Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Continuum Mech. Thermodyn.

DOI 10.1007/s00161-014-0357-6

O R I G I NA L A RT I C L E

Maria Laura De Bellis · Giuseppe C. Ruta · Isaac Elishakoff

A contribution to the stability of an overhanging pipe


conveying fluid

Received: 26 November 2013 / Accepted: 7 April 2014


© Springer-Verlag Berlin Heidelberg 2014

Abstract We investigate the dynamic stability of a pipe that conveys fluid, clamped or pinned at one end
and with an intermediate support, thus exhibiting an overhang. The model of the pipe incorporates both
Euler–Bernoulli and Bresse–Timoshenko schemes as well as transverse inertia. Material and external damping
mechanisms are taken into account, while the conveyed fluid is supposed to be in fully turbulent flow. The pipe
can rest on a linear elastic Winkler soil. The influence of all the physical quantities and of the overhang length
on the critical velocity of the fluid front is investigated. Some numerical results are presented and discussed.
Keywords Stability · Damping · Winkler foundation · Overhang

1 Introduction

Among the studies on mechanics of structures, the stability of equilibrium has received a lot of attention,
starting from the pioneering work by Euler [1]. With the ever growing use of lightweight slender structures,
especially thin-walled beams and shells, the subject has dragged much attention also in the past decades: after
the path-opening thesis by Koiter [2], a lot of literature has been published on the static loss of equilibrium
(buckling), as confirmed by also by comprehensive monographs [3–5]. Contributions on this field have been
given also by some of the authors [6–11] for various model of beams and axisymmetric plates. Another instance
of investigations is in [12] for a more general loss of stability.
The loss of equilibrium may assume a dynamic feature (flutter). In particular, the stability of structural
elements under non-conservative actions, some of which are labelled as ‘follower’, has been examined. Some
recent papers published on the subject are [13–17] and refer to the stability, in a general dynamic setting
(Lyapounov stability), of planar purely flexible beams on a damped soil, under both conservative and non-
conservative loads, subjected to both lumped and distributed damping: both the critical values of the load
parameter(s) and the post-critical path are investigated, and the effect of the various physical parameters is
examined. Other problems regarding dynamics and stability of beams on foundations are found in [18,19].

Communicated by Andreas Öchsner.


M. L. De Bellis · G. C. Ruta (B)
Dipartimento d’Ingegneria Strutturale e Geotecnica, “Sapienza” University, Rome, Italy
E-mail: giuseppe.ruta@uniroma1.it

M. L. De Bellis
E-mail: marialaura.debellis@uniroma1.it

I. Elishakoff
Department of Mechanical Engineering, Florida Atlantic University, Boca Raton, FL, USA
E-mail: elishako@fau.edu
M. L. De Bellis et al.

Even if Koiter in [20] claimed that statically applied follower forces were ‘unrealistic’, he recognized
that there are technical problems involving non-conservative forces, such as the pipe conveying a fluid under
pressure. This subject has been thoroughly studied by Païdoussis and co-workers since 1963 [21]; among more
recent investigations, we may quote [22]. Statics and dynamics of pipes conveying fluid are also considered,
among other problems of interest, in [23–26], where comprehensive expositions of models and, in particular,
beams are discussed. The pipe resting on an intermediate support and conveying pressurized fluid was investi-
gated, among others, by [27], and experimental results on the subject are presented in [28]. The pipe conveying
fluid resting on a Winkler elastic foundation was studied, inter alia, by Lottati and Kornecki [29].
There are paradoxical behaviours arising in some of these studies, and Elishakoff discussed the paradoxes
associated with columns under follower forces in [30]. In 1972, Smith and Herrmann [31] investigated the sta-
bility of a uniform Beck’s column attached to a uniform Winkler foundation and found that the flutter load does
not depend on the foundation stiffness, a paradox clarified later. Thus, it is useful to study the effect of elastic soil
and other physical parameters on the realistic problem of the stability of a pipe conveying a fluid under pressure.
Elishakoff and Impollonia [32] investigated a linear elastic Winkler soil, reacting to translation and rotation,
partially attached to a pipe, seen as a purely flexible (or Euler–Bernoulli) beam. The full elastic foundation
(i.e, the generalized Smith and Herrmann [31] problem) is, thus, the particular case when the attachment ratio
of the foundation equals unity. Elishakoff and Impollonia found that the dependence of the critical velocity
on the attachment ratio is non-monotonous, thus putting into evidence an additional seemingly paradoxical
behaviour. Ruta and Elishakoff [33] showed that if the pipe is a shear-deformable (or Bresse–Timoshenko)
beam, the majority of this non-monotonicity is not present. In order to justify the remaining cases of non-
monotonicity, De Bellis et al. [34] investigated the pipe attached to a different foundation, specifically the
Wieghardt [35] elastic soil.
Here, we study a pipe conveying a pressurized fluid, with one end either hinged or clamped, and resting on
an intermediate support, thus exhibiting an overhang. The pipe is modelled as a beam; suitable non-dimensional
control parameters provide either purely flexible or shear-deformable beams. Hence, we intend to generalize
studies by Elishakoff and Hollkamp [36], Ari-Gur and Elishakoff [37] and Elishakoff et al. [38] to a non-
conservative problem.
We linearize exact field equations, accounting for fluid–structure interaction; thus, we rationally prove the
validity and limits of the equations in former studies. Standard time-harmonic solutions are searched; the onset
of instability is reached when the real part of one root of the characteristic equation vanishes. A numerical
procedure is implemented and some approximate values for the fluid critical velocity inducing instability are
obtained. Material and external damping, and a full linear elastic Winkler foundation are considered.
Thus, we generalize the problems presented, for instance, in [24], Section 6.8, and [25], Section 9.3. Indeed,
we deal with a more general model, encompassing extensible, shear-deformable and flexible beams; in addition,
we consider a overhang, and two possible conditions at one end (clamped and hinged); finally, our pipe lies
on a continuous viscous-elastic foundation, providing both additional stiffness and damping, thus remarkably
broadening the possible bifurcation scenarios. Some generalization is provided also with respect to the results
provided in [15–17], because in those papers only purely flexible beams are considered, no intermediate
support is present, and only a fixed end is taken into account as an external constraint. On the other hand, always
comparing with [15–17], in this contribution, we limit ourselves to consider non-conservative actions and do not
perform multiple scales; in addition, we are happy with evaluating the critical velocity of the fluid and investigate
its variation with respect to the variation of the meaningful physical quantities affecting the phenomenon.

2 Linearized compatibility and balance equations

A slender pipe of length l and mass per unit volume ρ p has uniform hollow circular cross-section of area A p
and moment of inertia I p with respect to a diameter; hence, it has mass per unit length m p = ρ p A p and relevant
mass moment of inertia J p = ρ p I p . The pipe conveys an incompressible fluid in full turbulent flow with mass
per unit volume ρ f , per unit length m f = ρ f A f . The flow is a non-extensible rope, i.e. an infinitely flexible
rod travelling through the pipe, all points with the same velocity V . This approximation (plug flow [21]) is
valid when the pipe deflections are of long wavelength compared with the diameter of the pipe, considered
slender enough [21]. Inertia is the sum of that of the pipe and the fluid, respectively.
The pipe being a beam and the fluid flow an inextensible rope, all mechanical fields depend on the time
t and the axial coordinate x along the pipe, since we consider ‘small’ deflections about the straight initial
configuration. The pipe end at x = 0 can be either hinged or clamped, the other end at x = l is free and there
is an in between support at x = a, 0 ≤ a ≤ l, Fig. 1.
An overhanging pipe conveying fluid

Fig. 1 A pipe with overhang with two possible constraints at the left end

We derive the equations of motion linearizing compatibility, balance and constitutive equations around the
initial straight, pre-stressed shape, following a perturbation-like technique [5]. In a two-dimensional ambient
space, finite strain measures are [9,39]

u = R ⊤ p′ − q ′ , U = R⊤ R′ (1)

where q and p are the position vector fields of the pipe axis and of the rope-flow in the reference and the
present configuration, respectively, and R is the rotation common to the cross-section of the pipe and the fluid
front; an apex stands for the derivative of the indicated field with respect to x, i.e. (·)′ = ∂(·)/∂ x. Thus, in
Eq. (1), q′ is the tangent vector to the pipe axis in the reference configuration; p′ is the tangent vector to the
pipe axis in the present configuration; u is the strain, i.e. the elongation of the axis and the shearing between
the axis and the cross-section; and U is the change in curvature of the pipe axis and of the rope-flow.
If all fields regularly depend on an evolution parameter s ∈ [0, 1], s = 0, s = 1 characterizing the reference
and present shapes, respectively, we linearize Eq. (1) with respect to s:

u ≈ (I + sR1 )⊤ (q′ + sd1′ ) − q′ ≈ 0 + s(R1⊤ q′ + d1′ )


(2)
U ≈ (I + sR1 )⊤ (I + sR1 )′ ≈ 0 + sR1′

where the subscripts 0 and 1 indicate the relevant field in the reference configuration and its first-order increment.
Thus, R0 = I, the identity, and, if d = p − q is the displacement common to the pipe axis and the rope-flow,
p1 = d1 , since the first-order increment of the reference axis is obviously nil. Equation (2) confirm that, due
to their very definition of differences of geometrical characteristics between the present and the reference
configurations, strain in the reference configuration vanishes and first-order strain depends on first-order
rotation and displacement.
The pipe moves in a horizontal plane spanned by the Cartesian coordinates x, y, thus gravity is discarded;
u(x, t), v(x, t) are the relevant displacement components of the pipe axis and the rope-flow; θ (x, t) is the
rotation of both the pipe cross-section and the fluid front, because of the plug flow. Let e1 , e2 be an orthonormal
basis consistent with the chosen Cartesian frame. For simplicity, being understood that all fields depend on
x, t, when no confusion arises, we drop them from notation. If we perform, as above, the linearization with
respect to s,
   
cos θ (s) − sin θ (s) 0 −1
(R) = ⇒ R0 = I, R1 = θ1 = θ1 e1 ∧ e2 ,
sin θ (s) cos θ (s) 1 0
(3)
u1 = R1⊤ q′ + d1′ = (θ1 e1 ∧ e2 )e1 + u ′1 e1 + v1′ e2 = u ′1 e1 + (v1′ − θ1 )e2 ,
U1 = R1′ = θ1′ e1 ∧ e2

and we obtain the linearized scalar measures of strain. For the sake of simplicity, henceforth, we will drop the
subscript 1 from the fields u, v, θ , implicitly understanding that we are considering their first-order increments
from the reference configuration. Assuming that both the pipe and the rope-flow are inextensible, from Eq. (3),
we let the first-order strain increments be
∂u ∂v ∂θ
ε= = 0, α1 γ = − θ, χ= (4)
∂x ∂x ∂x
with ε the elongation, supposed negligible, γ the shearing strain, and χ the bending curvature; α1 is a control
parameter, equal to 0 for Euler–Bernoulli beams, 1 for Bresse–Timoshenko ones.
M. L. De Bellis et al.

Balance equations of force and moment, pulled back in the reference configurations in finite form, and in
a two-dimensional ambient space are [9,39]
s′ + Us + a = 0, S′ + u + q ′ ∧ s + A = 0
 
(5)
where s, S are contact force and couple measured in the reference configuration, while a, A are bulk force
and couple measured in the reference configuration, and u, U are the finite strains introduced in Eq. (1). If
we suppose, again, that all the fields regularly depend on the evolution parameter s ∈ [0, 1], as above, the
linearization of the finite balance, Eq. (5), leads to two couples of equations of balance of force and moment:
s0′ + a0 = 0, S′0 + q′ ∧ s0 + A0 = 0
(6)
s1′ + U1 s0 + a1 = 0, S′1 + q′ ∧ s1 + A1 = 0
The first two of Eq. (6) describe balance in the reference configuration, accounting for initial states of stress;
the second two of Eq. (6) describe first-order incremental balance.
Since the fluid flow is seen as a rope, only Eqs. (6-1), (6-3), the balance of force in the reference and
perturbed configuration, hold true. We denote by h the friction per unit length of the pipe axis between fluid
and pipe in the reference configuration and by n the reaction between fluid and pipe normal to the pipe wall
per unit length of the pipe axis in the perturbed configuration. Then, if p is the pressure, above atmospheric,
acting on the fluid, Eqs. (6-1) and (6-3) have the following scalar projections on the chosen basis:
0 = 0⎬

p′ A f + h = 0

∂ 2
 
order zero, ∂ order one (7)
0=0 −θ ′ p A f + n − m f +V v = 0⎭
∂t ∂x
Equations (7-1, 3) put into evidence that in the reference configuration, at the zeroth order of the formal s-
parametrization, the only forces supported by the rope-flow are along the axis of the pipe, and the pressure fall
is balanced by the wall friction. On the other hand, Eq. (7-2) states that in the possible perturbed configuration
there is no increment in axial force, since d’Alembert force in the longitudinal direction is neglected because
the fluid velocity V is constant in time. On the other hand, Eq. (7-4) states that there is a first-order increment
in transverse force, due to curvature effect (first addend), normal wall reaction (second addend) and inertia
forces (third addend). Such distinctions were not evident in textbooks such as [21] and not clearly declared
also in recent works such as [33,34]. The correct form of the first-order incremental field equations for the
considered problem in terms of displacement is found, for instance, in [40]. In that paper, however, the authors
adopted a direct variational approach, thus the different contributions of the linearization procedure to the sets
of kinematic, balance and constitutive equations are somewhat hidden in the procedure. Once performed the
necessary identification of symbols, it is immediate to check that Eq. (7) coincide with those present in [24,25]
for the static and dynamic case, respectively. It is also immediate to check that Eqs. (7) coincide with other
standard sets of balance equations for curved beams, for instance in [41].
Both balance equations of force and moment Eq. (6) hold for the pipe. We denote by N , Q, M the normal
and shearing components of the contact force and the (unique) bending component of the contact couple,
respectively. In addition, we denote by f the external force per unit length of the pipe axis. The balance of
force yields two couples of scalar equations:
N0′ + h = 0 N1′ − θ ′ Q 0 = 0
 
order zero, order one (8)
Q ′0 = 0 Q ′1 + θ ′ N0 − n − m p v̈ − f 1 = 0
˙ = ∂(·)/∂t. The balance of
where a superimposed dot stands for the derivative with respect to time, i.e. (·)
moment yields only two scalar equations in a two-dimensional ambient space:
M0′ + Q 0 = 0 order zero, M1′ − α1 γ N0 + Q 1 + α2 J p θ̈ = 0 order one (9)
α2 being another control parameter, equal to 0 or 1 if rotatory inertia is neglected or accounted for, respectively.
We neglect it here, henceforth α2 = 0 and no inertia appears in Eq. (9-2).
By Eqs. (8-1, 3), in the straight reference configuration, the only external source of force is the axial wall
friction between fluid and pipe and external transverse reactions arise only in a perturbed shape. By Eq. (8-2,
4), curvature enters the balance of force in the perturbed configuration (second addend in both equations).
The axial wall friction does not enter first-order incremental balance, as for the rope-flow; on the other hand,
An overhanging pipe conveying fluid

transverse incremental balance of force accounts for normal wall reaction (third addend of Eq. (8-4)), inertia
(fourth addend) and external reaction. The balance of moment in the reference and perturbed shape (9) looks
almost the same, apart from the contribution of the shearing strain in Eq. (9-2). These comments cannot be
found in [21] and are not widely discussed in [33,34]. Again, once performed the necessary identification of
symbols, it is immediate to check that Eqs. (8), (9) coincide with those present in [24,25] for the static and
dynamic case, respectively.
If the right end of the pipe is free and the flow is freely discharged, Eq. (7-1), (8-1) lead to

N0′ − p ′ A f = 0 ⇒ N0 − p A f = const. = 0 (10)

which equals the result of literature [21]. We remark that we obtained it in the reference, pre-stressed configu-
ration, while no such considerations are found in [21]. On the other hand, Eq. (8-3), (9-1) yield, since the right
end of the pipe is free,

Q ′0 = 0 ⇒ Q 0 = const. = 0, M0′ = 0 ⇒ M0 = const. = 0 (11)

and the only state of stress present in the reference shape is parallel to the pipe axis, which is not explicitly
said in [21]. Equations (11-1), (8-2) and the condition of the pipe free right end yield

N1′ = 0 ⇒ N1 = const. = 0 (12)

that is, the normal force in the pipe vanishes also in the possible perturbed configuration. This is not found in
the literature [21], and it is sometimes confused with what stated by (10).
Equations (7-4), (8-4), (10), (9-1), if we denote by m t = m f + m p , the total mass of the system (pipe and
fluid) per unit length of the axis, yield the balance of transverse force and moment

Q ′1 − m t v̈ − 2m f V v˙′ − m f V 2 v ′′ − f 1 = 0, M1′ + Q 1 = 0 (13)

in a possible perturbed configuration. Such equations can be found in the literature [21] as well as in previous
works [33,34], even if they were derived in a less rational way. Once performed the necessary identification
of symbols, and neglecting the unnecessary contributions (for instance, gravity, which is discarded here, or
inertia, when one wants to investigate static phenomena), it is easy to check that Eqs. (13) coincide with those
in [24,25] for the static and dynamic case, respectively.

3 Incremental constitutive relations, perturbed motion

The constitutive relations for the mechanical quantities in the possible perturbed configuration are linear,
because of the first-order incremental nature of the adjacent configuration. On the other hand, nonlinear
constitutive equations may be needed, even in a one-dimensional setting, for investigating other interesting
physical phenomena, such as vibro-acoustics, for instance for investigating guided waves [42] with monitoring
scopes, or in the case of large deformations, to measure coupling effects [43,44]. However, this is not the case,
and linearized constitutive equations suffice for our aim.
The normal force N vanishes because of balance and boundary conditions, see Eqs. (10), (12), and is in
any case purely reactive because of the constraint in Eq. (4-1). We assume linear visco-elastic constitutive
relations for the shearing force and bending moment in the pipe:
   
Ā p G 0 ∂ ∂
Q1 = 1 + c1 α1 γ , M1 = E 0 I p 1 + c2 χ (14)
α3 ∂t ∂t

with Ā p the area of the pipe cross-section modified by the shear shape factor, E 0 , G 0 the elastic moduli in
extension and shear, respectively, c1 , c2 the damping coefficients in translation and rotation, respectively, and
α3 a non-dimensional control parameter equal to 0 or 1, providing a purely flexible and a shear-deformable
beam, respectively. Note that if α3 = 0, the ratio α1 /α3 is undetermined, and Q is not constitutively prescribed
by Eq. (14), in accord with the vanishing shear constraint. The shearing force is then determined only by the
balance Eq. (8).
M. L. De Bellis et al.

The pipe rests on a Winkler [45] foundation, supposed linear visco-elastic with stiffness k and damping
coefficient d. In addition, the environment surrounding the pipe is viscous, with damping coefficient b. Thus,
the first-order total transverse external force per unit length is:
 
∂v ∂
f1 = b + k+d v. (15)
∂t ∂t
If we replace the constitutive relations, Eqs. (14), (15) into Eq. (13), we get the first-order field equations
in terms of the incremental displacement components

∂ 2 ∂ 2v
       
Ā p G 0 ∂ ∂ ∂v ∂ ∂v ∂
1+c1 −θ −m f +V v−m p 2 −b − k +d v =0
α3 ∂t ∂ x ∂ x ∂t ∂x ∂t ∂t ∂t (16)
  2   
∂ ∂ θ Ā p G 0 ∂ ∂v
E 0 I p 1 + c2 2
+ 1 + c1 −θ =0
∂t ∂ x α3 ∂t ∂x
supplemented by the boundary and continuity conditions
 
∂ ∂θ
v(0) = 0, θ (0) = 0, 1+c2 (0) = 0, v(αl − ) = v(αl + ) = 0, θ (αl − ) = θ (αl + ),
∂t ∂ x
∀t (17)
∂ ∂ 2θ
      
∂ ∂θ − ∂θ + ∂ ∂θ
1+c2 (αl )− (αl ) = 0, 1+c2 (l) = 0, 1+c2 (l) = 0
∂t ∂x ∂x ∂t ∂ x ∂t ∂ x 2
where Eq. (17-2) holds for a clamp and Eq. (17-3) for a simple support at the left end.
We search solutions for the perturbed motion around the initial stressed configuration as

v(x, t) = v̂(x) exp( t), θ (x, t) = θ̂ (x) exp( t) (18)

where is the natural angular frequency of the system. By introducing Eq. (18) into Eq. (17), the field equations
for the spatial modes v̂(x), θ̂ (x) are, letting α3 = 1,

Ā p G 0 (1 + c1 ) v̂ ′′ − θ̂ ′ − m t 2
v̂ − 2m f V v̂ ′ − m f V 2 v̂ ′′ − ((b + d) + k)v̂ = 0
(19)
) θ̂ ′′ + Ā p G 0 (1 + c1 ) v̂ ′ − θ = 0
 
E 0 I p (1 + c2

supplemented by the boundary and continuity conditions

v̂(0) = 0, θ̂(0) = 0, (1 + c2 ) θ̂ ′ (0) = 0, v̂(αl − ) = v̂(αl + ) = 0, θ̂ (αl − ) = θ̂ (αl + ),


(20)
(1 + c2 ) θ̂ ′ (αl − ) − θ̂ ′ (αl + ) = 0, (1 + c2 ) θ̂ ′ (l) = 0, (1 + c1 ) v̂ ′ (l) − θ̂(l) = 0

where Eq. (20-2) holds for a clamp and Eq. (20-3) for a simple support at the left end.
To obtain a non-dimensional form of Eqs. (19), (20), let us multiply Eq. (19-1) by l 3 /E 0 I p , Eq. (19-2) by
2
l /E 0 I p and introduce the following quantities:

x v̂ mf Ā p G 0 l 2 mt kl 4
ξ= ; η = ; ϑ = θ̂ ; μ = ; λ= ; ω= l 2; κ = ;
l l mt E0 I p E0 I p E0 I p
(21)
ci E0 I p dl 2 bl 2 mf
γi = 2 , i = 1, 2; δ =  ; β= ; υ = Vl .
l mt E0 I p m t E0 I p m t E0 I p

We denote ν = (E 0 − 2G 0 )/2G 0 , Re , Ri Poisson’s ratio and the outer and inner radii of the pipe and of
its cross-section, respectively. The following notations and results hold [21,33,34,46]

l Ri 48σ 2 (1 + ζ 2 )ζ 2 1
σ = ; ζ = ; λ= ; μ= . (22)
2Ri Re (7 + 6ν)(1 + ζ 2 )2 + (20 + 12ν)ζ 2 ρp 1
1+ ρf ζ2
−1
An overhanging pipe conveying fluid

For simplicity in notation, we keep denoting by a prime the derivative with respect to ξ . Replacing Eq. (21)
into Eq. (19), we get the dimensionless equations of perturbed motion

λ(1 + γ1 ω)(η′ − ϑ)′ − ω2 η − υ 2 η′′ − 2υμωη′ − [(β + δ)ω + κ]η = 0


(23)
(1 + γ2 ω)ϑ ′′ + λ(1 + γ1 ω)(η′ − ϑ) = 0

that can be seen as a particular case of those presented by the authors previously [33,34]; Eq. (23) are supple-
mented by the relevant dimensionless boundary and continuity conditions

η(0) = 0, ϑ(0) = 0, ϑ ′ (0) = 0, η(α − ) = η(α + ) = 0, ϑ(α − ) = ϑ(α + ),


(24)
ϑ ′ (α − ) = ϑ ′ (α + ), ϑ ′ (1) = 0, η′ (1) = ϑ(1)

where Eq. (24-2) holds for a clamp and Eq. (24-3) for a simple support at the left end.
Equations (23), (24) are non-self-adjoint: the symbolic differential operator providing the differential
equations and boundary conditions, Eqs. (23), (24), does not commute when a variational technique is adopted.
The behaviour of the solutions is known to be ruled by the real part of ω. If Re(ω) > 0, the displacements
amplitude increases, and flutter occurs. If Re(ω) < 0, the solutions are decaying oscillations and the system
returns to a stable configuration.
Let the physical quantity controlling the system be the dimensionless velocity υ, and suppose all the other
parameters fixed. When υ = 0, the system is stable, in the sense that no motion arises. If υ increases, the
onset of instability is reached, when damping is present, as the real part of one of the non-dimensional angular
frequencies attains zero from below. In an undamped system, it is well known that the bifurcation mechanism
is different from the previous one, since when υ increases, all the eigenvalues lie on the imaginary axis until
a critical value for υ is reached. The lowest value υcr inducing either static or dynamic bifurcation, known as
the critical velocity, is numerically searched in the following. The effect of the location of the intermediate
support on the boundaries of instability is also investigated.

4 Exact solution

Because of the intermediate support, we split the domain into two, introducing the notations
 
η1 (ξ ), 0 ≤ ξ < α ϑ1 (ξ ), 0 ≤ ξ < α
η(ξ ) = ϑ(ξ ) = (25)
η2 (ξ ), α < ξ ≤ 1 ϑ2 (ξ ), α < ξ ≤ 1.

into the modal equations provided by Eq. (23). We obtain

λ(1 + γ1 ω)(ηi′ − ϑi )′ − υ 2 ηi′′ − 2υμωηi′ − [(β + δ)ω + κ + ω2 ]ηi = 0


i = 1, 2 (26)
(1 + γ2 ω)ϑi′′ + λ(1 + γ1 ω)(ηi′ − ϑi ) = 0

where i = 1 denotes fields in the span, i.e. the domain 0 ≤ ξ < α, while i = 2 denotes fields in the overhang,
i.e. the domain α < ξ ≤ 1. The boundary conditions Eq. (24), supplemented by the continuity conditions of
displacement, rotation, force and moment at ξ = α, become

η1 (0) = 0, ϑ1 (0) = 0, ϑ1′ (0) = 0, η1 (α − ) = η2 (α + ) = 0, ϑ1 (α − ) = ϑ2 (α + ),


(27)
ϑ1′ (α − ) = ϑ2′ (α + ), ϑ2′ (1) = 0, η2′ (1) = ϑ2 (1)

where Eq. (27-2) holds for a clamp and Eq. (27-3) for a simple support at the left end.
We simplify Eq. (26) by obtaining ηi′ in Eq. (26-2) and replacing it into Eq. (26-1):

υ2
 
′′′′ 2υμω(1 + γ2 ω) ′′′
(1 + γ2 ω) 1− ϑ − ϑi
λ(1 + γ1 ω) i λ(1 + γ1 ω)
  i = 1, 2.
(1 + γ2 ω) ′′ ′
+ υ 2 + −[(β + δ)ω + κ + ω2 ] ϑi + 2ωυμϑi +[(β + δ)ω + κ + ω2 ]ϑi = 0,
λ(1 + γ1 ω)
(28)
M. L. De Bellis et al.

We obtain the same equations in the unknowns η1 , η2 by obtaining ϑ1′ in Eq. (26-1) and replacing it into
Eq. (26-2). Equation (28) coincide, modulo the different physical quantities involved, with Eq. (20) in [34].
Equation (28) is linear and admit, Θ1 , Θ2 being constants, the solutions
ϑ1 (ξ ) = Θ1 exp(r1 ξ ), ϑ2 (ξ ) = Θ2 exp(r2 ξ ), (29)
By substituting the positions of Eq. (29) into the modal Eq. (28), we get
υ2
 
2υμω(1 + γ2 ω) 3
(1 + γ2 ω) 1− ri4 − ri
λ(1 + γ1 ω) λ(1 + γ1 ω)
  i = 1, 2. (30)
(1 + γ2 ω)
+ υ 2 + −[(β + δ)ω + κ + ω2 ] ri2 + 2ωυμri + [(β + δ)ω + κ + ω2 ] = 0,
λ(1 + γ1 ω)
Each of the characteristic equations (30) admits four complex conjugate solutions ri j . The index i takes
the values 1 or 2. The index j = 1, 2, 3, 4 indicates the different roots of the ith equation. The solution of
Eq. (26) with the notations in Eq. (25) become
4

ϑi (ξ ) = Θi j exp(ri j ξ ), i = 1, 2 (31)
j=1

Θi j being constants. For each of the characteristic roots r1 j , r2 j , Eq. (26) yields
4  
 1 (1 + γ2 ω) 2
ηi (ξ ) = Hi j exp(ri j ξ ), Hi j = 1− ri j Θi j (32)
ri j λ(1 + γ1 ω)
j=1

Hi j being constants. Equation (32) can be shortened as


Hi j = L i j Θi j , i = 1, 2; j = 1, 2, 3, 4, no sum on indices. (33)
The boundary conditions (27), by the notations in Eqs. (31), (33), are
4
 4
 4

L 1 j Θ1 j = 0, Θ1 j = 0 (clamped end), Θ1 j r1 j = 0 (hinged end),
j=1 j=1 j=1
4 4

L 1 j Θ1 j exp(r1 j α) = L 2 j Θ2 j exp(r2 j α) = 0,
j=1 j=1
4 4 4 4 (34)
   
Θ1 j exp(r1 j α) = Θ2 j exp(r2 j α), Θ1 j r1 j exp(r1 j α) = Θ2 j r2 j exp(r2 j α),
j=1 j=1 j=1 j=1
4 4
  
Θ2 j r2 j exp(r2 j ) = 0, L 2 j r2 j − 1 Θ2 j exp(r2 j ) = 0,
j=1 j=1

a system of linear algebraic equations in the unknowns Θi j , i = 1, 2; j = 1, 2, 3, 4. Its matrices of coefficients


Ai j are in Tables 1 and 2, where K 2 j = (L 2 j r2 j − 1) exp(r2 j ), j = 1, 2, 3, 4

Table 1 Matrix of the coefficients of the boundary conditions (34), clamped left end

L 11 L 12 L 13 L 14 0 0 0 0
1 1 1 1 0 0 0 0
L 11 er11 α L 12 er12 α L 13 er13 α L 14 er14 α 0 0 0 0
0 0 0 0 L 21 er21 α L 22 er22 α L 23 er23 α L 24 er24 α
er11 α er12 α er13 α er14 α −er21 α −er22 α −er23 α −er24 α
r11 er11 α r12 er12 α r13 er13 α r14 er14 α −r21 er21 α −r22 er22 α −r23 er23 α −r24 er24 α
0 0 0 0 r21 er21 r22 er22 r23 er23 r24 er24
0 0 0 0 K 21 K 22 K 23 K 24
An overhanging pipe conveying fluid

Table 2 Matrix of the coefficients of the boundary conditions (34), hinged left end

L 11 L 12 L 13 L 14 0 0 0 0
r11 r12 r13 r14 0 0 0 0
L 11 er11 α L 12 er12 α L 13 er13 α L 14 er14 α 0 0 0 0
0 0 0 0 L 21 er21 α L 22 er22 α L 23 er23 α L 24 er24 α
er11 α er12 α er13 α er14 α −er21 α −er22 α −er23 α −er24 α
r11 er11 α r12 er12 α r13 er13 α r14 er14 α −r21 er21 α −r22 er22 α −r23 er23 α −r24 er24 α
0 0 0 0 r21 er21 r22 er22 r23 er23 r24 er24
0 0 0 0 K 21 K 22 K 23 K 24

In both cases, it is well known that to ensure non-trivial solutions, i.e. for a perturbed configuration to exist,
the determinant of Ai j must vanish:
det(Ai j )(ri j , α, β, γi , δ, κ, λ, μ, υ, ω) = 0, i = 1, 2, j = 1, 2, 3, 4. (35)
The solutions of the characteristic Eq. (35) will be the non-dimensional natural angular frequencies of the pipe
conveying fluid in terms of all the physical parameters involved. However, Eq. (35) is highly nonlinear and its
solutions cannot be found in closed form, so we adopt a numerical technique, which requires a starting value
of the critical velocity.

5 Approximate critical velocities

We adopt Galerkin’s technique to obtain approximate critical velocities of the fluid. The field Eq. (28) for the
rotation of the pipe cross-section and of the fluid front is shortened as
L(ϑi ) = 0, i = 1, 2 (36)
with L a suitable symbolic differential operator. We approximate the solution of Eq. (36) by the sum of a finite
number of comparison functions, for instance, the Duncan polynomials [47] for the corresponding purely
flexible, Euler–Bernoulli beam. Indeed, the approximate value is only a starting point for an exact evaluation
of the roots of Eq. (35); thus, we opt for a simpler way of determining the comparison functions. Duncan
polynomials for a purely flexible beam are, if h is an integer and D = (h + 1)(h + 2)(h + 3)(h + 4),

ξ 3 (h + 3)(h + 4)(h(α − 1) + 2α − 1) − 2α h+2


 
ξ h+4
ϑ1h (ξ ) = −
D 4Dα
ξ 2 (h + 3)(h + 4)(h(α − 1) + 2α − 1) − 6α h+2
 
+ ,
4D (37)
ξ (h + 3)(h + 4)α(h(α − 3) + 2α − 3) − 6α h+3
 
ξ h+4 ξ3 ξ2
ϑ2h (ξ ) = − + +
D 6(h + 1) 2(h + 2) 4D
2
(h + 3)(h + 4)α (h(α − 3) + 2α − 3) − 6α h+4
− .
12D
for a clamped left end, while for a hinged supported left end (whence the subscript s)

ξ h+4 ξ 3 (1 − αh + h − 2α) ξ (h + 3)(h + 4)α 2 (h(α − 1) + 2α − 1) − 6α h+4


 
s
ϑ1h = + + ,
D 6(h + 2)(h + 1)α 6Dα
s ξ h+4 ξ3 ξ2 α2
ϑ2h = − + + (38)
D 6( j + 1) 2h + 4 6(h + 2)
ξ (h + 3)(h + 4)α 2 (h(α − 4) + 2α − 4) − 6α h+4
 
+
6Dα
To save space, we present the procedure when the left end is clamped; the same goes for a hinge, utilizing
the different comparison functions, Eqs. (38) instead of Eqs. (37). Chosen n polynomials (37) for the two
sub-domains 0 ≤ ξ < α, α < ξ ≤ 1, we pose
M. L. De Bellis et al.


ϑ1h (ξ ), 0 ≤ ξ < α
ϑ̄h (ξ ) = (39)
ϑ2h (ξ ), α < ξ ≤ 1

so that, if Bh are unknown coefficients, we may write the approximate expression


n

Z (ξ ) = Bh ϑ̄h (ξ ). (40)
h=1

If we substitute the approximate sum (40) into Eq. (28), or into its symbolic expression (36), we do not get
zero, but a measure of the error ǫ made in the approximation (40):
n
 n

ǫ(ξ ) = L(Z (ξ )) = Bh L(ϑ̄h (ξ )) (41)
h=1 h=1

Following Galerkin technique, we cannot reduce the error function to zero, but rather its weighted mean.
The chosen weights are the comparison functions approximating the solutions:
1
ǫ(ξ )ϑ̄k (ξ )dξ = 0 ∀k = 1, 2, . . . , n (42)
0

By inserting the error, Eq. (41), in its weighted mean, Eq. (42), we get
n
 1
Bh L(ϑ̄h (ξ ))ϑ̄k (ξ )dξ = 0 ∀k = 1, 2, . . . , n ⇒ Ckh Bh = 0, (43)
h=1 0

a system of n algebraic equations in the n unknowns Bh , admitting non-trivial solutions only when the rank
of the n × n matrix of its coefficients is not maximum. This equals to admitting det(Ckh ) = 0, and let us find
the searched approximate value of the critical fluid velocity.

6 Numerical results: comments and discussion

We examine the influence of the non-dimensional parameters α, β, κ, λ, μ on the critical values of the fluid
velocity; for simplicity, we will consider the effect of γ1 , γ2 in future investigations. We show the results for
a clamped left end first, then those for a hinged left end.
Figure 2 exhibits the effect of the non-dimensional external damping factor β as the non-dimensional
location of the intermediate support varies. We pose δ = 0.005; γ1 = 0.001; γ2 = 0.00005; μ = 0.68; λ =
2000; κ = 100 for the other non-dimensional parameters involved. As known in the literature [36], the location

30
Critical flow velocity [-]

25

20 Flutter
β=20
β=25
15 β=30

Divergence
10 β=20-25-30

0 0.2 0.4 0.6


α
Fig. 2 Critical velocity versus support location when external damping varies—clamped left end
An overhanging pipe conveying fluid

Critical flow velocity [-]


30
Flutter
κ=50
κ=100
κ=200
20 Divergence
κ=50
κ=100
κ=200
10

0 0.2 0.4 0.6


α
Fig. 3 Critical velocity versus support location if the foundation stiffness varies—clamped left end

30
Critical flow velocity [-]

25
Flutter
λ=500
20
λ=2000
λ=4000
15 Divergence
λ=500
λ=2000
10 λ=4000

0 0.2 0.4 0.6


α
Fig. 4 Critical velocity versus support location if the tube stiffness ratio varies—clamped left end

of the intermediate support is a physical parameter affecting the possible instability of clamped beams under
non-conservative forces. We see from Fig. 2 that the damping factor does not affect the location of the support
αt inducing transition from flutter to buckling. The value of this non-dimensional location equals 0.351 and
is not much different from that provided by Elishakoff and Hollkamp [36] for a different, maybe less realistic,
problem. That β does not affect the limit value αt is reasonable from a physical point of view. Indeed, damp-
ing has influence on dynamics, while buckling and its boundary are a feature of the static behaviour of the
considered system. We clearly see from Fig. 2 that the lower bound of the buckling instability region is not
affected by the changing value of β and the three curves coincide. On the other hand, the dependence of the
critical velocity on the location of the intermediate support is non-monotonic, and the damping factor seems to
have a destabilizing effect: the critical velocity of the fluid decreases, for the same location of the intermediate
support, while β increases.
Figure 3 shows the effect of the non-dimensional foundation stiffness κ as the non-dimensional location
of the intermediate support varies. We pose β = 25; δ = 0.005; γ1 = 0.001; γ2 = 0.00005; μ = 0.68; λ =
2000. It is apparent that κ affects both the location of the intermediate support inducing transition and the bound-
aries of stability. As a matter of fact, when κ grows, αt moves from 0.352 to 0.349, thus slightly shortening the
amplitude of the domain of flutter: this is reasonable from the physical point of view, since stiffening the system
should basically enlarge the stability domain and, in particular, should also enlarge the domain of buckling. On
the other hand, increasing κ also means that the critical velocity inducing both flutter and buckling increases,
i.e. the foundation stiffness has a stabilizing effect, which is again reasonable from a physical point of view.
The dependence of the critical velocity with the location of the intermediate support remains non-monotonic.
Figure 4 shows the effect of the non-dimensional shearing-to-bending stiffness ratio λ as the non-
dimensional location of the intermediate support varies. We pose β = 25; δ = 0.0; γ1 = 0.0; γ2 = 0.0;
M. L. De Bellis et al.

30

Critical flow velocity [-]


25
Flutter
μ=0.57
20 μ=0.68
μ=0.78

15 Divergence
μ=0.57
μ=0.68
10 μ=0.78

0 0.2 0.4 0.6


α
Fig. 5 Critical velocity versus support location if the mass density ratio varies—clamped left end

μ = 0.68; κ = 100. Figure 4 shows that λ does not vary the value of αt , always equal to 0.351, which is
physically reasonable. Indeed, the ratio of the shearing and bending stiffnesses affects the deformation modes
of the pipe, i.e. if we may consider it a purely flexible, Euler–Bernoulli, beam or a shear-deformable, Bresse–
Timoshenko, beam. On the other hand, λ cannot affect the values of such external parameters as the location
of the intermediate support inducing transition from flutter to buckling. The dependence of the critical velocity
with the location of the intermediate support is always non-monotonic, and we may see that when λ grows
sensibly the two curves approach each other. As a matter of fact, the value of λ provided by Eq. (22) makes it
clear that, once fixed the material (i.e. Poisson’s ratio ν) of the pipe, the geometry of the pipe (i.e. the outer and
inner radii of its tubular cross-section, hence non-dimensional ratio ζ in Eq. (22-2)), the shearing-to-bending
stiffness ratio is proportional to the square of the non-dimensional ratio σ [see Eq. (22-1)]. Thus, a greater
value of λ means a greater value of σ and, if the inner radius is fixed, a slimmer pipe; we know, then, that in
such cases shearing strains become negligible, and Euler–Bernoulli beams suffice as models. Thus, the blue
and black line in Fig. 4 both approach to the common limit of an Euler–Bernoulli beam model; on the other
hand, the red curve in Fig. 4 shows the effect of Bresse–Timoshenko beam model corrections when the pipe is
not slim enough. The non-dimensional critical velocity sensibly decreases with respect to the values it assumes
in the other cases, and slimmer pipes seem to have larger stability domains: this can be explained by the fact
that Euler–Bernoulli beams are somehow stiffer than the corresponding shear-deformable ones.
Figure 5 shows the effect of the ratio of fluid-to-total mass density μ as the non-dimensional location of the
intermediate support varies. We pose β = 25; δ = 0.0; γ1 = 0.0; γ2 = 0.0; λ = 2000; κ = 100. We find that
actually μ slightly affects the value of αt , varying from 0.352 when μ = 0.57 to 0.351 when μ = 0.68 to 0.350
when μ = 0.78. This is physically reasonable, since the mass affects the inertial forces turning the possible
instability process on, and the bigger the fluid mass, the longer the overhang to pass from flutter to buckling.
The dependence of the critical velocity with the location of the intermediate support remains non-monotonic,
and we may see that when μ grows the boundary of buckling is unaffected, that is again physically reasonable,
since the mass should affect inertia, hence, the dynamic behaviour. On the other hand, in the domain of flutter,
the variation of the mass ratio strongly affects the value of the critical velocity, and a bigger value of μ seems
to have a stabilizing effect on the system.
Let us pass to a hinged left end; for all the cases, we will provide comments and compare the results with
the preceding ones obtained for a pipe clamped at the left end.
Figure 6 exhibits the effect of the non-dimensional external damping factor β as the non-dimensional
location of the intermediate support varies. As for the clamped pipe, we pose δ = 0.005; γ1 = 0.001; γ2 =
0.00005; μ = 0.68; λ = 2000; κ = 100 for the other non-dimensional parameters involved. We see from
Fig. 6 that: (a) the damping factor does not affect, again, the location of the support αt inducing transition
from flutter to buckling; (b) αt significantly changes with respect to the clamped case, moving to 0.505; (c) the
critical values of the fluid velocity globally decrease, sometimes dramatically, for both flutter and buckling.
This seems to reflect the physics of the problem, since this system is much softer, from the point of view of
global rigidity, than the clamped pipe: in particular, dealing with buckling, the critical load may reduce to
approximately one half of the corresponding one for the clamped pipe, keeping the length of the interval fixed.
Again, β not affecting αt is reasonable from a physical point of view, and the lower bound of the buckling
An overhanging pipe conveying fluid

14

Critical flow velocity [-]


12
Flutter
β=20
10 β=25
β=30
8
Divergence
β=20-25-30
6

4
0 0.2 0.4 0.6 0.8 1
α
Fig. 6 Critical velocity versus support location if external damping varies—hinged left end

14
Flutter
Critical flow velocity [-]

12 κ=50
κ=100
κ=200
10
Divergence
κ=50
8 κ=100
κ=200
6

4
0 0.2 0.4 0.6 0.8 1
α
Fig. 7 Critical velocity versus support location if the foundation stiffness varies—hinged left end

instability region is not affected by the changing value of β, so that the three curves coincide. We may also see
that, from α ≈ 0.4, the curves representing the boundaries of flutter tend also to coincide: thus, approaching
the transition value of the location of the intermediate support renders the system almost unaffected by the
damping factor, like the system itself prepares to the transition to buckling. The dependence of the critical
velocity on α is non-monotonic, and the damping seems to be destabilizing. Remark that, differently from
[37], when α → 0 the system is not a mechanism any more, since the pipe rests on an elastic foundation; thus,
the critical velocity does not start from zero, like there.
Figure 7 shows the effect of the non-dimensional foundation stiffness κ as the non-dimensional location
of the intermediate support varies. We pose β = 25; δ = 0.005; γ1 = 0.001; γ2 = 0.00005; μ = 0.68; λ =
2000. We see that κ affects the boundaries of stability, while it does not affect sensibly the value of the support
location inducing transition from flutter to buckling. As a matter of fact, and in a manner similar to what we
already saw in the case of the clamped left end, when κ grows, we are, in principle, stiffening the system.
Thus, as a result, we basically enlarge both the stability domain and the domain of buckling. Starting from
α ≈ 0.4, the curves representing the boundaries of flutter tend to coincide: a comment similar to that for
Fig. 6 holds. Indeed, this behaviour is physically justified in that, approaching the transition to buckling, the
difference in the behaviour of the system tends to reduce. On the other hand, increasing κ also means that
the critical velocity inducing both flutter and buckling increases, i.e. the foundation stiffness has a stabilizing
effect, which is again reasonable from a physical point of view. The dependence of the critical velocity on the
location of the intermediate support remains non-monotonic, and the critical values are always lower than the
corresponding ones for the clamped left end. Again, when α → 0, the critical velocity is not zero, but quickly
decreases with κ, as it is reasonable when comparing with the problem studied in [37].
M. L. De Bellis et al.

14
Flutter
=500

Critical flow velocity [-]


12 =2000
=4000
10
Divergence
=500
8
=2000

=4000
6

4
0 0.2 0.4 0.6 0.8 1
α
Fig. 8 Critical velocity versus support location if the tube stiffness ratio varies—hinged left end

14
Flutter
Critical flow velocity [-]

12 μ=0.57
μ=0.68

10 μ=0.78
Divergence
μ=0.57
8
μ=0.68
μ=0.78
6

4
0 0.2 0.4 0.6 0.8 1
α
Fig. 9 Critical velocity versus support location if the mass density varies—hinged left end

Figure 8 shows the effect of the non-dimensional shearing-to-bending stiffness ratio λ as the non-
dimensional location of the intermediate support varies. We pose β = 25; δ = 0.0; γ1 = 0.0; γ2 = 0.0; μ =
0.68; κ = 100. Figure 8 shows that λ does not vary the value of αt , always equal to 0.505, which is physically
reasonable. We may provide the same comments on the dependence on λ of the beam model adopted as we
did for the pipe with the left clamped end. The dependence of the critical velocity on the location of the inter-
mediate support is always non-monotonic, and the non-dimensional critical velocity sensibly decreases with
respect to the values it assumes for the clamped left end. Again, slimmer pipes seem to have larger stability
domains, because Euler–Bernoulli beams are somehow stiffer than the shear-deformable ones; in addition,
when α → 0, the critical velocity is not zero.
Figure 9 shows the effect of the ratio of fluid-to-total mass density μ as the non-dimensional location of
the intermediate support varies. We pose β = 25; δ = 0.0; γ1 = 0.0; γ2 = 0.0; λ = 2000; κ = 100. It seems
that μ does not sensibly affect the transition value αt : thus, this softer system does not seem affected by the
considered physical parameters with respect to the pipe with the clamped left end. The dependence of the
critical velocity on α remains non-monotonic, and when μ grows, the boundary of buckling is unaffected, like
in the case of a clamped left end, that is physically reasonable; again, a bigger value of μ seems to stabilize
the system, and when α → 0, the critical velocity is not zero.
In [37], the effect of transverse shear deformation on the buckling of columns with overhang is discussed.
The authors found that when the location of the intermediate support tends to zero, i.e. when the support
approaches the left end hinge, thus realizing a clamped left end, the critical value of the buckling load approaches
zero from above for a very shear-deformable beam. We wish to extend those results, and we show them in Fig. 10,
where, for a fixed, very low value of the location of the intermediate support, we evaluate the critical velocity
versus the non-dimensional stiffness ratio. It is apparent that, for a nonzero value of the non-dimensional
An overhanging pipe conveying fluid

Critical flow velocity [-]


6

4
κ=100

2 κ=0

0
0 1000 2000 3000 4000
λ
Fig. 10 Effect of stiffness ratio as the intermediate support tends to the hinged end

30
Critical flow velocity [m/s]

25

20

15
EXPER.

NUMER.
10

0 0.1 0.2 0.3


l/L
Fig. 11 Comparison of numerical and experimental results

foundation stiffness (blue line), when the pipe stiffness ratio increases (i.e. when the beam is slim), the critical
load presents a plateau, corresponding to the purely flexible, Euler–Bernoulli model. On the other hand, when
the stiffness ratio goes to zero (i.e. when the beam is stout), the critical load decreases, yet it does not go to
zero, since, as already remarked, the system is not a mechanism any more because of the foundation. When
no foundation is present (red line), the critical load goes to zero, fully recovering the results provided by [37].
To end with, we make a comparison of the results numerically obtained here with those find by experiments
in [28], and, in particular, the experiments about the hinged pipe with the intermediate support. To this aim,
we simply inserted a vanishing foundation in our equations and adopted a minimal damping coefficient; the
other non-dimensional coefficients were obtained by properly inserting the dimensional values suggested by
[28] into Eqs. (21, 22), with the suitable conversion coefficients.
The results of the comparison are shown in Fig. 11 and are limited to a short range of the possible location
of the intermediate support because of the poor validity of experimental results beyond a limit value of this
location, as remarked in [28]. From Fig. 11, it is apparent that, within the limits of the precision of the model
and of the numerical code, the two sets of results are in good agreement.

7 Concluding remarks

We presented an investigation of the dynamic stability for a tube conveying a pressurized fluid, clamped or
hinged at one end and exhibiting an overhang, supported by a visco-elastic Winkler foundation. We derived
the field equations of the problem in a rational way by a perturbation technique and turned them into a non-
dimensional form. We decided to find the solution in a numerical form and adopted a Galerkin technique with
polynomial comparison functions called after Duncan. We presented the lower bounding regions of flutter and
M. L. De Bellis et al.

buckling instability as the intermediate support location and other physical parameters vary, thus providing
some insight on the phenomenon and generalizing the results by Elishakoff et al. [37,38].
Further investigations on the dependence on other boundary conditions and on different foundation models
are in due course and will be reported elsewhere.

References

1. Euler, L.: De Curvis Elasticis. Additamentum Primum, Methodus Inveniendi Lineas Curvas Maximi Minimive Proprietate
Gaudentes, …. Bousquet & Co., Lausanne & Geneve (1744)
2. Koiter, W.T.: Over de stabiliteit van het elastisch evenwicht, Thesis, Delft, Amsterdam, 1945. English translation: On the
Stability of Elastic Equilibrium, NASA Technical Translation F-10, 833, Clearinghouse, US Dept. of Commerce/Nat. Bur.
of Standards N6725033 (1967)
3. Timoshenko, S.P., Gere, J.M.: Theory of Elastic Stability. McGraw-Hill, New York (1961)
4. Budiansky, B.: Theory of buckling and postbuckling behavior of elastic structures. In: Yih, C.S. (ed.) Advances in Applied
Mechanics, vol. 14, pp. 1–65. Academic Press, New York (1974)
5. Pignataro, M., Rizzi, N.L., Luongo, A.: Stability, Bifurcation and Postcritical Behaviour of Elastic Structures. Elsevier, Ams-
terdam (1991)
6. Pignataro, M., Ruta, G.: Coupled instabilities in thin-walled beams: a qualitative approach. Eur. J. Mech. A/Solids 22, 139–
149 (2002)
7. Elishakoff, I., Ruta, G., Stavsky, Y.: A novel formulation leading to closed-form solutions for buckling of circular plates. Acta
Mechanica 185, 81–88 (2006)
8. Elishakoff, I., Ruta, G.: Buckling of a beam on a Wieghardt foundation. ZAMM (Zeitschrift für Angewandte Mathematik
Und Mechanik) 86, 617–627 (2006)
9. Ruta, G., Pignataro, M., Rizzi, N.: A direct one-dimensional beam model for the flexural-torsional buckling of thin-walled
beams. J. Mech. Mater. Struct. 1:8, 1479–1496 (2006)
10. Ruta, G., Pignataro, M., Rizzi, N.: A beam model for the flexural–torsional buckling of thin-walled members. Thin-Walled
Struct. 46, 816–822 (2008)
11. Ruta, G., Varano, V., Pignataro, M., Rizzi, N.: The effects of warping constraints on the buckling of thin-walled structures. J.
Mech. Mater. Struct. 4, 1711–1727 (2009)
12. Eremeyev, V.A., Freidin, A.B., Sharipova, L.L.: The stability of the equilibrium of two-phase elastic solids. J. Appl. Math.
Mech. 71(1), 61–84 (2007)
13. Paolone, A., Vasta, M., Luongo, A.: Flexural–torsional bifurcations of a cantilever beam under potential and circulatory
forces: part I. Nonlinear model and stability analysis. Int. J. Nonlinear Mech. 41(4), 586–594 (2006)
14. Paolone, A., Vasta, M., Luongo, A.: Flexural–torsional bifurcations of a cantilever beam under potential and circulatory
forces: part II. Post-critical analysis. Int. J. Non-Linear Mech. 41(4), 595–604 (2006)
15. Di Egidio, A., Luongo, A., Paolone, A.: Linear and nonlinear interactions between static and dynamic bifurcations of damped
planar beams. Int. J. Non-Linear Mech. 42(1), 88–98 (2007)
16. Luongo, A., D’Annibale, F.: Bifurcation analysis of damped visco-elastic planar beams under simultaneous gravitational
and follower forces. Int. J. Modern Phys. B 26(25), (2012)
17. Luongo, A., D’Annibale, F.: Double zero bifurcation of non-linear viscoelastic beams under conservative and non-
conservative loads. Int. J. Non-Linear Mech. 55, 128–139 (2013)
18. Cazzani, A.: On the dynamics of a beam partially supported by an elastic foundation: an exact solution-set. Int. J. Struct.
Stab. Dyn. 13(8), Art. no. 1350045 (2013)
19. Lofrano, E., Paolone, A., Ruta, G.: Stability of non-trivial equilibrium paths of beams on partial visco-elastic foundation. Acta
Mechanica 223, 2183–2195 (2012)
20. Koiter, W.T.: Unrealistic follower forces. J. Sound Vib. 194(4), 636–638 (1996)
21. Païdoussis, M.P.: Fluid-Structure Interactions. Volume 1: Slender Structures and Axial Flow. Academic Press, New
York (1998)
22. Tornabene, F., Marzani, A., Viola, E., Elishakoff, I.: Critical flow speeds of pipes conveying fluid by the generalized
quadrature method. Adv. Theor. Appl. Mech. 3(3), 121–138 (2010)
23. Feodosiev, V.I.: Advanced Stress and Stability Analysis. Worked Examples. Springer, Berlin (2005)
24. Svetlitsky, V.A.: Statics of Rods. Springer, Berlin (2000)
25. Svetlitsky, V.A.: Dynamics of Rods. Springer, Berlin (2005)
26. Perelmuter, A.V., Slivker, V.: Handbook of Mechanical Stability in Engineering (3 vols). World Scientific, New Jersey (2013)
27. Edelstein, W.S., Chen, S.S.: Flow-induced instability of an elastic tube with a variable support. Nuclear Eng. Des. 84, 1–
11 (1985)
28. Jendrzejczyk, J.A., Chen, S.S.: Experiments on tubes conveying fluids. Thin-Walled Struct. 3, 109–134 (1985)
29. Lottati, I., Kornecki, A.: The effect of an elastic foundation and of dissipative forces on the stability of fluid-conveying
pipes. J. Sound Vib. 109(2), 327–338 (1986)
30. Elishakoff, I.: Controversy associated with the so-called “follower forces”: critical overview. Appl. Mech. Rev. 58, 117–
142 (2005)
31. Smith, T.E., Herrmann, G.: Stability of a beam on an elastic foundation subjected to a follower force. ASME J. Appl.
Mech. 39, 628–629 (1972)
32. Elishakoff, I., Impollonia, N.: Does a partial elastic foundation increase the flutter velocity of a pipe co)nveying fluid?. ASME
J. Appl. Mech. 68, 206–212 (2001)
33. Ruta, G., Elishakoff, I.: Towards the resolution of the Smith–Herrmann paradox. Acta Mechanica 173, 89–105 (2004)
An overhanging pipe conveying fluid

34. De Bellis, M.L., Ruta, G., Elishakoff, I.: Influence of a Wieghardt foundation on the dynamic stability of a fluid conveying
pipe. Arch. Appl. Mech. 80, 785–801 (2010)
35. Wieghardt, K.: Über den balken auf nachgiebiger Unterlage. ZAMM 2(3), 165–184 (1922); (in German)
36. Elishakoff, I., Hollkamp, J.: Computerized symbolic solution for a nonconservative system in which instability occurs by
flutter in one range of the parameter and by divergence in the other. Comp. Meth. Appl. Mech. Eng. 62, 27–46 (1987)
37. Ari-Gur, J., Elishakoff, I.: On the effect of shear deformation on buckling of columns with overhang. J. Sound Vib. 139, 165–
169 (1990)
38. Elishakoff, I., Ari-Gur, J., Das, P.S.: Refined theories maybe needed for vibration analysis of structures with overhang. Int.
J. Solids Struct. 36, 3581–3589 (1999)
39. Antman, S.S.: The theory of rods. In: Truesdell C. (ed.) Handbuch der Physik, vol. VIa/2, pp. 641–703. Springer, Berlin
(1972)
40. Laithier, B.E., Païdoussis, M.P.: The equations of motion of initially stressed Timoshenko tubular beams conveying fluid. J.
Sound Vib. 79(2), 175–195 (1981)
41. Cerri, M.N., Dilena, M., Ruta, G.: Vibration and damage detection in undamaged and cracked circular arches: experimental
and analytical results. J. Sound Vib. 314, 83–94 (2008)
42. Pau, A., Ruta, G., Vestroni, F.: Wave propagation for stress evaluation. In: Zingoni, A. (ed.) SEMC 2013: The Fifth Interna-
tional Conference on Structural Engineering, Mechanics and Computation. Cape Town, South Africa, 2–4 Sept 2013. Taylor
& Francis Group, London (2013). doi:10.1201/b15963-14
43. dell’Isola, F., Ruta, G., Batra, R.: Generalized Poyntings effects in a prestressed bar. J. Elast. 50, 181–196 (1998)
44. dell’Isola, F., Ruta, G., Batra, R.: A second-order solution of Saint–Venants problem for an elastic bar predeformed in
flexure. Int. J. Non-Linear Mech. 40, 411–422 (2005)
45. Winkler, E.: Die Lehre von der Elastizität und Festigkeit. Prague, Dominicus (1867); (in German)
46. Cowper, G.R.: The shear coefficient in Timoshenko’s beam theory. ASME J. Appl. Mech. 33, 335–340 (1966)
47. Duncan, W.J.: Galerkin’s method in mechanics and differential equations. Aeron. Res. Comm. Rep. Mem. 1798 (1937)

You might also like