result that is proved, for instance, in [32]. Therefore, $ \Pi $ admits a convergent power series expansion at the origin finishing the proof.

    Proof. Let the origin be a center. Then Eq (1.10) must have the solution $ \Pi(\rho_0) = \rho_0 $ and therefore $ \hat{F}(0, \rho_0) = \hat{F}(0, \rho_0) \, \exp(-\alpha_{pq}(\rho_0)) $ holds. We claim that $ \hat{F}(0, \rho_0) \not\equiv 0 $ so it must occur $ \alpha_{pq}(\rho_0) \equiv 0 $ proving the proposition.

    To prove the claim, we use the $ (p, q) $-quasihomogeneous expansion $ F(x, y) = F_{s}(x, y) + \cdots $ with $ F_s(x, y) \not\equiv 0 $ a $ (p, q) $-quasihomogeneous polynomial of degree $ s $ and the dots are $ (p, q) $-quasihomogeneous terms of higher degree. Notice that $ s \geq 1 $ because $ (x, y) = (0, 0) $ is an isolated real zero of $ F $. Now we consider the expression

    $ ˆF(φ,ρ)=F(ρpcosφ,ρqsinφ)=ρs[Fs(cosφ,sinφ)+O(ρ)],
    $
    (3.6)

    and we observe that $ \hat{G}(\varphi, \rho) = \hat{F}(\varphi, \rho) / \rho^s = 0 $ is an invariant curve $ \hat{\mathcal{X}} $. Thus either

    $ \hat{G}^{-1}(0) \cap \{\rho = 0\} = \emptyset $

    or

    $ \hat{G}^{-1}(0) \cap \{\rho = 0\} \subset \{ (\varphi, \rho) = (\varphi^*, 0) : \varphi^* \in \Omega_{pq} \} $

    by uniqueness of solutions of $ \hat{\mathcal{X}} $. But $ \hat{G}^{-1}(0) \backslash \{\rho = 0\} = \emptyset $ by the monodromy of the polycycle $ \{\rho = 0\} $. Therefore $ \hat{F}(\varphi, \rho) $ has an isolated zero at $ \rho = 0 $ for any $ \varphi \in \mathbb{S}^1 $, and, in particular, the claim follows.

    The second part, that $ \zeta_{pq}(\rho_0) \equiv 0 $, is a trivial consequence of the relation (1.9).

    Proof. The first part is straightforward since $ \mathcal{X}(F) = K F $ holds because $ \mathcal{X}_n $ is a $ (p, q) $-quasihomogeneous polynomial vector field of degree $ n $; hence its components $ P_{p+n} $ and $ Q_{q+n} $ satisfy Euler relations

    $ XE(Pp+n)=(p+n)Pn+p,  XE(Qq+n)=(q+n)Qq+n.
    $
    (3.7)

    The second part is also straightforward since $ \mathcal{X}(V) = {\rm div}(\mathcal{X}) V $ holds, taking into account Euler relations (3.7) and $ \mathcal{X}_E(A) = (m-1) A $, together with the fact that system $ \mathcal{X}_n(H) = 0 $ and $ \mathcal{X}_E(H) = d H $ can be solved as $ \partial_x H = d Q_{q+n} H/F $ and $ \partial_y H = - d P_{p+n} H/F $.

    Proof. In $ (p, q) $-weighted polar coordinates, the angular speed of vector field (1.12) is $ \dot\varphi = G_r(\varphi) = F(\cos\varphi, \sin\varphi)/(p \cos^2\varphi + q \sin^2\varphi) $. Therefore, a monodromic necessary condition for the origin of $ \mathcal{X} $ is that $ F $ has no real factors in $ \mathbb{R}[x, y] $ so that $ G_r \neq 0 $ in $ \mathbb{S}^1 $. In particular, $ F $ has an isolated singularity at the origin, $ \mathcal{X} \in {\rm Mo}^{(p, q)} $ so $ \Pi $ is analytic at the origin, $ \Lambda_{pq} = \emptyset $ and the origin is the unique real finite singularity of $ \mathcal{X} $. We observe that $ \mathcal{X}_n $ has a monodromic singular point at the origin when $ \mathcal{X} $ has it because both vector fields share the same angular speed $ \dot\varphi $. Indeed, the origin is a center of $ \mathcal{X}_n $ under our hypothesis that implies the existence of the first integral $ H $ of Proposition 1.10, see Remark 1.12.

    The monodromy of the origin implies that $ \mathbf{N}(\mathcal{X}) = \mathbf{N}(\mathcal{X}_n) $ has only one edge and that when we only vary the parameters of $ \mathcal{X} $ in the monodromic space $ \Lambda $, then the Newton diagram $ \mathbf{N}(\mathcal{X}|_\Lambda) $ of the restricted vector field $ \mathcal{X}|_\Lambda $ is fixed. The inverse integrating factor

    $ V(\varphi, \rho) = \rho^{m - n} \left(H(\cos\varphi, \sin\varphi)\right)^{(m - n - 1)/d}. $

    The explicit Taylor expansion of $ \Pi $ at the origin is just a consequence of statement (ⅱ) of Theorem 4 in [6]. Finally, we can use Theorem 7 in [6] to conclude that

    $ \eta_{m-n} = \int_0^{2 \pi} \frac{\mathcal{F}(\varphi, r)}{V(\varphi, r)} \, d \varphi = I_1 + r^{n-m+1} \, I_2, $

    where $ I_i $ are integrals independents of $ r $ and

    $ I_1 = \int_0^{2 \pi} \frac{(p \cos^2\varphi + q \sin^2\varphi) A(\cos\varphi, \sin\varphi)}{F(\cos\varphi, \sin\varphi) \left( H(\cos\varphi, \sin\varphi) \right)^{(m-n-1)/d}} \, d \varphi. $

    Since the expression of $ \eta_{m-n} $ must be independent of $ r > 0 $ and sufficiently small, by Theorem 7 in [6], we deduce that $ I_2 = 0 $, and the proposition follows.

    In this paper we have considered planar analytic vector fields $ \mathcal{X} $ having a monodromic singular point with Poincaré map $ \Pi $. Using the fact that always exists a real analytic invariant curve $ F = 0 $ of $ \mathcal{X} $ in a neighborhood of that singularity in the paper are given the relations between $ \Pi $ and $ F $ that can be used to determine new conditions in order to guarantee the analyticity of $ \Pi $ at the singularity.

    The special case when $ F $ is inverse integrating factor of $ \mathcal{X} $ we rediscover the formula obtained previously in [25] by an other method. Finally some applications to the center-focus problem and also to vector fields with degenerate infinity are given.

    All authors carried out the main results of this article, drafted the manuscript, and read and approved the final manuscript. All authors have read and approved the final version of the manuscript for publication.

    The authors declare they have used Artificial Intelligence (AI) tools in the creation of this article.

    The authors are partially supported by the Agencia Estatal de Investigación grant PID2020-113758GB-I00 and an AGAUR (Generalitat de Catalunya) grant number 2021SGR 01618.

    The author is partially supported by a MICIN grant number PID2020-113758GB-I00 and an AGAUR grant number 2017SGR-1276.

    Prof. Jaume Giné is the Guest Editor of special issue "Advances in Qualitative Theory of Differential Equations" for AIMS Mathematics. Prof. Jaume Giné was not involved in the editorial review and the decision to publish this article. The authors declare no conflicts of interest.

    [1] WHO (1948) Preamble to the Constitution of WHO as adopted by the International Health Conference, In: Official Records of the World Health Organization, New York: 19 June–22 July 1946 signed on 22 July 1946 by the representatives of 61 States and entered into force on 7 April 1948, 100.
    [2] Münzel T, Daiber A, Steven S, et al. (2017) Effects of noise on vascular function, oxidative stress, and inflammation: Mechanistic insight from studies in mice. Eur Heart J 38: 2838–2849. doi: 10.1093/eurheartj/ehx081
    [3] Kemppainen LM, Kemppainen TT, Reippainen JA, et al. (2017) Use of complementary and alternative medicine in Europe: Health-related and sociodemographic determinants. Scand J Public Health 2017: 1403494817733869.
    [4] Rice BI (2001) Mind-body interventions. Diabetes Spectrum 14: 213–217. doi: 10.2337/diaspect.14.4.213
    [5] Bazarko D, Cate RA, Azocar F, et al. (2013) The impact of an innovative Mindfulness-Based Stress Reduction Program
on the Health and Well-Being of Nurses Employed in a Corporate Setting. J Workplace Behav Health 28: 107–133.
    [6] Greeson JM (2009) Mindfulness research update: 2008. Complement Health Pract Rev 14: 10–18. doi: 10.1177/1533210108329862
    [7] Barnes PM, Bloom B, Nahin RL (2008) Complementary and alternative medicine use among adults and children: United States, 2007. Natl Health Stat Report 12: 1–23.
    [8] National Center for Complementary and Integrative Health. Available from: https://nccih.nih.gov/health/yoga.
    [9] AAP Section of integrative medicine (2016) Mind-body therapies in children and youth. Pediatr 138: e20161896. doi: 10.1542/peds.2016-1896
    [10] Montgomery GH, Hallquist MN, Schnur JB, et al. (2010) Mediators of a brief hypnosis intervention to control side effects in breast surgery patients: Response expectancies and emotional distress. J Consult Clin Psychol 78: 80–88. doi: 10.1037/a0017392
    [11] The relaxation response. Benson H, Kliper MZ. Avon Books Pub1975;p:1-221
    [12] Mayden KD (2012) Mind-body therapies: Evidence and implications in advanced oncology practice. J Adv Pract Oncol 3: 357–373.
    [13] Piepoli MF, Hoes AW, Agewall S, et al. (2017) 2016 European guidelines on cardiovascular disease prevention in clinical practice. The Sixth Joint Task Force of the European Society of Cardiology and Other Societies on Cardiovascular Disease Prevention in Clinical Practice. Eur Heart J 37: 2315–2381.
    [14] Goff DC Jr, Lloyd-Jones DM, Bennett G, et al. (2014) 2013 ACC/AHA guideline on the assessment of cardiovascular risk: A Report of the American College of Cardiology/American Heart Association Task Force on Practice Guidelines. Circulation 129: S49–S73. doi: 10.1161/01.cir.0000437741.48606.98
    [15] Lee M, Pittler M, Guo R, et al. (2007) Qigong for hypertension: A systematic review of randomized clinical trials. J Hypertens 25: 1525–1532. doi: 10.1097/HJH.0b013e328092ee18
    [16] Guo X, Zhou B, Nishimura T, et al. (2008) Clinical effect of qigong practice on essential hypertension: A meta-analysis of randomized controlled trials. J Altern Complement Med 14: 27–37. doi: 10.1089/acm.2007.7213
    [17] Elliot WJ, Izzo JL, White WB, et al. (2004) Graded blood pressure reduction in hypertensive outpatients associated with use of a device to assist with slow breathing. J Clin Hypertens 6: 553–559. doi: 10.1111/j.1524-6175.2004.03553.x
    [18] Schein MH, Gavish B, Herz M, et al. (2001) Treating hypertension with a device that slows and regularises breathing: A randomised, double-blind controlled study. J Hum Hypertens 15: 271. doi: 10.1038/sj.jhh.1001148
    [19] Anderson JW, Liu C, Kryscio RJ (2008) Blood pressure response to transcendental meditation: A meta-analysis. Am J Hypertens 21: 310–316. doi: 10.1038/ajh.2007.65
    [20] Greenhalgh J, Dickson R, Dundar Y (2010) Biofeedback for hypertension: A systematic review. J Hypertens 28: 644–652. doi: 10.1097/HJH.0b013e3283370e20
    [21] Yucha CB, Clark L, Smith M, et al. (2001) The effect of biofeedback in hypertension. Appl Nurs Res 14: 29–35. doi: 10.1053/apnr.2001.21078
    [22] Nakao M, Yano E, Nomura S, et al. (2003) Blood pressure-lowering effects of biofeedback treatment in hypertension: A meta-analysis of randomized controlled trials. Hypertens Res 26: 37–46. doi: 10.1291/hypres.26.37
    [23] Paul-Labrador M, Polk D, Dwyer JH, et al. (2006) Effects of a randomized controlled trial of transcendental meditation on components of the metabolic syndrome in subjects with coronary heart disease. Arch Intern Med 166: 1218–1224. doi: 10.1001/archinte.166.11.1218
    [24] Khatri D, Mathur KC, Gahlot S, et al. (2007) Effects of yoga and meditation on clinical and biochemical parameters of metabolic syndrome. Diabetes Res Clin Pract 78: 9–10. doi: 10.1016/j.diabres.2007.05.002
    [25] Cohen BE, Chang AA, Grady D, et al. (2008) Restorative yoga in adults with metabolic syndrome: A randomized, controlled pilot trial. Metab Syndr Relat Disord 6: 223–229. doi: 10.1089/met.2008.0016
    [26] Liu X, Miller YD, Burton NW, et al. (2011) Qi-gong mind-body therapy and diabetes control. A randomized controlled trial. Am J Prev Med 41: 152–158.
    [27] Weigensberg MJ, Lane CJ, Ávila Q, et al. (2014) Imagine HEALTH: Results from a randomized pilot lifestyle intervention for obese Latino adolescents using interactive guided imagery SM. BMC Complementary Altern Med 14: 28. doi: 10.1186/1472-6882-14-28
    [28] Younge JO, Gotink RA, Baena CP, et al. (2015) Mind–body practices for patients with cardiac disease: A systematic review and meta-analysis. Eur J Prev Cardiol 22: 1385–1389. doi: 10.1177/2047487314549927
    [29] Krucoff MW, Crater SW, Gallup D, et al. (2005) Music, imagery, touch and prayer as adjuncts to interventional cardiac care: The Monitoring and Actualization of Noetic Trainings (MANTRA) II randomized study. Lancet 366: 211–217. doi: 10.1016/S0140-6736(05)66910-3
    [30] Wang X, Pi Y, Chen P, et al. (2016) Traditional Chinese exercise for cardiovascular diseases: Systematic review and meta-analysis of randomized controlled trials. J Am Heart Assoc 5: e002562. doi: 10.1161/JAHA.115.002562
    [31] Cramer H, Lauche R, Paul A, et al. (2015) Mind–body medicine in the secondary prevention of coronary heart disease: A systematic review and meta-analysis. Dtsch Arztebl Int 112: 759–767.
    [32] Schneider RH, Grim CE, Rainforth MV, et al. (2012) Stress reduction in the secondary prevention of cardiovascular disease: Randomized, controlled trial of transcendental meditation and health education in Blacks. Circ Cardiovasc Qual Outcomes 5: 750–758. doi: 10.1161/CIRCOUTCOMES.112.967406
    [33] Yeh GY, Wood MJ, Lorell BH, et al. (2004) Effects of tai chi mind-body movement therapy on functional status and exercise capacity in patients with chronic heart failure: a randomized controlled trial. Am J Med 117: 541–548. doi: 10.1016/j.amjmed.2004.04.016
  • Reader Comments
  • © 2018 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(6263) PDF downloads(955) Cited by(1)

/

DownLoad:  Full-Size Img  PowerPoint
Return
Return

Catalog