【82】

Download as pdf or txt
Download as pdf or txt
You are on page 1of 23

Physica D 401 (2020) 132172

Contents lists available at ScienceDirect

Physica D
journal homepage: www.elsevier.com/locate/physd

Swirling fluid flow in flexible, expandable elastic tubes: Variational


approach, reductions and integrability

Rossen Ivanov a , Vakhtang Putkaradze b,c ,
a
School of Mathematical Sciences, Technological University Dublin, City Campus, Kevin Street, Dublin, D08 NF82, Ireland
b
Department of Mathematical and Statistical Sciences 632 CAB University of Alberta Edmonton, AB, T6G 2G1, Canada
c
ATCO Space Lab, 5302 Forand St SW, Calgary, AB T3E 7S5, Canada

article info a b s t r a c t

Article history: Many engineering and physiological applications deal with situations when a fluid is moving in flexible
Received 24 May 2019 tubes with elastic walls. In real-life applications like blood flow, a swirl in the fluid often plays an
Received in revised form 25 July 2019 important role, presenting an additional complexity not described by previous theoretical models. We
Accepted 23 August 2019
present a theory for the dynamics of the interaction between elastic tubes and swirling fluid flow.
Available online 29 August 2019
The equations are derived using a variational principle, with the incompressibility constraint of the
Communicated by C. Josserand
fluid giving rise to a pressure-like term. In order to connect this work with the previous literature,
Keywords: we consider the case of inextensible and unshearable tube with a straight centerline. In the absence
Variational methods of vorticity, our model reduces to previous models considered in the literature, yielding the equations
Fluid–structure interaction of conservation of fluid momentum, wall momentum and the fluid volume. We pay special attention
Collapsible tubes carrying fluid to the case when the vorticity is present but kept at a constant value. We show the conservation
Spiral blood flow of energy-like quality and find an additional momentum-like conserved quantity. Next, we develop
Boussinesq equation
an alternative formulation, reducing the system of three conservation equations to a single compact
Monge–Ampère equation
equation for the back-to-labels map. That single equation shows interesting instability in solutions
when the velocity exceeds a critical value. Furthermore, the equation in stable regime can be reduced
to Boussinesq-type, KdV and Monge–Ampère equations in several appropriate limits, namely, the first
two in the limit of a long time and length scales and the third one in the additional limit of the small
cross-sectional area. For the unstable regime, the numerical solutions demonstrate the spontaneous
appearance of large oscillations in the cross-sectional area.
© 2019 Elsevier B.V. All rights reserved.

1. Introduction

Models of flexible tubes conveying fluid have been under investigation for many decades, because of their importance to industrial
and biomedical applications. Broadly speaking, the research can be put in two different categories. The first category of papers focused
on the instability of centerline of the tubes due to the complex interaction of the fluid inside the tube and the tube’s elasticity, most
often treated as an elastic rod. For such systems, an instability appears when the flow rate through the tube exceeds a certain critical
value, sometimes referred to as the garden hose instability. While this phenomenon has been known for a very long time, the quantitative
research in the field started around 1950 [1]. We believe that Benjamin [2,3] was the first to formulate a quantitative theory for the
2D dynamics of the initially straight tubes by considering a linked chain of tubes conveying fluids and using an augmented Hamilton
principle of critical action that takes into account the momentum of the jet leaving the tube. A continuum equation for the linear
disturbances was then derived as the limit of the discrete system. Interestingly, Benjamin’s approach is the inverse to the derivation
of variational integrators for continuum equations [4–6] This linearized equation for the initially straight tubes was further considered
by Gregory and Païdoussis [7].
These initial developments formed the basis for further stability analysis of this problem for finite, initially straight tubes [8–17]. The
linear stability theory has shown a reasonable agreement with experimentally observed onset of the instability [10,18–21]. Nonlinear
deflection models were also considered in [14,22–24], and the compressible (acoustic) effects in the flowing fluid in [25]. Alternatively, a

∗ Corresponding author at: Department of Mathematical and Statistical Sciences 632 CAB University of Alberta Edmonton, AB, T6G 2G1, Canada.
E-mail addresses: [email protected] (R. Ivanov), [email protected] (V. Putkaradze).

https://doi.org/10.1016/j.physd.2019.132172
0167-2789/© 2019 Elsevier B.V. All rights reserved.
2 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

more detailed 3D theory of motion was developed in [26] and extended in [27], based on a modification of the Cosserat rod treatment for
the description of elastic dynamics of the tube, while keeping the cross-section of the tube constant and orthogonal to the centerline. In
particular, [27] analyzes several non-straight configurations, such as tube hanging under the influence of gravity, both from the point of
view of linear stability and nonlinear behavior. Unfortunately, this Cosserat-based theory could not easily incorporate the effects of the
cross-sectional changes in the dynamics. Some authors have treated the instability from the point of view of the follower force approach,
which treats the system as an elastic beam, ignoring the fluid motion, with a force that is always tangent to the end of the tube. Such
a force models the effect of the jet leaving the nozzle [28], showing interesting dynamics of solutions. However, once the length of
the tube becomes large, the validity of the follower force approach has been questioned, see [29] for a lively and thorough discussion.
For the history of the development of this problem in the Soviet/Russian literature, we refer the reader to the monograph [30] (still
only available in Russian). To briefly touch upon the developments in Russian literature that have been published in parallel with their
western counterparts, and perhaps less known in the west, we refer the reader to the selection of papers, with most of them now
available in English [31–40].
It is also worth noting the developments in the theory of initially curved pipes conveying fluid, which has been considered in some
detail in the literature. The equations of motion for such theory were initially derived using the balance of elastic forces from tube’s
deformation and fluid forces acting on the tube when the fluid is moving along a curved line in space. In the western literature, we
shall mention the earlier work [41], followed with more detailed studies [42–44] which developed the theory suited for both extensible
and inextensible tubes and discussed the finite-element method realization of the problem. We shall also mention [45,46] deriving a
variational approach for the planar motions of initially circular tubes, although the effect of curved fluid motion was still introduced
as extra forces through the Lagrange–d’Alembert principle. In the Soviet/Russian literature, [36] developed the rod-based theory of
oscillations and [39] considered an improved treatment of forces acting on the tubes. Most of the work has been geared towards the
understanding of the planar cases with in-plane vibrations as the simplest and most practically relevant situations (still, however,
leading to quite complex formulas).
In spite of substantial amount of literature in previous work in the area, many questions remain difficult to answer using the
traditional approach. Most importantly, it is very difficult (and perhaps impossible) to extend the previous theory to accurately take
into account the changes in the cross-sectional area of the tube, also called the collapsible tube case. In many previous works, the
effects of cross-sectional changes have been considered through the quasi-static approximation: if A(s, t) is the local cross-section area,
and u(s, t) is the local velocity of the fluid, with s being the coordinate along the tube and t the time, then the quasi-static assumption
states that uA = const, see, e.g., [22,47]. Unfortunately, this simple law is not correct in general and should only be used for steady
flows. This problem has been solved in the fully variational derivation [48,49], where a geometrically exact setting for dealing with a
variable cross-section depending on tube’s deformation was developed and studied, showing the important effects of the cross-sectional
changes on both linear and nonlinear dynamics. The nonlinear theory was derived from a variational principle in a rigorous geometric
setting and for general Lagrangians. It can incorporate general boundary conditions and arbitrary deviations from equilibrium in the
three-dimensional space. From a mathematical point of view, the Lagrangian description of these systems involves both left-invariant
(elastic) and right-invariant (fluid) quantities. This method has since been applied to the linear stability of initially curved (helical) tubes
in [50], where it was shown that the geometric method of [48,49] is guaranteed to yield an equation with constant coefficients for the
linear stability analysis for any helical initial state. This theory further allowed consistent variational approximations of the solutions,
both from the point of view of deriving simplified reduced models and developing structure preserving numerical schemes [51], echoing
the original development of the equations by Benjamin [2,3] for constant cross-section.
The second category of papers concerns with the flow of fluid through tubes which can dynamically change its cross-section. In these
works, the cross-section, or its radius, in the case of circular cross-sections, serves as an additional dynamic variable. The equations of
motion then involve the balance of fluid momentum, wall momentum and conservation law for the fluid. The theory for such tubes is
indispensable in biomedical applications, such as arterial flows, and the key initial progress in theoretical understanding is based on
balancing the conservation laws in a collapsible tube, see [52]. Further work in this field explored applications to arterial [53–59] and
lung flows [60–62]. Analytical studies for such flows are usually limited to cases when the centerline of the tube is straight. In addition,
for analytical progress, one needs to further assume that material particles on the wall can only move normally to the centerline during
the dynamics (see, however, [63] for non-trivial longitudinal displacement). We should also mention the instability through the neck
formation and self-sustaining flow pulsations suggested by [64] and experimentally measured in [65], the treatment of forward and
backward running waves in arteries [66,67] and the treatment of poroelastic walls [68]. While substantial progress in the analysis
of the flow has been achieved so far, it was difficult to describe analytically the general dynamics of 3D deformations of the tube
involving e.g. the combination of shear, transversal deformation, and extensions. A lively discussion of the dichotomy between the
needs of numerical solutions of Navier–Stokes equations in realistic geometries, and the importance of theoretical developments based
on simplified models can be found in [69]. For more informations about the application of collapsible tubes to biological flows, we refer
the reader to the reviews [70–72].
A fully variational theory of motion of tubes with expandable walls in three dimensions was derived in [73], both for compressible
and incompressible flow motion inside the tube. The major difference with the earlier models described by the variational approach [48,
49] was the ability to treat compressible flows, and incorporate the coupled dynamics of the walls, the tube and the fluid. This model can
treat arbitrary motion of the tubes in three dimensions, using arbitrary elasticities and mass distribution of the walls. For compressible
flows, the model is also capable to incorporate the motion of shock waves in moving tubes with expandable walls. It was shown that
all simplified analytical models of flows with expandable walls can be obtained as particular cases of the more generalized model,
with the introduction of appropriate friction terms. As far as we are aware, there is no way to achieve such generality of motion
with non-variational models. Indeed, one can guess the right terms in simplified geometries using force balance, but trying to balance
forces in more complex problems, such as deforming centerline and expandable walls, is fraught with possibilities for errors. In contrast,
variational methods provide a systematic way of deriving the equations and their subsequent analysis, such as Hamiltonian formulation
and the existence of Poisson brackets. We thus believe that the problems in fluid–structure interactions can benefit greatly from the
application of modern variational methods and geometric mechanics.
Most of the analytical aspects of previous works discussed above have treated the flow of fluid in a simplified manner, treating
one characteristic streamwise velocity for each point of the centerline. However, the fluid flow regimes with vorticity are frequently
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 3

encountered around stents in arteries, have been observed both experimentally and numerically. They are also known as ‘spiral flows’,
or ’spiral laminar flows’, and are believed to play a crucial role in arterial dynamics. The study of these flows with applications to
blood flow dynamics started with the pioneering observations by Stonebridge & Drophy [74] and continued by many authors [75–80].
Numerical studies of vorticity transport in complex geometries relevant to arteries can be found in e.g. [81], but as far as we are aware,
there have been no theoretical studies including the wall dynamics, fluid dynamics and vorticity, especially from the variational point
of view. The goal of this paper is precisely to fill this theoretical gap. More precisely, we shall derive a fully three-dimensional flow of
mean vorticity transport based on Kelvin’s theorem of circulation in fluid mechanics. We will mostly focus on the variational derivation,
and thus most of our considerations will be dedicated to the inviscid models. The friction terms essential for blood flow applications
can be added later as they have to be incorporated in the variational principle using the Lagrange–d’Alembert’s method of external
forces. Our model will be valid for arbitrary three-dimensional deformations of the centerline and arbitrary elasticity laws satisfied by
the wall. As we show, the simplified mathematical treatment for the flow can produce interesting mechanical effects and instability
caused by the interaction of wall and fluid. As far as practical applications of the theory presented here, we view it as a relatively
complete model of industrial flows in flexible and expandable tubes with high vorticity, and a first step towards understanding the
complete structure of inertial terms in arterial models, to which terms containing the internal viscous friction and wall shear can later
be added.
A major role in this manuscript will be played by the back-to-label Lagrangian map for the fluid part, which maps the location of fluid
particle at a given point on the tube at the time t to its original location at t = 0. The back-to-labels map has an important role in the
development of continuum mechanics. Of particular value to our studies here is the Clebsch representation of back-to-labels map which
was used to derive variational and Hamiltonian structure of continuum mechanics. One of the pioneering examples of this approach
was developed in [82], where Poisson brackets and Clebsch representation for fluid mechanics, magnetohydrodynamics (MHD), and
elasticity theory were derived. Nonlinear stability theory of fluid flows using back-to-labels maps was investigated in [83,84], and
applications to quasi-geostrophic flows was studied in [85]. A multi-symplectic formulation of fluid mechanics for ideal fluids using
back-to-labels maps was further derived in [86]. In these papers, variational derivatives with respect to back-to-labels map was crucial
in obtaining the laws of motion and proving the Hamiltonian structure of the flow. Because of the features of 3D hydrodynamics,
the expression of the incompressibility condition in terms of back-to-labels map is rather awkward, and thus the works cited here
used variations with constrained Lagrangian and needed the pressure as the corresponding Lagrange multiplier. Our problem has the
advantage that one can explicitly connect the cross-sectional area and the spatial derivatives of back-to-labels map, and substitute into
the Lagrangian directly, leading to the unconstrained Euler–Lagrange equations.
Structure of the paper and main results. The paper is structured as follows. In Section 2, we derive the full three-dimensional equations
of motion for flexible tubes carrying incompressible fluid which has streamwise velocity and vorticity. We have tried to keep the
derivation self-contained and pedagogical, and build up our theory from the theory of exact geometric rods with no fluid, to the
theory of fluid-conveying tubes with expandable walls, introducing the swirling motion of the fluid in the process. The fluid volume
incompressibility condition leads to the appearance of a pressure-like term in the equations of motion. In Section 3, we consider the
case of the inextensible and unshearable tube, and a static straight centerline when the walls can only move normal to the centerline,
which is the case often considered in the literature for theoretical studies. The equations of motion then reduce to a generalization
of the equations familiar from the previous literature, with four equations of motion describing the conservation of the fluid and wall
momenta and the fluid’s mass, and the transport of vorticity (or, more precisely, velocity circulation), with the corrections due to the
vorticity terms. These equations explicitly contain the fluid pressure as the Lagrange multiplier for incompressibility. We show the
conservation of energy for these equations and also derive a new constant of motion linear in momenta in (3.9), that seems to have not
been observed in the previous works on the subject. By using an alternative derivation using the Lagrangian back-to-labels map, we
show how to reduce this equation to a single nonlinear scalar equation of motion in Section 3.3. In the linearized form, the equation
shows interesting stability properties, in particular, instability for a sufficiently large velocity in the tube. We perform the numerical
studies of the solution in both linearly stable and unstable regimes. For the linearly stable regime, in Section 4 we utilize the slow-time
and large-wavelength approximation to first develop a reduction to a Monge–Ampére’s equation, then a Boussinesq-like equation and
finally the KdV equation. We compare the numerics of the approximate models with the full equations and discuss the relevance of
KdV approximations for the left- and right-running waves.

2. Exact geometric theory for flexible tubes conveying incompressible fluid

In this section, we first quickly review the Lagrangian variational formulation for geometrically exact rods without fluid motion
as the foundation of the theory. Then, we extend this variational formulation to incorporate the motion of the incompressible fluid
inside the tube and the motion of the wall, and finish by introducing the vorticity. To achieve this goal we need to first identify the
configuration manifold of the system (which is infinite-dimensional), as well as the convective and spatial variables for the tube and
the fluid. Our model is then obtained by an application of the Hamilton principle, reformulated in convective variables for the tube and
in spatial variables for the fluid. For rigorous justification of the variational approach to fluid–structure interactions see [49,73].

2.1. Background for geometrically exact rod theory

Here we briefly review the theory of geometrically exact rods following the approach developed, on the Hamiltonian side, in [87] and
subsequently in the Lagrangian framework more appropriate to this article in [88,89]. A more comprehensive introduction is contained
in [49,50] to which we refer the reader for the details. The purely elastic (i.e., rods carrying no fluid) geometrically exact theory is
equivalent to the Cosserat’s rods [90], the equivalence of these approaches was shown in [87,89].
The configuration of the rod deforming in the ambient space R3 is defined by specifying the position of its line of centroids by means
of a map r(t , s) ∈ R3 , and by giving the orientation of the cross-section at that point. Here t is the time and s ∈ [0, L] is a parameter
along the strand that does not need to be arclength. The orientation of the cross-section is given by a moving basis {ei (t , s) | i = 1, 2, 3}
attached to the cross section relative to a fixed frame {Ei | i = 1, 2, 3}. The moving basis is described by an orthogonal transformation
4 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

Λ(t , s) ∈ SO(3) such that ei (t , s) = Λ(t , s)Ei . We interpret the maps Λ(t , s) and r(t , s) as a curve t ↦→ (Λ(t), r(t)) ∈ G in the infinite
dimensional Lie group G = F ([0, L], SO(3)×R3 ) of SO(3)×R3 -valued smooth maps defined on [0, L].1 The Lie group G is the configuration
manifold for the geometrically exact rod.
Following Hamilton’s principle, given a Lagrangian function

L = L(Λ, Λ̇, r, ṙ) : TG → R,

defined on the tangent bundle TG of the configuration Lie group G defined above, the equations of motion are the Euler–Lagrange
equations obtained by the critical action principle
∫ T
δ L(Λ, Λ̇, r, ṙ)dt = 0, (2.1)
0

for arbitrary variations δ Λ and δ r vanishing at t = 0, T . It turns out that the Lagrangian of geometrically exact rods can be exclusively
expressed in terms of the convective variables

γ = Λ−1 ṙ , ω = Λ−1 Λ̇ ,
(2.2)
Γ = Λ−1 r′ , Ω = Λ−1 Λ′ ,
see [87], where γ (t), ω(t) ∈ F ([0, L], R3 ) are the linear and angular convective velocities and Γ (t), Ω (t) ∈ F ([0, L], R3 ) are the linear
and angular convective strains. This gives rise to a Lagrangian ℓ = ℓ(ω, γ, Ω, Γ ) : F ([0, L], R3 )4 → R written exclusively in terms of
convective variables. Here, we treat all four variables as infinite-dimensional functions of t, which are themselves R3 -valued functions
of s for any fixed t, see [89] for the detailed mathematical exposition of the method. For the moment, we leave the Lagrangian function
unspecified, we will give its explicit expression later in Section 2.3 for the case of fluid-conveying tubes.
The equations of motion in convective description are obtained by writing the critical action principle (2.1) in terms of the Lagrangian
ℓ. This is accomplished by computing the constrained variations of ω, γ, Ω, Γ induced by the free variations δ Λ, δ r via the definitions
(2.2). We find
∂Σ ∂Ψ
δω = + ω × Σ, δγ = +γ ×Σ +ω×Ψ, (2.3)
∂t ∂t
∂Σ ∂Ψ
δΩ = + Ω × Σ, δΓ = + Γ × Σ + Ω × Ψ, (2.4)
∂s ∂s
where Σ (t , s) = Λ(t , s)−1 δ Λ(t , s) ∈ R3 and Ψ (t , s) = Λ(t , s)−1 δ r(t , s) ∈ R3 are arbitrary functions vanishing at t = 0, T . Hamilton’s
principle (2.1) induces the variational principle
∫ T
δ ℓ(ω, γ, Ω, Γ )dt = 0, (2.5)
0

with respect to the constrained variations (2.3), (2.4), which in turn yields the reduced Euler–Lagrange equations
δℓ δℓ δℓ δℓ

⎨ (∂t + ω×)
⎪ +γ × + (∂s + Ω×) +Γ × =0
δω δγ δΩ δΓ (2.6)
δℓ δℓ
⎩ (∂t + ω×)
⎪ + (∂s + Ω×) = 0,
δγ δΓ
together with the boundary conditions
δℓ ⏐⏐ δℓ ⏐⏐
⏐ ⏐
= 0, = 0. (2.7)
δΩ ⏐s=0,L δΓ ⏐s=0,L
If one of the extremity (say s = 0) of the rod is kept fixed, i.e., r(t , 0) = r0 , Λ(t , 0) = Λ0 for all t, then only the boundary condition at
s = L arises above.
From their definition (2.2), the convective variables verify the compatibility conditions

∂t Ω = ω × Ω + ∂s ω and ∂t Γ + ω × Γ = ∂s γ + Ω × γ. (2.8)

Remark 2.1 (Lagrangian Reduction by Symmetry). The process of passing from the Lagrangian (or material) representation in terms
of (Λ, Λ̇, r, ṙ) with variational principle (2.1) to the convective representation in terms of (ω, γ, Ω, Γ ) with constrained variational
principle (2.5) can be understood via a Lagrangian reduction process by symmetries. It has been carried out in [89] and is based on the
affine Euler–Poincaré reduction theory of [91].

Remark 2.2 (On the Functional Form of the Lagrangian). Note that in all our theoretical considerations we will keep the Lagrangian in
the general form, as we are interested in the symmetry-reduction approach to the fully three dimensional problem, rather than in the
derivation of the equations of motion in a particular reduced setting, e.g., restricted to two dimensions, straight line, etc. We believe that
such an approach based on Lagrangian mechanics yields the simplest possible treatment of the elastic, three-dimensional deformation
of the tube.

1 The notation F ([0, L], V ) refers to the set of functions defined on the interval s ∈ [0, L] taking the values in the set V .
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 5

2.2. Definition of the configuration space for the tube with expandable walls conveying fluid

We now incorporate the motion of the fluid inside the tube, the vorticity and the motion of the wall of the tube by extending the
geometrically exact framework. Recall that the geometrically exact rod (without fluid) consists of a left invariant system and, therefore,
can be written in terms of convective variables. On the other hand, the fluid is a right invariant system, naturally written in terms
of spatial variables. The coupling of these two systems therefore yields the interesting combined fluid–structure interaction involving
both convective and spatial variables but whose left and right invariances are broken by the coupling constraint. Without the vorticity,
our derivation in this chapter is equivalent to [73], see also [92] for a more detailed and pedagogical exposition.
In addition to the rod variables (Λ, r) ∈ F ([0, L], SO(3) × R3 ) considered above, the configuration manifold for the fluid-conveying
tube also contains the Lagrangian description of the fluid. It is easier to start by defining the back-to-label map, which is an embedding
ψ : [0, L] → R, assigning to a current fluid label particle s ∈ [0, L] located at r(s) in the tube, its Lagrangian label s0 ∈ R. Its inverse
ϕ := ψ −1 gives the current configuration of the fluid in the tube. A time dependent curve of such maps thus describes the fluid motion
in the tube, i.e.,
s = ϕ (t , s0 ), s ∈ [0, L].
We now include the motion of the wall of the tube, as a reaction to the fluid motion and pressure. In order to incorporate this effect
in the simplest possible case of a tube with a circular cross-section, let us consider the tube radius R(t , s) to be a free variable. In this
case, the Lagrangian depends on R, as well as on its time and space derivatives Ṙ and R′ , respectively. Let us assume that R can lie on
an interval IR , for example, IR = R+ (the set of positive numbers). As we will see below, the presence of the swirl will add an additional
scalar function to the variables of the Lagrangian, namely, the circulation of fluid velocity measured at a given cross-section. Then, the
configuration manifold for the fluid-conveying tube is given by the infinite dimensional manifold
Q := F [0, L], SO(3) × R3 × IR × R
( )
(2.9)
× ϕ : ϕ −1 [0, L] → [0, L] | ϕ diffeomorphism .
{ }

Note that the domain of definition of the fluid motion s = ϕ (t , s0 ) is time dependent, i.e., we have ϕ (t) : [a(t), b(t)] → [0, L], for
ϕ (t , a(t)) = 0 and ϕ (t , b(t)) = L. The time dependent interval [a(t), b(t)] contains the labels of all the fluid particles that are present in
the tube at time t.

2.3. Definition of the Lagrangian

Let us now turn our attention to the derivation of the Lagrangian of the fluid-conveying geometrically exact tube with expandable
walls. The Lagrangian will be computed as the sum of kinetic energy of rod and fluid and negative of the potential energy of the rod’s
deformation.
Kinetic energy of the rod. The kinetic energy of the elastic rod is the function Krod given by
∫ L
1
α|γ|2 + aṘ2 + I(R)ω · ω |Γ |ds,
( )
Krod =
2 0

where α is the linear density of the tube and I(R) is the local moment of inertia of the tube. The term 12 aṘ2 describes the kinetic energy
of the radial motion of the tube.
We now derive the total kinetic energy of the fluid. In material representation, the total velocity of the fluid particle with label s0
is given by
d
r(t , ϕ (t , s0 )) = ∂t r(t , ϕ (t , s0 )) + ∂s r(t , ϕ (t , s0 ))∂t ϕ (t , s0 )
dt (2.10)
= ∂t r(t , ϕ (t , s0 )) + ∂s r(t , ϕ (t , s0 ))u(t , ϕ (t , s0 )),
where the Eulerian velocity is defined by
u(t , s) = ∂t ϕ ◦ ϕ −1 (t , s), s ∈ [0, L].
( )
(2.11)
Therefore, the kinetic energy of the translational motion of the fluid reads
ϕ −1 (L,t)
∫ ⏐ ⏐2
1 ⏐d
ρ Q0 (s0 ) ⏐ r(t , ϕ (t , s0 ))⏐⏐ ds0 ,

Kfluid,transl = ⏐
2 ϕ −1 (0,t) dt
where Q0 (s0 )ds0 is the initial infinitesimal volume of fluid.
Circulation of velocity and Kelvin’s theorem. In order to go beyond the theory derived in [73], we introduce the swirling motion in the
Lagrangian. We will introduce the simplest model of swirling. We assume that the shape of the tube at cross-section is approximately
circular, and there is an axisymmetric swirling motion in addition to the streamline velocity u(s, t). The trajectories of the fluid particles
are locally helical, although they could be quite complex globally, i.e. extended for long t.
Take Cα to be a contour that circles a given cross-section close to the boundary at a Lagrangian point α . Assuming that the
deformations of the centerline remain small, the cross-section roughly circular, and the viscosity can be neglected. the cross-section
remains normal to the centerline. Under this assumption, C (t) will always be close to the boundary of a cross-section at all other times
t at the point s = ϕ (α, t). For an inviscid incompressible fluid considered here, Kelvin’s theorem states that the circulation λ(s, t) is
conserved along the flow, in other words, for a three-dimensional fluid velocity v,


= (∂t + u∂s )λ = 0 , λ := v · dl = 0 , (2.12)
Dt C (t)
6 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

where D/Dt is the full derivative of the integral quantity. A more general version of Kelvin’s theorem uses the fluid momentum instead
of [93]. In what follows we shall use λ as an additional dynamic quantity in the Lagrangian to model the swirling motion of the flow.
We thus obtain from (2.12) the following approximate conservation law for the swirling motion.

(∂t + u∂s )λ = 0 ⇔ λ = λ0 ◦ ϕ −1 (s, t) . (2.13)

Remark 2.3 (On the Description of Swirling Motion Through Vorticity). One could be tempted to introduce a ‘typical’ vorticity in the
streamwise direction ξ (s, t) instead of circulation λ(s, t) introduced above. Then, the vorticity advection written in the local coordinates
will state

∂t ξ + u∂s ξ − ξ ∂s u = 0 . (2.14)

If the velocity field is assumed to be smooth in every cross-section, then we can write λ ≃ ξ Q . Differentiating λ defined that way, we
obtain exactly (2.13).
While this calculation is formally correct, it is difficult to justify it physically, since the swirling motion in inviscid fluids may have a
singularity in the velocities at the centerline of the tube. Thus, the concept of a ‘typical’ vorticity for a cross-section is ill-defined even
for the simplest problem of a perfectly cylindrical straight tube with the flow independent of s. In that case, for every cross-section, the
problem reduces to the problem of the motion of a point vortex in a circle inviscid fluid, and a point vortex at the center with a given
circulation value is a solution [94,95]. However, Kelvin’s theorem is still valid, even when the vorticity is infinite at a set of points, as
long as the circulation of the vorticity is finite. Thus, we believe that the circulation λ(s, t) defined by (2.12) with the evolution equation
given by (2.13) is superior characterization of swirl as compared to the notion of a ‘typical’ vorticity. Note that physical cut-offs must
be introduced to utilize the notion of the kinetic energy, as, technically speaking, the kinetic energy of the fluid driven by a point vortex
is infinite.

It is worth noting that when the viscosity is present, the conservation of circulation is no longer valid. Indeed, the rotational velocity
component will be diminished as the fluid is propagating along the tube because of the friction forces. If the radius of the tube is R,
and the kinematic viscosity of fluid is ν , then the rate of change of the typical rotational component of velocity Vrot is given by
dVrot ν
≃− Vrot . (2.15)
dt R2
If the typical axial velocity of the fluid is U, then after traveling the distance L, from (2.15) the axis component will diminish by a factor
given by

− ν2L −L 1
Vrot (s = L, t) ≃ Vrot (s = 0, t)e R U = Vrot (s = 0, t)e R ReR (2.16)

where ReR is the Reynolds number based on the cross-sectional radius of the tube R. Thus, for the decay of the azimuthal velocity to
be negligible, Reynolds number ReR has to be much larger than the tube’s aspect ratio L/R.
Even when the vorticity is sufficiently small, but not negligible, one has to be careful interpreting the notion of circulation. Indeed,
because of the no-slip condition at the boundary, the circulation on the contour equal to the boundary is zero. Typically, in the limit
of small vorticity, the viscous effects will take place close to the wall in the narrow boundary layer. We thus take the contour used in
circulation to be sufficiently far from the walls to avoid the boundary layer effects, and yet sufficiently large to encode the information
about the kinetic energy. For example, the circulation contour can be taken to be at r = R/2 at the inlet.
In addition, it has been shown [96] that even for finite viscosity, there is a direct correspondence between the Navier–Stokes
equations and Euler’s equation taken as a stochastic process. There is also the corresponding conservation of circulation, which is
exact, when taken in the stochastic sense. Thus, we believe that even when the flow becomes turbulent, the circulation variable λ(s, t)
introduced here will capture enough physics to describe the rotational motion of fluid.
The rotational component of kinetic energy can be then approximated by integrating the rotational component of velocity over a
cross-sectional area, and then over the whole tube. Note that dimensionally, λ2 ≃ R2 vrot 2
, so ρλ2 already has the dimensions of the
kinetic energy of rotation per unit length of the tube. The exact value of the kinetic energy depends on the velocity profile of the radial
velocity as a function of r, namely, vrot (r , s, t). In general, this velocity profile is a complex function of the tube’s shape and the local
vorticity profile of the fluid. We thus define the profile shape function for every cross-section Cs
∫R
2π ρ r vrot
2
dr
Φ= 0
|Γ | , Φ = Φ (Ω, Γ , R, R′ , . . .). (2.17)
λ(s, t)2
In general, the shape function Φ is dimensionless with values of order 1. It describes the deviation of the velocity profile from its
given equilibrium value during the dynamics. The profile shape function Φ can only be obtained through either numerical simulations or
experiments, and is thus a modeling component of the theory. Our equations of motion will be derived by the variational principle and
will be valid for any Φ . While the exact nature of Φ and its dependence on the flow is to be investigated in future work, although some
conclusions can be drawn from simple considerations. For example, for a point vortex on the plane with circulation λ, the rotational
velocity is vrot = 2πλ r , so the kinetic energy of the fluid’s rotation in a circle of radius R is infinite.2 However, if we introduce a smoothing
cut-off of radius ϵ around the vortex attributed to viscosity, we arrive to
∫ ∫ R
1 1 R
Φ= 2π ρ r vrot
2
dr = log (2.18)
λ2 ϵ 2π ϵ

2 Here, v is the true rotational component of velocity which is a function of (r , s, t). In contrast, the notation V used above is the typical, or averaged, value
rot rot
of the rotational component of velocity for every cross-section, and is thus a function of (s, t). We hope no confusion will arise from this notation.
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 7

Fig. 2.1. Left panel: a sketch of conservation law (2.21). Right panel: a tube (shown in red) and particle trajectories (shown in blue). (For interpretation of the
references to color in this figure legend, the reader is referred to the web version of this article.)

In the numerical simulations in this paper we will use (2.18) for a fixed value of ϵ . Then,
π

Kfluid,rot = ρ Φ λ2 ds . (2.19)
2
Using (2.10) together with the change of variables s = ϕ (t , s0 ), we can rewrite Kfluid as
∫ L
1
Kfluid = ρ Q |γ + Γ u|2 + ρ Φ λ2 ds,
2 0

where ρ is the mass density of the fluid per unit volume, in units Mass/Length3 , and Q (t , s) is the area of the tube’s cross section, in
units Length2 .

Elastic energy. The potential energy due to elastic deformation is a function of Ω , Γ and R. While the equations will be derived for an
arbitrary potential energy, we shall assume the simplest possible quadratic expression for the calculations, namely,
∫ L(
1 )
Erod = JΩ ·Ω + λ(R)|Γ − χ|2 + 2F (R, R′ , R′′ ) |Γ |ds , (2.20)
2 0

where χ ∈ R3 is a fixed vector denoting the axis of the tube in the reference configuration, J is a symmetric positive definite 3 × 3
matrix, which may depend on R, R′ and R′′ , and λ(R) is the stretching rigidity of the tube. The stretching term proportional to λ(R)
can take the more general form K(Γ − χ) · (Γ − χ), where K is a 3 × 3 tensor. The part of this expression for the elastic energy
containing the first two terms in (2.20) is commonly used for a Cosserat elastic rod, but more general functions of deformations Γ are
possible, see [50] for a more detailed discussion of possible forms of the potential energy. A particular case is a quadratic function of Γ
leading to a linear dependence between stresses and strains. We have also introduced the elastic energy of wall F (R, R′ , R′′ ) which can
be explicitly computed for simple elastic tubes. In general F depends on higher derivatives, such as R′′ . We shall derive the equations
of motion in their general form for arbitrary F , but for particular examples used in simulations we will use F (R, R′ ) for simplicity.

Mass conservation. Before we write the final expression for the Lagrangian, let us discuss the question of the mass conservation since
it will be used as a constraint to derive the equations of motion, with the appropriate Lagrange multiplier playing the role of fluid’s
pressure. We shall assume that the fluid fills the tube completely, and the fluid velocity at each given cross-section is aligned with the
axis of the tube. Since we are assuming a one-dimensional approximation for the fluid motion inside the tube, the mass density per
unit length ρ Q = ρ A|Γ |, has to verify

ρ Q = (ρ Q0 ◦ ϕ −1 )∂s ϕ −1 . (2.21)

Since ρ = const, and the fluid is incompressible, we deduce the conservation law

Q = Q0 ◦ ϕ −1 ∂s ϕ −1 ⇒ ∂t Q + ∂s (Qu) = 0. (2.22)

The origin of this conservation law is illustrated on the left panel of Fig. 2.1. A Lagrangian point a is mapped to s = ϕ (a, t), and
the Lagrangian point a′ is mapped to s′ = ϕ (a′ , t). Inversely, a point on a centerline s will be mapped to the initial Lagrangian point
a = ϕ −1 (s, t). Taking a′ = a + da, we see that the infinitesimal physical volume is described at the physical space by Q ds, and the
corresponding volume at t = 0 by Q0 (a)∂s ϕ −1 (s, t), with a = ϕ −1 (s, t), giving exactly (2.22).
The key new feature of this Section is in the introduction of the swirling motion in the kinetic energy part of the Lagrangian. The
particle trajectories for such a swirling motion are approximately helical, as illustrated by a sketch on the right panel of Fig. 2.1.

Lagrangian. From all the expressions given above, we obtain the Lagrangian of the fluid-conveying given by

L = L Λ, Λ̇, r, ṙ, ϕ, ϕ̇, R, Ṙ : T Q → R, L = Krod + Kfluid − Erod − Eint ,


( )
(2.23)

and defined on the tangent bundle T Q of the configuration space Q, see (2.9). Note that all the arguments of L are functions of s, so
we do not need to include the spatial derivatives of (Λ, r, R) explicitly as variables in L. These spatial derivatives appear explicitly in
8 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

the expression of the integrand of the reduced Lagrangian (2.24) below. Assuming there is a uniform external pressure pext acting on
the tube, the Lagrangian expressed in terms of the variables ω, γ, Ω, Γ , u, Q , λ reads

ℓ(ω, γ, Ω, Γ , u, Q , λ, R, Ṙ)
∫ L [(
1 1 1 1
= α|γ|2 + I(R)ω·ω + aṘ2 − F (R, R′ ) − JΩ ·Ω
0 2 2 2 2
1 ) 1
− S(R)|Γ − χ| 2
|Γ | + ρ Q |γ + Γ u| 2
(2.24)
2 2
1 ]
+ ρ Φ λ2 s − pext Q ds
2
∫ L [ ]
=: ℓ0 (ω, γ, Ω, Γ , u, Q , R, Ṙ, R′ , λ) − pext Q ds .
0

where S(R) is the function describing the potential energy of tube’s stretching along its centerline as a function of R. Our derivation
of the equations of motion (2.28) below will be valid for the Lagrangian ℓ being an arbitrary function of its variables. Of course, the
numerical simulations will require specifying the exact functional form of the Lagrangian ℓ for every particular case.

2.4. Variational principle and equations of motion

The equations of motion are obtained from the Hamilton principle applied to the Lagrangian (2.23), namely
∫ T
δ L(Λ, Λ̇, r, ṙ, ϕ, ϕ̇, λ, R, Ṙ)dt = 0, (2.25)
0

for arbitrary variations δ Λ, δ r, δϕ, δ R vanishing at t = 0, T . In terms of the symmetry reduced Lagrangian ℓ, this variational principle
becomes
∫ T
δ ℓ(ω, γ, Ω, Γ , u, ξ , λ, R, Ṙ)dt = 0 , (2.26)
0

for variations (2.3), (2.4), and δ u, δξ and δλ computed as

δ u = ∂t η + u∂s η − η∂s u δλ = −η∂s λ, δ (Q0 ◦ ϕ −1 ∂s ϕ −1 ) = −∂s (ηQ ) (2.27)

where η = δϕ ◦ ϕ −1 . Note that η(t , s) is an arbitrary function vanishing at t = 0, T . Note that the Lagrangian function in (2.26) will
contain the incompressibility constraint (2.22) with the Lagrange multiplier p, which we drop from the notation in (2.26) for simplicity.
A lengthy computation (see [73] for details) yields the system

D δℓ δℓ D δℓ δℓ

⎪ +γ ×
+ +Γ × =0
Dt δω δγ Ds δΩ δΓ




D δℓ D δℓ



+ =0


Dt δγ Ds δΓ



δℓ δℓ δℓ δℓ δℓ



∂t + u∂s + 2 ∂s u + ∂s λ = Q ∂s

δu δu δu δλ δQ (2.28)
δℓ δℓ


∂t

⎪ − =0
δ Ṙ δ R






∂t Ω = Ω × ω + ∂s ω, ∂t Γ + ω × Γ = ∂s γ + Ω × γ






∂t Q + ∂s (Qu) = 0, ∂t λ + u∂s λ = 0,

where the symbols δℓ/δω, δℓ/δΓ , . . . denote the functional derivatives of ℓ relative to the L2 pairing, and we introduced the notations
D D
= ∂t + ω × and = ∂s + Ω × .
Dt Ds
Note that the first equation arises from the terms proportional to Σ in the variation of the action functional and thus describes the
conservation of angular momentum. The second equation arises from the terms proportional to ψ and describes the conservation
of linear momentum. We remind the reader that the quantities Σ and ψ are defined by Σ = (ΛT δ Λ)∨ and ψ = ΛT δ r, with
(Σ (s, t), ψ (s, t)) for fixed s and t taking values in the Lie algebra se(3).
The third equation is obtained by collecting the terms proportional to η and describes the conservation of fluid momentum. The
fourth equation comes from collecting the terms proportional to δ R and describes the elastic deformation of the walls due to the
pressure. Finally, the last four equations arise from the four definitions Ω = (Λ−1 Λs )∨ , Γ = Λ−1 r′ , Q = (Q0 ◦ ϕ −1 )∂s ϕ −1 , and
λ = λ0 ◦ ϕ −1 .
We will derive the equations for general Q = Q (Ω, Γ , R), keeping in mind the fully three-dimensional dynamics, but will only use
it for the tubes satisfying Q = π R2 |Γ | which is a good approximation for mostly straight tubes. For the choice
∫ L[ ]
ℓ= ℓ0 (ω, γ, Ω, Γ , u, λ, R, Ṙ, R′ ) − p(Q − Q0 ◦ ϕ −1 ∂s ϕ −1 ) − pext Q ds ,
0
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 9

the functional derivatives are computed as


δℓ ∂ℓ0 ∂Q
= + (p − pext )
δΩ ∂Ω ∂Ω
δℓ ∂ℓ0 ∂Q
= + (p − pext )
δΓ ∂Γ ∂Γ (2.29)
δℓ ∂ℓ0 ∂ℓ0 ∂ℓ0 ∂Q
= − ∂s ′ + ∂s2 ′′ + (p − pext )
δR ∂R ∂R ∂R ∂R
δℓ ∂ℓ0
= − (p − pext ),
δQ ∂Q
where ∂ℓ0 /∂Ω, ∂ℓ0 /∂Γ , . . . denote the ordinary partial derivatives of ℓ0 , whose explicit form can be directly computed from the
expression of ℓ0 in (2.24).

Theorem 2.4. For the Lagrangian ℓ in (2.24), the variational principle (2.26) with constrained variations (2.3), (2.4), (2.27) yields the
equations of motion
D ∂ℓ0 ∂ℓ0 D ∂ℓ0 ∂Q ∂ℓ0 ∂Q
⎧ ( ) ( )
⎪ +γ × + + (p − pext ) +Γ × + (p − pext ) =0
Dt ∂ω ∂γ Ds ∂Ω ∂Ω ∂Γ ∂Γ





D ∂ℓ0 D ∂ℓ0 ∂Q

⎪ ( )

⎪ + + (p − pext ) =0
Dt ∂γ Ds ∂Γ ∂Γ





∂ℓ0 ∂ℓ0 ∂ℓ0 ∂ℓ0

∂t + u∂s +2 ∂s u + ∂s λ = −Q ∂s p (2.30)
⎪ ∂u ∂u ∂u ∂λ
∂ℓ0 ∂ℓ0 ∂ℓ0 ∂ℓ0 ∂Q



∂t − ∂s2 ′′ + ∂s ′ −

⎪ − (p − pext ) =0
∂ Ṙ ∂R ∂R ∂R ∂R




∂t Ω = Ω × ω + ∂s ω, ∂t Γ + ω × Γ = ∂s γ + Ω × γ





∂t Q + ∂s (Qu) = 0, ∂t λ + u∂s λ = 0

together with appropriate boundary conditions enforcing vanishing of the variations of the boundary terms.

Proof. The system is obtained by replacing the expression of the functional derivatives (2.29) in the system (2.28). ■
δℓ0
Note that the term δR
contains both the terms coming from the elastic energy and the derivatives of the shape function Φ (R). Thus,
the vorticity contributes to the pressure term and momentum balance of the wall, as expected.

Remark 2.5 (On the Formal Correspondence Between the Entropy and Circulation). One can notice the formal correspondence between
(2.28) and the corresponding equations for the incompressible fluid with thermal energy derived in [73]. Mathematically, the appearance
of equations is equivalent when the circulation λ is associated with entropy S in [73], which is due to the same way the circulation
and entropy are advected. That association is purely mathematical: there is no physical correspondence between these quantities, and
they enter the Lagrangian in a very different way. The circulation enters the kinetic energy part of the Lagrangian, and entropy enters
the internal energy part. Thus, in spite of the apparent formal similarity between these equations, the use of any particular physically
relevant Lagrangian leads to very different equations for the fluids containing swirling vs thermal energy.

Eqs. (2.30) have to be solved as a system of nonlinear partial differential equations, since all equations are coupled through the
functional form of the Lagrangian (2.24). This can be seen, for example, from computing the derivative
∂ℓ0
= ρ A(R)|Γ | γ + Γ u · Γ ,
( )
(2.31)
∂u
which appears in the balance of fluid momentum, i.e., the third equation in (2.30). Thus, the fluid momentum involves the radius,
stretch Γ , and linear velocity γ , as well as the fluid’s velocity.
If we define the rescaled fluid momentum as
1 δℓ0
m :=
ρ Q δu
the third equation in (2.30) can be (deceptively) simply written as
1 ∂ℓ0 1 ∂ℓ0
( )
1
∂t m + ∂s mu − + λs = − ∂s p , . (2.32)
ρ ∂Q ρ Q ∂λ ρ
which is reminiscent of the equations of motion for one-dimensional fluid with the additional circulation terms describing the evolution
of vorticity.
For Q (R, Γ ) = A(R)|Γ |, Eqs. (2.30) can be further simplified since
∂Q Γ Γ ∂Q
= A(R) =Q and Γ× = 0.
∂Γ |Γ | |Γ |2 ∂Γ
Notice also that for A(R) = π R2 , we have dA/dR = 2π R, the circumference of a circle.
10 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

3. Tubes with a straight centerline: dynamics and reduction to a single equation

In this section we shall specify our model to the case of an inextensible and unshearable expandable tube. We then focus on straight
expandable tubes with no rotational motion and compare our model with previous works. This case has been studied extensively in
the literature, mostly in the context of blood flow involving incompressible fluid. We also show that these simplified models arise from
a variational principle using the back-to-labels map. The explicit use of back-to-labels map allows to derive a single equation of motion
without the need for the pressure-like Lagrange multiplier term p.

3.1. Equations of motion for inextensible and unshearable tubes

The dynamics of an inextensible and unshearable, but expandable, tube conveying compressible fluid is obtained by imposing
the
∫ T ∫constraint Γ (t , s) = χ, for all t , s. If we denote by z the Lagrange multiplier associated to this constraint and add the term
L
0 0
z · ( Γ − χ )ds dt to the action functional in our variational principle (2.26), the first two equations in (2.30) will change to
D ∂ℓ0 ∂ℓ0 D ∂ℓ0 ∂Q ∂ℓ0 ∂Q
⎧ ( ) ( )
+γ × + + (p − pext ) +Γ × + (p − pext ) +z =0


Dt ∂ω (∂γ Ds ∂Ω ) ∂Ω ∂Γ ∂Γ

⎪ D ∂ℓ0 + D ∂ℓ0 + (p − pext ) ∂ Q + z = 0.



Dt ∂γ Ds ∂Γ ∂Γ

The physical meaning of z is the reaction force enforcing the inextensibility constraint.
Particular simple solutions of this system can be obtained by assuming the axis of the tube being straight and no rotational motion.
In that case, the inextensibility constraint leads to z = z χ, the direction along the axis. For such particular solutions, the angular
momentum equation is satisfied identically, and the linear momentum equation reduces to a one dimensional equation for the reaction
force z. Therefore, we only need to compute the fluid momentum equation and the Euler–Lagrange equations for R. From the third and
fourth equations in (2.30), we get for A = A(R) = π R2 :
∂t ρ Au + ∂s ρ Au2 + ρ Φ λλs = −A∂s p
{ ( ) ( )
∂F ∂F 1 ∂Φ 2 (3.1)
aR̈ − ∂s ′ + − λ = 2π R (p − pext ) ,
∂R ∂R 2 ∂R
together with the conservation of mass and vorticity
∂t A + ∂s (Au) = 0, ∂t λ + u∂s λ = 0, (3.2)
This gives a system of four equations for the four variables u, R, λ, and p.
It is often more convenient to use the cross-sectional area A instead of the radius R, and express the Lagrangian in terms of A and its
∫ λ. To explicitly denote the change of variables, we shall denote that Lagrangian as ℓA , and the Lagrangian
derivatives, as well as u and
function as ℓ0,A , so ℓA = ℓ0,A ds. The equations of motion are then:
∂ℓ0,A ∂ℓ0,A ∂ℓ0,A

⎪ ∂t + u∂ s +2 us = −Aps
∂u ∂u ∂u




∂ℓA ∂ℓA ∂ℓA ∂ℓ (3.3)
∂t + ∂s − + λs = (p − pext ) ,
∂ A ∂ A ∂ A ∂λ

t s


At + ∂s (Au) = 0 , λt + uλs = 0 .

Without the vorticity, (3.3) represent the familiar model of tube with elastic walls incorporating the wall inertia [56,57]. For example,
one can consider the following form of the Lagrangian function ℓ0,A :
1 1
ℓ0,A = KA (A, At ) + ρ Au2 + ΦA (A)λ2 − 2FA (A, As ) , (3.4)
2 2
where KA (A, At ) is the kinetic energy of the wall motion expressed in terms of A and its time derivatives, FA (A, As ) is the elastic energy
function expressed in terms of A and its spatial derivatives, and ΦA is the vorticity shape function Φ (R) expressed in terms of A. If the
dependence of the Lagrangian function ℓ0,A on At and As is neglected, the second equation of (3.3) yields the so-called wall pressure
law giving an explicit connection between pressure and area, see [72] and references therein. This wall pressure law p(A) can then be
substituted in the first equation of (3.3), and, in conjunction with the conservation law (the first equation of (3.2)) allows to obtain
a closed set of two coupled equations for A and u. We shall not drop the inertia terms in (3.3) in what follows and pursue the more
general case of a general Lagrangian dependence on A and its time and s-derivatives.

3.2. The case of uniform circulation, energy, and the new constant of motion

Let us now analyze the equations of motion (3.3) in the case when λ = λ∗ = const. Then, λ = λ∗ everywhere in the flow. Physically,
such system can be realized by producing a constant vortex circulation at tube’s entry using a mechanical device, e.g. an idealized
propeller. In that case, we can define ℓ0,A = ℓ0,A with λ set to be equal to the value of λ∗ . Equations of motion for the general Lagrangian
function ℓ0,A (A, At , As , u) are then given as
∂ℓ0,A ∂ℓ0,A ∂ℓ0,A

⎪ ∂t + u∂s +2 us = −Aps
∂u ∂u ∂u




∂ℓ0,A ∂ℓ0,A ∂ℓ0,A (3.5)
∂t + ∂s − = (p − pext ) ,
∂ At ∂ As ∂A




At + ∂s (Au) = 0 .

R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 11

One would expect that there is an energy-like quantity E which is conserved on solutions of (3.5). Indeed, we can prove the following

Lemma 3.1 (Energy Conservation). The quantity


∂ℓ0,A ∂ℓ0,A

E= At + u − ℓ0,A ds . (3.6)
∂ At ∂u
is conserved on solutions of (3.5).

Proof. This result is proven by computing the time derivative of E and integrating by parts, using the equations of motion (3.5),
assuming that the boundary terms vanish when integrating by parts, and utilizing the constraint, i.e., the last equation of (3.5). This
proof closely follows that of Lemma 3.4, and it is relatively straightforward, so we do not present it here. ■

Remark 3.2 (On the Physical Meaning of E). One can notice that the quantity E defined by (3.35) is, in general, not equal to the kinetic
plus potential energy (also sometimes called the ‘physical’ energy). For example, for the Lagrangian defined by (3.4),

1 1 1
Ea = KA (A, At ) + ρ Au2 - ΦA (A)λ2 + 2FA (A, As )ds , (3.7)
2 2 2
where as the kinetic plus potential energy will be given by the expression

1 1 1
Kin + Pot Energy = KA (A, At ) + ρ Au2 + ΦA (A)λ2 + 2FA (A, As )ds . (3.8)
2 2 2
The energy defined by (3.7) is conserved, whereas the quantity defined by 3.8 is not. Changing time derivatives to the corresponding
momenta in (3.7) will give the Hamiltonian of the system. The non-conservation of ‘physical’ energy, i.e., the kinetic plus potential
energy, is typical for systems with rotation, see [97]. We believe it is difficult to guess the rather non-intuitive form (3.7) without
appealing to the variational principle used in the derivation of (3.33).
What is rather surprising, however, is that one can find there is another constant of motion. We prove the following

Theorem 3.3 (Additional Constant of Motion). The following quantity is conserved on the solutions of (3.5)
∂ℓ0,A ∂ℓ0,A

I= − As ds . (3.9)
∂u ∂ At

Proof. Differentiating the quantity defined by (3.9) with respect to time, we obtain:
∂ℓ0,A ∂ℓ0,A ∂ℓ0,A

dI
= ∂t
− Ast − As ∂t ds
dt ∂u ∂ At ∂ At
∂ℓ0,A ∂ℓ0,A ∂ℓ0,A

= −u∂s −2 us − Aps − Ast
∂u ∂u ∂ At
∂ℓ0,A ∂ℓ0,A
( )
− As −∂s + + (p − pext ) ds (3.10)
∂ As ∂A
∂ℓ0,A ∂ℓ0,A

= − us − Ass
∂u ∂ As
∂ℓ0,A ∂ℓ0,A
− Ats − As − ∂s [(p − pext ) A] ds = 0
∂ At ∂A
In the second line of (3.10) we have used Eqs. (3.5) for time derivatives of momenta, the third line is obtained by the integration by parts
and neglecting the boundary terms, and noticing that the expression under the integrand in the last line is (minus) the full derivative
of the quantity ℓ0,A + (p − pext )A with respect to s. ■

The expression is valid for an arbitrary Lagrangian functions, and, apart from the kinetic energy coming from the rotation and being
proportional to the parameter λ2∗ , which can be set to 0, is quite general to describe a majority of particular choices of kinetic and
potential energy for models in the literature, see e.g. [69,72] and references within. We shall compute the physical meaning of the
new conserved quantity (3.9) as the generalized linear momentum conservation, expressed in terms of back-to-labels map below in
Section 3.3. It was thus rather surprising to us that in spite of a large body of literature on the topic, we have failed to find any mention
of the conserved quantity (3.9). This is probably due to the fact that the back-to-labels map description of this conservation law is
rather non-intuitive.
We now turn our attention to the alternative description of motion utilizing the back-to-labels map. As it turns out, this description
is highly beneficial from the mathematical point of view, as it allows to combine several conservation laws into a single equation.

3.3. Equations of motion for the back-to-labels map

Let us now show how to reduce the fluid and wall momenta equations, and the conservation laws, to a single differential equation
using the back-to-labels map. The procedure of this method can be explained as follows. Let us denote, for shortness, ψ (s, t) = ϕ −1 (s, t)
to be the back-to-labels map, giving the Lagrangian label of a fluid particle at the point s at time t. For a given initial cross-section A0 (a),
the pressure term is entering the equations as the Lagrange multiplier for the constraint A(s, t) = A0 ◦ ψ∂s ψ = A0 (ψ )∂s ψ . Instead of
12 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

enforcing this holonomic constraint using the Lagrange multiplier p, in this Section, we will use ψ (s, t) as the free variable, express all
the other variables, such as the cross-section, in terms of ψ in the Lagrangian, and take the variations with respect to ψ . It will allow
us to compute the equations of motion as the explicit Euler–Lagrange equations in terms of variables ψ , and not the Euler–Poincaré
equations as we used in (2.30).
We can define the function Ψ (α ) as the antiderivative of A0 (α ) and rewrite the conservation law as
∫ a
Ψ (α ) = A0 (α ′ )dα ′ , A = ∂s Ψ (ψ (s, t)) (3.11)

In general, Ψ (α ) is a given function given by the initial profile of the cross-section. Since A0 (α ) > 0, Ψ (α ) is a monotonically increasing
function of α . We then use (3.11) to rewrite the conservation law
At + ∂s (Au) = 0 ⇔ Ψst + ∂s (Ψs u) = 0 . ⇔ Ψt + uΨs = C (t) . (3.12)
Setting the integration constant C (t) = 0 in the above formula, for example, due to the boundary conditions, we find the expression
of u in terms of the back-to-labels map ψ as
∂t Ψ (ψ ) ψt
u=− =− (3.13)
∂s Ψ (ψ ) ψs
The Lagrangian ℓA and the corresponding action SA for an inextensible, unshearable tube with the straight centerline are written as
∫ ∫
ℓA = ℓ0,A (A, At , As , u, λ)ds , SA = ℓA dt , (3.14)

for example, see the explicit version of ℓA given in (3.4). We have denoted, as before, the Lagrangian itself as ℓA , which involves the
integral over the s-domain. The function that is being integrated, i.e., the integrand of the Lagrangian, is denoted with a subscript 0,
i.e., ℓ0,A .
We are now ready to derive the conservation law (3.9) as the consequence of s-independence of the Lagrangian. Let us express the
velocity u in the Lagrangian ℓA through (3.11) and (3.13). We also assume that the initial cross-section is uniform, and set λ = λ∗ = const
to conform with the assumptions in the derivation of (3.9). We obtain the new Lagrangian

ℓψ = ℓ0,Ψ (A, At , As , ψs , ψt )ds
(3.15)
ψt
∫ ( )
= ℓ0,A A, At , As , u = − , λ = λ∗ ds
ψs
The variations of A and Ψ are now arbitrary. Since the Lagrangian ℓψ is independent of s, from general principles of Noether’s theorem
applied to fields [98], we have the conservation of the following quantity:
∂ℓ0,ψ ∂ℓ0,ψ ∂ℓ0,A ∂ℓ0,A
∫ ∫
P = As + ψs ds = As − ds = −I = const . (3.16)
∂ At ∂ψt ∂ At ∂u
Thus, the conservation of quantity I introduced in (3.9) is simply the conservation for the generalized linear momentum for the
Lagrangian written in terms of back-to-labels map (3.15). Thus, the conservation laws (3.9), as well as (3.38) below, arise from the
shift invariance with respect to s of the Lagrangian formulated in terms of back-to-labels maps. For reference, the energy invariance
(3.6) and, correspondingly, (3.35) below, arise from the t-invariance of the Lagrangian, as also described in [98].
Since, according to (2.22), A = A0 (ψ )∂s ψ , using (3.13) and the constraint A = A0 (ψ )∂s ψ , we can express the Lagrangian for the tube
(3.14) as a function of ψ only, and derive the equations of motion as a single Euler–Lagrange equation obtained by variations of the
action with respect to ψ . Let us illustrate this method in more detail, including the vorticity in the tube.
In order to explicitly separate the effect of fluid’s motion, let us define b = ψ (s, t) − s, since for the absence of fluid’s motion
ψ (s, t) = s. Then, from (3.13), the fluid velocity is expressed as
bt
u=− . (3.17)
1 + bs
In what follows we consider, for simplicity, the initial cross-section to be constant and the tube being initially uniform, i.e, A0 (α ) =
A0 = const, corresponding to Ψ (α ) = α . In principle, we can extend the theory for an arbitrary initial cross-section A0 (α ) > 0, although
this will introduce unnecessary algebraic complexity into equations.
Instead of the full cross-section A, it is more convenient to work with the relative deviation of the cross-section from the equilibrium,
as this approach leads to the most algebraically simple equations. More precisely, let us define a to be the relative deviation from the
equilibrium value according to A = A0 (1 + a). Then, the constraint A = A0 ◦ ϕ −1 ∂s ϕ −1 becomes
a = bs . (3.18)
In general, we will have ℓa = ℓ0,a (a, at , as , u, λ)ds. Just as in (3.14), we have dropped the dependence of the Lagrangian on ass , the

inclusion of this dependence is trivial and only leads to unnecessary algebraic complexity. We re-define the Lagrangian using (3.18),
by substituting a = bs , at = bts , as = bss into ℓa , and further employing (3.17) to express the velocity in terms of derivatives of b:
ℓ0,b = ℓ0,a (a, at , as , u, λ) where a = bs , u given by (3.17). (3.19)
ℓb dsdt We notice that the variations of the velocity in terms of

We then take variations of the corresponding action defined as Sb =
δ b are given as
bt δ bt bt 1
u=− , ⇒ δu = − + δ bs = − (δ bt + uδ bs ) (3.20)
1 + bs 1 + bs (1 + bs )2 1 + bs
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 13

and hence
bt δb
∂t λ + u∂s λ = 0 , ⇒ ∂t λ − ∂s λ = 0 , ⇒ δλ = ∂s λ . (3.21)
1 + bs 1 + bs
Starting with the Lagrangian ℓa = ℓ0,a (a, at , as , u, λ)ds, and remembering that a = bs , we compute the Euler–Lagrange equations

from the critical action principle as
∂ℓ0,a ∂ℓ0,a ∂ℓ0,a ∂ℓ0,a ∂ℓ0,a
∫ ( )
δS = δ bst + δ bss + δ bs + δλ + δ u dsdt
∂ at ∂ as ∂a ∂λ ∂u
∂ℓ0,a ∂ℓa ∂ℓ0,a
∫ (
= ∂st2 + ∂ss2 − ∂s
∂ at ∂ as ∂a
(3.22)
λs ∂ℓ0,a
)
+∂t m1 + ∂s (m1 u) + δ bdsdt
1 + bs ∂λ
1 ∂ℓ0,a
where m1 :=
1 + bs ∂ u
Hence, since δ b is arbitrary, the Euler–Lagrange equation corresponding to the variations with respect to b gives
∂ℓ0,a ∂ℓ0,a ∂ℓ0,a 1 ∂ℓ0,a
∂st2 + ∂ss2 − ∂s + + ∂t m1 + ∂s (m1 u) = 0 ,
∂ at ∂ as ∂a 1 + bs ∂λ
(3.23)
1 ∂ℓa
m1 = .
1 + bs ∂ u
We can simplify (3.23) further and bring it to a more compact form. Notice that due to the conservation law for volume (3.2) for
A = A0 (1 + a) = A0 (1 + bs ), we have

∂t (1 + bs ) + ∂s u(1 + bs ) = 0 , ⇒ ∂t b + u(1 + bs ) = C (t) , (3.24)

see also (3.12) for more general derivation of a non-constant initial cross-section A0 . In the expression above, we will set the integration
function C (t) = 0 assuming appropriate boundary conditions. This is the simplest choice we shall use in the remainder of the paper.
Multiplying equation (3.23) by (1 + bs ) and using (3.23), we can write the complete system as
∂ℓ0,a ∂ℓ0,a ∂ℓ0,a
( )
∂ts + ∂ss − ∂s (1 + bs )
∂ at ∂ as ∂a
∂ℓ0,a
+ λs + mt + ums + 2mus = 0
∂λ (3.25)
∂ℓ0,a bt
a = bs , m := , u=−
∂u 1 + bs
λt + uλs = 0 .
We can now make approximations in the Lagrangian directly, and choosing any physically relevant Lagrangian function ℓ0,a in (3.25)
will produce a consistent system with the Lagrangian structure. To be concrete, let us take the potential energy of the tube’s wall to be
1
2
(Pa2s + Qa2 + F (a)), where F (a) is a smooth function describing the higher-order terms in the elastic potential, for example, F (a) = Q2 a4 .
Thus, F (a) = o(a2 ) as a → 0. The first two terms of this potential energy are in fact the elastic part of the potential energy used in the
generalized string model, see e.g. [56,57], which is well-accepted in the literature. The last term F (a), corresponding to the nonlinear
elasticity, is one way to stabilize the dynamics in the unstable regime, as we shall see below. For further calculations, we take the
Lagrangian to be

1
ℓ0,a = Ka2t − Pa2s − Qa2 − 2F (a) + ρ A0 (1 + a)u2 + ρ Φ (a)λ2 ds . (3.26)
2
This Lagrangian is, in essence, a particular version of (3.4) reformulated in terms of relative deviation of the area a instead of the area
A itself. The first term in (3.26) is the kinetic energy of the wall deformation; the second and third terms are the potential energy of
the wall deformation. The fourth term is the kinetic energy of the fluid 21 ρ Au2 and the last term is the kinetic energy of fluid rotation,
with Φ (a) being a dimensionless quantity (shape function) of order 1. Technically, K = K (a), P = P(a) and Q = Q (a) are functions of
a, however, in the approximation (3.26) we have put them to be constants for simplicity since we assumed a to be sufficiently small
during the evolution. In that case, Eqs. (3.25) become
[ ]
1
Kbttss − Pbssss + Qbss + (F ′ (a))s − ρ∂s A0 u2 + Φ ′ (a)λ2 (1 + bs )
( )
2
+ ρ Φ (a)λλs + mt + ums + 2mus = 0 ,
(3.27)
bt δℓa
u=− , m := = ρ A0 (1 + bs )u = −ρ A0 bt ,
1 + bs δu
λt + uλs = 0 .
Notice that we do not have a pressure appearing in (3.27) since the constraint (3.18) is enforced explicitly by substitution in the
Lagrangian. The fluid momentum equation, i.e., the first equation of (3.27), now becomes a single equation for the quantity b, which is
14 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

coupled with the evolution equation for λ. After some algebra, we obtain

Kbttss − Pbssss + Q − ρ Φ ′′ (bs )λ20 bss + (F ′ (bs ))s (1 + bs )


( ( ) )

b2
( )
+ ρ A0 −btt + ∂s t + ρ Φ (bs ) − Φ ′ (bs )(1 + bs ) λs = 0
( )
1 + bs (3.28)
bt
λt = λs .
1 + bs
System (3.28) defines two coupled PDEs for the unknowns b and λ. It is interesting to compare this system with the one obtained by the
Euler–Poincaré method (3.3), which explicitly involves pressure, and has to be solved as a system of four equations for (u, λ, A, p). Notice
also that (3.3) does not define an evolution equation for pressure p, which has to be found in order to enforce the incompressibility.
In our opinion, Eq. (3.28) is advantageous for both theoretical and numerical studies, since there is no need to compute the pressure
term.

3.4. The case of constant vorticity

Let us now consider the case of Section 3.2, where the circulation generated by the source at the origin of the tube is held constant
by a mechanical device, so λ = λ0 . The general Lagrangian with λ = λ∗ = const in (3.23) gives

∂ℓ0,a ∂ℓ0,a ∂ℓ0,a 1 ∂ℓ0,a


∂st2 + ∂ss2 − ∂s + ∂t m1 + ∂s (m1 u) = 0 , m1 = . (3.29)
∂ at ∂ as ∂a 1 + bs ∂ u

where, as before, we have denoted ℓ0,a = ℓ0,a to be the Lagrangian function ℓ0,a evaluated at λ = λ0 . Similarly, for a particular
choice of the Lagrangian ℓ0,a given by (3.26), the last equation of (3.28) describing the evolution of λ is satisfied identically, and the
equation reduces to a single PDE for the variable b(s, t). We shall mostly analyze a particular form (3.28) in what follows; however, the
conservation of energy we will derive will be valid for an arbitrary functions ℓ0,a .
If we choose the relevant time and length scales T and L respectively, and defined the non-dimensionalized variables according to
K PT 2 QT 2
K∗ = , P∗ = , Q∗ := ,
ρ A0 L2 ρ A0 L4 ρ A 0 L2 (3.30)
A0 T F ′ (bs )
Φ∗ (a) = Φ (a) , λ∗ = λ0 , f (bs ) :=
L2 L2 KT 2 L2
We will choose the length and time scales so that K∗ = 1 and P∗ = 1 in (3.30), although other choices of time and length scales are
also possible. Combine the coefficients containing the background vorticity λ∗ in a single coefficient ζ defined as

ζ (a) := Q∗ − Φ∗′′ (a)λ2∗ . (3.31)

For example, for the cross-section approximation of the point vortex (2.18), we obtain
λ2∗
ζ (a) = Q∗ + >0 (3.32)
2π (1 + a)2
Keeping the notations b, t and s for the non-dimensionalized b, time and s, and performing some algebra, we obtain
b2t
(bttss − bssss + ζ (bs )bss + ∂s f (bs )) (1 + bs ) − btt + ∂s = 0. (3.33)
1 + bs
Eq. (3.33) gives, as far as we know, a new result expressing the evolution of the collapsible tube with compliant walls carrying vorticity in
terms of back-to-labels mab b(s, t). While it is difficult to observe the variable b in an experiment, its spatial derivative a = bs can readily
be observed. The rest of the paper is dedicated to the analysis of (3.33), both in general case, and in the asymptotic approximations of
long wavelengths and large times.
We start with the analysis of the full equation (3.33) and its general counterpart (3.23) first.

3.5. Energy conservation and additional constant of motion

Energy conservation. Let us define ℓ0,b to be ℓ0,a , i.e., the Lagrangian function ℓ0,a with the substitution a = bs , at = bst , as = bss . Since
(3.29) is derived as an Euler–Lagrange equation for the Lagrangian

ℓb = ℓ0,b (bst , bt , bs , bss )ds , (3.34)

it conserves the energy-like function as stated in the following

Lemma 3.4 (Energy Conservation). For appropriate periodic boundary conditions, Eq. (3.33) conserves the quantity
∂ℓb ∂ℓb

E= eds , e := bst + bt − ℓb . (3.35)
∂ bst ∂ bt
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 15

Proof. We prove this theorem by computing the evolution of energy density e defined by (3.35).
∂ℓb ∂ℓb ∂ℓb ∂ℓb d ℓb
( ) ( )
∂t e = bstt + btt + bst ∂t + bt ∂t −
∂ bst ∂ bt ∂ bst ∂ bt dt
∂ℓb ∂ℓb ∂ℓb ∂ℓb
[ ]
= bt −∂st + ∂t + ∂s − ∂ss (3.36)
∂ bst ∂ bt ∂ bs ∂ bss
∂ℓb ∂ℓb ∂ℓb
( )
+ ∂s bt ∂t − bt − bst
∂ bst ∂ bs ∂ bss
The expression in the square brackets is exactly the Euler–Lagrange equation for the Lagrangian (3.34), so it vanishes identically.∫Thus,
we have ∂t e + ∂s Je = 0 where the energy flux Je is defined by the equation in the parenthesis on the last line of (3.36). Thus, et ds
vanishes for periodic boundary conditions of the conditions where Je = 0 on the boundaries, for example, s ∈ (−∞, ∞) with all
variables tending to 0 as s → ±∞. ■

In the case when at the undisturbed state the fluid is moving with a constant velocity U, it is convenient to write (3.33) in the form
b(s, t) = −Ut + B(s, t) so B(s, t) = 0 if there are no deviations of the flow from the steady flow u = U. Then, (3.33) becomes
(Bt − U)2
(Bttss − Bssss + ζ (Bs )Bss + ∂s f (Bs )) (1 + Bs ) − Btt + ∂s = 0. (3.37)
1 + Bs
In addition to the energy conservation, one can prove the appropriate equivalent of (3.9) reformulated in terms of the back-to-labels
map B. Namely, we have the following

Lemma 3.5 (Additional Constant of Motion). For periodic boundary conditions, or boundary conditions where B → 0 when s → ±∞,
Eq. (3.37) conserves the following quantity:

I= Bt + Bss Bts ds . (3.38)

Proof. This can be seen by re-writing (3.37) in the form ∂t I + ∂s J for some quantity J depending on B, Bt , Bs , Bss etc. The result is valid
for arbitrary functions ζ (a) and f (a), and arbitrary velocity U. ■

3.6. Linear stability and instability

Let us now turn our attention to the linear stability analysis of Eq. (3.33). Since we assume that f (bs ) = o(bs ), we can write the
linearized version of that equation, by assuming small amplitude of b with moderate values for t- and s-derivatives. For convenience,
we also define ζ0 := ζ (0) to be the linearized value of ζ at b = 0. Then, (3.33) gives

(b − bss )tt = (ζ0 b − bss )ss . (3.39)

Note that on the left of that equation, we obtain exactly the Helmholtz operator acting on b. On the right of that equation, we have a
Helmholtz operator for ζ0 > 0 and anti-Helmholtz operator for ζ0 < 0. We will consider cases of ζ0 being possibly positive or negative,
keeping in mind that for physical reasons, ζ0 > 0 due to the expression (3.32). While the systems values ζ0 < 0 may exist, their
physical origin still remains uncertain, and thus all our numerical solutions will be computed for ζ0 > 0 (3.32) unless explicitly stated.
To compute the dispersion relation for no external flow, assume b = b0 ei(ks−ωt) and substitute into (3.39). We obtain:

ζ0 + k2
ω (1 + k ) = k (ζ0 + k ) ,
2 2 2 2
ω(k) = ±k . (3.40)
1 + k2
For large k, the waves propagate with the√ constant speed (phase velocity) of ±1. If ζ ≥ 0, all wavelengths are stable. If ζ0 < 0, then
there is a band of instability 0 < k <√ −ζ where the waves in that band grow exponentially. Long wavelengths k → 0 propagate
with the phase and group velocities ζ0 and short wavelengths k → ∞ propagate with the speed 1.
Let us now compute the stability for the case when the base flow has the circulation λ∗ and velocity U in the dimensionless variables.
The linearization of the terms in Eq. (3.37) proceeds as follows:

b = −Ut + ϵ B , bt = −U + ϵ Bt , bs = Bs , ϵ ≪ 1. (3.41)

Then, the linearization of (3.33) up to order ϵ is given by

Bttss − Bssss + (ζ0 − U 2 )Bss − Btt − 2UBst = 0 . (3.42)


ikx−iωt
Substitution of B = e gives the dispersion relation connecting ω and k

ω (1 + k ) + 2Ukω − k2 (ζ0 − U 2 + k2 ) = 0 .
2 2
(3.43)

We can thus solve for ω = ω(k) as


k ( √ )
ω= −U ± U 2 + (1 + k2 )(ζ0 − U 2 + k2 ) . (3.44)
1+ k2
We want to find the areas of stability and instability, i.e., the values of parameters U and ζ0 such that Imω = 0 (stability) or Imω > 0
(instability). Because of the ± sign in (3.44), it is sufficient to find the values of parameters when ω becomes complex for instability. The
16 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

Fig. 3.1. Examples of ω(k) computed from (3.44), with ζ0 = 1. Left panels: U = 1 (stable case). Right panels: ζ0 = 0 (unstable case). Solid red and blue lines
correspond to the respective choices of the + and − sign in (3.44). (For interpretation of the references to color in this figure legend, the reader is referred to the
web version of this article.)

square root (discriminant) in (3.44) is non-negative if and only if the quadratic polynomial f (z) = −U 2 z + (1 + z)(ζ0 + z) is non-negative
for all z = k2 ≥ 0. Therefore, the system is unstable if and only if
(1 + k2 )(ζ0 + k2 )
U 2 ≥ min (3.45)
k k2

Since the minimum of the function on the right-hand side of (3.45) is achieved at k2 = ζ0 (we remind that ζ0 is assumed to be
positive), then the condition of instability is given as

|U | > U∗ = 1 + ζ0 . (3.46)

For values of U satisfying (3.46), there is a band of unstable wavelengths of k satisfying (3.45). When U is slightly above√
U∗ , the instability
band of k is centered around k∗ = ζ0 . Thus, at the bifurcation point U = U∗ , the typical unstable wavelength is k ∼ ζ0 .√
We shall also note that for the long wavelength k → 0, Eq. (3.43) gives ω ∼ v± k, where the wave speed v± = −U ± ζ0 . The long
wavelength is thus always stable for ζ0 > 0. In the short wave limit k → ∞, Eq. (3.43) gives ω =∼ ±k, so the short waves propagate
with speeds of ±1.
The results of linear stability analysis presented above are summarized on Fig. 3.1. We take ζ0 = 1 corresponding to the critical
velocity U∗ = 2. The top and the bottom panels of this figure show, respectively, the real part of ω(k) computed from (3.44). On the left
panels, we show the results with U = 1, which is the stable case. For the stable case, Imω(k) = 0 for all k. The right panels show the
results for U = 3 which is the unstable case. There is a band of instability centered around k ∼ ζ0 = 1 as expected from the discussion
after (3.45).
Physically, the swirl makes it more difficult for the instability to occur. This is demonstrated by the formula (3.32) which shows
the increase of ζ0 with the circulation strength λ∗ . The higher value of ζ0 will correspond to the higher threshold for the instability for
velocity U given by (3.46).

3.7. Numerical simulations: stable case

The results of simulations of (3.37) with periodic boundary conditions for a(s, t) = Bs (s, t) are presented in Fig. 3.2. We have taken
Q∗ = 1, and ζ (a) given by (3.32). For the stable case, we have also set the nonlinear terms to 0, i.e., f (a) = 0. The parameters of the
simulations on the Figure are U = 1 and λ∗ = 1, with ζ0 ≃ 1.1592, corresponding to the stable case. We present a = Bs since a(s, t) is
the physical observable, being the deviation of the relative area from its equilibrium value.
On the left panel we present a = Bs inferred from the solution B(s, t) of the full equation (3.37) starting with a localized disturbance
B(s, 0) = e−0.1s , Bt (s, 0) = 0. On the right panel, we present a = Bs inferred from the solution for the Boussinesq-like approximation
2

of the full equation (4.13), derived below in Section 4.2 with the same parameter values and initial conditions as on the left panel. To
check the accuracy of our calculations, in Fig. 3.3, we also show the energy defined by (3.7) (left panel) and the conserved quantity I
defined by (3.38) (right panel). The energy defined in (3.7) is preserved to the relative accuracy of 10−4 . For comparison, on the right
panel we also present the terms Bt and Bss Bts , (green dotted and blue dashed lines) which are of order one. The constant of motion
−4
defined by (3.38), √ shown as a solid red line, is preserved to expected precision, about 10 .
If |U | > 1 + (ζ0 ), the solution becomes unstable, and if there is no potential energy to limit the growth of the solution, it quickly
reaches the singular (collapsed) state a = −1. In reality, any physical system will have a potential energy preventing the singularity. One
method is to introduce higher derivatives in the elasticity function F , and another way is to consider the potential energy preventing
both indefinite growth of a and collapse to a = −1. We chose the same parameters as in 3.2 except
√ d a4 2a3 (2 + a)
U = 1.25 + ζ0 , and f (a) = = (3.47)
da (1 + a)2 (1 + a)3
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 17

Fig. 3.2. Examples of solutions for a = Bs (s, t). Left: full equation (3.37). Right: Boussinesq-like approximation (4.13).

Fig. 3.3. Computed value∫of the conserved quantities


∫ for simulation shown in Fig. 3.2. Left panel: energy ∫ E as defined in (3.35). Right panel: quantity defined by
(3.38). Dashed blue line: Bt ds. Dotted green line: Bss Bts ds. Solid red line: the conserved quantity I = Bt + Bss Bts ds.


Fig. 3.4. Solution a = Bs for the unstable case U > 1 + ζ0 . Initial conditions are given by small random perturbations of order 10−3 to the equilibrium state B = 0.

The initial conditions at t = 0 are taken to be a random perturbation for B(s, 0) with the amplitude 10−3 and Bt (s, 0) = 0. The results
of the simulations for a = Bs are demonstrated in Fig. 3.4. The solution grows exponentially from the unstable equilibrium state a = 0,
saturating at an amplitude |a| ∼ 0.5 and yielding complex non-steady dynamics.
18 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

4. Asymptotic analysis of Eq. (3.37)

4.1. Small cross section (collapse) and reduction to Monge–Ampère equation

One of the most interesting regimes of evolution occurs when the cross-section almost collapses to 0, in other words 1 + Bs > 0
and |1 + Bs | ≪ 1. This regime is important for physiological flow applications and is studied quite extensively in the literature, see, for
example, recent review [99] and the references therein. We show that in this regime, our Eq. (3.37) is approximated by the celebrated
Monge–Ampère equation. We proceed as follows. For 1 + Bs being small and positive, and Bss , Bsstt , Bssss remaining finite, the asymptotic
equation following from (3.37) is
(Bt − U)2
Btt = ∂s . (4.1)
1 + Bs
We now use the substitution
B = Ks (s, t) − s + Ut . (4.2)
With this substitution, (4.1) integrates once with respect to s. Dropping the arbitrary function of time appearing under this integration,
we obtain the following equation for K (s, t):
Kss Ktt = Kst2 , (4.3)
which is the (real) homogeneous Monge–Ampère equation with (s, t) being independent variables. For the approximation (4.1) to be
valid, we assume K to be order 1, and since Bs + 1 is small, we need that Kss should be small. We can thus assume that K = K (ϵ S , ϵτ )
where ϵ ≪ 1 is a small parameter and the variables S , τ are of order 1. Then Eq. (4.3) transforms to
KSS Kτ τ = KS2τ , (4.4)
i.e. satisfies the same Monge–Ampère equation. Thus, any solution K (S , τ ) of (4.4) will generate a solution K (s, t) describing the collapse
of the tube. For physical reasons, we need KSS > 0 for all solutions since 1 + Bs = Kss > 0. For that solution, the neglected terms in
(3.37) are order ϵ 3 or smaller in order than the terms kept in (4.1), so the approximation is consistent.
As is known from the literature, Eq. (4.4) possesses a rich structure of solutions and is integrable [100–102]. For example, for any
smooth function g(x) with g ′′ (x) > 0, K (S , τ ) = g(S − c τ ) is a solution of (4.4) for any constant c. Thus, the collapse of the tube is
strongly dependent on the initial conditions and is unlikely to be described by a universal law.

4.2. Approximation by Boussinesq-type equations

In this section, we are interested in a consistent approximation of (3.37) for small amplitude and long wavelength approximation
which is a common approximation for hydrodynamic system with moving boundaries. Perhaps the most celebrated examples are the
Korteweg–de Vries (KdV) and Boussinesq equations. We start with the reduction of our system to the Boussinesq-type equation. Such
reduced models have been used in the analysis of pulse propagation in arteries, see [103,104].
In order to keep the description self-contained, let us briefly introduce the integrable Boussinesq equation (BE),
utt = uxx + 3(u2 )xx + uxxxx . (4.5)
Note that the notations u, ψ, λ in this section are not related to variables from the previous sections. We hope no confusion arises from
the clash of notations, since these are the variables used commonly in the field of integrable equations.
Eq. (4.5) arises in many various physical applications, most notably in describing propagation of the long waves in shallow
water [105,106]. The BE is integrable by the inverse scattering method, with the Lax pair representation that originates from
Zakharov [107]:

4ψxxx + (6u + 1)ψx + 3ux ψ ± 3i(∂x−1 ut )ψ = λψ,

ψt = ∓i 3(ψxx + uψ ). (4.6)
Here ψ is the eigenfunction of the spectral problem and λ is the spectral parameter. The solutions are explicitly described in terms of
solitons which are computed by Hirota [108]. It is more convenient to rescale the variables in (4.5)
u → α u, ∂t → β∂t , ∂x → β∂x (4.7)
where α and β are some constants. The BE equation (4.5) then becomes
utt = uxx + 3α (u2 )xx + β 2 uxxxx . (4.8)
Let us now turn our attention to Eq. (3.37) for ζ = const and no nonlinear terms in the potential, i.e f (a) = 0. We get the following
nonlinear equation, written here for reference:
(−U + Bt )2
( )
(Bttss − Bssss + ζ Bss )(1 + Bs ) − Btt + = 0. (4.9)
1 + Bs s

This equation will form the basis of derivation of approximate limiting cases in this chapter.
Let us consider a small amplitude and large wavelength propagation regime, where B → ε B, ∂t → ε∂t , ∂x → ε∂x where ε is a small
parameter that defines the scale for the corresponding quantities. By keeping terms up to and including ε 2 in (4.9) we obtain
ζ
( )
(ζ − U 2 )Bss − 2UBts − Btt + ε 2 (Bttss − Bssss ) + ε 2 B2s + B2t + 2UBt Bs = 0. (4.10)
2 s
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 19

The leading order terms in this equation are

(ζ − U 2 )Bss − 2UBts − Btt = 0 (4.11)


giving an operator equality to the leading order
√ √
∂t = (± ζ − U)∂s + O(ε ) = −c± ∂s + O(ε ) , c± := U ± ζ (4.12)
when the operators are acting on the solution B of (4.10) . The plus sign is for the right-running waves, the minus sign is for the
left running waves. Then, c+ is, to the leading order, the propagation speed of the right-running waves and c− is the speed of the
left-running waves.
Next, we use the expression for ∂t from (4.12) to make an approximation of the last term in (4.10)
ζ ζ √ √
B2s + B2t + 2UBt Bs = B2s + (ζ ∓ 2 ζ U + U 2 )B2s + 2U(± ζ − U)B2s + O(ε )
2 ( 2

)
= −U 2
B2s + O(ε )
2
Then, this approximation of (4.10) gives an equation simultaneously valid for both left- and right-running waves:

( )
(ζ − U 2 )Bss − 2UBts − Btt + ε 2 (Bttss − Bssss ) + ε 2 − U 2 (B2s )s = 0. (4.13)
2
This equation provides a universal, Boussinesq-type approximation of Eq. (3.37). We present a numerical solution of this equation on
the right panel of Fig. 3.2 to compare with the solution of full equation (3.37) with the same initial conditions. We observe that there is
a very good quantitative agreement between these solutions for this propagation regime. Note that (4.13) is substantially algebraically
simpler than (3.37). We do not yet know whether Eq. (4.13) is integrable, but from the algebraic structure of this equation, we believe
that it is much more amenable to analytical studies than the full (3.37).
While it is not known whether (4.13) is integrable, a further approximation can be made to reduce it to an integrable Boussinesq
case. We can make the following approximation using (4.12):

Bttss − Bssss = [( ζ ± U)2 − 1]Bssss + O(ε ) . (4.14)
This approximation, substituted in (4.13) gives the following equation:

( )

(ζ − U 2 )Bss − 2UBts − Btt + ε 2 [( ζ ± U)2 − 1]Bssss + ε 2 − U 2 (B2s )s = 0. (4.15)
2
Remembering that a := Bs is the relative deviation of the cross-sectional area, we can write (4.15) as

( )

(ζ − U 2 )ass − 2Uats − att + ε 2 [( ζ ± U)2 − 1]assss + ε 2 − U 2 (a2 )ss = 0. (4.16)
2
In order to write (4.16) in the canonical form (4.5), we make the following change of variables:
s − Ut
x= √ , τ =t (4.17)
ζ
U 1
∂t = ∂τ − √ ∂x , ∂s = √ ∂x . (4.18)
ζ ζ
In the new variables, (4.16) transforms to

( )
1 √ 1
axx − aτ τ + ε 2 [( ζ ± U)2 − 1]axxxx + ε 2 − U 2 (a2 )xx = 0, (4.19)
ζ 2 ζ 2
which is of the form
utt = uxx + 3α (u2 )xx ± β 2 uxxxx . (4.20)
The + sign in front of β term corresponds to the classical Boussinesq equation (4.8), and the − sign corresponds to the alternative
2

sign Boussinesq equation obtained from (4.8) by the substitution x → ix and t → it, which is also integrable. The constants α and β
from (4.20) are given by

( )
1 1 ⏐√
⏐ ⏐
α≡ε 2
−U 2
, β±2 ≡ ε 2 ⏐( ζ ± U)2 − 1⏐ . (4.21)

ζ 2 ζ2

Note that including α = 0 makes Eq. (4.8) a linear equation. While this reduction is interesting, to describe both left- and right-running
waves at the same time, we believe that Eq. (4.13) is more universal since it does not contain the ± signs.

4.3. Reduction to Korteweg–de Vries (KdV) equation

It is also interesting to investigate the appropriateness of the use of Korteweg–de Vries (KdV) approximation for this problem. While
the universal Boussinesq-like equation (4.13) is valid for both left and right-traveling waves, its integrable counterpart (4.19) is written
separately for the left- and right-running waves. However, each version of Eq. (4.19) itself has both the left- and the right-running
waves. Thus, considering both left- and right version of (4.19) generates too much spurious behavior compared to the full system. In
20 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

order to obtain an adequate integrable model for those waves, we reduce (3.37) to the KdV equation for waves running in each direction
as follows. The KdV approximation for the fluid flow with elastic walls has been considered in the literature before, especially in the
context of solitary wave propagation and variable properties of the arteries [109,110] and reflection of the waves from the branching
points [111]. Thus, we believe it is important to consider the appropriateness of KdV approximation in our case as well.
We are going to proceed as follows. Suppose that B satisfies a KdV equation of the form
B2s
( )
Bt = −cBs + ε 2 pBsss + ε 2 q (4.22)
2

Here p, q, c are yet undetermined constants. A subindex ± will denote right/left running wave, and we shall enforce c = c± = U ± ζ
as defined by (4.12). We will drop the ± sign in the derivation below for brevity. From (4.22) we necessarily have Bt = −cBs + O(ε 2 ).
Differentiating (4.22) with respect to t gives
B2s B2s
( ) ( )
Btt = −cBst + ε pBssst + ε q
2 2
= −cBst − ε pcBssss − ε qc
2 2
(4.23)
2 t 2 s

Next, differentiating (4.22) with respect to s gives


B2s
( )
Bts = −cBss + ε pBssss + ε q
2 2
(4.24)
2 s

From (4.23), (4.24) we arrive to


Btt − c 2 Bss + ε 2 2pcBssss + ε 2 qc B2s
( )
s
=0
(4.25)
2UBts + 2UcBss − ε 2UpBssss − ε Uq Bs
2 2
( 2)
s
=0
Finally, adding equations in (4.25) we obtain

(c 2 − 2Uc)Bss − 2UBts − Btt − ε 2 2(c − U)pBssss − ε 2 (c − U)q B2s


( )
=0 s
(4.26)

which matches (4.15). Indeed, we take c 2 − 2Uc = ζ − U 2 , or c = c± = U ± ζ , as expected for the speeds of the left- and right-running
waves (4.12). For the parameters p and q, we obtain
2
c2 − 1 c± −1
−2p(c − U) = c 2 − 1, p=− , p± = − √ ,
2(c − U) ±2 ζ
3ζ 3ζ
(4.27)
3ζ − U2 − U2
−(c − U)q = −U , 2
q=− , 2
q± = − 2 √ .
2 c−U ± ζ
The equation for B is then

c2 − 1 − U 2 B2s √
Bt + cBs + ε 2 Bsss + ε 2 2
= 0, c = c± = U ± ζ, (4.28)
2(c − U) c−U 2
where we have implicitly assumed that c has the label of c± for the left (−) and right (+) running waves. The KdV equation for a = Bs
is

c2 − 1 − U2 a2
( )

at + cas + ε 2 asss + ε 2 2
= 0, c = c± = U ± ζ. (4.29)
2(c − U) c−U 2 s

In order to compare the accuracy of Eq. (4.28), we perform the simulation with the same initial conditions and parameters as in Fig. 3.2,
separating the solutions into the left- and right-running waves, i.e., taking the plus and minus signs in the corresponding equation (4.28).
The results for a = Bs are presented in Fig. 4.1. Note that the speed of propagation is computed quite well as compared to the left panel
of Fig. 3.2, i.e., solution of the full equation (3.37). However, the details of the evolution do not quite match, especially because of the
contribution of short wavelength ripples in the KdV equation, which is supposed to be the long-wavelength approximation only. Still,
we believe that it is useful to consider the KdV for the propagation of the wave pattern in either direction.

5. Conclusions

In this paper, we have derived several novel results for the description of fluid with vorticity flowing in tubes with expandable walls.
As our approach is variational in nature, it is applicable to the flow of essentially inviscid fluids. The friction terms can be introduced
later in the model by using the Lagrange–d’Alembert’s principle of external forces. We have intentionally avoided discussion of the
friction terms here, as the nature of fluid friction in blood is very complex and is certainly beyond the scope of this paper. That being
said, in order to apply our theory successfully to the blood flow, the friction between the fluid and the walls will need to be introduced.
This is an important feature of the practical applicability of the theory developed here and should be considered later.
On a more theoretical side, in this manuscript we have concentrated on the purely Lagrangian description of the dynamics. An
equivalent Hamiltonian formulation can be derived by constructing an appropriate Poisson bracket. Since the description through the
back-to-labels map comes as the Euler–Lagrange equation, the Poisson bracket for the back-to-label description will be canonical. We
have not touched upon the Hamiltonian description here to keep the paper concise; the Hamiltonian description will be considered in
the follow-up paper.
Another interesting direction of study, having both a theoretical and a practical value, is a more detailed incorporation of the
dynamics of the media outside the tube. In our paper, we have modeled this media as simply exerting a constant pressure, but having
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 21

Fig. 4.1. Solutions for a = Bs the left-running and right-running KdV equations (4.28) a(s, t). Left panel: Left-running KdV equation, corresponding to the − sign in
(4.28). Right panel: Right-running KdV equation corresponding to the + sign in (4.28).

no dynamics of its own. This may be a very good approximation for tubes conveying fluid when the outside media is an ideal gas kept
at constant pressure. However, in reality, the dynamics of tubes submerged in a resisting media will cause a non-trivial response from
the media beyond constant pressure, and thus affect the dynamics in a different way. This response could be treated in the variational
framework if, for example, the outside media is modeled either as an elastic body, or an ideal fluid. This interesting and challenging
direction of study will be also considered in the future work.

Acknowledgments

We acknowledge inspiration from fruitful and productive discussions with D.D. Holm, F. Gay-Balmaz and T.S. Ratiu. The research of
VP was partially supported by the NSERC Discovery Grant program and the University of Alberta. This work has also been made possible
by the awarding of a James M Flaherty Visiting Professorship from the Ireland Canada University Foundation, with the assistance of the
Government of Canada/avec l’appui du gouvernement du Canada.

References

[1] H. Ashley, G. Haviland, Bending vibrations of a pipe line containing flowing fluid, J. Appl. Mech. 17 (1950) 229–232.
[2] B.T. Benjamin, Dynamics of a system of articulated pipes conveying fluid I. Theory, Proc. Roy. Soc. A 261 (1961) 457–486.
[3] B.T. Benjamin, Dynamics of a system of articulated pipes conveying fluid II Experiments, Proc. Roy. Soc. A 261 (1961) 487–499.
[4] A.P. Veselov, Integrable discrete-time systems and difference operators, Funkt. Anal. Prilozhen. 22 (2) (1988) 1–26 (in Russian).
[5] J. Moser, A. Veselov, Discrete versions of some classical integrable systems and factorization of matrix polynomials, Comm. Math. Phys. 139 (1991) 217–243.
[6] J.E. Marsden, M. West, Discrete mechanics and variational integrators, Acta Numer. (2001) 1–158.
[7] R.W. Gregory, M.P. Païdoussis, Unstable oscillation of tubular cantilevers conveying fluid I. Theory, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 293 (1966)
512–527.
[8] M.P. Païdoussis, Dynamics of tubular cantilevers conveying fluid, Int. J. Mech. Eng. Sci. 12 (1970) 85–103.
[9] M.P. Païdoussis, N.T. Issid, Dynamic stability of pipes conveying fluid, J. Sound Vibrations 33 (1974) 267–294.
[10] M.P. Païdoussis, Fluid-Structure Interactions. Slender Structures and Axial Flow, Vol. 1, Academic Press, London, 1998.
[11] S. Shima, T. Mizuguchi, Dynamics of a tube conveying fluid, 2001, arxiv:nlin.CD/0105038.
[12] O. Doaré, E. de Langre, The flow-induced instability of long hanging pipes, Eur. J. Mech. A Solids 21 (2002) 857–867.
[13] M.P. Païdoussis, G.X. Li, Pipes conveying fluid: A model dynamical problem, J. Fluids Struct. 7 (1993) 137–204.
[14] M.P. Païdoussis, Fluid-Structure Interactions. Slender Structures and Axial Flow, Vol. 2, Academic Press, London, 2004.
[15] L.D. Akulenko, M.I. Ivanov, L.I. Korovina, S.V. Nesterov, Basic properties of natural vibrations of an extended segment of a pipeline, Mech. Solids 48 (2013)
458–472.
[16] L.D. Akulenko, D.V. Georgievskii, S.V. Nesterov, Transverse vibration spectrum of a part of a moving rod under a longitudinal load, Mech. Solids 50 (2015)
227–231.
[17] L.D. Akulenko, D.V. Georgievskii, S.V. Nesterov, Spectrum of transverse vibrations of a pipeline element under longitudinal load, Dokl. Akad. Nauk 467 (2016)
36–39.
[18] R.W. Gregory, M.P. Païdoussis, Unstable oscillation of tubular cantilevers conveying fluid II. Experiments, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 293
(1966) 528–542.
[19] S. Kuronuma, M. Sato, Stability and bifurcations of tube conveying flow, J. Phys. Soc. Japan 72 (2003) 3106–3112.
[20] F. Castillo Flores, A. Cros, Transition to chaos of a vertical collapsible tube conveying air flow, J. Phys. Conf. Ser. 166 (2009) 012017.
[21] A. Cros, J.A.R. Romero, F. Castillo Flores, Sky dancer: A complex fluid-structure interaction, in: Experimental and Theoretical Advances in Fluid Dynamics:
Environmental Science and Engineering, Springer, 2012, pp. 15–24.
[22] G.X. Li C. Semler, M.P. Païdoussis, The non-linear equations of motion of pipes conveying fluid, J. Sound Vib. 169 (1994) 577–599.
[23] Y. Modarres-Sadeghi, M.P. Païdoussis, Nonlinear dynamics of extensible fluid-conveying pipes supported at both ends, J. Fluids Struct. 25 (2009) 535–543.
[24] M. Ghayesh, M.P. Païdoussis, M. Amabili, Nonlinear dynamics of cantilevered extensible pipes conveying fluid, J. Sound Vib. 332 (2013) 6405–6418.
[25] V.N. Zhermolenko, Application of the method of extremal deviations to the study of forced parametric bend oscillations of a pipeline, Autom. Telemech. 9
(2008) 10–32 (in Russian).
[26] M.A. Beauregard, A. Goriely, M. Tabor, The nonlinear dynamics of elastic tubes conveying a fluid, Int. J. Solids Struct. 47 (2010) 161–168.
[27] J. Rivero-Rodriguez, M. Perez-Saborid, Numerical investigation of the influence of gravity on flutter of cantilevered pipes conveying fluid, J. Fluids Struct. 55
(2015) 106–121.
[28] N. Bou-Rabee, L. Romero, A. Salinger, A multiparameter, numerical stability analysis of a standing Cantilever conveying fluid, SIAM J. Appl. Dyn. Sys. 1 (2002)
190–214.
22 R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172

[29] I. Elishakoff, Controversy associated with the so-called follower forces: Critical overview, Appl. Mech. Rev. 58 (2005) 117–142.
[30] V.A. Svetlitskii, Mechanics of rods, vol. 2, Vyssh. shkola, Moscow, 1987 (in Russian).
[31] A.A. Movchan, On one problem of stability of a pipe with moving fluid, J. Appl. Math. Mech. 29 (1965) 760–762 (in Russian).
[32] A.A. Mukhin, Dynamic criterium of stability of a pipeline with moving fluid, Izv. Akad. Nauk USSR Ser. Mekh. 3 (1965) 154–155 (in Russian).
[33] M.A. Ilgamov, Oscillations of Elastic Shells Containing Fluid and Gas, Nauka, 1969 (in Russian).
[34] T.E. Anni, E.L. Martin, R.N. Duby, Hydroelastic instability of pipes with constant radius of curvature with fluid, Appl. Mech. Prikl. Mekh. 6 (1970) 244–249
(in Russian).
[35] A.S. Vol’mir, M.S. Gratch, Oscillations of a shell with moving fluid, Izv. Akad. Nauk USSR Ser. Mekh. Tverd. Tela 6 (1973) 162–166 (in Russian).
[36] V.A. Svetlitskii, Small oscillations of spatially curved pipelines, Appl. Mecha. Prikl. Mekh. 14 (1978) 70–75 (in Russian).
[37] P.D. Dotsenko, Some studies of auto-oscillations of straight pipelines with fluid, Appl. Mech. Prikl. Mekh. 15 (1979) 69–75.
[38] S.V. Chelomey, On dynamical stability of elastic systems conveying moving pulsating fluid, Izv. Akad. Nauk USSR Ser. Mekh. Tverd. Tela 5 (1984) 170–174
(in Russian).
[39] V.G. Sokolov, A.V. Bereznev, Solution of the problem for free oscillations of curved pipelines with moving fluid, Izv. Vuzov Oil Gas (2005) 80–84 (in Russian).
[40] R.Yu. Amenzade, A.B. Aliev, Propagation of waves in fluid moving in an elastic tube taking into account viscoelastic friction of surrounding media, Education
4 (2015) 6–9.
[41] S-S Chen, Vibration and stability of a uniformly curved tube conveying fluid, J. Acoust. Soc. Am. 51 (1B) (1972) 223–232.
[42] A.K. Misra, M.P. Païdoussis, K.S. Van, On the dynamics of curved pipes transporting fluid. Part I: inextensible theory, J. Fluids Struct. 2 (3) (1988) 221–244.
[43] A.K. Misra, M.P. Païdoussis, K.S. Van, On the dynamics of curved pipes transporting fluid Part II: Extensible theory, J. Fluids Struct. 2 (3) (1988) 245–261.
[44] C. Dupuis, J. Rousselet, The equations of motion of curved pipes conveying fluid, J. Sound Vib. 153 (3) (1992) 473–489.
[45] R.W. Doll, C.D. Mote, On the dynamic analysis of curved and twisted cylinders transporting fluids, J. Press. Vessel Technol. 98 (2) (1976) 143–150.
[46] R. Aithal, G.S. Gipson, Instability of internally damped curved pipes, J. Eng. Mech. 116 (1) (1990) 77–90.
[47] M.H. Ghayesh, M.P. Païdoussis, M. Amabili, Nonlinear dynamics of cantilevered extensible pipes conveying fluid, J. Sound Vib. (2013) 6405–6418.
[48] F. Gay-Balmaz, V. Putkaradze, Exact geometric theory for flexible, fluid-conducting tubes, C.R. Acad. Sci., Paris II 342 (2014) 79–84.
[49] F. Gay-Balmaz, V. Putkaradze, On flexible tubes conducting fluid: geometric nonlinear theory, stability and dynamics, J. Nonlinear Sci. 25 (2015) 889–936.
[50] F. Gay-Balmaz, D. Georgievskii, V. Putkaradze, Stability of helical tubes conveying fluid, J. Fluids Struct. 78 (2018) 146–174.
[51] F. Gay-Balmaz, V. Putkaradze, Variational discretizations for the dynamics of flexible tubes conveying fluid, C. R. Méc. 344 (2016) 769–775.
[52] T. Pedley, The Fluid Mechanics of Large Blood Vessels, Cambridge University Press, 1980.
[53] O.E. Jensen, Instabilities of flow in a collapsed tube, J. Fluid Mech. 220 (1990) 623–659.
[54] T.J. Pedley, X.Y. Luo, The effects of wall inertia on flow in a two-dimensional collapsible channel, J. Fluid Mech. 363 (1998) 253–280.
[55] X.Y. Luo, T.J. Pedley, Modelling flow and oscillations in collapsible tubes, Theor. Comput. Fluid Dyn. 10 (1998) 277–294.
[56] A. Quarteroni, M. Tuveri, A. Veneziani, Computational vascular fluid dynamics: problems, models and methods, Comput. Visual Sci. 2 (2000) 163–197.
[57] L. Formaggia, D. Lamponi, A. Quarteroni, One-dimensional models for blood flow in arteries, J. Eng. Math. 47 (2003) 251–276.
[58] P.S. Stewart, S.L. Waters, O.E. Jensen, Local and global instabilities of flow in a flexible-walled channel, Eur. J. Mech. B Fluids 28 (2009) 541–557.
[59] D. Tang, Y. Yang, C. Yang, D.N. Ku, A nonlinear axisymmetric model with fluid-wall interactions for steady viscous flow in stenotic elastic tubes, Trans. ASME
121 (2009) 494–501.
[60] D. Elad, R.D. Kamm, A.H. Shapiro, Steady compressible flow in collapsible tubes: application to forced expiration, J. Appl. Physiol. 203 (1989) 401–418.
[61] E. Marchandise, P. Flaud, Accurate modelling of unsteady flows in collapsible tubes, Comput. Methods Biomech. Biomed. Eng. 13 (2010) 279.
[62] G. Donovan, Systems-level airway models of bronchoconstriction, Wiley Interdiscip. Rev. Syst. Biol. Med. 8 (2016) 459.
[63] M. Bukac, S. Canic, R. Glowinski, J. Tambaca, A. Quaini, Fluid-structure interaction in blood flow capturing non-zero longitudinal structure displacement, J.
Comput. Phys. 235 (2013) 515–541.
[64] T.J. Pedley, Longitudinal tension variation in collapsible channels: a new mechanism for the breakdown of steady flow, J. Biomech. Eng. 114 (1992) 60–67.
[65] K. Kounanis, D.S. Mathioulakis, Experimental flow study within a self oscillating collapsible tube, J. Fluids Struct. 13 (1999) 61–73.
[66] Kim Parker, C.J.H. Jones, Forward and backward running waves in the arteries: analysis using the method of characteristics, J. Biomech. Eng. 112 (1990)
322–326.
[67] C.J.H. Jones, K.H. Parker, R. Hughes, D.J. Sheridan, Nonlinearity of human arterial pulse wave transmission, J. Biomech. Eng. 114 (1992) 10–14.
[68] M. Bukac, A loosely-coupled scheme for the interaction between a fluid, elastic structure and poroelastic material, J. Comput. Phys. 313 (2016) 377–399.
[69] T.J. Pedley, Mathematical modelling of arterial fluid dynamics, J. Eng. Math. 47 (2003) 419–444.
[70] J.B. Grotberg, O.E. Jensen, Biofluid mechanics in flexible tubes, Ann. Rev. Fluid Mech. 36 (2004) 121–147.
[71] M. Heil, A.L. Hazel, Fluid-structure interaction in internal physiological flows, Ann. Rev. Fluid Mech. 43 (2011) 141–162.
[72] Timothy W. Secomb, Hemodynamics, Compr. Physiol. 6 (2016) 975–1003.
[73] F. Gay-Balmaz, V. Putkaradze, Geometric theory of flexible and expandable tubes conveying fluid: equations, solutions and shock waves, J. Nonlinear Sci. 29
(2) (2019) 377–414.
[74] P.A. Stonebridge, C.M. Brophy, Spiral laminar flow in arteries? Lancet 338 (8779) (1991) 1360–1361.
[75] P.A. Stonebridge, P.R. Hoskins, P.L. Allan, J.F.F. Belch, Spiral laminar flow in vivo, Clin. Sci. 91 (1) (1996) 17–21.
[76] R. Mittal, S.P. Simmons, H.S. Udaykumar, Application of large-eddy simulation to the study of pulsatile flow in a modeled arterial stenosis, J. Biomech. Eng.
123 (4) (2001) 325–332.
[77] M. Grigioni, C. Daniele, U. Morbiducci, C. Del Gaudio, G. DAvenio, A. Balducci, V. Barbaro, A mathematical description of blood spiral flow in vessels: application
to a numerical study of flow in arterial bending, J. Biomech. 38 (7) (2005) 1375–1386.
[78] M. Paul, A. Larman, Investigation of spiral blood flow in a model of arterial stenosis, Med. Eng. Phys. 31 (9) (2009) 1195–1203.
[79] J. Biasetti, F. Hussain, T. Christian Gasser, Blood flow and coherent vortices in the normal and aneurysmatic aortas: a fluid dynamical approach to intra-
luminal thrombus formation, J. R. Soc. Interface 8 (2011) 1449–1461.
[80] G. Reiter, U. Reiter, G. Kovacs, H. Olschewski, M. Fuchsjäger, Blood flow vortices along the main pulmonary artery measured with MR imaging for diagnosis
of pulmonary hypertension, Radiology 275 (1) (2014) 71–79.
[81] D.J. Doorly, S.J. Sherwin, P.T. Franke, J. Peiró, Vortical flow structure identification and flow transport in arteries, Comput. Methods Biomech. Biomed. Eng. 5
(3) (2002) 261–273.
[82] D.D. Holm, B.A. Kupershmidt, Poisson brackets and clebsch representations for magnetohydrodynamics, multifluid plasmas, and elasticity, Physica D 6 (1983)
347–363.
[83] Darryl D. Holm, Jerrold E. Marsden, Tudor Ratiu, Alan Weinstein, Nonlinear stability of fluid and plasma equilibria, Phys. Rep. 123 (1–2) (1985) 1–116.
[84] Henry H.D.I. Abarbanel, D.D. Holm, Nonlinear stability analysis of inviscid flows in three dimensions: incompressible fluids and barotropic fluids, Phys. Fluids
30 (11) (1987) 3369–3382.
[85] J.S. Allen, D.D. Holm, Extended-geostrophic hamiltonian models for rotating shallow water motion, Physica D 98 (2–4) (1996) 229–248.
[86] C.J. Cotter, D.D. Holm, P.E. Hydon, Multisymplectic formulation of fluid dynamics using the inverse map, Proc. R. Soc. A 463 (2086) (2007) 2671–2687.
[87] J.C. Simó, J.E. Marsden, P.S. Krishnaprasad, The Hamiltonian structure of nonlinear elasticity: The material and convective representations of solids, rods, and
plates, Arch. Ration. Mech. Anal. 104 (1988) 125–183.
[88] D.D. Holm, V. Putkaradze, Nonlocal orientation-dependent dynamics of charged strands and ribbons, C. R. Acad. Sci., Paris I 347 (2009) 1093–1098.
[89] D. Ellis, D.D. Holm, F. Gay-Balmaz, V. Putkaradze, T. Ratiu, Symmetry reduced dynamics of charged molecular strands, Arch. Ration. Mech. Anal. 197 (2010)
811–902.
[90] E. Cosserat, F. Cosserat, Théorie des corps déformables, Hermann, Paris, 1909.
[91] F. Gay-Balmaz, T.S. Ratiu, The geometric structure of complex fluids, Adv. Appl. Math. 42 (2009) 176–275.
[92] F. Gay-Balmaz, V. Putkaradze, in: M. Holler, P. Grohs, A. Weinmann (Eds.), Variational Methods for Fluid-Structure Interactions, in: Springer Handbook, Springer,
2019.
R. Ivanov and V. Putkaradze / Physica D 401 (2020) 132172 23

[93] D.D. Holm, T. Schmah, C. Stoica, Geometric Mechanics and Symmetry: From Finite to Infinite Dimensions, Vol. 12, Oxford University Press, 2009.
[94] P.G. Saffman, Vortex Dynamics, Cambridge University Press, 1992.
[95] P.K. Newton, The N-Vortex Problem: Analytical Techniques, Vol. 145, Springer Science & Business Media, 2013.
[96] P. Constantin, G. Iyer, A stochastic Lagrangian representation of the three-dimensional incompressible Navier–Stokes equations, Comm. Pure Appl. Math. 61
(3) (2008) 330–345.
[97] J. José, E. Saletan, Classical Dynamics: A Contemporary Approach, AAPT, 2000.
[98] N.N. Bogoliubov, D.V. Shirkov, Quantum Fields, Benjamin/Cummings, London, 1982.
[99] C.-H Han, J.K.W. Chesnutt, J.R. Garcia, Q. Liu, Qi Wen, Artery buckling: new phenotypes, models, and applications, Ann. Biomed. Eng. 41 (7) (2013) 1399–1410.
[100] T. Aubin, Nonlinear Analysis on Manifolds. Monge–Ampère Equations, Vol. 252, Springer Science & Business Media, 1982.
[101] L.A. Caffarelli, Some regularity properties of solutions of Monge–Ampère equation, Comm. Pure Appl. Math. 44 (8–9) (1991) 965–969.
[102] C.E. Gutiérrez, H. Brezis, The Monge–Ampère Equation, Vol. 44, Springer, 2001.
[103] R.C. Cascaval, A Boussinesq model for pressure and flow velocity waves in arterial segments, Math. Comput. Simulation 82 (6) (2012) 1047–1055.
[104] A. Liberson, J.S. Lillie, D.A. Borkholder, Numerical solution for the Boussinesq type models with application to arterial flow, J. Fluid Flow Heat Mass Transf. 1
(2014) 9–15.
[105] J. Boussinesq, Théorie de l’intumescence appelée onde solitaire ou de translation se propagente dans un canal rectangulaire, Comptes Rendus 72 (1871)
755–759.
[106] J. Boussinesq, Théorie des ondes et des remous qui se propagent le long d’un canal rectangulaire horizontal, en communiquant au liquide contenu dans ce
canal des vitesses sensiblemant parielles de la surface au fond, J. Math. Pures Appl. Deuxième Série 17 (1872) 55–108.
[107] V. Zakharov, On stochastization of one-dimensional chains of nonlinear oscillators, Sov. Phys.–JETP 38 (1974) 108–110.
[108] R. Hirota, Exact N-soliton solution of the wave equation of long waves in shallow-water and in nonlinear lattices, J. Math. Phys. 14 (1973) 810–815.
[109] H. Demiray, Waves in fluid-filled elastic tubes with a stenosis: Variable coefficients KdV equations, J. Comput. Appl. Math. 202 (2) (2007) 328–338.
[110] H. Demiray, Variable coefficient modified KdV equation in fluid-filled elastic tubes with stenosis: Solitary waves, Chaos Solitons Fractals 42 (1) (2009) 358–364.
[111] W. Duan, B. Wang, R. Wei, Reflection and transmission of nonlinear blood waves due to arterial branching, Phys. Rev. E 55 (2) (1997) 1773.

You might also like