Drag Coefficient of A Sphere in A Non-Stationary Ow: New Results

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233845224

Drag coefficient of a sphere in a non-stationary flow: New


results

Article  in  Proceedings of The Royal Society A Mathematical Physical and Engineering Sciences · December 2007
DOI: 10.1098/rspa.2007.0058

CITATIONS READS

55 2,385

6 authors, including:

Georges Jourdan Ozer Igra


Aix-Marseille Université Ben-Gurion University of the Negev
92 PUBLICATIONS   1,060 CITATIONS    178 PUBLICATIONS   2,350 CITATIONS   

SEE PROFILE SEE PROFILE

Jean-Luc Estivalèzes Christophe Devals


The French Aerospace Lab ONERA Polytechnique Montréal
110 PUBLICATIONS   942 CITATIONS    61 PUBLICATIONS   481 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

multiscale interface tracking View project

Phase Inversion Project View project

All content following this page was uploaded by Jean-Luc Estivalèzes on 08 July 2014.

The user has requested enhancement of the downloaded file.


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere in a non-stationary


flow: new results
G Jourdan, L Houas, O Igra, J.-L Estivalezes, C Devals and E.E Meshkov

Proc. R. Soc. A 2007 463, doi: 10.1098/rspa.2007.0058, published 8


December 2007

References This article cites 12 articles, 1 of which can be accessed free


http://rspa.royalsocietypublishing.org/content/463/2088/3323.full.
html#ref-list-1

Email alerting service Receive free email alerts when new articles cite this article - sign up in
the box at the top right-hand corner of the article or click here

To subscribe to Proc. R. Soc. A go to: http://rspa.royalsocietypublishing.org/subscriptions


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Proc. R. Soc. A (2007) 463, 3323–3345


doi:10.1098/rspa.2007.0058
Published online 2 October 2007

Drag coefficient of a sphere in a non-stationary


flow: new results
B Y G. J OURDAN 1, * , L. H OUAS 1 , O. I GRA 2 , J.-L. E STIVALEZES 3 ,
C. D EVALS 3 AND E. E. M ESHKOV 4
1
Département de Mécanique Energétique, IUSTI, UMR CNRS 6595,
Polytech’ Marseille, Université de Provence, 5 rue Enrico Fermi,
13013 Marseille, France
2
Department of Mechanical Engineering, The Pearlstone Center for
Aeronautical Engineering Studies, Ben Gurion University,
Beer Sheva 84105, Israel
3
ONERA CERT DMAE, 2 Avenue Edouard Belin, B. P. 4025, 31055 Toulouse,
Cedex 4, France
4
Russian Federal Nuclear Center, Institute of Experimental Physics,
607190 Sarov, Russia

The drag coefficient of a sphere placed in a non-stationary flow is studied experimentally


over a wide range of Reynolds numbers in subsonic and supersonic flows. Experiments
were conducted in a shock tube where the investigated balls were suspended, far from all
the tube walls, on a very thin wire taken from a spider web. During each experiment,
many shadowgraph photos were taken to enable an accurate construction of the sphere’s
trajectory. Based on the sphere’s trajectory, its drag coefficient was evaluated. It was
shown that a large difference exists between the sphere drag coefficient in steady and
non-steady flows. In the investigated range of Reynolds numbers, the difference exceeds
50%. Based on the obtained results, a correlation for the non-stationary drag coefficient
of a sphere is given. This correlation can be used safely in simulating two-phase flows
composed of small spherical particles immersed in a gaseous medium.
Keywords: drag coefficient; non-stationary flow; shock tube experiments

1. Introduction

For evaluating the motion of a solid object in a gaseous medium, one has to know
the drag coefficient of the considered object. It is therefore not surprising that
significant research has been carried out and published on this matter during the
past century. In experimental investigations aimed at finding the drag coefficient
of a solid sphere moving through a fluid, various techniques were employed.
These different techniques were used for evaluating the drag coefficient over a
wide range of Reynolds and Mach numbers. Among the frequently used methods,
one should mention freely falling spheres in a liquid or air, spheres placed in wind
tunnels, spheres mounted on flying aircraft, spheres towed in water channels and
* Author for correspondence (georges.jourdan@polytech.univ-mrs.fr).

Received 30 May 2007


Accepted 7 September 2007 3323 This journal is q 2007 The Royal Society
Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3324 G. Jourdan et al.

spheres flying in aero-ballistic range. In many of these experiments, the relative


velocity of the sphere was constant or almost constant. Based on the results from
experiments conducted using the above-mentioned techniques, a ‘standard drag
coefficient’ curve has been derived for a sphere. For low subsonic speeds, one
curve represents the sphere drag coefficient over a large range of Reynolds
numbers; this curve is available in almost all fluid mechanics textbooks (see
Olson 1980). For high subsonic and supersonic flows, the flow Mach number
appears as a parameter (Bailey & Hiatt 1972; Bailey & Starr 1976; Henderson
1976). This detailed and accurate knowledge available for the drag coefficient of
a sphere is limited to the case where the sphere’s relative velocity is constant, i.e.
to a steady flow condition. However, in many engineering applications, the
sphere’s motion through the fluid in which it is immersed is not steady, i.e. it
experiences acceleration or deceleration. Examples of this are the nozzle flow of a
solid propellant rocket or flows behind shock/blast waves propagating into a
dusty gas, volcanic eruptions, etc. Using the standard drag curve for such flows
is, at least, questionable. To have a reliable simulation of a non-stationary two-
phase flow (solid particles immersed in a gaseous medium), an accurate
knowledge of the particle drag coefficient is essential. Measurements of sphere
drag coefficients for non-stationary flows were conducted during the past four
decades. Among the published results, one should mention the work of Buckley
(1968), Selberg & Nicholls (1968), Rudinger (1970), Karanfilian & Kotas (1978),
Outa et al. (1981), Sommerfeld (1985), Igra & Takayama (1993), Boiko et al.
(1997), Tano et al. (2003) and Sun et al. (2004). Selberg & Nicholls (1968) used a
shock tube for accelerating single particles and photographed their displacement
from which the particle trajectory was reconstructed and the drag coefficient was
deduced. The particle trajectory, which serves as the basis for obtaining the
particle velocity and acceleration, was obtained through a curve fit to the
observed particle displacements. Because the curve fit was derived from only five
observed particle positions, it is obvious that the deduced particle velocity and
acceleration are only a rough estimate of the real values and so is the particle
drag coefficient, CD. Furthermore, in the technique used, by injecting particles
into the shock tube channel, one observes many particles on the recorded film. In
such a case, some particles are flying in the wake produced by particles flying
ahead (closer to the incident shock wave), and thereby experiencing a drag force
which is different from that experienced by a single particle under similar
conditions. Their results were limited to Reynolds numbers within the range
200!Re!1700. Buckley (1968) investigated the drag force acting on a particles
cloud that was accelerated to a transonic slip flow, using a light-extinction
technique. Rudinger (1970) accelerated a particles cloud in a vertical shock tube
and deduced the particles’ velocity from their trajectories recorded by streak
photography. Since it was hard to establish an accurate particle trajectory, which
is the basis for obtaining the particles’ velocity and acceleration needed for
evaluating the particles’ CD, a different method was adopted. In the new
approach, Rudinger (1970) evaluated the particle velocity via the deflection of a
light beam passing through the particle cloud. From the ratio of the scattered
light intensity before the passage of the incident shock wave to that during the
passage of the particle cloud, the suspension density ratio across the shock can be
estimated. This in turn, for a steady one-dimensional flow, is equal to the ratio
between the shock velocity and the particles’ velocity. This, at best, would yield

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3325

an average velocity of the dust cloud, not the velocity of a specific particle.
Furthermore, the assumption that the post-shock flow is steady and one-
dimensional is questionable. In order that this assumption will not be grossly
violated, Rudinger (1970) was forced to limit his experiments to very weak
incident shock waves. His suggested correlation for CD was therefore limited to
the range 50!Re!300. Henderson (1976) proposed three different correlations
for CD: (i) for a subsonic flow, (ii) for a supersonic flow at Mach numbers MisO
1.75, and (iii) a linear interpolation equation for the intervening region. While
considering the effects of physical variables, such as Reynolds number, Mach
number and whether or not the flow is continuum or free molecular flow, on the
sphere drag coefficient, no attention was given to whether or not the relative flow
is steady. Karanfilian & Kotas (1978) measured the drag force on a sphere
experiencing a simple harmonic (unsteady) motion in a liquid at rest. Their
measurements were limited to the range 10!Re!6000. Outa et al. (1981)
deduced the particle drag from a streak record of dispersed glass spheres in the
air inside a shock tube. Sommerfeld (1985) essentially repeated Rudinger’s
experiments. He used laser Doppler anemometry to find a drag law describing the
particle acceleration behind an incident shock wave, in a vertical shock tube
filled with a uniform gas–particle suspension. His proposed correlation for the
particle CD was limited to the range 50!Re!500. An attempt to propose a
correlation for the particle drag coefficient over a wide range of Reynolds
numbers is found in the paper by Igra & Takayama (1993). In their study, a
shock tube was used for inducing a relatively high acceleration of small spheres
laid on the shock tube floor. Using double-exposure holography, the spheres’
trajectory could be reconstructed accurately while they flew along the 260 mm
windows of the tube test section. Based on such trajectories, the sphere drag
coefficient was evaluated for a range of 6000!Re!101 000. As was the case in
previous studies, the results in Igra & Takayama (1993) also suggested that the
drag coefficients obtained for a sphere in a non-stationary flow are significantly
larger than those obtained in a similar steady flow. Boiko et al. (1997) embarked
on an ambitious investigation, studying experimentally and theoretically the
shock wave interaction and propagation through a dust cloud. The dust volume
concentration was within the range of 0.1–3%. Obviously, when dealing with a
dust cloud, constructing a particle trajectory is impossible; instead, the frontal
boundary of the cloud was reconstructed from the recorded shadowgraphs. For
the theoretical part, a correlation relating the particle’s drag coefficient to the
appropriate flow Re and Mach numbers is needed. The required correlation was
obtained by repeating experiments with a single particle. These experiments
were limited to a narrow range of Reynolds numbers, 1.5!104!Re!2.5!104.
The main findings of this investigation are that a reflected shock wave is formed
ahead of the particles cloud and a compression wave arises inside the cloud due to
the deceleration of the gas by the particles.
In all the shock tube experiments described so far, the diameter of the spheres
used was very small in comparison with the dimension of the shock tube test
section. Therefore, the time to complete the shock wave diffraction over the
sphere was negligibly small and, during most of the investigated time, the sphere
acceleration was due mostly to the relative velocity existing between the sphere
and the post-shock gas flow. This was not the case in the studies carried out by
Tano et al. (2003) and Sun et al. (2004), where a large sphere (80 mm in

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3326 G. Jourdan et al.

diameter) was tested inside a 300 mm!300 mm cross-section shock tube. In this
study, the sphere drag coefficient was evaluated experimentally and numerically
during its interaction process with the incident shock wave. During the process
starting when the incident shock wave collides with the sphere and continues
while it diffracts over it, the sphere experiences a steep increase in the applied drag
force. This, almost instantaneous, rise in the drag coefficient reaches CDZ10.
Thereafter, the drag coefficient drops down quickly as the shock wave moves to
the rear side of the sphere. Once the shock wave propagated downstream from the
sphere, the drag coefficient reaches a low and stable value (see figs 5 and 6 of Sun
et al. (2004)). This stable value is the drag experienced by the sphere during its
non-stationary motion in the post-shock gas flow. The duration of the very quick
rise and decline in the sphere’s drag is approximately equal to the time taken by
the shock wave to travel a distance equal to the sphere’s diameter. For spheres
tested in the present work, where the largest diameter was 6.5 mm, the time of
shock diffraction over the sphere was negligibly small.
Igra & Takayama’s (1993) work provided a significant improvement to the
previously available data regarding the drag coefficient of a sphere in a non-
stationary flow because the evaluation of the sphere’s trajectory was based on a
large number of recorded sphere locations taken during a relatively long time
(optical field of view of 260 mm!150 mm). It also covered a much wider range of
Reynolds numbers than in the previously published works. However, it suffered
from two basic setbacks. Before the arrival of the incident shock wave, the
investigated small spheres were placed on the shock tube floor. When they
started their motion, after the incident shock wave had passed them, they were
partly submerged inside the boundary layer developed along the shock tube
walls. It was argued by Igra & Takayama (1993) that, during the time when the
particle starts its motion, the prevailing boundary-layer thickness is very small
and most of the investigated particles were out of this layer. Nevertheless, one
could not blindly ignore the effect of the boundary layer and the friction between
the shock tube wall and the particle on the evolved sphere’s trajectory and
thereby on the deduced CD. It is crystal clear that starting the particle motion
when it is far away from all walls would yield more reliable results. The second
disadvantage lies in the following facts.

— In all of the Igra & Takayama (1993) results, the post-shock gas velocity was
subsonic. To have a reliable correlation for CD in a non-steady flow, one should
include a significant amount of both subsonic and supersonic post-shock flows.
— To obtain the particle trajectory, many runs had to be repeated with the same
initial conditions but photographing the sphere at different times during its
motion. This was needed since, in each run, only one photograph was taken. Even
with a high repeatability, such a procedure is a source for unavoidable errors in
the construction of the sphere’s trajectory.

Later, Rodriguez et al. (1993) investigated both the steady and unsteady drag
coefficients of a sphere initially free-falling in a vertical shock tube, using a rapid
camera shadowgraph technique. Their work investigated the drag coefficient of a
sphere, but, in each experiment, 20–30 particles were dropped down the vertical
shock tube in order to ensure that at least one of them would be present in the
field of view during its interaction with the incident shock wave. No correction

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3327

was made to account for the wake interference between the falling particles.
More recently, Suzuki et al. (1999) developed a technique enabling the injection
of a spherical particle into the middle of a horizontal shock tube test section, just
before the incident shock wave reached the injected sphere. A triple-exposure
photographic technique was used for recording the particle displacement caused
by the incident shock wave. The test section field of view in their experiments
was only 200 mm in the flow direction and the incident shock wave Mach number
was within the range of 1.10%Mis%1.40, resulting in the subsonic post-shock
flow in all their experiments.
The significant achievement of the present experimental study described
subsequently is the method used for the drag force measurement by tracing the
trajectory of a single spherical particle suspended in mid-air by a fine filament
(spider web) from the roof of a shock tube. Multiple shadowgraphs (several tens)
were acquired at each shot, eliminating the serious repeatability-related
inaccuracy encountered in former studies where only a single or a few
shadowgraphs were recorded per shot. Moreover, by employing multiple-sphere
shots (up to four different particles), the repeatability-associated inaccuracy was
further reduced. In nine experiments, the post-shock flow was subsonic and in
seven supersonic. In four of the subsonic experiments, two different spheres were
tested simultaneously and, in one, three different spheres were tested
simultaneously. In one of the supersonic experiments, two different spheres were
tested simultaneously; in two experiments three different spheres were tested
simultaneously; and, in one run, four different spheres were tested simultaneously.
Furthermore, in the presently reported results, the field of view was 300 mm long,
enabling a very accurate construction of the investigated particle trajectory
during each test.

2. Experimental facilities

Experiments were conducted in the 80!80 mm cross-section, multi-phase


variable inclination shock tube of the University of Provence. This shock tube,
in its horizontal position, and the diagnostics used are shown in figure 1. The
total length of the tube is 3.75 m, 75 cm is the driver’s length and 3 m is
the driven section length. The driver pressure can be raised up to 20 bars and the
driven section pressure can be reduced down to 0.5 mbar to result in a maximum
pressure ratio of approximately 40 000. Consequently, experiments with the
incident shock wave Mach number within the range 1.1!Mis!5 can be easily
conducted. This in turn allows the present investigations to cover both subsonic
and supersonic post-shock flow cases. As was done in previous experiments
(Igra & Takayama 1993; Rodriguez et al. 1993; Suzuki et al. 1999), here too the
sphere drag coefficient, CD, was deduced from its trajectory. To obtain an
accurate reconstruction of the investigated sphere’s trajectory, the windows in
the presently used test section had a field of view of 80!300 mm (300 mm in the
flow direction). The spatial locations of the sphere during its motion, induced by
the post-shock flow, were recorded using a shadowgraph technique coupled with
a high-speed camera and a stroboscopic Nanolite flash lamp that was
synchronized with the rotating-drum camera. This optical facility enables
photographing every 70 ms during each experiment (i.e. having approx. 60 and

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3328 G. Jourdan et al.

flash lamp scope gas pumping


TDS 510 and filling

pressure triggering
transducers’ amplification high pressure
power supply
M2
pressure transducers

M1
f=300 mm
f =3 m
p
flash lam
delayed
flash lamp triggering
control

came
ra co
ntrol

le)
'Marseil
(Polytech d
odrum high-speeamera
aederlé strob
L. Schw c

Figure 1. A layout of the experimental facility.

approx. 35 photos per run in the subsonic and supersonic cases, respectively),
thereby ensuring the accurate construction of the tested sphere’s trajectory. PCB
piezoelectric transducers were used for recording pressure histories, deducing the
incident shock wave Mach number and triggering optical and recording facilities.
Recording the post-shock pressure history is essential since we limited the
reconstruction of the sphere’s trajectory to the duration in which a uniform post-
shock flow prevails. During this period of time, the post-shock flow properties can
be easily predicted using the Rankine–Hugoniot shock relations.
The diameter (f), the material density (rp) and the type of spheres used in the
present experiments are listed in table 1.
A crucial point in the experiments was the problem of how to keep the tested
spheres away from the test section walls. After checking various options, the
following method was chosen. The tested sphere was suspended from the tube
ceiling, close behind the entrance to the test section, on a wire taken from a
spider web. The spider wire was strong enough to keep the small sphere
suspended in the air until the arrival of the incident shock wave. This technique,
although very delicate, is made possible thanks to the sticky substance that
covers the spider wire. Furthermore, the spider wire quickly accelerates to the
post-shock gas flow velocity before disintegrating. Figure 2a–c contains three
shadowgraphs taken, respectively, 36 ms before the shock collision with the
sphere, 34 ms after the incident shock wave hit the sphere and 104 ms after its
interaction with the 1.92 mm diameter nylon sphere suspended on the spider
wire. As can be seen in figure 2, the process of the spider wire acceleration and
disintegration is completed very quickly after it collides, head-on, with the
incident shock wave. Disturbances produced by the spider wire and its breaking

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3329

(a)

t = –36 µs
(b)
t = 34 µs

(c)
t = 104 µs

Figure 2. Three spheres suspended on a spider wire. (a) Before the arrival of the incident shock
wave, (b) just after and (c) shortly after the head-on collision between the spheres and the incident
shock wave. In the present case, the flow behind the shock is supersonic and the incident shock
wave moves from right to left.

Table 1. Diameter (f), density (rp), geometry and material of the tested spheres.

f (mm) rp (kg mK3) type material

6.5 25 expanded foam polystyrene


2.3 25 expanded foam polystyrene
2.6 25 expanded foam polystyrene
1.5 25 expanded foam polystyrene
2.55 25 expanded foam polystyrene
2.6 25 expanded foam polystyrene
1.3 300 hollow sphere polymer ATECA
1.4 300 hollow sphere polymer ATECA
1.55 300 hollow sphere polymer ATECA
1.63 300 hollow sphere polymer ATECA
0.35 600 hollow sphere polypropylene
0.29 600 hollow sphere polypropylene
1.92 1130 solid (full) sphere nylon
1.96 1204 solid (full) sphere nylon
0.66 1550 solid (full) sphere nylon
0.5 1550 solid (full) sphere nylon
0.64 1270 solid (full) sphere nylon
0.68 1270 solid (full) sphere nylon
2.9 900 solid (full) sphere polystyrene
1.64 1056 solid (full) sphere polystyrene
0.61 1180 solid (full) sphere polystyrene
1.53 1064 solid (full) sphere polystyrene
0.62 1096 solid (full) sphere polystyrene

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3330 G. Jourdan et al.

were hardly noticeable. On the other hand, given that in the considered case the
post-shock flow is supersonic, the detached shock wave from the sphere is easily
noted in figure 2c.

3. Theoretical background

When a solid particle is exposed to a gas flow, its response depends on the
relative velocity that exists between the particle and the flow. For a low
concentration of solid particles in a suspension (and surely in the case when only
a few small spheres are immersed in the flowing gas), one can ignore both the
interaction between solid particles and their contribution to the suspension
pressure. In such a case, the drag force acting on the solid particle(s) is the sole
meaningful force that determines the particle motion. In experiments conducted
in shock tubes with relatively small (and therefore light) particles, the particles
experience a very large acceleration due to the very fast post-shock gas flow.
Until the particles reach the post-shock flow velocity, the relative velocity
between the particle and the gas flow changes and the particle motion is non-
stationary. Should the particle trajectory be recorded accurately, its drag
coefficient could be evaluated as follows. The equation of motion of a solid
particle accelerated by the gas flow is
dU p 3 rg CD
ZgC kU p KU g kðU p KU g Þ; ð3:1Þ
dt 4 rp f
where Up, f and rp are the solid sphere’s velocity, diameter and material density,
respectively. Ug and rg are the gas velocity and density, respectively. It was
shown by Igra & Takayama (1993) that, based on equation (3.1), the particle
drag coefficient and the appropriate Reynolds number can be expressed as
follows:
 2
dup
r
4 p dt
CD Z rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
 2  2 and ð3:2Þ
3 rg dup dvp
ðug Kup Þ dt C dt Kg

 ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
s   2
rg f  ug Kup  dup 2 dvp
Rep Z Kg ; ð3:3Þ
mg  dup 
C
dt dt
dt

where u and v are the components of the velocity vector U in the x and y
directions, respectively. g is the gravity acceleration and mg is the gas viscosity.
Similarly, the sphere’s Mach number, based on the relative velocity, is
 sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
   2
1  ug Kup  dup 2 dvp
Mp Z pffiffiffiffiffiffiffiffiffiffiffi  du  C Kg ; ð3:4Þ
gRT p dt dt
dt

where g and R are the gas specific heat ratio and the gas constant, respectively.
In the following, the trajectory of a sphere suspended in the entrance to the
shock tube test section will be reconstructed accurately using the optical system
described earlier. Each trajectory is a curve fit passing, on average, through

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3331

35–60 recorded locations passed by the considered particle during its motion
imposed by the incident shock wave. Once the particle trajectory is available, its
velocity and acceleration can be easily obtained by the first differentiation of its
trajectory (velocity) and the second differentiation yields its acceleration.
Substituting the obtained values for the sphere’s velocity and acceleration in
equations (3.2)–(3.4) provides the sphere’s CD, Reynolds and Mach numbers.

4. Results and discussion

Unlike in the experimental work of Igra & Takayama (1993), where many
experiments were repeated in order to reconstruct the sphere’s trajectory, in the
present case, approximately 35–60 photos with a time difference of 70 ms between
successive frames were taken in a single test. Therefore, the obtained temporal
locations of the flying sphere were more than needed for an accurate
reconstruction of the sphere’s trajectory. In the present results, the accuracy
of measuring the sphere location, from the recorded shadowgraphs, is
within G1 mm. A summary of all the conducted experiments is given in
table 2. It provides information about the tested spheres, the initial and the
prevailing conditions in the shock tube. In some experiments, a few particles were
tested in the same run. In such cases, the different spheres were suspended either
along a line perpendicular to the flow direction for minimizing the wake
interference between the tested spheres, or with a significant longitudinal gap
between them. The different spheres tested in the same run are marked by letters
‘a’, ‘b’, ‘c’, etc. appearing after the run number in table 2. In all the experiments
where the post-shock flow was subsonic, the flying sphere experiences a uniform
post-shock flow until the arrival of the reflected shock wave from the tube end
wall. (x–t) diagrams (computed using the random choice method), showing the
lines of constant density, appear in figure 3. Numbers appearing in the uniform
flow zones indicate the local density in kg mK3. A typical case where the post-
shock flow is subsonic is shown in figure 3a. In all the experiments where the post-
shock flow was supersonic, the flying sphere experiences a uniform post-shock flow
until the arrival of the contact surface as shown in figure 3b. In evaluating the
sphere drag coefficient, only the part of the sphere’s trajectory reconstructed
inside the uniform post-shock flow zone was used. In this zone, the post-shock flow
conditions were derived using the Rankine–Hugoniot shock relations and the
measured incident shock wave Mach number. In the following, the procedure used
for deducing the drag coefficient of spheres exposed to the shock wave-induced
flow is outlined for four different cases. In two cases, the post-shock-induced flow
was subsonic and, in the other two, supersonic.
The sample of the obtained shadowgraphs for the subsonic cases is shown in
figures 4 (run no. 148, where 14 out of the 60 recorded shadowgraphs are shown)
and 5 (run no. 166, where 14 out of the 63 recorded shadowgraphs are shown).
The results obtained for the supersonic cases are shown in figures 6 (run no. 142,
where 14 out of the 36 recorded shadowgraphs are shown) and 7 (run no. 184,
where 12 out of the 15 recorded shadowgraphs are shown). The distance between
the two vertical reference lines appearing on all shadowgraphs was 21.3 cm
except in run no. 142, where it was 19.7 cm.

Proc. R. Soc. A (2007)


Proc. R. Soc. A (2007)

3332
Table 2. Summary of all results. (Pi is the initial pressure (atm.), Mis is the incident shock wave Mach number, ug is the flow velocity (m sK1), upmax is
the maximal velocity reached by the sphere (m sK1), f (mm) and rp (kg mK3) are the sphere’s diameter and the density, Mp and Rep are the particular
Mach and Reynolds numbers, respectively, Ac is the acceleration factor and CD is the sphere drag coefficient.)

upmax =ug
run no. driver/test gases Pi Mis ug upmax (%) f rp Mp Rep Ac CD

Downloaded from rspa.royalsocietypublishing.org on March 12, 2014


subsonic flows
148 air/air 1 1.56 271 85(3300 ms) 31 1.92 1130 0.47–0.66 40 000–56 000 9.2!10K4 0.60
149 air/air 1 1.56 271 70(2250 ms) 27 1.92 1130 0.47–0.64 37 000–50 000 1.2!10K3 0.80
163 air/air 1 1.49 236 65(3350 ms) 27 1.64 1056 0.42–0.59 27 000–38 000 7.5!10K4 0.50
165 air/air 1 1.49 236 50(3500 ms) 21 2.9 900 0.49–0.60 54 500–68 500 8.5!10K4 0.48
166a air/air 1 1.51 244 60(3500 ms) 25 1.96 1204 0.47–0.61 37 000–48 000 7.9!10K4 0.54
166b air/air 1 1.51 244 200(2500 ms) 82 6.5 25 0.10–0.16 27 000–41 000 3.5!10K2 0.60

G. Jourdan et al.
173a air/air 1 1.5 236 115(3250 ms) 49 1.55 300 0.32–0.60 20 000–36 000 1.7!10K3 0.30
173b air/air 1 1.5 236 105(3250 ms) 45 1.63 300 0.34–0.60 22 000–38 000 1.6!10K3 0.30
182a air/air 1 1.5 236 100(3000 ms) 42 0.66 1180 0.34–0.58 8200–14 000 6.5!10K4 0.46
182b air/air 1 1.5 236 100(3000 ms) 42 0.61 1550 0.33–0.43 9000–15 000 6.5!10K4 0.60
186a air/heliumC5% air 0.1 1.4 512 130(1250 ms) 25 2.3 25 0.33–0.43 950–1250 1.3!10K3 1.05
186b air/heliumC5% air 0.1 1.4 512 110(1250 ms) 22 2.6 25 0.34–0.43 1130–1430 1.0!10K3 1.03
186c air/heliumC5% air 0.1 1.4 512 20(1250 ms) 4 1.4 300 0.41–0.43 735–760 6.0!10K5 1.02
187a air/heliumC24% air 0.05 1.84 632 110(1250 ms) 17 1.5 100 0.64–0.75 1270–1470 3.3!10K4 1.00
187b air/heliumC24% air 0.05 1.84 632 30(1250 ms) 5 0.5 1550 0.72–0.74 471–485 2.5!10K5 1.02
supersonic flows
142 helium/argon 0.07 3.88 845 60(1500 ms) 7 1.92 1130 1.09–1.15 12 000–13 000 2.3!10K4 0.70
145 helium/argon 0.05 4.81 1101 750(800 ms) 68 0.35 600 0.59–1.16 660–1350 4.5!10K4 1.01
179a helium/air 0.024 4.35 1176 250(1250 ms) 21 1.4 300 1.29–1.59 3600–4500 2.5!10K4 0.70
179b helium/air 0.024 4.35 1176 105(1500 ms) 9 1.92 1130 1.46–1.60 5600–6100 1.0!10K4 1.02
180a helium/air 0.027 4.23 1142 280(1400 ms) 25 1.3 300 1.19–1.58 2950–3950 2.4!10K4 0.90
180b helium/air 0.027 4.23 1142 140(1500 ms) 12 1.53 1064 1.40–1.58 4040–4600 1.3!10K4 1.04
180c helium/air 0.027 4.23 1142 100(1500 ms) 9 2.9 900 1.46–1.58 8000–8800 1.6!10K4 1.05
181a helium/air 0.023 4.23 1140 250(1500 ms) 22 0.62 1096 1.25–1.59 1450–1900 1.0!10K4 1.03
181b helium/air 0.023 4.23 1140 250(1500 ms) 22 0.64 1270 1.27–1.60 1540–1950 1.0!10K4 1.03
(Continued.)
Proc. R. Soc. A (2007)

Downloaded from rspa.royalsocietypublishing.org on March 12, 2014


Table 2. (Continued.)

Drag coefficient of a sphere


upmax =ug
run no. driver/test gases Pi Mis ug upmax (%) f rp Mp Rep Ac CD

181c helium/air 0.023 4.23 1140 250(1500 ms) 22 0.68 1270 1.28–1.60 1650–2100 1.0!10K4 1.03
183av helium/air 0.029 4.06 1089 400(700 ms) 37 0.29 600 1.17–1.58 750–1150 2.5!10K4 1.02
183ar helium/air 0.029 4.06 1089 250(500 ms) 23 0.29 600 1.17–1.58 750–1150 2.5!10K4 1.02
184a helium/air 0.031 4.06 1089 610(500 ms) 56 1.6 25 0.70–1.53 3000–6000 4.2!10K3 0.85
184b helium/air 0.031 4.06 1089 580(500 ms) 53 2.3 25 0.88–1.49 5000–9000 3.5!10K3 0.74
184c helium/air 0.031 4.06 1089 410(500 ms) 38 2.55 25 0.80–1.49 5000–10 000 5.2!10K3 1.00
184d helium/air 0.031 4.06 1089 390(500 ms) 36 2.3 25 0.78–1.51 5000–9000 3.5!10K3 0.71

3333
Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3334 G. Jourdan et al.

(a) (b)
12.0 5.0
10.0 4.0
8.0
t (×10–3 s)

4.28 reflected 0.898


shock wave 3.0
6.0
the sphere's 2.0 the sphere's
trajectory 0.388
4.0 2.37 trajectory
4.42 Incident contact incident
0.422 shock wave
2.0 shock wave surface 1.0 0.115
1.19 contact
10.18 1.153 surface
0 1 2 3 0 1 2 3
x (m) x (m)

Figure 3. Theoretical (x–t) diagram and experimental trajectory of a nylon particle


(rpZ1130 kg mK3) of 1.92 mm in diameter (the initial x-coordinate of the particle is xZ2.97 m):
(a) in a subsonic post-shock flow, run no. 148 and (b) in a supersonic flow, run no. 142.

While in run no. 148 (figure 4) only one sphere was present in the test section,
in run no. 166 (figure 5) two different spheres were employed. One, a nylon sphere
(fZ1.96 mm and rpZ1204 kg mK3), was suspended on a spider wire just at the
entrance to the test section (see the frame taken at tZ0 ms in figure 5), the second
sphere, made of polystyrene foam (fZ6.5 mm and rpZ25 kg mK3), was laid on
the shock tube floor far upstream of the nylon sphere, and therefore is not seen in
figure 5 until tZ1109 ms. It is apparent from the shadowgraphs shown in figures
4–7 that the spheres move along horizontal lines, and therefore one can safely
neglect the gravity effect and a possible sphere’s rotation. As expected, the sphere
response to the shock-induced flow is not instantaneous. It is apparent from
figure 4 that, for the first 300 ms after the collision between the incident shock
wave and the nylon sphere, the sphere hardly moves. The same delay is observed
in figure 5 for the nylon ball. The 6.5 mm polystyrene foam ball is significantly
lighter than the 1.96 mm nylon ball. Although it was placed far upstream of the
nylon ball, it reaches the smaller ball at approximately tZ1110 ms and soon
thereafter passes it (figure 5). As seen in figure 5, shortly before tZ2500 ms, the
large light polystyrene foam ball reaches the end of the field of view and soon
thereafter the reflected shock wave from the shock tube end wall enters the field
of view. This significant difference in velocity between the two balls, tested in run
no. 166, will be seen clearly in the reconstructed sphere’s trajectories.
Based on all the recorded shadowgraphs, the sphere’s trajectories during
experiment nos. 148 and 166 were reconstructed; they are shown in figure 8a,b,
respectively, and those deduced from the shadowgraphs taken during experiment
nos. 142 and 184 are shown in figure 9a,b, respectively. The straight lines
appearing in figures 8 and 9 represent the incident shock wave and the shock
wave reflected from the shock tube end wall. The appropriate wave’s velocity is
indicated near these lines. The sphere’s velocity can be obtained from the results
shown in figures 8 and 9 by either of the following two options.

— Executing a curve fit through the recorded points that indicate the different
sphere locations during the considered experiment. The first differentiation of
this curve fit provides the sphere’s velocity.

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3335

fixed lines 21.3 cm apart

t = 1786 µs
t<0

incident

t = 2276 µs
t = 36 µs

shock wave

t = 2626 µs
t = 316 µs

t = 2836 µs
t = 456 µs

t = 3116 µs
t = 736 µs

reflected
shock wave
t = 1086 µs

t = 3326 µs
t = 1506 µs

t = 3676 µs

Figure 4. Sample of shadowgraphs from experiment no. 148 in which the post-shock flow is
subsonic carrying a nylon sphere, 1.92 mm in diameter and rpZ1130 kg mK3. The incident shock
wave moves from right to left.

— Conducting numerical differentiation of the sphere’s path as represented by


the recorded discrete sphere displacements (shown in figures 8 and 9).

Both options were checked. In the cases where the recorded sphere locations
exhibit a smooth, monotonic change, the difference between the two results
obtained for the sphere’s velocity was very small (less than 2%). The results
obtained for the sphere’s velocities via numerical differentiation of the recorded
sphere locations (shown in figures 8 and 9) appear as discrete points in figures 10
and 11, respectively. The line passing through these discrete points is the
sphere’s velocity obtained by the differentiation of the curve fit to the recorded
sphere locations. The straight solid lines in figures 10 and 11 indicate the gas
velocity ahead and behind the incident shock wave. The sphere’s acceleration,

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3336 G. Jourdan et al.

fixed lines 21.3 cm apart

t = 1809 µs
t<0

t = 2159 µs
incident
t = 59 µs

shock wave

t = 2369 µs
t = 339 µs

t = 2509 µs
t = 689 µs

t = 3139 µs
t = 899 µs

reflected
shock wave
t = 1109 µs

t = 3489 µs
t = 1459 µs

t = 3909 µs

Figure 5. Sample of shadowgraphs from experiment no. 166 in which the post-shock flow is
subsonic carrying two boles: a nylon sphere, 1.96 mm diameter and rpZ1204 kg mK3, and a foam
sphere, 6.5 mm diameter and rpZ25 kg mK3. The incident shock wave moves from right to left.

needed for evaluating its drag coefficient, can be obtained by differentiating its
velocity curve. Since numerical differentiation yields a very noisy acceleration
curve, we used the polynomial curve fit to the recorded sphere locations as a basis
for evaluating the sphere’s velocity and acceleration. Polynomials of different
orders could be fitted to the given set of sphere locations for constructing its
trajectory. Choosing a high-order polynomial fit resulted in a relatively noisy
acceleration curve. Therefore, a third-order polynomial fit was used for
constructing the sphere’s trajectory. The values obtained for the sphere’s
velocity and acceleration (by first and second differentiations of the third-order
polynomial fit to the sphere’s trajectory) were substituted into equation (3.2) for
deducing the sphere drag coefficient. The appropriate sphere’s Reynolds number
was obtained from equation (3.3). The results obtained for the sphere’s CD versus

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3337

fixed lines 19.7 cm apart

t = 1518 µs
reflected
shock wave
t<0

incident

t = 1618 µs
t = 18 µs

shock wave

t = 1718 µs
t = 118 µs

t = 2318 µs
t = 418 µs

t = 2618 µs
t = 718 µs

t = 2818 µs
t = 818 µs

t = 3318 µs
t = 918 µs

Figure 6. Sample of shadowgraphs from experiment no. 142 in which the post-shock flow is
supersonic carrying a nylon sphere, 1.92 mm in diameter and rpZ1130 kg mK3. The incident shock
wave moves from right to left.

Reynolds number, for the subsonic cases (run nos. 148 and 166), are shown in
figure 12a,b, respectively. The results obtained for the cases where the post-shock
gas flow is supersonic (run nos. 142 and 184) are shown in figure 13a,b,
respectively. Also shown in figures 12 and 13 are the sphere’s Mach number
(deduced from equation (3.4)) and the appropriate ‘standard drag curve’. It is
clear from the results shown in figures 12 and 13 that, for the covered range of
Re, appearing in each figure, the standard drag curve is practically constant and
the changes in the deduced unsteady sphere drag coefficient are moderate.
Therefore, an average value of CD was deduced from the results shown in figures
12 and 13 to be associated with the appropriate average value of Re. This
average value of CD appears in table 2 and figure 15, where the results obtained
from a specific run appear as a single point. For example, the results shown in

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3338 G. Jourdan et al.

fixed lines 21.3 cm apart

t = 364 µs
t< 0

incident
t = 14 µs

t = 434 µs
shock wave

t = 504 µs
t = 84 µs
t = 154 µs

t = 574 µs
t = 224 µs

t = 644 µs
t = 294 µs

t = 714 µs

Figure 7. Sample of shadowgraphs from run no. 184 in which the post-shock flow is supersonic
carrying four foam spheres whose diameters are 1.6, 2.2, 2.55 and 2.3 mm. All have a density of
25 kg mK3. The incident shock wave moves from right to left.

(a) T80 no. 148 (b) T80 no.166


0.8
f = 1.92 mm f = 1.96 mm
0.7 (r = 1130 kg m–3) (r = 1204 kg m–3)
0.6 f = 6.5 mm
(r = 25 kg m–3)
s –1
1
ms –

0.5
W rs

W rs
m
x (m)

=–

20

0.4
45

=–
32

=5
=5

32
5m

0.3
4m
is
is

W
s–
W

0.2
s
–1

0.1

0 1.0 2.0 3.0 4.0 0 1.0 2.0 3.0 4.0


t (×10 –3 s) t (×10 –3 s)

Figure 8. Trajectories of spheres tested in run nos. (a) 148 and (b) 166.

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3339

T80 no. 142 T80 no. 184


(a) 0.8 (b)

391 m s –1
f = 1.92 mm ( r = 1130 kg m–3) f = 1.60 mm (r = 25 kg m–3)

W rs
m s –1
0.7 f = 2.20 mm (r = 25 kg m–3)

W rs

=–
0.6 f = 2.30 mm (r = 25 kg m–3)
is 1210

520
f = 2.55 mm (r = 25 kg m–3)

=–

Wis = 1
0.5

m
647
W =
x (m)

s–
0.4

1
m s –1
ms
–1
0.3

090
0.2

ug = 1
0.1

0 1.0 2.0 3.0 4.0 0 1.0 2.0 3.0 4.0


t (×10 –3 s) t (×10 –3 s)

Figure 9. Trajectories of spheres tested in run nos. (a) 142 and (b) 184.

(a) T80 no. 148 (b) T80 no. 166


300 u = 271 m s–1
g f = 1.92 mm (r = 1130 kg m–3) f = 6.5 mm (r = 25 kg m–3)
ug = 244 m s–1 f = 1.96 mm (r = 1204 kg m–3)
250

200
up (m s–1)

150

100

50
ug = 0 ug = 0
0 1.0 2.0 3.0 4.0 0 1.0 2.0 3.0 4.0
t (×10 –3 s) t (×10–3 s)

Figure 10. (a,b) The sphere’s velocity obtained by numerical differentiation of the sphere’s
trajectory shown in figure 8a,b, respectively.

(a) (b)
T80 no. 142 T80 no. 184
1200 u = 1089 m s–1
800 u = 845 m s–1 f = 1.92 mm g
f = 1.6 mm
g (r = 1130 kg m–3) 1000 (r = 25 kg m–3)
600 800 f = 2.2 mm
up (m s–1)

( r = 25 kg m–3)
600 f = 2.3 mm
400 ( r = 25 kg m–3)
400 f = 2.55 mm
200 ( r = 25 kg m–3)
ug = 0 200
ug = 0
0 1.0 2.0 3.0 4.0 0 1.0 2.0 3.0 4.0
t (×10 –3 s) t (×10–3 s)

Figure 11. (a,b) The sphere’s velocity obtained by numerical differentiation of the sphere’s
trajectory shown in figure 9a,b, respectively.

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3340 G. Jourdan et al.

(a) 10 T80 no. 148 2.0


f = 1.92 mm (r = 1130 kg m–3)
CD
MP 1.5
standard drag curve
CD 1 1.0 Mp

0.5

0.1 0
4.0 4.5 5.0 5.5
Rep (×10 4)

(b)
T80 no. 166 T80 no. 166
10 2.0
f = 1.96 mm (r = 1204 kg m–3) f = 6.5 mm (r = 25 kg m–3)
CD CD
MP MP 1.5
standard drag curve standard drag curve
CD 1 1.0 Mp

0.5

0.1 0
3.5 4.0 4.5 5.0 2.5 3.0 3.5 4.0
4
Rep (×10 ) 4
Rep (×10 )

Figure 12. (a,b) The sphere drag coefficient and Mach number derived from the results shown in
figure 8a,b, respectively.

figure 12a will appear as CDZ0.6 and ReZ48 000. Furthermore, it is clearly
visible from figures 12 and 13 that the obtained values of CD are higher than
those obtained for a similar CD in a steady flow.
The two spheres in run no. 166 had very different characteristics: one
(polystyrene foam) was relatively large and very light while the other (a nylon
ball) was relatively small and heavy. The polystyrene foam ball experiences a
relatively large drag force and therefore it is quickly accelerated towards the
prevailing post-shock gas flow velocity. As a result, its relative velocity u gKu p
was smaller than that of the nylon sphere. This explains the lower particle Mach
number seen in figure 12b for the polystyrene foam ball.
The sphere drag coefficient deduced from the present experiments is based on
the total drag force acting on the considered sphere. In the present non-steady
flow, this drag is a summation of a few different contributions, as follows: drag
due to friction (flow viscosity) and form (pressure distribution on the immersed
body); added mass; shock diffraction effect during the incident shock passage
over the particle; and history (Basset force). The contribution of the added mass
to the total drag coefficient depends on the ratio between the immersed object’s
material density and the density of the fluid in which the object is immersed. The
closer the object material density is to that of the fluid, the larger will be the
added mass contribution to the total drag coefficient. Note that, in the present
work, the density ratio ranges from one to several hundreds. Moreover, in light of

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3341

(a) T80 no. 142


10 2.0
f = 1.92 mm (r = 1130 kg m–3)
CD
MP 1.5
standard drag curve
CD 1.0 Mp
1

0.5

0.1 0
1.20 1.22 1.24 1.26 1.28 1.30
Rep (×10 4)
(b) T80 no. 184 T80 no. 184
10 2.0
f = 1.6 mm (r = 25 kg m–3) f = 2.2 mm (r = 25 kg m–3)
CD CD
MP MP 1.5
standard drag curve standard drag curve
CD 1 1.0 Mp

0.5

0.1 0
2.5 3.5 4.5 5.5 6.5 4.5 5.5 6.5 7.5 8.5 9.5
T80 no. 184 T80 no. 184
10 2.0
f = 2.3 mm (r = 25 kg m–3) f = 2.55 mm (r = 25 kg m–3)
CD CD
MP MP 1.5
standard drag standard drag curve
curve
CD 1 1.0 Mp

0.5

0.1 0
4.5 5.5 6.5 7.5 8.5 9.5 5.0 6.0 7.0 8.0 9.0 10.0
Rep (×10 3) Rep (×10 3)

Figure 13. (a,b) The sphere drag coefficient and Mach number derived from the results shown in
figure 9a,b, respectively.

this, it is of interest to note that, in run nos. 142 and 184, spheres of similar
diameter were tested: 1.92 mm in run no. 142, 2.2 mm (‘a’) and 2.3 mm (‘d’) in
run no. 184. However, there is a significant difference in their material densities;
while in run no. 142, rpZ1130 kg mK3, in run no. 184 the sphere’s material
density is only 25 kg mK3. Therefore, if the added mass contribution in the
present experiments would have been of any importance, one would expect to
find a significant difference between the drag coefficients obtained in these two
runs. Inspecting the results shown in table 2 indicates that this is not the case;
the average CD for the sphere in run no. 142 is 0.7 while for the two similar
spheres ‘a’ and ‘d’, in run no. 184, CDZ0.74 and 0.71, respectively. It is therefore

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3342 G. Jourdan et al.

3 subsonic Mp < 1
run no. 148
run no. 149
run no. 163
run no. 165
run no. 166
run no. 173
run no. 182
2 run no. 186
run no. 187
supersonic Mp > 1
run no. 142
CD run no. 145
run no. 179
run no. 180
run no. 181
run no. 183
1 run no. 184
standard curve
Igra and Takayama correlation
present work

0
10 2 10 3 10 4 10 5
Rep

Figure 14. Summary of all results in a CD–Re plane.


safe to assume that in the present experiments, the contribution of the added
mass to the total drag coefficient is small.
It was mentioned that the results for the sphere’s CD shown in figures 12 and
13, and those to be shown subsequently, are based on a third-order polynomial fit
to the recorded sphere locations. This curve fit was differentiated to obtain the
sphere’s velocity and its second differentiation yields the sphere acceleration. It is
reasonable to ask whether the choice of a third-order polynomial fit is physically
justified, as such a choice dictates a linearly declining acceleration. As expected,
and observed in the recorded shadowgraphs, after the passage of the incident
shock wave, the sphere starts moving; it increases its velocity from zero towards
the post-shock flow velocity. The sphere’s velocity increases monotonically as is
evident from its recorded trajectory. Hence, it is an acceptable and reasonable
practice to select the lowest order that provides the requested fit to the available
data. In the present case, a third-order polynomial fit provided a very good fit to
the recorded sphere locations.
A summary of all the results obtained in the present research is shown in figure 14
in the CDKRe plane. Three lines appear in this figure. The solid line is a curve fit to
the present results. The dotted line is the correlations proposed for CD by Igra &
Takayama (1993) and the dashed line is the standard drag curve proposed for a
similar steady flow. The correlation describing the present results, which appears as
a solid line in figure 14, is a third-order polynomial fit, similar to that proposed by
Igra & Takayama (1993). This correlation is given in the following equation:
log10 CD ZK0:696C1:259!log10 Rep K0:465!ðlog10 Rep Þ2 C0:045!ðlog10 Rep Þ3 :
ð4:1Þ
In this correlation, Rep stands for the Reynolds number based on the particle
relative velocity.

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3343

3
subsonic Mp < 1
run no. 148
run no. 149
run no. 163
run no. 165
run no. 166
2 run no. 173
run no. 182
run no. 186
run no. 187
CD supersonic Mp > 1
run no. 142
run no. 145
run no. 179
1 run no. 180
run no. 181
run no. 183
run no. 184
standard curve
Igra and Takayama correlation
present work

0
10 2 10 3 10 4 10 5
Rep

Figure 15. CD versus Re based on the average values for CD.

A clearer picture of the obtained results is achieved when in each run an


average value of CD is assigned to the appropriate average value of Rep. The
obtained results are shown in figure 15. Also shown in this figure are the above-
mentioned correlations proposed for the sphere’s CD (ours and that of Igra &
Takayama (1993)) and the standard drag curve proposed in the literature for a
sphere in a steady flow. It is clear from this figure that there is a significant
difference between the drag coefficient of a sphere in a steady flow (the standard
drag curve) and that obtained in non-stationary flow conditions. The proposed
correlations for the non-stationary drag coefficient suggest values of CD which are
significantly higher than that obtained in a similar steady flow case for Reynolds
numbers within the range of 500!Re!104. Inspection of figure 15 shows that the
results obtained for CD in this range for subsonic and supersonic flows can be
found above and below the correlations given by equation (4.1), indicating that,
in the considered range of Mp, the Mach number of the relative velocity probably
does not play a dominant role in the resulting sphere drag coefficient. It may be
the case that the relative low number of points in this range does not allow us to
conclude definitively on the compressibility effects.
It was suggested by Crowe et al. (1963), Selberg & Nicholls (1968), Karanfilian &
Kotas (1978) and Temkin & Metha (1982) that the flow unsteadiness contribution
to the sphere’s drag may be expressed in terms of a non-dimensional parameter
AcZ ððdup =dtÞfÞ=ðug Kup Þ2 . This parameter can be deduced from equation (3.2).
In the present experiments, this parameter is quite small, it is within the range of
2.5!10K5%Ac%3.5!10K2 as is evident from table 2. The fact that Ac is so small
should not be surprising since, in spite of the sphere’s high initial acceleration, its
diameter is very small and its initial relative velocity is relatively high. Therefore,
one could conclude that the parameter Ac is not a characteristic parameter for the

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

3344 G. Jourdan et al.

considered flows. It should be noted that, in Karanfilian & Kotas (1978), the
investigated spheres were immersed in liquid and in such a case the added mass
contribution to the unsteady drag coefficient is significant (water density is almost
three orders of magnitude higher than that of the air). Therefore, it is not surprising
that, in their results, Ac varies within the range of 0.1%Ac%0.5; this is significantly
higher than the Ac values obtained in the present investigation. We therefore
believe that, in the considered experiments, the dominant mechanism is the flow
establishment around the investigated sphere; similar to the flow establishment
around an aerofoil.

5. Conclusions

The present results obtained for the sphere drag coefficient, CD, strengthen past
assessments that a significant difference exists between the values obtained for
CD in steady and non-steady flows. The correlations proposed in equation (4.1)
can be used safely in numerical simulations of dusty flows where the dust
concentration is not high, and for Reynolds numbers within the range covered in
the present investigation. Furthermore, throughout most of the investigated
range of Reynolds numbers, the obtained non-steady values are over 50% higher
than those obtained in a similar steady flow case. The gap between the two
increases with decreasing Reynolds numbers. Since the drag force in this range is
dominated by viscous flow effects, the dynamic drag coefficient obtained in the
present study seems to indicate a type of unsteady ‘shear waves’ flow effect. But,
at this time, we do not have an appropriate explanation for the significant
increase in the sphere drag coefficient in non-steady flows and particularly at the
lower range of Reynolds numbers, and we think that this point will be a challenge
to researchers in theoretical or computational fluid mechanics.

References
Bailey, A. B. & Hiatt, J. 1972 Sphere drag coefficient for a broad range of Mach and Reynolds
numbers. AIAA J. 10, 1436–1440.
Bailey, A. B. & Starr, R. F. 1976 Sphere drag at transonic speeds and high Reynolds numbers.
AIAA J. 14, 1631–1635.
Boiko, V. M., Kiselev, V. P., Kiselev, S. P., Papyrin, A. N., Popalavsky, S. V. & Fomin, V. M.
1997 Shock wave interaction with a cloud of particles. Shock Waves J. 7, 275–285. (doi:10.1007/
s001930050082)
Buckley Jr, F. T. 1968 An experimental investigation of the drag coefficient of particles
accelerating within particulate clouds at transonic slip flow conditions by a light extinction
technique. DPhil thesis, University of Maryland, USA.
Crowe, C. T., Nicholls, J. A. & Morison, R. B. 1963 Drag coefficient of inert and burning particles
accelerating in gas streams. In Proc. 9th Symp. on Combustion, New York, USA, pp. 395–406.
New York, NY: Academic Press.
Henderson, C. B. 1976 Drag coefficient of spheres in continuum and rarefied flows. AIAA J. 14,
707–708.
Igra, O. & Takayama, K. 1993 Shock tube study of the drag coefficient of a sphere in a non-
stationary flow. Proc. R. Soc. A 442, 231–247. (doi:10.1098/rspa.1993.0102)
Karanfilian, S. K. & Kotas, T. J. 1978 Drag on a sphere in unsteady motion in a liquid at rest.
J. Fluid Mech. 87, 85–96. (doi:10.1017/S0022112078002943)
Olson, R. M. 1980 Essentials of engineering fluid mechanics, 4th edn. New York, NY: Harper & Row.

Proc. R. Soc. A (2007)


Downloaded from rspa.royalsocietypublishing.org on March 12, 2014

Drag coefficient of a sphere 3345

Outa E., Tajima K. & Suzuki S. 1981 Cross sectional concentration of particles during shock
process propagating through a gas–particles mixture in a shock tube. In Proc. 13th Int. Symp.
on Shock Waves, Niagara Falls, NY, USA, pp. 655–663. New York, NY: State University of New
York Press.
Rodriguez, G., Grandeboeuf, P., Kaelifi, M. & Haas, J. F. 1993 Drag coefficient measurement of
spheres in a vertical shock tube and numerical simulations. In Proc. 19th Int. Symp. on Shock
Waves, Marseille, France, vol. III, pp. 43–48. Berlin, Germany: Springer.
Rudinger, G. 1970 Effective drag coefficient for gas-particle flow in shock tubes. ASME J. Basic
Eng. D 92, 165–172.
Selberg, B. P. & Nicholls, J. A. 1968 Drag coefficient of small spherical particles. AIAA J. 6,
401–407.
Sommerfeld, M. 1985 The unsteadiness of shock wave propagating through gas particle mixtures.
Exp. Fluid 3, 197–206.
Sun, M., Saito, T., Takayama, K. & Tanno, H. 2004 Unsteady drag on a sphere by shock wave
loading. Shock Waves J. 14, 3–9. (doi:10.1007/s00193-004-0235-4)
Suzuki, T., Sakamura, Y., Adachi, T. & Igra, O. 1999 Shock tube study of particles’ motion behind
a planar shock wave. In Proc. 22nd Int. Symp. on Shock Waves, London, UK, pp. 1411–1416.
Southampton, UK: Southampton University Media Press.
Tano, H., Itho, K., Saito, T., Abe, A. & Takayama, K. 2003 Interaction of shock with a sphere suspended
in a vertical shock tube. Shock Waves J. 13, 191–200. (doi:10.1007/s00193-003-0209-y)
Temkin, S. & Metha, H. K. 1982 Droplet drag in an accelerating and decelerating flow. J. Fluid
Mech. 116, 297–313. (doi:10.1017/S0022112082000470)

Proc. R. Soc. A (2007)

View publication stats

You might also like