Sie sind auf Seite 1von 14

Proceedings of the Institution of Mechanical

Engineers, Part P: Journal of Sports


Engineering and Technology
http://pip.sagepub.com/

Fluid dynamics of cricket ball swing


James A Scobie, Simon G Pickering, Darryl P Almond and Gary D Lock
Proceedings of the Institution of Mechanical Engineers, Part P: Journal of Sports Engineering and Technology published
online 30 October 2012
DOI: 10.1177/1754337112462320

The online version of this article can be found at:


http://pip.sagepub.com/content/early/2012/10/30/1754337112462320
A more recent version of this article was published on - Aug 19, 2013

Published by:

http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for Proceedings of the Institution of Mechanical Engineers, Part P: Journal of Sports Engineering and
Technology can be found at:

Email Alerts: http://pip.sagepub.com/cgi/alerts

Subscriptions: http://pip.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Version of Record - Aug 19, 2013

>> OnlineFirst Version of Record - Oct 30, 2012

What is This?

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Original Article

Proc IMechE Part P:


J Sports Engineering and Technology
0(0) 1–13
Fluid dynamics of cricket ball swing Ó IMechE 2012
Reprints and permissions:
sagepub.co.uk/journalsPermissions.nav
DOI: 10.1177/1754337112462320
pip.sagepub.com

James A Scobie, Simon G Pickering, Darryl P Almond and


Gary D Lock

Abstract
Swing describes the lateral deviation of a cricket ball in its trajectory towards the batsman. Conventional swing is effective
with a new, or well-preserved, ball, and the fluid dynamics governing this phenomenon was first explained in 1957. In
2012, many test-match fast bowlers are able to swing, at high speed, an older ball in the reverse direction. This reverse
swing of a ball aged under match conditions has never been explained fully. A cricket ball is asymmetric with six seams of
80–90 encircling stitches, protruding approximately 1 mm proud of the surface. Both conventional and reverse swings
are a consequence of asymmetrical flow separation leading to a skewed wake and a net pressure force on the ball per-
pendicular to the flight trajectory. Here, experimental evidence is presented for the first time showing that the forma-
tion of a laminar separation bubble is the prominent flow feature creating the flow asymmetry for reverse swing. A new
flow visualisation technique to capture the fluid dynamics of boundary-layer separation using an infrared camera is also
introduced here.

Keywords
Cricket ball, reverse swing, boundary layer, flow visualisation, fluid dynamics

Date received: 25 April 2012; accepted: 28 August 2012

Introduction Conventional and reverse swing


In the opening overs of a cricket match, much of the In cricket, swing is a consequence of the asymmetry of
radio and television commentary will discuss the swing the ball. In other sports (e.g. football, tennis and golf),
of the new, shiny ball. As the match progresses and the a virtually symmetric sphere under the influence of the
ball ages, this conventional form of swing will tend to Magnus effect can be made to swerve in flight due to
reduce until it no longer poses a threat to the batsmen. significant spin. The asymmetry of a cricket ball is
However, after approximately 25–30 overs, a skilful inherently created by the six seams of 80–90 encircling
bowler will find that an older, worn ball can be made to stitches protruding about 1 mm proud of the surface.
swing significantly, at high speed, in the reverse direc- This seam is exploited by the bowler to create a net
tion. This reverse swing (RS) occurs only under condi- transverse pressure force resulting from the asymmetric
tions where the surface of the ball has been carefully separation of the fluid dynamic boundary layer5 on the
manipulated by the fielding side. two sides of the ball. In the boundary layer, the viscous
The fluid dynamics of conventional swing (CS) was fluid (i.e. air) sticks to the ball; in this thin region near
first explained in 1957.1 This well-understood phenom- the ball, the relative air velocity increases from zero at
enon applies to the new, or well-preserved, ball and has the surface to the undisturbed freestream velocity. For
been demonstrated in experiments (see Barton,2 Bentley typical bowling speeds in a cricket match, this bound-
et al.3 and Mehta et al.4). RS was introduced controver- ary-layer thickness is approximately a millimeter.3
sially to game of cricket by Pakistan’s quicker bowlers Figure 1(a) illustrates the fluid dynamics of CS in a
during the 1990s. In 2012, many test-match fast bowlers polar view, that is, one looking in the direction of
are able to swing, at high speed, an older ball in the
reverse direction. In contrast to CS, the reverse form
has never been fully understood by cricketers or fluid Department of Mechanical Engineering, University of Bath, Bath, UK
dynamicists. This article will demonstrate, with refer-
Corresponding author:
ence to the fluid dynamic boundary layer and by means Gary D Lock, Department of Mechanical Engineering, University of Bath,
of novel flow visualisation experiments, how RS arises Bath BA2 7AY, UK.
and why it occurs at high speed. Email: ensgdl@bath.ac.uk

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


2 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

Figure 1. Current theories for (a) conventional and (b) reverse swing (boundary layer greatly exaggerated for clarity).

gravity. The seam is inclined at a small angle (typically on the seam side, the turbulent boundary layer remains
15°) to the airflow. By imparting a small backspin, the attached up to a greater angle (’120°) compared to the
bowler can create enough gyroscopic inertia to stabilise laminar side (’80°) where no seam is present to trip
this seam position during the trajectory towards the the boundary layer. The asymmetrical separation leads
batsman. As the flow accelerates around the ball, the to an angled wake and a net pressure force on the ball
thin boundary layer (greatly exaggerated in thickness perpendicular to the flight trajectory. At the start of a
for clarity) separates from the surface. The separation match, the new ball has a shiny surface, which
angles are asymmetric as the encircling seam trips this encourages the laminar flow on the side opposite of the
boundary layer into turbulence, which would otherwise seam. During the course of play, the fielding side will
remain laminar, on one side of the ball. The turbulent vigorously polish the ball to help maintain this smooth
boundary layer, by virtue of its increased momentum, condition. The ball will wear and roughen as the match
is more robust than the laminar layer. Consequently, progresses, causing natural transition to turbulent flow

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 3

on side of the ball without the seam. Match experience sphere. As shown in Figure 2(a) and (b), an amount of
suggests that this typically occurs after 10–15 overs, roughness (k/d ’ 0.0025) was shown to significantly
when asymmetry and CS are lost. reduce the critical Reynolds number. The roughness ele-
Fluid dynamicists (see Brown and Mehta,6 Barrett ments disturb the boundary layer causing transition to
RS and Wood,7 Sayers8 and Mehta,9 including the turbulence at lower Reynolds number. The separation
Lock et al.10) have previously described RS as illu- angle is a maximum at the critical Reynolds number and
strated in Figure 1(b). RS is encountered when the ball then significantly decreases in the transcritical region.
is bowled at speeds greater than 35.8 m s21 (80 mile/h) The prominent fluid dynamic feature at the critical
with turbulent flow present on both sides of the ball. Reynolds number is a separation bubble,14 which
This explanation (proved in this article to be incom- Achenbach’s experiments (at 5° resolution) were not
plete) suggests that the seam thickens the turbulent sensitive enough to determine. Separation bubbles
boundary layer on one hemisphere causing the separa- occur when the laminar boundary layer, under the
tion angle to move upstream relative to the side with influence of the adverse pressure gradient, separates
the undisturbed boundary layer.6 The reverse in pres- from the surface and then undergoes a transition to
sure asymmetry will create a net side force in the oppo- turbulence with characteristic rapid thickening. This
site direction to that for CS. thickening can cause the now turbulent shear layer to
Here, a novel method of flow visualisation (using an come back into contact with the surface and reattach
infrared camera) has provided new insight into the as a turbulent boundary layer. A bubble of circulatory
structure of the separating flow around the cricket ball fluid is consequently trapped under the separated shear
under conditions of RS. These flow visualisation experi- layer between the separation and reattachment points,
ments were linked to simultaneous measurements of as shown in Figure 3.
pressure distribution and side force. Large-scale cricket Figure 4(a) to (c) illustrates the flow patterns over a
balls as well as those aged in first-class match condi- sphere in the subcritical, critical and transcritical
tions were examined. Experimental evidence is pre- regions. In Figure 4(b), uL, uR and uT indicate the lami-
sented for the first time that reveals that the formation nar separation, turbulent reattachment and turbulent
of a separation bubble with late turbulent reattachment separation points, respectively, for the separation bub-
is the primary fluid dynamic feature associated with ble over a sphere at the critical Reynolds number. The
RS. It should be stated that this explanation of RS was second point of separation, which takes place in a tur-
first postulated by Grant et al.,11 though no clear evi- bulent layer, is located significantly suddenly down-
dence was produced. stream due to the better resistance to the adverse
pressure gradient by the turbulent layer, resulting in a
narrow wake and a significant reduction in drag.
Boundary-layer separation around spheres Taneda14 measured 92° \ uL, \ 110°, 107° \ uR \ 127°
and 123° \ uT \ 147° (depending on Re) for smooth
Achenbach12 investigated experimentally the fluid
spheres. In the transcritical region, the width of the
dynamic boundary layer governing the drag coefficient
wake increases as the turbulent separation points move
of a hydraulically smooth metal sphere. These experi-
upstream towards uT ’ 120° round the rear surface
mental results are reproduced as curves in Figure 2.
with a consequential increase in drag.
Figure 2(a) shows the drag coefficient (Cd) of the
Achenbach13 measured the critical Reynolds number
smooth sphere as a function of the Reynolds number
on the smooth sphere to be Re ’ 150–175 3 103. This
(Re), obtained by direct measurement of the drag force
value will be less for the double-scaled cricket ball (as
by strain gauges; Figure 2(b) shows the position of
well as the balls aged under match conditions) due to the
boundary-layer separation obtained from skin-friction
seam; even on the non-seam side, the inherent surface
measurements. For Re \ 2 3 105 (subcritical flow),
roughness will reduce this critical Reynolds number.
Achenbach showed that the boundary layer is laminar
and separates at uL ’ 80°. At increased Re, the bound-
ary layer encounters a critical condition where the drag
Experiments
coefficient reduces dramatically at Re ’ 4 3 105; here
the attached boundary layer is now turbulent and the Two wind tunnels were used in the experiments
separation points move downstream to uT ’ 120°. As reported here. The University of Bath wind tunnel,
the Reynolds number is increased further, the boundary with a maximum speed of 20 m s21, featured a circular
layer enters the supercritical and ultimately a transcriti- open-jet test section 0.77 m in diameter and working
cal region. The attached boundary layer is predomi- length of 1.5 m. Additional experiments were con-
nantly turbulent, and the angle of transition translates ducted in the University of Bristol’s larger wind tunnel
upstream; the turbulent flow thickens the boundary (octagonal test section of 2.1 m 3 1.5 m), which was
layer, causing a gradual reduction in the angle of capable of reaching speeds of up to 60 m s21. The free-
separation (uT \ 120°). stream turbulence intensity in both tunnels was less
Achenbach13 also investigated the effect of surface than 1%. A range of cricket balls (71.3 mm \ d \ 71.9
roughness on the fluid dynamics of flow around a mm, measured not including the seam), used for 8–80

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


4 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

Figure 2. (a) Variation of the drag coefficient with the Reynolds number for a smooth and slightly roughened sphere and (b)
separation angles (with respect to the stagnation point) for a smooth and slightly roughened sphere.
Adapted from Achenbach.12,13

overs under first-class match conditions, was supplied rapid prototype nylon. This model was complete with
by Warwickshire County Cricket Club. To assess the an accurate, scaled representation of the seam.
static pressure distribution at 44 locations around the The experimental set-up is illustrated in Figure 5. A
circumference, a hollow, scaled cricket ball with twice strain gauge was used to measure the side force S per-
the diameter (i.e. 142 mm) was manufactured from pendicular to the flow. The gauge was located on a

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 5

Figure 3. Laminar separation bubble.


Adapted from Houghton and Carpenter.15

supporting cantilever beam, which acted as an aerody-


namic sting. The side force was non-dimensionalised
by the projected area of the ball and the dynamic
head of the fluid (see CS in Appendix 1). In this form,
the actual force is proportional to the square of the
air speed. The aerodynamic sting was positioned
within the wake with the vertical beam 250 mm
behind the ball in order to ensure minimal influence
on the flow. The dimensions of the cantilever (13 mm
3 6.5 mm) were designed so that the beam was stiff
enough to avoid resonance and severe vibration but
able to provide sensitivity to very low forces.
Vibrational effects were taken into account by sam-
pling measurements at 10 Hz over a time period of
1 min and calculating a mean.
A new, qualitative, flow visualisation technique to
capture the fluid dynamics of boundary-layer separa-
tion using an infrared camera is introduced here. As
illustrated in Figure 6, heated air was injected upstream
at the rear of the ball in a similar manner to which
smoke is used in flow visualisation experiments. The
heated air is entrained into the separated boundary
layer near the surface of the ball. In the steady state,
the heat propagates upstream until it is met by the cold
mainstream air near the separation point. An infrared
camera was used to capture thermal images of the sur-
face of the ball. The flow visualisation technique used
here is qualitative, rather than quantitative, as there is a
competing mechanism between upstream heat conduc-
tion along the surface and convective cooling. Ideally,
the surface of the ball would be perfectly insulating; the
double-sized ball used here is made of nylon, which is a
good insulator so the separation points can only be
determined approximately.
The thermal imaging camera was a Cedip Jade III
mid-wave infrared that used a cooled indium antimo-
Figure 4. Flow patterns over a sphere in the (a) subcritical nide focal plane array with a resolution of 320 3 256
(400 \ Re \ 3 3 105), (b) critical (3 3 105 \ Re \ 3 3 106)
pixels and a noise-equivalent temperature difference of
and (c) transcritical regions (Re . 3 3 106).
Adapted from Houghton and Carpenter.15
approximately 21 mK.

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


6 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

Figure 5. Experimental set-up.

Figure 6. Flow visualisation technique implemented during the study (see online version for colour illustration).

Experimental results ball. At low Re (flight velocity \ 20 mile/h), laminar


separation is expected on both sides of the ball, with a
Side force measurements symmetric pressure distribution and zero side force. As
Figure 7 shows the non-dimensional side force mea- Re increases, the seam trips the boundary layer to tur-
sured using both the double-sized and actual cricket bulent flow (critical Re ’ 70 3 103), while the bound-
balls aged under match conditions plotted against ary layer on the non-seam side remains laminar. Thus,
Reynolds number. Also, the equivalent bowling speed between 8.9 and 33.5 m s21 (20 and 75 mile/h), the ball
based on the actual ball diameter in miles per hour is experiences CS, with a positive side force of non-
shown, which allows the data to be presented in the dimensional magnitude approximately 0.3 \ CS \ 0.4.
context of the game and in terms the cricket audience The critical Reynolds number for the non-seam side is
would be familiar. For consistency, all the experiments Re ’ 170 3 103 (’75 mile/h); here, both sides of the
were performed with the seam angle set at 15° to the ball experience turbulent flow and the CS is lost. RS is
air stream. measured for Re . 170 3 103. Figure 7 shows the non-
The measurements, over the range of bowling dimensional side force; the RS is significant because the
speeds, demonstrate the behaviour expected from expe- lateral force is proportional to the square of the
rience. Consider first the data from the double-sized velocity.

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 7

Figure 7. Variation of non-dimensional side force with speed for cricket balls at different stages of wear.

Consider now the data obtained from the cricket data for three Reynolds numbers are Re = 20 3 103,
balls used in first-class matches. The results are again 150 3 103 and 200 3 103, which correspond to full-sized
consistent with expectations. The balls begin to experi- cricket ball velocities of ’4, 30 and 40 m s21, respec-
ence CS at approximately the same bowling speed. The tively, that is, ’9, 67 and 89 mile/h. These three speeds
magnitude of the side force and speed at which each fall in the regime for no-swing (NS), CS and RS, respec-
ball begins to swing in the reverse direction are then tively. Although the 9 mile/h case has no context in
directly related to the age of the ball and hence surface terms of the game, it was performed in order to demon-
roughness. Considering a speed of 35.8 m s21 (80 mile/ strate the NS condition. Also, the pressure distribution
h), the approximate bowling speed of a fast- to expected from an inviscid fluid where there is no flow
medium-paced bowler, the ball aged between 8 and 10 separation is shown.16
overs swings the most in the conventional manner. In Figure 8, the constant-pressure region on both
Once the ball has aged sufficiently, that is, approxi- sides of the ball at angles of u . 120° is an indication
mately after 25–40 overs, CS cannot be obtained at this of separated flow at the downstream rear. The angle of
speed. However, as the ball wears further, negative side boundary-layer separation is indicated by the end of
force measurements are produced, indicating RS. the rise in pressure (i.e. the adverse pressure gradient)
Significant RS can only be achieved at bowling speeds to this constant level. At low Re (9 mile/h), laminar
greater than 38–40.2 m s21 (85–90 mile/h) with a used separation occurs before transition to turbulent flow,
or worn ball. Based on the speed at which RS first even on the seam side. Here, the pressure distribution is
appears, the results suggest that the roughness of the symmetric with separation angles uL ’ 80°, and the
double-sized model was equivalent to an actual ball side force (see Figure 7) is measured to be zero with
aged between 25 and 50 overs. NS. As Re increases, the seam trips the boundary layer
The data confirm that it is the medium- to fast-paced to turbulent flow (critical Re ’ 70 3 103) while the
bowlers who generate CS using a new or well-preserved boundary layer on the non-seam side remains laminar.
polished ball. It is only the quicker bowlers who achieve At Re = 150 3 103 (67 mile/h), there is relatively lower
RS. A ball that is scuffed, worn or conditioned (legally pressure on the hemisphere with the seam, which cre-
or illegally) during the course of play will exacerbate ates a net side force, or the CS illustrated in Figure
the effect of RS. 1(a). The separation angle on the non-seam side of the
ball is still uL ’ 80°, but this angle has been delayed to
Pressure measurements uT ’ 120° on the seam side. The critical Reynolds num-
Figure 8 illustrates the variation of pressure coefficient, ber for the non-seam side is Re ’ 170 3 103 (’75 mile/
Cp, with angle u from the stagnation point for the h), where there is natural turbulent flow on both sides
double-sized cricket ball; the seam was positioned at of the ball and the CS is lost. These angles follow the
15° to the mainstream. Note that there is a region on trends measured by Achenbach,12,13 as shown in
the seam side where no data could be recorded (because Figure 2.
of the seam itself) and the data do not show Cp = 1, as At Re = 200 3 103 (89 mile/h), turbulent separation
the stagnation point lies on the seam of the model. The still occurs on the seam side of the ball, although in this

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


8 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

transcritical flow, the separation point has moved (the angles of laminar separation, reattachment and
upstream to uT ’ 110°. This again accords with the turbulent separation, respectively) are marked in
observations of Achenbach. On the non-seam side, Figure 8. These angles agree well with those measured
however, a dramatic change has taken place: the pres- by Taneda14 for a separation bubble on a smooth
sure recovers to the base level at uT ’ 135°. (Note that sphere.
it was not possible to measure pressure beyond this
angle). Thus, the angles of separation on the seam and
non-seam side are 110° and 135°, respectively. There
Flow visualisation
has been an inversion of the pressure asymmetry, with Thermal flow visualisation images were acquired using
relatively lower pressure on the hemisphere without the the double-sized cricket ball model using the three
seam; this again creates a net side force, but now this is orientations, as shown in Figure 9. The temperature
the RS. An explanation for the delay in separation on distributions are presented in Figure 10 for the NS, CS
the non-seam side is indicated by the region of rela- and RS conditions. These conditions correspond to
tively constant pressure between approximately equivalent speeds of 9, 67 and 89 mile/h, respectively,
95° \ u \ 110°, followed by a sharp downstream pres- consistent with the speeds presented in the pressure dis-
sure recovery. This is evidence for the formation of a tribution in Figure 8.
laminar separation bubble, as discussed earlier and is In Figure 10, the mainstream air is cold and heated air
consistent with the findings of Fage.16 With reference has been injected upstream. As discussed with reference
to Figures 3 and 4, uL ’ 95°, uR ’ 110° and uT ’ 135° to Figure 6, the flow visualisation provides qualitative

Figure 8. Variation of pressure coefficient with the angle from the stagnation point.
NS: no-swing; CS: conventional swing; RS: reverse swing.

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 9

Figure 9. Orientation of flow visualisation experiments; air is moving into the page.

Figure 10. Flow visualisation for NS, CS and RS conditions; airflow is from right to left with heated air from the opposite direction;
red indicates hot and blue cold (see online version for colour illustration).
NS: no-swing; CS: conventional swing; RS: reverse swing.

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


10 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

Figure 11. Flow visualisation for polar view with superimposed boundary layer: (a) NS, (b) CS and (c) RS conditions.
NS: no-swing; CS: conventional swing; RS: reverse swing.

information, and boundary between the hot (red) and Conversely, blue indicates that the cool air from the
cold (blue) flows gives an indication of the angle of mainstream is attached to the surface. The polar view in
boundary-layer separation (see online version for colour Figure 10 (i.e. Figure 10(d) to (f)) corresponds to the view
illustration). Red in the thermal images indicates the presented in Figure 1, and the expected fluid dynamic
regions where the flow is not attached, that is, the bound- boundary layer (again exaggerated for clarity) is superim-
ary layer has separated from the surface of the ball. posed on this polar view in Figure 11.

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 11

Figure 12. Flow visualisation repeated using an actual cricket ball aged 25 overs.
NS: no-swing; CS: conventional swing; RS: reverse swing.

At the NS condition (9 mile/h), symmetrical laminar trailing edge of the seam. The separation bubble has
separation is indicated on both sides of the ball. At a trapped a pocket of hot air in the manner indicated in
larger Re (67 mile/h), the seam trips the boundary layer Figure 4. The break in the centre of the bubble is attrib-
to turbulent flow, while the boundary layer on the non- uted to the increased roughness caused by the pressure
seam side remains laminar. The separation of the turbu- tapping holes arranged around the circumference; this
lent boundary layer on the seam side is delayed result- disturbance locally trips the flow into turbulence. It
ing in the asymmetry causing CS. should be noted that the separation bubble is only visible
At 89 mile/h, the ball experiences RS. On the seam to the infrared camera when the downstream heated
side of the ball, the angle of separation moves marginally source is activated before the cold mainstream flow is ini-
upstream relative to the CS condition. On the non-seam tiated and the surface is pre-heated; if the heated flow is
side, the laminar separation bubble indicated in the pres- activated after the mainstream flow has been activated,
sure measurements is clearly visible (especially Figure the bubble is not detected by the camera as the heat is
10(i)). The isolated region of heated surface indicates unable to propagate forward to the trapped region.
that laminar separation, followed by turbulent reattach- With reference to Figure 7, it was noted that the
ment, has occurred. The second point of (turbulent) roughness of the double-sized model was equivalent to
separation is not clear but appears to be located near the an actual ball aged between 25 and 50 overs. To ensure

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


12 Proc IMechE Part P: J Sports Engineering and Technology 0(0)

that the fluid dynamics observed on the scaled ball Engineering/Exxon-Mobil. A series of cricket balls,
takes place with realistic surface roughness created worn under known first-class match conditions, were
under match conditions, the flow visualisation experi- supplied by Dougie Brown of Warwickshire County
ments were repeated using an actual cricket ball aged Cricket Club. The authors also thank Dr Raf
25 overs in a first-class match. The flow visualisation is Theunissen of the University of Bristol.
shown in Figure 12. The faster wind tunnel at the
University of Bristol was used for these experiments References
and, relative to the images collected in Figure 10, the
1. Lyttelton RA. The swing of a cricket ball. Discovery
infrared camera has less resolution due to the larger
1957; 18: 186–191.
distance to the ball and the smaller diameter of the ball 2. Barton NG. On the swing of a cricket ball in flight. P
itself. Although the results in Figure 12 are less clear Roy Soc Lond A Mat 1982; 379: 109–131.
than the double-sized model, they indicate a consistent 3. Bentley K, Varty P, Proudlove M, et al. An experimental
pattern in terms of flow visualisation and hence fluid study of cricket ball swing. Imperial College Aero. Tech-
dynamics. At the highest speed on the non-seam side, nical note 82-106, 1982.
the laminar separation bubble is observed. 4. Mehta RD, Bentley K, Proudlove M, et al. Factors
As a final point, it should be noted that during the affecting cricket ball swing. Nature 1983; 303: 787–788.
course of play, a swing bowler, as discussed earlier, usu- 5. Prandtl L. Über Flüssigkeitsbewegung bei sehr kleiner
ally imparts a small degree of backspin on the ball. The reibung [On the motion of fluids of very small viscosity].
In: Verhandlungen des III internationalen Mathematiker-
influence of this rotation on the formation and stability
Kongresses (ed A Krazer), Heidelberg, Germany, 8–13
of the separation bubble was not tested in the experi- August 1904, pp.484–491. Leipzig, Teubner.
ments reported here. 6. Brown W and Mehta RD. The seamy side of swing bowl-
ing. New Sci 1993; 139: 21–24.
7. Barrett RS and Wood DH. The theory and practice of
Conclusion reverse swing. Sports Coach 1996; 18: 28–30.
This article provides new insight into boundary-layer 8. Sayers AT. On the reverse swing of a cricket ball – mod-
elling and measurements. Proc IMechE, Part C: J
separation for the flow around a cricket ball, clearly
Mechanical Engineering Science 2001; 215: 45–55.
demonstrating the fluid dynamics of both conventional 9. Mehta RD. An overview of cricket ball swing. Sports
and reverse swing. A novel method of flow visualisation Eng 2005; 8: 181–192.
(using an infrared camera) was linked to simultaneous 10. Lock GD, Edwards S and Almond DP. Flow visualiza-
measurements of pressure distribution and side force on tion experiments demonstrating the reverse swing of a
large-scale cricket balls as well as those aged in first- cricket ball. Proc IMechE, Part P: J Sports Engineering
class match conditions. A cricket ball is asymmetric due and Technology 2010; 224: 191–199.
to the six seams of 80–90 encircling stitches protruding 11. Grant C, Anderson A and Anderson JA. Cricket ball
about 1 mm proud of the surface. This seam is exploited swing-the Cooke-Lyttleton theory revisited (Proceedings
of the second international conference on the engineering of
by the bowler to create a net transverse pressure force
sport, Sheffield, UK, July 1998). In: Haake S (ed.) The
resulting from the asymmetric separation of the fluid
engineering of sport. Oxford: Blackwell Science, pp.371–
dynamic boundary layer on the two sides of the ball. 378.
CS is experienced for Reynolds numbers in the range of 12. Achenbach E. Experiments on the flow past spheres at
70–170 3 103 (’35–75 mile/h). On the seam side of the high Reynolds numbers. J Fluid Mech 1972; 54(3): 565–
ball, the boundary layer is tripped into turbulent flow 575.
and remains attached to an angle of ’120°; the flow on 13. Achenbach E. The effects of surface roughness and tun-
the other side of the ball separates at an angle of ’80° nel blockage on the flow past spheres. J Fluid Mech 1974;
as there is no seam to trip the laminar boundary layer. 65(1): 113–125.
RS is a consequence of the asymmetry inverting at 14. Taneda S. Visual observations of the flow past a sphere
at Reynolds numbers between 104 and 106. J Fluid Mech
Reynolds numbers . 170 3 103. Experimental evi-
1977; 85(1): 187–192.
dence is presented for the first time that reveals that the 15. Houghton EL and Carpenter PW. Aerodynamics for engi-
formation of a separation bubble with late turbulent neering students. 4th ed. London, UK: Edward Arnold,
reattachment is the primary fluid dynamic feature asso- 1993.
ciated with RS. 16. Fage A. Experiments on a sphere at critical Reynolds num-
bers. Aeronautical Research Committee reports and
memoranda, no. 1766, 1936.
Funding
This research received no specific grant from any funding
agency in the public, commercial, or not-for-profit sectors. Appendix 1
Notation
Acknowledgements d diameter
Professor GD Lock has been awarded a prize for k size of roughness element
Excellence in Teaching by the Royal Academy of pN static pressure in the freestream

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014


Scobie et al. 13

pu static pressure at angle u m dynamic viscosity


vN freestream velocity fs separation angle
Cp pressure coefficient, r density
Cp = (pu p‘ )=(1=2(rv2‘ )) u angle from the stagnation point
Cs non-dimensional side force, uL laminar separation angle
Cs = 8S=rv2‘ pd2 uR turbulent reattachment angle
Re Reynolds number, Re = rv‘ d=m uT turbulent separation angle
S side force

Downloaded from pip.sagepub.com at INDIAN INST SCI on August 24, 2014

Das könnte Ihnen auch gefallen