You are on page 1of 20

Mathematical and Computer Modelling of Dynamical

Systems
Methods, Tools and Applications in Engineering and Related Sciences

ISSN: 1387-3954 (Print) 1744-5051 (Online) Journal homepage: http://www.tandfonline.com/loi/nmcm20

Numerical study of forebody wake effect on


axisymmetric parachute opening shock and drag
reduction

Jianhua Tang & Linfang Qian

To cite this article: Jianhua Tang & Linfang Qian (2016) Numerical study of forebody
wake effect on axisymmetric parachute opening shock and drag reduction,
Mathematical and Computer Modelling of Dynamical Systems, 22:2, 141-159, DOI:
10.1080/13873954.2016.1149492

To link to this article: http://dx.doi.org/10.1080/13873954.2016.1149492

Published online: 16 Mar 2016.

Submit your article to this journal

Article views: 3

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=nmcm20

Download by: [University of California, San Diego] Date: 22 March 2016, At: 03:44
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS, 2016
VOL. 22, NO. 2, 141–159
http://dx.doi.org/10.1080/13873954.2016.1149492

Numerical study of forebody wake effect on axisymmetric


parachute opening shock and drag reduction
Jianhua Tang and Linfang Qian
School of Mechanical Engineering, Nanjing University of Science and Technology, Nanjing, China

ABSTRACT ARTICLE HISTORY


Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Parachute–forebody distance is a parameter which is amongst the most Received 9 July 2015
critical factors to be considered in forebody wake effect. In this study, a Accepted 29 January 2016
new axisymmetric parachute–forebody coupling model is developed. KEYWORDS
Axisymmetric wrinkling membrane element is built to assess the dynamic Parachute inflation; axis
response of the parachute canopy membrane under fluid pressure. symmetric; forebody wake
Besides, fluid model and its further implementation on the fluid structure effect; drag reduction
analysis are discussed. With the proposed method, the wake effect on
both the opening shock during inflation state and the drag reduction
during steady state can be obtained efficiently. Finally, numerical model is
validated with published experimental result and further employed to
investigate the influence of distance parameters on fluid–parachute cou-
pling behaviour. On the basis of numerical results, failure distance during
the inflation process and critical forebody–parachute distance are deter-
mined. The results show that forebody–parachute distance has a strong
influence on flow behaviour around the parachute in both inflation state
and steady descent state.

Nomenclature
Dp: Constructed canopy diameter
Lfp: Distance between the forebody and the leading edge of the parachute
u: Point displacement vector
X: Point position vector
Cd: Drag coefficient of the parachute with forebody
Cd,∞: Drag coefficient of the parachute without forebody
Cf: Drag coefficient of the forebody
v∞: Free stream velocity
Closs: Ratio of drag coefficient with forebody to drag coefficient without forebody
Rf: Forebody radius
Lf: Trailing length of the forebody
Lp: Trailing distance ahead of the parachute
m: Gores number of the parachute
h: Thickness of the parachute membrane
k: Turbulent kinetic energy
ε: Turbulent dissipation
~v: Fluid mesh moving velocity
v: Average velocity of the fluid

CONTACT Linfang Qian lfqian@vip.163.com


© 2016 Taylor & Francis
142 J. TANG AND L. QIAN

v0 : Fluctuations velocity of the fluid


υ: Fluid kinematic viscosity

1. Introduction
Since parachutes are much simpler and more effective than other aero decelerators, deceleration
systems based on parachutes systems are widely used as an effective air drop for deploying
personnel as well as cargo during a combat operation or even otherwise. The main purpose of
a parachute is to decelerate a payload to a lower speed aerodynamically. In the deceleration
application, the payload, also known as the forebody, will affect the flow field and bring about a
new vortex distribution over the canopy.
The change of flow field resulting from forebody is termed forebody wake effect. According to
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Knacke’s study [1], the reduction in parachute drag due to the forebody influence ranges from 2%
to 70%. This range of parachute drag reduction is primarily dependent on the forebody size and
the distance between the forebody and the leading edge of the parachute, which is termed
forebody–parachute distance for the sake of simplicity. The design for the suspension line length
is dependent on the forebody–parachute distance, as depicted in Figure 1. Therefore, the forebody
wake effect should be fully considered before the parachute designing and its optimization.
In an airdrop performance, a decelerator system usually experiences three divided states:the
deployment state, the inflation state and the steadily decent state. The forebody wake has a
significant influence on the inflation state as well as on the steady descent state. Herein the
study on the forebody wake effect should focus on two phases during the parachute performance.
And these two phases include the inflation phase and the steady descent phase which will be
discussed in details as follows.
Parachute inflation phase is a very special one when parachute fabric materials can undergo
large and arbitrary deformation in a very short time. The inflation characteristics investigation is
carried out through experiments in the early study [2–4]. With the continuous development of
numerical algorithms and the computer hardware, it brings tremendous savings in time and cost
of parachute simulating and thereby FSI problem of the parachute has been attracting scholars’
attention. Irvin Aerospace initiated a strategy to develop an analysis methodology that enabled
parachutes to be simulated in a virtual fluid environment. And models developed in finite element
code LS-DYNA were performed in both infinite and finite mass scenario [5–7]. In addition,
similar investigation can be found at Kim’s works. At his works, parachute inflation was simulated
with immersed boundary method [8–10]. Yu [11] and Cheng [12] studied the inflation of a
parachute with SALE (simple ALE) method and experimentally validated the results. Fan

Figure 1. Parachute–forebody structure.


MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 143

numerically studied the parachute based on non-linear finite element method and precondition-
ing finite volume method. The results obtained with these methods agreed with the experiments
well.
However, there exists one shortcoming for these methods, namely, Eulerian mesh must span
the entire range of the activity associated with the Lagrangian structure while the number of these
Eulerian elements must be sufficient to permit the fluid to flow into the closed mouth of the
canopy in packed condition. This would surmount a very large number of elements and hence a
high computing cost. For instance, the parachutes test ranging from 3.5 ft to 9 ft takes around 0.9
million elements to fulfil the task [7]. For this reason, although the foregoing mentioned
successful studies have been made to analyse the inflation state of parachute, the wake effect
during this state is usually omitted for the sake of computational expediency. Therefore, the wake
effect on the inflating parachute is rarely studied and reported. In this article, however, in order to
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

study the wake effect, a axisymmetric model is developed, which is suggested by the inherent
axisymmetric geometry of the parachute. This study assumes a circumferentially averaged axisym-
metric flow leading to the solution of a two-dimensional problem in the meridional plane. The
axisymmetric model can be considered as the ‘’bridge‘’ between one- and three-dimensional
design procedures for the parachute, and provides an optimal compromise between computa-
tional cost and fidelity of the results [13–15]. Based on this simplification, the forebody wake
effects during the inflation process are studied by virtue of this efficient developed model.
After parachute inflation process terminates, the fluid keeps on interacting with the parachute
until an equivalent state is achieved, which is termed steady descent state. During this state, the
presence of the wake vortex dramatically makes the dynamic pressure inside canopy decrease.
Many researchers have proposed various approaches to describe the wake effect in steady descent
state. Heinrich et al. [16] originally described the reduction in parachute drag through giving an
empirical velocity distribution. Peterson and Johnson further improved the method through
experiments study [17]. In addition to these empirical results, Anita Sengupta provided a cost-
efficient aerodynamic performance database that captured the wake effect of the Orion crew
module on the parachute through subscale model experiment [18]. McQuilling et al. [19]
presented time-resolved and time-averaged results from simulations using a Reynolds-averaged
Navier-Stokes (RANS) flow solver to study the effects of a forebody wake on the aerodynamics of
the double annular parachute. Based on a simple ‘immersed boundary technique’, Xiao et al.
investigated the effects of suspension lines on the supersonic flow field [20]. These methods
provide a tool to study the forebody wake effect in steady descent phase. However, the effect of
suspension line and forebody are not considered in the inflation state.
In this article, the forebody wake effect is studied numerically through establishing the
axisymmetric forebody–parachute coupling model. Both inflation case and steady decent case
can be reflected in the present analysis. With this method, not only can the flow behaviour of the
forebody wake be clearly observed, but also new criteria that help to judge the fluid topology
around the forebody and parachute are provided.
This article is organized as follows. In the first section, the axis symmetric finite element model
is developed to facilitate opening a parachute. In the second section, an axisymmetric finite
membrane structure element is established and discussed. In the third section, the fluid model
as well as the coupling method will be outlined. And the inflated method will be presented.
Moreover, the challenge such as the efficiency problem can be solved in this axisymmetric
configuration. Finally, the forebody wake effect in both inflation state and steady descent state
is discussed.

2. Mathematical equation of the axisymmetric membrane


In this section, an axisymmetric finite membrane structure element is established and discussed.
With this element, the canopy is modelled by a series of elements whose geometries are
144 J. TANG AND L. QIAN

axisymmetric. These elements are subjected to axially symmetric loading conditions. A dynamic
model of the canopy is established based on the following assumptions:

(a) Thickness of the membrane is invariant.


(b) The membrane cannot stand any bending.
(c) The forebody is fixed.
(d) The suspension line effect is neglected.
(e) The parachute shape is invariable after it is fully inflated.

The forebody shape is determined by engineering practice [17,18,21,22]. In this model, the
forebody is modelled as a rotating disk for the sake of simplicity, as depicted in Figure 2.
After the parachute is fully inflated, the steady descent state is achieved. Breathing phenom-
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

enon as well as the fly out angles variant will totally bring about 5%–20% reduction in parachute
drag. However, according to Knacke’s study, the reduction in parachute drag due to the forebody
influence ranged from 2% to 70%. Since this article is more concerned about the trailing distance,
the parachute shape is assumed as invariable after it is fully inflated for the sake of simplicity and
stability. And the effect of the breathing phenomenon is cancelled numerically.
According to the foregoing assumptions, the axisymmetric parachute can be modelled by rings
of constant cross-sections. The centre of all nodal circles is on the z-axis, which is the axis of
revolution. The directions of mechanical displacements are represented by r, θ, z, which are the
radial, circumferential and axial coordinates, respectively. The origin is located at the centre of the
disk. The structure deformation is outlined in terms of this inertial frame. As shown in Figure 2,
let er, ez be unit vectors in radial, and axial directions. A is an arbitrary point of the canopy. u and
X denote displacement and position vector of this arbitrary point, respectively.
The position of the arbitrary point A can be represented as

X ¼ Rer þ Zez (1)

Figure 2. Position and displacement of the parachute.


MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 145

And the displacement of the point can be represented as


u ¼ uer þ wez (2)

The general theory of thin membrane is used to represent the dynamics of very flexible thin
membrane structures. Let V denote a volume occupied by a part of the body in the current
configuration. The displacement formulation for stress analysis is based on the energy concept of
finding u that both satisfies the essential boundary conditions on u and minimizes the total
energy. The virtual strain and kinetic energies are written as
ð
ðδU þ δT þ δW ÞdV ¼ 0 (3)
V

In the expression above, T and U are kinetic and internal energy densities per unit volume,
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

respectively, and W is the virtual work of applied loads per unit volume. The weak form can be
written as
ð ð
 T 
uÞ dV þ δuTfdS ¼ 0
δε σ þ δuT ðf  ρ€ (4)
ν sσ

The equilibrium formulation requires the expression of strain and stress. The relations of strain
and displacements are given by the linear part of Green–Lagrange strain tensor in cylindrical
coordinates, which take the following form

ε ¼ ð εs εθ ÞT ¼ Bu (5)

σ ¼ Dε (6)

A piece of membrane can support large tensile forces but is weak at supporting torque. As the
membrane cannot stand any bending, shear stress is neglected, which greatly simplifies the model.
Thus the shear strain no longer exists in this inflation model on purpose. With the foregoing
assumptions, the membrane equation matrix B can be written as
 @ @

cos ϕ @s sin ϕ @s
B¼ (7)
0 1=r

cos ϕ ¼ dZ=ds and sin ϕ ¼ dR=ds (8)

where s is the meridian length along the canopy.


Researchers do come up with a relatively simple and meanwhile a productive parachute
structural model by taking advantage of some specific construction features of a circular para-
chute. When the membrane is wrinkling, tension is lost in one principal direction and canopy
fabric cannot support any stresses. Thus it’s assumed that the canopy membrane piece cannot
bear circular tensile stress until it is fully inflated. Taut stresses only engage when the current
distance between the axis of symmetry and the radius position of point is greater than the final
shape function of the parachute.
  
1 cν 0 RðsÞ  ηðsÞ
D ¼ 1ν2
Eh
c¼ (9)
ν c p RðsÞ > ηðsÞ

E is the Young’s modulus of the membrane, ν is Poisson’s ratio and p is the penalty factor of
the parachute expansion. The parachute ceases stretching with large penalty factor after inflation.
η is the radius between the node and the symmetry axis of an inflated parachute, which describes
the fully inflated shape of the parachute. h is the thickness of the membrane. The fully inflated
146 J. TANG AND L. QIAN
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 3. Undeformed Gore profile of a round parachute.

shape is determined by the structure of the parachute. Without using the CALA (CANOPY
LOADS ANALYSIS) logic proposed by Sundberg [23], the final shape in this article is calculated
with FEM code by imposing a constant pressure. After the FEM code has found the values of
nodes displacement, the approximation value of the η(s) at node i is obtained, which is denoted as
~
ηðSi Þ. The field variable η(s) can be approximated with NURBS interpolation [24].
The membrane mass is obtained through the undeformed Gore profile, as depicted in Figure 3.
Therefore, the virtual work of inertia force is written as followed
ð ð
δuT ρ€
udV ¼ δuT 2m ρhLðsÞ€ uds (10)
V s
where L is the distance between the seam and the symmetric axis of the Gore; m is the number of
gores; The virtual work of the aerodynamic force is obtained as followed
ð ð
δu fdV ¼ δuT 2πRðsÞPðsÞds
T
(11)
V s

where P is the pressure difference across the membrane.

P ¼ ½ P sin ϕ P cos ϕ T (12)

To solve the above equations the canopy domain is discretized into a number of elements. The
Lagrange interpolation is used to construct the shape function of the element. The position and
displacement is written as

u ¼ N j uj (13)

where Nj is the shape function matrix of node j.


Equation (8) can be written as

cos ϕ ¼ 1J dZ
d and sin ϕ ¼ 1J dR
d (14)
qffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
J ¼ ds=d ¼ ðdR=dÞ2 þ ðdZ=dÞ2 (15)

where  2 ½1; 1 is normalized coordinate of an element.


Based on this analysis, an updated parachute position can be obtained when the pressure is
given. For the purpose of comparing with the aerodynamic loads distribution of the elastic
parachute, the parachute is divided into many segments along the meridian shape.
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 147

3. Mathematical equations of the fluid coupling model


3.1 Formulation of fluid model
The structure model requires a pressure distribution along the meridional length of the canopy as
input. It is more efficient to obtain a reliable solution with numerical computational fluid method.
The flow field is governed by the continuity and momentum equations. The incompressible
NavierStokes (NS) equations are written as
Ñv ¼ 0 (16)
@v 1
þ Ñvv ¼  Ñp þ υÑ2 v (17)
@t ρ

where ρ is the fluid density, v, p, υ are velocity, static pressure and kinematic viscosity of fluid
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

respectively. Ormieres [25] discussed the relationship between Reynolds number and turbulent
wake of a sphere. The critical Reynolds number is about 800. Namely, when an external flows
around spherical obstacles, its Reynolds number is beyond this critical number, and the
turbulent wake will appear after the sphere. As depicted, the inflated shape of a typical circular
parachute can be assumed as an hemispherical shell, and its Reynolds number is over 1000.
Therefore, the flow characteristic is considered as turbulent, and turbulence model is
added [26].
Decomposing the N-S equations into the RANS equations makes it possible to simulate
engineering fluid dynamic problems. In this article, the Standard k-epsilon model is implemented
to simulate this engineering fluid dynamic problem. This model is RANS equations-based
turbulence models.
Decompose the pressure and the velocity into mean and fluctuations
v ¼ v þ v0 (18)
where v is the average velocity and v0 is the fluctuations velocity
The fluctuations are lumped into the kinetic energy
k ¼ kv0 k=2 (19)
The turbulent kinematic eddy viscosity is introduced, which is given as

υT ¼ C1 k2 =ε (20)
where k is the turbulent kinetic energy, and ε is the dissipation.
The average velocity and the average pressure are solved by
@ 1
v þ Ñ  vv ¼  Ñp þ ðυ þ υT ÞÑ2 v (21)
@t ρ
In order to enclose the equations, turbulent kinetic energy and dissipation are solved by
@k=@t ¼ Ñ  ðυ þ C2 υT ÞÑk  τ ij @v=@xj  ε (22)

@ε=@t ¼ Ñ  ðυ þ C3 υT ÞÑε þ C4 ðε=kÞτ ij ð@vi =@xj Þ  C5 ε2 =k (23)



2 @vi @vj
τ ij ¼ kδij  υT þ (24)
3 @xj @xi

where constants are given as C1 = 0.09, C2 = 1, C3 = 1.3, C4 = 1.44, C5 = 1.92 [27]. The
formulation is applied to the forebody fluid simulation.
148 J. TANG AND L. QIAN

Figure 4. ALE formulation and mesh moving.

ALE formulation of the parachute is necessary due to the large deformation taking placed in the
parachute inflation state. The moving grid methods are performed [28–30], as depicted in Figure 4.
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

@ 1
v þ Ñ  vðv  ~vÞ ¼  Ñp þ ðυ þ υT ÞÑ2 v (25)
@t ρ
where ~v is the fluid nodes moving velocity. This velocity stems from the structure boundary
motion. The axisymmetric incompressible turbulent N-S equations is modified as below

1@ @vz
ðrvr Þ þ ¼0 (26)
r @r @z

  
@vr @ðvr  ~vr Þ @ðvr  ~vr Þ @p 1 @ @vr @ 2vr vr
þ ur þ uz ¼  þ ðυ þ υT Þ ðr Þþ  2 (27)
@t @r @z @r r @r @r @z r

   
@vz @ðvz  ~vz Þ @ðvz  ~vz Þ @p 1@ @vz @ 2vr
þ ur þ uz ¼  þ ðυ þ υT Þ r þ 2 (28)
@t @r @z @z r @r @r @z

In this article, the far-field free stream condition is applied and the normal velocity boundary
condition is set. The solution method is implicit formulation finite volume method. This method
is used to spatial discrete the space of fluid field. Second-order upwind scheme is used on
convection and turbulent viscosity terms. Through computation, the aerodynamic parameters of
every parachute segment are obtained.

3.2 Coupling solution procedure


After the fluid model is established, it will be coupled with the structural model discussed in
Section 2. In this article, a two-way coupling method is applied [31]. The coupling approach in the
parachute model is explicit in time. During each time step in the coupled run the CFD and
structural dynamic computations are advanced one time step. First of all, fluid meshes are
updated based on the updated canopy surface representation. The meshes are defined by repo-
sitioning the surface boundary as well as the outer boundary of the parachute, and then updating
interior nodes and nodes along the axisymmetric axis. Next, pressure distribution on two sides of
the parachute is updated. Second, with the new pressure difference, the motion of the membrane
can be obtained. The process is repeated over and over again, until a termination condition is
achieved. With this procedure, additional insight into parachute–fluid behaviour is achieved using
the proposed finite element model.
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 149
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 5. Time history of the vent hole parachute force.

4. Example and discussion


4.1 Inflation state analysis
4.1.1 Validation of numerical inflation process
With application of the numerical model developed, simulations of the flow around the parachute
are now carried out. The structure code and the coupling system are implemented with ANSI C
[32]. And we resort to the commercial CFD solver Fluent to complete the fluid dynamic
computation.
In order to validate the inflation process, comparison is made with respect to experimental data
provided by Desabrais’s experiments [33]. The experiments were performed in a water tunnel
which has internal dimensions of 0.6 m wide by 0.6 m deep by 2.4 m long. Diameter of parachute
is 30.5 cm. The weight of cloth is 37.3 g/m2, and the thickness of fabric is 500 μm. The material of
the fabric and suspension line is rip-stop nylon and nylon, respectively. The opening force is
compared with the experiment results provided by Desabrais’s experiments [33], as depicted in
Figure 5.
The opening shock obtained with the simulation process accords with the tendency of the
experiments. Parachute FSI simulation results show agreement with experimental results and
previous literature, and thus qualify the method developed in this article to be applicable with
large-scale parachute. This validated model is applied to the following full-scale parachute
analysis.

4.1.2 Forebody wake effect during the inflation process


When the parachute inflation process terminates, steady decent state begins and the parachute is
found to be fully inflated. From the initial state to fully inflated state, the parachute experiences
large deformation. During this inflation process, volume expansion of the parachute leads to the
pressure plummeting. The pressure difference between inside and outside of the parachute forces
the fluid around the parachute to get into the canopy. When the forebody is placed ahead of the
parachute, the fluid accumulates ahead of the forebody instead of flowing into the canopy. Some
of the fluid bypasses obstacles and fills the low-pressure region inside the canopy. However, this
part of fluid is quite limited when the forebody stays close to the parachute. In this example, a
Df = 1.6 m diameters round disk is chosen and an Dp = 5.6Df diameter parachute is selected. The
150 J. TANG AND L. QIAN
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 6. Parachute inflation process.

surrounding fluid is air and the free stream velocity is 20 m/s. The material of the computational
domain is air. The computational domain is 12 m along radius direction and 96 m along the axial
direction. The fluid domain is discredited with about 5e4 cells.
When the forebody disk stays close to the parachute, the fluid is difficult to flow into the
canopy and the canopy may fail to inflate, as depicted in Figure 6. When no forebody is placed
ahead of the parachute, the parachute is inflated successfully. However, when the distance between
the disk plane and the parachute skirt is 2.8Df, the parachute shrinks and fails to expand.
Figure 6 shows the failure process of a parachute with a placed 2.8Df from the skirt. Owing to
this obstacle, most of the fluid cannot successfully fills the quasi-vacuum region and the low
pressure inside the canopy forces the parachute to shrink at the opening mouth. However, when
the parachute begins its shrink at the opening mouth, it becomes even more difficult to fill the
canopy. All in all, to guarantee the safety of the decelerator system is the primary task of airdrop
system design, and the distance between the parachute and the forebody cannot be too close.
Non-dimensional forebody–parachute distance varies from 5.6Df to 11.2Df to explore the effect
of wake coupling on the parachute flow. The free stream velocity varies from 10 m/s to 30 m/s.
Figure 7 shows time histories of parachute opening forces under free stream velocity condition
are 10 m/s, 20 m/s, and 30 m/s, respectively. In order to provide a thorough comparison of the
wake reduction, the nominal shock is applied. The nominal shock is present in non-dimensional
form

Fn ¼ 2FðtÞ=ðρv2 πD2p Þ (29)

The non-dimensionalize time is given as followed

Figure 7. Opening shock with different free stream velocity.


MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 151
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 8. Variant of the flow topology according to time.

tnd ¼ tV1 =Dp (30)

According to the assumption (5) in the second section, canopy structures do not change
after it has fully expanded. Namely, the influence of breathing phenomenon has been
eliminated and the forebody placement is the main reason that results in the difference
between opening forces. Through the results, the parachute achieves a stable flow field and
drags force quickly without forebody, while with the forebody, the parachute opening force
will experience plummeting. The topology [34] of the vortexes can also be observed in
Figure 8. The flow field drastically changes with time when the forebody is placed near the
parachute skirt.
The main source of instabilities originates in the turbulence interaction between the
vortexes of the forebody and the canopy [35]. Through Figure 8 (a)–(d), when there is no
disk placed ahead of the parachute the drag converges quickly and varies slowly after it is fully
inflated. However, the drag caused by the canopy behaves differently when a forebody is
placed ahead of the parachute. It is shown that the opening force varies enormously when the
forebody is closer to the parachute and the flow field becomes less stable. Vortices begin
shedding and it takes a long time for the vortices to separate and combine with other vortices
until it achieves the balance state.
152 J. TANG AND L. QIAN

4.2 Steady-state analysis of forebody–parachute structure


After the parachute is fully inflated, the parachute achieves a new balance state descent steadily. In
this state, parachute drag reduction is primarily dependent on the ratio of forebody diameter to
parachute diameter as well as the distance between the forebody and the leading edge of the
parachute. With the turbulent model mentioned above, the predominate flow features could be
identified and obtained numerically. In order to study relationship between these parameters and
the drag loss effect, the flow behaviour around the forebody as well as the parachute in steady
descent phase will be presented and discussed in this example.

4.2.1 Steady-state analysis of forebody


Since the ratio of forebody diameter to parachute diameter is an important parameter that affects
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

the flow topology, this study will focus on the wake effect determination of forebody size.
A single rotating disk suffering constant velocity stream is investigated numerically. The basic
radius of this disk is 1.0 m, as depicted in Figure 9.
The forebody is considered as a bluff body. As depicted in Figure 10, the flow topology of a
single forebody is present. It can be observed that there is a vortex ring placed behind these bluff
bodies. There is a stagnation point located at the topic of vortexes rings. At stagnation point, the
velocity as well as the pressure is zero. The fluid begins to separate at this point which is located at
the axisymmetric axis. Therefore, the relationship between the forebody size and its stagnation
point position is investigated through pressure and velocity distribution along the axisymmetric
axis. These relations are present to manifest the flow characteristic and its corresponding
structure.
In Figure 11, pressure and velocity distribution with different forebody radius along the
axisymmetric axis are present. Non-dimensional distance which varies from 0 to 5Df is present

Figure 9. Structure of the forebody.

Figure 10. Flow topology of a single forebody.


MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 153
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 11. Pressure and velocity distribution with different forebody radius.

Figure 12. Pressure and velocity distribution with different velocity.

to discuss the forebody wake effect. Computations are conducted for Df from 1 m to 4 m. The free
stream velocity is 20 m/s.
Through Figure 11, it is shown that the pressure increases slowly after the initial
decrease. The velocity is almost zero at stagnation point. Therefore, it can be observed
that stagnation point is located at approximately Lf = 1.75Df to forebody. The forebody
effect decays according to distance and can almost be neglected at 5Df. In this example, it
can be observed that there is a linear relation between the forebody size and the stagnation
point distance.
The relationship between stagnation point distance and the free stream velocity is also present.
Figure 12 shows pressure and velocity distribution along the axis symmetric axis with different
free stream velocity ranging from 10 m/s to 30 m/s. It is obvious that the location of the
stagnation point is independent from the free stream velocity. In this example the stagnation
points are also located at approximately 1.75Df to the round disk.
154 J. TANG AND L. QIAN
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 13. Fluid topology of a single parachute.

Figure 14. Velocity and pressure distribution along the parachute.

According to these results, it can be concluded that the free stream velocity has little effect on
the stagnation point location. The trailing distance decays with distance and has a linear relation
to forebody size.

4.2.2 Steady state of parachute


The flow topology of one single parachute is present in Figure 13 [36,37]. The pressure and
velocity along parachute axis under free stream velocity 10 m/s, 20 m/s, 30 m/s are shown in
Figure 14. The fluid jets at vent hole and an obvious vortex ring are formed behind the canopy.
Similar to the single forebody analysis, the investigations of flow characteristic are carried out
through studying its pressure and velocity distribution along the axisymmetric axis. As a bluff
body, the pressure increases, the fluid velocity decelerates inside the canopy, and the fluid is
accumulated.
The effect that the parachute acts on the fluid can be neglected at the point located at
approximately Lp = 1.0Dp distance ahead of the canopy skirt.

4.2.3 Steady state of parachute–forebody structure


First of all, the velocity along the radius axis is compared with the velocity profiles obtained
through the experiment with wind tunnel [17]. Figure 15 shows that the computed velocity
distribution accords well with the experimental distribution during the steady state. CFD model
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 155

Figure 15. Experimental and CFD velocity distributions behind forebody.


Downloaded by [University of California, San Diego] at 03:44 22 March 2016

can be applied to further study the system parameters, such as the forebody diameter and the
distance between forebody and parachute leading edge.
When the fluid is blocked by the forebody, the flow behaviour around the parachute with block
is quite different from that around the forebody free parachute. Researchers have employed
different approaches to describe the forebody effect. For instance, Heinrich studied the measured
values of parachute drag loss and gave simplified turbulent wake velocity distribution equations
[17]. Peterson studied the experiment results and gave the modified expressions for forebody
configurations.

Cd 2 D1 2 D1 2 D1
Closs ¼ ¼1þ 2 1  eD2 Rf þ e2D2 Rf  (31)
Cd;1 R f D2 4 4

where the parameter is given as D1 ¼ a=ðLfp =2Rf Þm , D2 ¼ 1=ð0:435b2 ÞðLfp =2Rf Þ2n ; a and b are
exponential functions of forebody drag coefficient.

a ¼ 0:42e0:99Cf b ¼ 0:54e0:84Cf (32)

Fix the forebody–parachute distance Lfp = 11.2Df, and the ratio of forebody diameter to
parachute diameter which varies from 0.05 to 0.25Dp is studied numerically. The drag loss results
are compared with the empirical function results, as shown in Figure 16.
The results achieved by CFD method have a fine agreement with those obtained through the
empirical function. It can be concluded that both empirical function and CFD method can be
applied to predict the drag loss feature when the forebody is far away from the parachute.
However, when the forebody is getting close to the parachute, the drag loss manifests great
distinctions between two methods, as shown in Figure 17, the result is obtained as free stream
velocity is 20 m/s and the forebody radius is 0.7 m. It is assumed that parachutes are fully inflated
and no vortexes shedding takes place. Runs are made as forebody–parachute distances range from
5.6 to 11.2Df.
The tendency of CFD result is more align to comparative experiment data obtained by
Knacke. When the forebody is approaching the parachute, large drag loss occurs. However,
the results obtained through empirical data cannot provide a similar tendency, as shown in
Figure 17. The poor agreement between present CFD data and empirical predictions points
out that the empirical analysis may be insufficient to cover all the drag loss cases when the
forebody is approaching the parachute. In this study, a generalized correlation for the
variation of drag loss with Lfp has been developed for the cases in which the forebody
approaches the parachute and the flow changes. Subsequently, the critical value of Lp at
which the onset of vortex separation takes place is identified with both the length of
156 J. TANG AND L. QIAN
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 16. Drag loss vs. forebody radius.

Figure 17. Drag loss vs. forebody–parachute distance.

stagnation point vortex Lf and parachute distance Lp. The criteria distance Lfp can be
approximated by the summation of these two values.
~fp ¼ Lf þ Lp
L

When the parachute–forebody distance is much larger than this criteria distance, the vortexes
ahead of the parachute and vortex behind the forebody will separate. The pressure loss effect is
minor. However, when the forebody gets close to the parachute and the distance is smaller than
the criteria value, these vortexes will combine and lead to the parachute drag loss. The fluid
stream is present in Figure 18. The vortex begins to separate with distance L ~fp . The topology
evolution according to the distance is present in Figure 19.
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 157
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

Figure 18. Stream path with different forebody–parachute distance.

Figure 19. Flow topology with different forebody–parachute distance.

In this example, the drag loss will be alleviated with the larger distance. However, when the
distance is greater than 11.2Df, the forebody wake effect is quite limited. And the drag force may
be difficult to benefit from the increase of the distance Lfp.

5. Conclusions
A forebody–parachute model is developed and applied to the numerical analysis. The parachute in
both inflation and steady descent state is discussed. In addition, the fluid response of forebody–
parachute structure for various distances is considered. Through computation results, there are
several conclusions drawn as follows.

(1) With the simplified axisymmetric forebody–parachute model, the forebody wake effect in
both inflation and steady state can be obtained efficiently.
(2) When the parachute is inflating, if the parachute is close to the forebody, the parachute
may fail to inflate, or experience an unstable state for a long time.
(3) There is a separating point located on the forebody–parachute system. When vortex
shedding does not take place, the parachute will experience considerable drag loss due to
158 J. TANG AND L. QIAN

vortex combination. And when the forebody–parachute distance is larger than the criteria
distance, the drag loss effect will decay or alleviate.

Disclosure statement
No potential conflict of interest was reported by the author.

Funding
This study is supported by the National Natural Science Foundation of China [No. 11472137].
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

References
[1] T.W. Knacke, Parachute Recovery Systems: Design Manual, Para Pub, Santa Barbara, CA, 1992.
[2] P.S. Jason and A. Hart, Canned telemetry system for sensor data on parachute system, AIAA. 15 (2005), pp. l3.
[3] Y. Li, Z. Xinhua, and L. Shuisheng, Experimental on canopy payload performance of parachute, J. Beijing
Univ Aeronaut. Astronaut. 33 (10) (2007), pp. 1178.
[4] L. Yu, X. Ming, and B. Hu, Experimental investigation in parachute opening process, J. Nanjing Univ.
Aeronaut. Astronaut 38 (2) (2006), pp. 176–180.
[5] B.A. Tutt and A.P. Taylor. The use of LS-DYNA to simulate the inflation of a parachute canopy.18th AIAA
Aerodynamic Decelerator Systems Technology Conference and Seminar. 2005.
[6] Y. Fan and J. Xia, Simulation of 3D parachute fluid–structure interaction based on nonlinear finite element
method and preconditioning finite volume method, Chin. J. Aeronautics 27 (6) (2014), pp. 1373–1383.
doi:10.1016/j.cja.2014.10.003
[7] B. Tutt, S. Roland, R.D. Charles, and G. Noetscher, Finite mass simulation techniques in LS-DYNA, 21st
AIAA Aerodynamic Decelerator Systems Technology Conference and Seminar, Dublin, 2011, pp. 2592.
[8] J. Kim, Y. Li, and X. Li, Simulation of parachute FSI using the front tracking method, J. Fluids Struct. 37
(2013), pp. 100–119. doi:10.1016/j.jfluidstructs.2012.08.011
[9] Y. Kim and C.S. Peskin, 3-D parachute simulation by the immersed boundary method, Comput. Fluids 38 (6)
(2009), pp. 1080–1090. doi:10.1016/j.compfluid.2008.11.002
[10] Y. Kim and C.S. Peskin, 2-D parachute simulation by the immersed boundary method, SIAM J. Scientific
Comput. 28 (6) (2006), pp. 2294–2312. doi:10.1137/S1064827501389060
[11] L. Yu, H. Cheng, Y. Zhan, and S. Li, Study of parachute inflation process using fluid–structure interaction
method, Chin. J. Aeronautics 27 (2) (2014), pp. 272–279. doi:10.1016/j.cja.2014.02.021
[12] H. Cheng, X.-H. Zhang, L. Yu, and M. Chen, Study of velocity effects on parachute inflation performance
based on fluid–structure interaction method, Appl. Math. Mech. 35 (9) (2014), pp. 1177–1188. doi:10.1007/
s10483-014-1852-6
[13] K.R. Stein and R.J. Benney, Parachute inflation: A problem in aeroelasticity, DTIC Document, 1994.
[14] K.R. Stein, R.J. Benney, and E.C. Steeves, A computational model that couples aerodynamic and structural
dynamic behavior of parachutes during the opening process, DTIC Document, 1993.
[15] L. Yu and X. Ming, Study on transient aerodynamic characteristics of parachute opening process, Acta Mech
Sin 23 (6) (2007), pp. 627–633. doi:10.1007/s10409-007-0112-3
[16] H.G. Heinrich and T. Riabokin, “Analytical and experimental considerations of the velocity distribution in
the wake of a body of revolution”. 1959, DTIC Document.
[17] C.W. Peterson and D.W. Johnson, Reductions in parachute drag due to forebody wake effects, J. Aircr. 20 (1)
(1983), pp. 42–49. doi:10.2514/3.44826
[18] A. Sengupta, B. Goree, E.B. White, J. Guthery, R. Machin, G. Bourland, J. Laguna, R. Sinclair, and E.
Hennings, Performance of a Subscale CPAS Conical Ribbon Drogue Parachute in a Turbulent Wake, AIAA
Aerodynamic Decelerator Systems (ADS) Conference, Reston, VA, 2013, pp. 1307.
[19] M. McQuilling and J. Potvin. Forebody wake effects on the aerodynamics of an annular parachute. 42nd AIAA
Fluid Dynamics Conference and Exhibit. 2012.
[20] X. Xue, H. Koyama, Y. Nakamura, and C.-Y. Wen, Effects of suspension line on flow field around a supersonic
parachute, Aerosp. Sci. Technol. 43 (2015), pp. 63–70. doi:10.1016/j.ast.2015.02.014
[21] E.S. Ray and A.L. Morris. Measurement of CPAS main parachute rate of descent. 21st AIAA Aerodynamic
Decelerator Systems Technology Conference. 2011.
[22] G. Guglieri, Parachute–payload system flight dynamics and trajectory simulation, Int. J. Aerosp. Eng. 2012
(2012), pp. 1–17. doi:10.1155/2012/182907
MATHEMATICAL AND COMPUTER MODELLING OF DYNAMICAL SYSTEMS 159

[23] W.D. Sundberg, New solution method for steady-state canopy structural loads, J. Aircr. 25 (11) (1988), pp.
1045–1051. doi:10.2514/3.45701
[24] L.A. Piegl and W. Tiller, The NURBS Book, Springer, Berlin, 1997.
[25] D. Ormières and M. Provansal, Transition to turbulence in the wake of a sphere, Phys. Rev. Lett. 83 (1)
(1999), pp. 80–83. doi:10.1103/PhysRevLett.83.80
[26] A. Prosperetti and G. Tryggvason, Computational Methods for Multiphase Flow, Cambridge University Press,
New York, 2007.
[27] H.K. Versteeg and W. Malalasekera, An Introduction to Computational Fluid Dynamics: The Finite Volume
Method, Pearson Education, Edinburgh Gate, Harlow, 2007.
[28] E. Lefrançois, A simple mesh deformation technique for fluid–structure interaction based on a submesh
approach, Int. J. Numer. Methods. Eng. 75 (9) (2008), pp. 1085–1101. doi:10.1002/nme.v75:9
[29] F.J. Blom, Considerations on the spring analogy, Int. J. Numer. Methods Fluids 32 (6) (2000), pp. 647–668.
doi:10.1002/(ISSN)1097-0363
[30] K. Stein, T. Tezduyar, and R. Benney, Mesh moving techniques for fluid–structure interactions with large
displacements, J. Appl. Mech. 70 (1) (2003), pp. 58–63. doi:10.1115/1.1530635
Downloaded by [University of California, San Diego] at 03:44 22 March 2016

[31] L.K. Abbas, D. Chen, and X. Rui, Numerical calculation of effect of elastic deformation on aerodynamic
characteristics of a rocket, Int. J. Aerosp. Eng. 2014 (2014), pp. 1–11. doi:10.1155/2014/478534
[32] N.M. Josuttis, The C++ Standard Library: A Tutorial and Reference, Addison-Wesley, Upper Saddle River,
NJ, 2012.
[33] K.J. Desabrais, Velocity field measurements in the near wake of a parachute canopy. ProQuest dissertations
and theses; Thesis–Worcester Polytechnic Institute, 2002.
[34] X. Pan, L. Hu, and Y. Cao, Analysis of dynamic simulation and fluid field of parachute in inflation stage,
J. Aerosp. Power. 1 (2008), pp. 015.
[35] R. Ross and F. Nebiker, Survey of aerodynamic deceleration systems, 3rd Annual Meeting, Boston, MA, 1966,
pp. 988.
[36] C. Jiang, Influence of low angle of attack on the flowfield characteristics of an axisymmetric parachute in
terminal descent. Chin. Space Sci. Technol. 27 (2007), pp. 59–65.
[37] J.C.C.Y. Wenhan, The flowfield characteristics of an axisymmetric parachute in terminal descent, Spacecr.
Recovery Remote Sens. 3 (2005), pp. 004.

You might also like