You are on page 1of 9

191

Flow visualization experiments demonstrating


the reverse swing of a cricket ballg
G D Lock* , S Edwards, and D P Almond
Department of Mechanical Engineering, University of Bath, Bath, UK

The manuscript was received on 2 March 2010 and was accepted after revision for publication on 26 May 2010.

DOI: 10.1243/17543371JSET73

Abstract: Swing, the sideways deviation of the cricket ball in flight, is a phenomenon which can
be explained in terms of fluid dynamics. Conventional swing has been used since the beginning
of the twentieth century and is effective with a new or well-preserved ball. A controversial form
of swing emerged in Pakistan in the 1980s, featuring an older worn ball which, at a high speed,
swung in the reverse direction. A cricket ball is asymmetric because of the presence of a seam,
which is made up of rows of prominent encircling stitches. For conventional swing, this seam
trips into turbulence the boundary layer adjacent to one hemisphere of the ball which remains
attached to a greater angle (about 120◦ ) than does that on the other side (about 80◦ ) where no
seam is present to trip the laminar boundary layer. The result is asymmetrical separation, leading
to a skewed wake and a net pressure force on the ball perpendicular to the flight trajectory.
Reverse swing is thought to be a consequence of the fact that asymmetry inverts at a high
speed if the seam thickens the turbulent boundary layer on one side of the ball. Although
the fluid dynamics causes of both conventional and reverse swing are well known, this paper
demonstrates clearly, by means of flow visualization and pressure measurements, the inversion
of this pressure asymmetry at Reynolds numbers greater than 170 × 103 .

Keywords: cricket ball, reverse swing, boundary layer

1 INTRODUCTION a vital blow for his side.’ Australia batted Bradman


at number three to protect him from the shiny new
At the Oval in 1948 Don Bradman (Fig. 1) required ball in these treacherous opening overs. Conven-
just four runs in his final innings to retire with a Test tional swing bowling and why it is that the new or
batting average of 100, but was bowled, second ball, well-preserved ball is most likely to swing was first
for nought. His final average of 99.94, with its tinc- explained in 1957, and the fluid dynamics causes
ture of human fallibility, could not have been more are well understood (see, for example, references [3]
exquisitely contrived [1]. The fatal ball, from Eric to [6]). Reverse swing, which emerged controver-
Hollies, was a googly, one which spins in the reverse sially in Pakistan in the 1980s, features an older worn
direction to that expected. The modern googly has ball which, at a high speed, swings in the reverse
been reverse swing, a phenomenon which Bradman direction. It was a decided advantage to Pakistan’s
never had the opportunity to confront. He would, quicker bowlers in the late 1980s and early 1990s
however, have been very familiar with the conven- but increasingly exoteric since then. It was a skill
tional variety about which the 1952 MCC coaching mastered by England, and not by Australia, in 2005
book [2] provides a clear warning: ‘Nearly all great and proved the greatest factor in why England won
batsmen are agreed that it is the ball which swings the Ashes that year. In contrast with conventional
away from them late in its flight which is the most swing, the reverse form remained a mystery to both
dangerous, and the bowler who can command it, cricketers and fluid dynamicists until quite recently
even if only in his opening overs, may well strike (see, for example, references [7] to [9]). This paper
will demonstrate, with reference to the fluid dynamic
*Corresponding author: Department of Mechanical Engineer- boundary layer and by means of flow visualization
ing, University of Bath, Bath BA2 7AY, UK. experiments, how reverse swing arises and why it
email: ensgdl@bath.ac.uk occurs at a high speed.

JSET73 Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology
192 G D Lock, S Edwards, and D P Almond

pressure distribution which increases with the direc-


tion of flow, and the boundary layer is subjected to an
aerodynamically adverse condition. Here, the relative
decrease in the velocity in the strata near the surface
is more marked than for the outer surface, and with a
large enough pressure rise the flow close to the surface
may reverse in direction and move upstream. Such a
development is illustrated in Fig. 2(a), where it can
be seen that the boundary layer (greatly exaggerated
in thickness for clarity) separates from the surface.
However, the angles of separation are asymmetric as
the prominent seam trips into turbulence the bound-
ary layer, which would otherwise remain laminar,
adjacent to one hemisphere of the ball. Mixing and
turbulence cause the characteristics of a turbulent
boundary layer to differ profoundly from that of a
laminar boundary layer. A turbulent boundary layer
Fig. 1 D. G. Bradman in full flow
is more robust and consequently it can better resist
an adverse pressure gradient. On the seam side then,
the turbulent boundary layer sticks to a greater angle
2 CONVENTIONAL AND REVERSE SWING from stagnation (about 120◦ ) than does that on the
other side (about 80◦ ) where no seam is present
Swing is the sideways deviation of the cricket ball to trip the laminar boundary layer. The result is
during its trajectory towards the batsman. The swing asymmetrical separation, leading to a skewed wake
or swerve in flight of symmetric spheres, such as and a net pressure force on the ball perpendicular
the tennis and golf balls studied by Newton [10], to the flight trajectory. A new ball with a smooth
Rayleigh [11], and Thomson [12], is achieved by surface will help to maintain laminar flow on the
imparting significant spin, giving rise to the Magnus side opposite the seam and the effect is augmented
effect. A cricket ball, in contrast, is asymmetric owing during the course of play if this side of the ball
to the presence of a seam, which is made up of is polished vigorously. Once the surface is slightly
six rows of 80–90 prominent stitches which encircle worn, natural transition to turbulent flow on this
the ball, protruding about 1 mm above the surface. side of the ball will occur and the conventional swing
This seam can be exploited to create a net transverse is lost.
pressure force resulting from the asymmetric sepa- In Fig. 2(b), the polar view of a cricket ball coated
ration of boundary layers on the lateral hemispheres in surface-shear flow visualization paint (discussed
of the ball. The boundary layer, first proposed by in section 4) is superimposed on to the image,
Prandtl [13], is one of the most important concepts demonstrating conventional swing. Reverse swing is
in fluid mechanics. The theory is based on the obser- illustrated in Fig. 2(c). This occurs when the ball
vation that at the surface of a body (such as a cricket is bowled at a high speed (typically greater than
ball) the air is at rest relative to the wall, and in a 80 mile/h). Here, there is natural turbulent flow on
thin region near the body (i.e. the boundary layer) both sides of the ball but the seam thickens the tur-
the relative air velocity increases from zero at the bulent boundary layer on one hemisphere, causing
surface to the undisturbed mainstream velocity with the separation angle to move upstream relative to
distance normal to the surface. the side with the undisturbed boundary layer. The
The aerodynamics of conventional swing are illus- corresponding pressure asymmetry (and wake orien-
trated in Fig. 2(a), which is a polar view, i.e. a view tation) will reverse and a net side force will be created
looking in the direction of gravity. The bowling tech- in the opposite direction to that for conventional
nique is to incline the seam at a slight angle to the air swing.
flow, here shown as 15◦ . By imparting a small back
spin, the bowler can create enough gyroscopic iner-
tia to stabilize this seam position during the trajectory 3 BOUNDARY-LAYER SEPARATION AROUND
to the batsman. At the stagnation point the flow is SPHERES
brought to rest and there is a point of maximum pres-
sure. As the flow accelerates around the ball, there Achenbach [14] conducted a series of seminal experi-
is a corresponding reduction in pressure, creating ments to investigate how the boundary layer governed
a thin boundary layer characterized by a favourable the drag coefficient of a polished, hydraulically
pressure gradient. The flow eventually encounters a smooth metal sphere. These experimental results

Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology JSET73
Flow visualization experiments demonstrating the reverse swing of a cricket ball 193

Fig. 2 (a) Conventional swing; (b) conventional swing with superimposed flow visualization; (c) reverse swing

JSET73 Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology
194 G D Lock, S Edwards, and D P Almond

Fig. 3 (a) Variation in the drag coefficient with the Reynolds number for a smooth sphere; (b) sepa-
ration and transition angles (with respect to the stagnation point) for smooth and roughened
spheres [14, 16]

are reproduced as curves in Fig. 3 and an excellent separation reaches a peak at the critical Reynolds
summary of this research has been provided by Haake number and then decreases significantly in the tran-
et al. [15]. Achenbach showed that, for Reynolds num- scritical region. The rate of decrease increases with
bers Re < 2 × 105 , the flow in the boundary layer is increasing roughness. The diameter of a cricket ball
laminar and the boundary layer separates at θ ≈ 80◦ is between 71 mm and 72.5 mm, and in air the
in this subcritical region. As the Reynolds number critical speed for a smooth ball would be approx-
increases, the boundary layer enters the critical region imately 75 m/s. However, in practice, it is found
and the drag coefficient suddenly decreases dramat- that transition to turbulence for the seam-free side
ically at Re ≈ 4 × 105 . Here, the attached boundary occurs at speeds of about 30–35 m/s, because of
layer is turbulent and the separation points move inaccuracies in the spherical shape and minor surface
downstream so that the boundary layer leaves the irregularities [4].
surface of the sphere at θ ≈ 120◦ . With a further
increase in the Reynolds number, the boundary layer
is supercritical and eventually enters a transcritical 4 EXPERIMENTS
region in which the majority of the attached bound-
ary layer is turbulent. Here, the angle of transi- In the experiments reported here, the University of
tion from laminar to turbulent flow moves upstream Bath wind tunnel was used to test a range of cricket
towards the stagnation point and the turbulent flow balls supplied by Warwickshire County Cricket Club.
over the surface of the sphere thickens the bound- Balls used for 15–80 overs under first-class match
ary layer, which causes a marginally earlier separation conditions were available but the fluid dynamics data
(θ < 120◦ ). Achenbach [16] extended the work to presented here are for new balls only. However,
investigate the effect of surface roughness, showing the cricket balls were used to assess typical surface
that only a small amount of roughness (k/d ≈ 0.0025, roughness, and measurements obtaining using an
where k/d is the ratio of the typical roughness size optical scanner revealed that the surface roughness
to the sphere diameter) reduces the critical Reynolds increased linearly with increasing number of overs
number appreciably. An increase in the roughness (hence deliveries) bowled. A double-sized cricket ball
causes the transition to turbulence to occur at a (diameter, 142 mm) with appropriately scaled seam
lower Reynolds number because of the chaotic dis- was manufactured from rapid-prototype nylon. The
turbances introduced in the boundary layer by the surface was varnished so that the roughness, rela-
roughness elements. The angle of boundary layer tive to the diameter, was equal to that of a new

Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology JSET73
Flow visualization experiments demonstrating the reverse swing of a cricket ball 195

Fig. 4 (a) Scaled cricket ball with pressure tappings; (b) cantilevered beam in the ball wake for side-force
measurements

leather ball (i.e. that of one with zero overs bowled). critical Reynolds numbers for the two sides of the
The scaled ball was hollow with 44 pressure tappings cricket ball are expected to be less than those
around the circumference to record the static pressure measured on the smooth sphere by Achenbach,
distribution, as shown in Fig. 4(a). Re ≈ (150–175) × 103 according to reference [4]. As
The side force (i.e. the force perpendicular to the discussed above, this is partly due to the inher-
flow) was measured simultaneously using a strain ent surface roughness and seam but also because
gauge on a cantilevered beam, as illustrated in of inaccuracies in the general spherical shape of a
Fig. 4(b). The side force S was made non-dimensional manufactured cricket ball.
by the dynamic head of the fluid and projected Figure 5(a) illustrates the variation in the pres-
area of the ball to give Cs (see the Appendix), and sure coefficient Cp (see the Appendix), with the
in this form the actual force is proportional to the angle from the stagnation point for the scaled cricket
square of the air speed. Previous studies by Sayers ball with seam positioned at 15◦ to the air. In
and Hill [17] showed the maximum side force to be Fig. 5(b) the data for the seam side (open symbols)
only 1 N at a wind speed of 34 m/s. The force mea- and the non-seam side (full symbols) are superim-
surement system used in this work was designed posed on to one side of the graph for comparison
to have a sensitivity of 0.01 N. The double-sized ball purposes. Note that there is a region on the seam
was attached to a sting, whose length was five times side where no data could be recorded (because of
the ball diameter, and this was held by a verti- the seam itself). Two Reynolds numbers are shown:
cal post that passed through the upper and lower Re = 92 × 103 and Re = 215 × 103 , which correspond
walls of the wind tunnel to a pair of aluminium to full-sized cricket ball velocities of about 18 m/s and
cantilever beams, one of which was used for the side- 42 m/s (i.e. about 41 mile/h and 94 mile/h) respec-
force measurement. The dimensions of the cantilever tively. At Re = 92 × 103 there is a relatively lower
beams (50 mm × 15 mm × 1 mm) were obtained by an pressure on the hemisphere with the seam which
optimization of the conflicting requirements of low creates a net side force, or the conventional swing
stiffness, to provide sensitivity to very low forces, illustrated in Fig. 2(a). As discussed below, the sep-
and the need to maximize the beam resonant fre- aration angle on the non-seam side of the ball is
quency to minimize the amplitudes of vibrations. about 80◦ but this angle has been delayed to about
A strain gauge was used to monitor the deflec- 120◦ on the seam side. At Re = 215 × 103 there has
tion of the beam caused by the lateral forces on been an inversion of the pressure asymmetry, with
the ball. For each force measurement, the strain relatively lower pressure on the hemisphere with-
gauge was monitored for 1 min using a 10 Hz sam- out the seam; this again creates a net side force,
pling frequency, providing 600 readings. The vibra- but now this is the reverse swing illustrated in
tions of the ball in the wind tunnel air stream Fig. 2(c).
were dampened by a foam pad pressed up against Figure 6 shows an equatorial view of typical surface-
the end of the stain-gauged cantilever beam. The shear flow visualization results on the seam side
measurements of side force obtained by this sys- and the non-seam side of a new cricket ball. The
tem were estimated to have a typical uncertainty of seam was angled at 15◦ to the airflow at Re = 70 × 103 ,
0.004 N. corresponding to a flight velocity of about 31 mile/h,
Surface-shear oil-flow visualization was also used i.e. within the range of conventional swing. Prior
to help to identify the separation angles. The to the experiment the ball was covered in paint and

JSET73 Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology
196 G D Lock, S Edwards, and D P Almond

Fig. 5 (a) Variation in the pressure coefficient with the angle from the stagnation point for Re = 28 × 103 ,
92 × 103 , and 215 × 103 ; (b) data repeated with the seam side and the non-seam side superimposed

Fig. 6 Typical surface-shear flow visualization results on (a) the non-seam side and (b) the seam side
of a new cricket ball

the separation points are identified by the regions of are plotted against the Reynolds number, with the
low shear where the paint remains. In Fig. 2(b), a corresponding velocity shown in miles per hour.
polar view is superimposed on to the image demon- The data follow the trends predicted by Achenbach
strating conventional swing. The seam-side (turbu- [14, 16]. At low Reynolds numbers (flight veloc-
lent) and non-seam-side (laminar) separation points ity, less than 20 mile/h), laminar separation occurs
are seen to be close to the 120◦ and 80◦ angles before transition, even on the seam side. Here,
expected from the work of Achenbach. Figure 7 the pressure distribution is symmetric and the side
shows the seam-side separation angle θS and non- force is measured to be zero. As the Reynolds
seam side separation angle θNS measured from a number increases, the seam trips the boundary
combination of flow visualization and pressure mea- layer to turbulent flow (critical Reynolds number,
surements using both full-size and double-scaled about 70 × 103 ) while the boundary layer on the
cricket balls. Also shown is the difference between non-seam side remains laminar. The correspond-
these two angles, θS − θNS , and the measured non- ing angles of separation are 120◦ and 80◦ , leading
dimensional side force (right abscissa). The data to a skewed wake and a net transverse pressure

Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology JSET73
Flow visualization experiments demonstrating the reverse swing of a cricket ball 197

Fig. 7 The seam-side separation angle θS , the non-seam side separation angle θNS , and the difference
θS − θNS against the Reynolds number (or the flight speed). The non-dimensional side force
(right abscissa) is also shown

force on the ball. This is conventional swing, with (approximately 75–95 mile/h), the separation points
θS − θNS ≈ 40◦ . The critical Reynolds number for the on the two sides of the ball have reversed
non-seam side is Re ≈ 170 × 103 (about 75 mile/h) (θS − θNS ≈ −18◦ ) to those encountered in conven-
where there is natural turbulent flow on both sides of tional swing. This is illustrated in Fig. 2(c). Here, the
the ball, with θNS ≈ 120◦ and the conventional swing corresponding pressure asymmetry (and wake ori-
is lost. entation) will also reverse and a net side force is
Reverse swing occurs at Re > 170 × 103 . The measured in the opposite direction. Note that Fig. 7
boundary layer on the seam side enters the tran- shows the non-dimensional side force Cs . The mag-
scritical regime at Re > 120 × 103 and the separa- nitude of the actual side force is proportional to the
tion angle is observed to move upstream to 100◦ square of the velocity, and so the reverse swing is
at Re ≈ 220 × 103 . For 170 × 103 < Re < 220 × 103 significant.

JSET73 Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology
198 G D Lock, S Edwards, and D P Almond

Thus the range of flight speeds over which con- for supplying a range of worn cricket balls at known
ventional swing occurs corresponds to those of the match conditions.
medium-pace to medium–fast-pace bowler using
a new or well-preserved polished ball. It is only © Authors 2010
the quicker bowlers who achieve reverse swing.
Although the experiments conducted here used REFERENCES
balls with a smooth surface, a ball which is scuffed, or
worn (or illegally tampered with), during the course 1 Haigh, G. Inside out: writings on cricket culture, 2009
of play on the seam side will exacerbate the effect of (Aurum, London).
reverse swing. 2 Anon. The MCC cricket coaching book, 1952 (Naldrett
As a final note, the MCC coaching book has warned Press, London).
of the perils of late swing in particular. The speed 3 Lyttelton, R. A. The swing of a cricket ball. Discovery,
of a cricket ball over the course of its trajectory to 1957, 18, 186–191.
the batsman is only marginally reduced by drag. The 4 Barton, N. G. On the swing of a cricket ball in flight.
lateness of any lateral movement is a natural conse- Proc. R. Soc. A, 1982, 379, 109–131.
5 Mehta, R. D., Bentley, K., Proudlove, M., and Varty, P.
quence of the fact that the displacement, for constant
Factors affecting cricket ball swing. Nature, 1983, 303,
acceleration, is proportional to the square of the
787–788.
time in flight. Thus, the deviation takes a greater pro- 6 Mehta, R. D. Aerodynamics of sports balls. A. Rev.
portion of the total sideways movement as the ball Fluid Mech., 1985, 17, 151–189.
approaches the batsman. 7 Barrett, R. S. and Wood, D. H. The theory and practice
of reverse swing. Sports Coach, 1996, 18, 28–30.
8 Sawyers, A. T. On the reverse swing of a cricket ball –
5 CONCLUSIONS
modelling and measurements. Proc. IMechE, Part C:
J. Mechanical Engineering Science, 2001, 215(1), 45–55.
This paper demonstrates clearly, by means of flow DOI: 10.1243/0954406011520508.
visualization and pressure measurements, the fluid 9 Methra, R. D. An overview of cricket ball swing. Sports
dynamics causes of conventional and reverse swing. Engng, 2005, 8, 181–192.
A cricket ball is asymmetric because of the pres- 10 Newton, I. New theory of light and colours. Phil. Trans.
ence of a seam, which is made up of rows of pro- R. Soc., 1672, 1, 678–688.
minent encircling stitches. For conventional swing, 11 Rayleigh, Lord On the irregular flight of a tennis ball.
this seam trips into turbulence the boundary layer Mess. Math., 1899, 7, 14–16.
adjacent to one hemisphere of the ball which 12 Thompson, J. J. The dynamics of a golf ball. Nature,
1910, 85, 2151–2157.
remains attached to a greater angle (about 120◦ )
13 Prandtl, L. Über Flüssigkeitsbewegung bei sehr kleiner
than does that on the other side (about 80◦ ) where
reibung. In Verhandlungen des III internationalen
no seam is present to trip the laminar boundary Mathematiker-Kongresses (Ed. A. Krazer), Heidelberg,
layer. The result is asymmetrical separation, lead- Germany, 8–13 August 1904, pp. 484–491 (Teubner,
ing to a skewed wake and a net pressure force Leipzig).
on the ball perpendicular to the flight trajectory. 14 Achenbach, E. Experiments on the flow past spheres
Such circumstances occur for Reynolds numbers at high Reynolds numbers. J. Fluid. Mech., 1972, 54(3),
in the range (70–170) × 103 (about 35–75 mile/h). 565–575.
Flow visualization and pressure measurements have 15 Haake, S. J., Goodwill, S. R., and Carre, M. J.
shown that reverse swing is a consequence of the A new measure of roughness for defining the aero-
fact that the asymmetry inverts at Reynolds num- dynamic performance of sports balls. Proc. IMechE,
bers greater than 170 × 103 . Here, there is natural Part C: J. Mechanical Engineering Science, 2007, 221(7),
789–806. DOI: 10.1243/0954406JMES414.
turbulent flow on both sides of the ball but the
16 Achenbach, E. The effects of surface roughness and
seam thickens the turbulent boundary layer on one tunnel blockage on the flow past spheres. J. Fluid. Mech.,
hemisphere, causing the separation angle to move 1974, 65(1), 113–125.
upstream relative to the side with the undisturbed 17 Sayers, A. T. and Hill, A. Aerodynamics of a cricket ball.
boundary layer. J. Wind Engng Ind. Aerodynamics, 1999, 79, 169–182.

ACKNOWLEDGEMENTS APPENDIX
This project was funded by the Royal Academy of Notation
Engineering and Exxon-Mobil Prize for Teaching
Excellence (Dr G. D. Lock). The authors are grate- A projected area of sphere = πd 2 /4 (m2 )
ful to G. Green for helping with the figures and Cd drag coefficient = 2D/(ρv∞2 A)
Mr D. Brown of Warwickshire County Cricket Club Cp pressure coefficient = (pθ − p∞ )/(ρv∞ 2 )

Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology JSET73
Flow visualization experiments demonstrating the reverse swing of a cricket ball 199

Cs non-dimensional side-force = 2S/(ρv∞ 2 A) θ angle from the stagnation point (deg)


d diameter (m) θNS angle from the stagnation point on the
D drag force (N) non-seam side (deg)
k size of roughness element (m) θS angle from the stagnation point on the
pθ static pressure at angle θ (Pa) seam side (deg)
p∞ static pressure in the free-stream (Pa) µ dynamic viscosity (kg/m s)
Re Reynolds number = ρv∞ d/µ ρ density (kg/m3 )
S side force (N)
v∞ free-stream velocity (m/s)

JSET73 Proc. IMechE Vol. 224 Part P: J. Sports Engineering and Technology

You might also like