You are on page 1of 11

Experiments in Fluids 29 (2000) 553±563 Ó Springer-Verlag 2000

Experimental study of flow past a square cylinder at high Reynolds numbers


A. K. Saha, K. Muralidhar, G. Biswas

553
Abstract Measurements of two-components of velocity in D drag per unit length of the square cylinder
the wake of a square cylinder using a hot-wire anemometer f frequency of vortex shedding, Hz
are reported. Two Reynolds numbers, namely 8700 and k dimensionless turbulence kinetic energy
17,625, have been considered. The measurements were Re Reynolds number, uavB/l
carried out in a low-speed, low-turbulence wind tunnel. S Strouhal number, f B/uav
Benchmark experiments at much lower Reynolds numbers u, v streamwise and transverse components of velocity
show good agreement between the present experiments (dimensionless)
and published results. At higher Reynolds numbers, the uav average incoming velocity, also used as a velocity
experimental data reveal anticipated trends in terms of scale
wake recovery and turbulence decay. Both velocity and ; v
u time-averaged streamwise and transverse
velocity ¯uctuations show symmetry about the wake axis. components of velocity
The experimental data have been compared with the large u¢, v¢ streamwise and transverse components of velocity
eddy simulation (LES) calculation reported by Wang et al. ¯uctuations
[University of Illinois at Urbana ± Champaign (1996) Re- u02 †1=2 time average of streamwise ¯uctuation (rms)
port CFD 96-03] and LDV measurements of Lyn et al. v02 †1=2 time average of transverse ¯uctuation (rms)
[J Fluid Mech (1995) 304: 285±319]. The agreement among u0 v0 † time-averaged shear stress
the three sets is generally acceptable in terms of the time- x, y streamwise and transverse coordinates measured
averaged velocity components, but not the velocity ¯uc- from the centre of the cylinder (dimensionless)
tuations. The turbulence ¯uctuations in the present ex-
periments are seen to be lower than in the referred work.
Greek symbols
The differences have been traced to factors such as the
l dynamic viscosity, kg/m-s
aspect ratio, blockage ratio and upstream turbulence.
m kinematic viscosity, m2/s
Experiments with increased upstream turbulence did show
q density, kg/m3
a reduction in the discrepancy between the present ex-
x dimensionless spanwise vorticity
periments and the published data. An assessment of the
experimental data in terms of physical mechanisms re-
vealed that (a) streamwise normal stresses were correlated
1
with the vortex centers, and (b) the turbulence kinetic
Introduction
Motion of non-aerodynamic objects through ¯uids is
energy pro®les are similar to the turbulence shear stress.
encountered extensively in engineering. Submarines,
Spectral analysis of the velocity signals was carried out in
ships, passenger aircraft and cars are examples where
the present work. Energy transfer from the mean ¯ow to
the object is in motion through a stationary ¯uid me-
the streamwise velocity ¯uctuation was con®rmed in the
dium. High-rise buildings, chimneys and tube banks in
near wake. A redistribution of the kinetic energy between
heat exchangers are examples where the ¯uid is in
the streamwise and transverse components of velocity over
motion. In each class, the wake of the bluff object
a longer distance downstream was subsequently observed.
predominantly determines the distribution of forces on
it as well as ¯ow-induced vibration and heat transfer
List of symbols
rates. Recent applications such as the detection of un-
B width of the square cylinder, m
2 observed objects through signature analysis and the
CD drag coef®cient, D/(0.5 q uav B)
dispersion of pollution have further intensi®ed the need
for understanding wake dynamics. The present study is
concerned with high Reynolds number ¯ow past a
Received: 17 May 1999/Accepted: 29 December 1999 square cylinder.
Flow past bluff bodies has been simulated numerically
A. K. Saha, K. Muralidhar, G. Biswas (&) by different researchers. There exist two as well as three-
Department of Mechanical Engineering dimensional computations (Mukhopadhyay et al. 1992;
Indian Institute of Technology
Kanpur, UP ± 208 016, India Tamura et al. 1990) on the ¯ow past a square cylinder.
Two-dimensional calculations are not appropriate at
This investigation has been supported by the Aeronautics R & D higher Reynolds numbers. These must be computed using
Board, New Delhi, grant number AERO/RD-134/100/10/98-99/1032. the unsteady three-dimensional Navier±Stokes equations
on ®ne grids. Engineering applications have, however, been cross-section. The square cylinder used was made of
computed at high Reynolds number in two-dimensions Perspex, 25 ´ 25 mm in cross-section. The use of Perspex
using various eddy viscosity models of turbulence. Owing was advantageous since (1) good quality smooth surfaces
to certain drawbacks associated with the eddy viscosity were attainable over the entire cylinder, and (2) precise
models, the bluff body ¯ows are presently being computed sharpness of the cylinder edges could be maintained. The
using large eddy simulation (LES) (Wang et al. 1996). cylinder length was 40 cm, thus spanning the entire width of
Okajima (1982) carried out an experimental study of the wind tunnel. Upstream velocities of 5.2 and 10.53 m/s
¯ow past a square as well as a rectangular cylinder for a were utilized to produce Reynolds numbers of 8700 and
wide range of Reynolds numbers, namely 70±(2 ´ 104). 17,625, respectively, in the experiments. The room tem-
This study established the variation in Strouhal number, perature during the all experiments was 26(‹2) °C. A few
that is the non-dimensional vortex-shedding frequency, selected low Reynolds number experiments were carried out
554
with Reynolds number. These experiments have shown for benchmarking the instrumentation in a smaller but
that there is an abrupt change in Strouhal number when similar facility. Here, lower velocities and a smaller cylinder
the aspect ratio of the cylinder is reduced to the range 2±3. could be employed.
This is re¯ected as a corresponding change in the wake Velocity was measured using a two-channel hot-wire
¯ow pattern. DuraÄo et al. (1988) have conducted laser anemometer along with an X-wire probe. The X was
Doppler velocimetry (LDV) measurements for ¯ow past a formed in the vertical plane, with the cylinder placed in a
square cylinder in a water tunnel at a Reynolds number of horizontal position. The probe was mounted on a tra-
14,000. They have separated the periodic and random versing mechanism that facilitated all three orthogonal
components of velocity ¯uctuations. These measurements movements, to a positional accuracy in the most signi®-
show that the kinetic energy associated with the random cant direction, namely the vertical at ‹0.1 mm. The
components is about 40% of the total. Lyn et al. (1995) commercially available Dantec anemometer and probes
have reported an LDV study of turbulent ¯ow past a were employed in the present work. The two wires of the
square cylinder with emphasis on the ensemble-averaged probe were calibrated in the wind tunnel itself. Small
characteristics of the ¯ow behaviour. The Reynolds num- changes in room temperature (‹1 °C) were compensated
ber considered in their study was 21,400. The experiments through the use of a correction formula that assumes a
were carried out in a closed and constant head water constant heat transfer coef®cient. The probe was recali-
tunnel. Their results showed a relationship between the brated for larger changes in temperature. Both wires were
¯ow topology and the turbulence distribution. Vorticity operated at 200 °C, and their calibration curve were seen
saddles and streamlines saddles were clearly distinguished. to be almost identical. The assumption of equal sensitivity
A distinction was identi®ed between the ¯ow in the base coef®cients of the two wires was occasionally employed
region and the near wake. Differences in the length and during data reduction. The calibration curves were
velocity scales, and celerity of the vortices of ¯ows around smoothed using a ®fth-order polynomial. During calibra-
a circular cylinder and a square cylinder have also been tion, a pitot tube connected to a Furnes Controls
compared. 19.99 mm of H2O digital manometer was utilized. Both DC
In the present investigation, an experimental study of and rms values of voltages were recorded using ``true''
¯ow in the near wake of a square cylinder at Reynolds voltmeters supplied by Dantec. Integration times of typi-
numbers of 8700 and 17,625 was carried out using a hot- cally 300 s were used to obtain all time-averaged quanti-
wire anemometer. Measurements were conducted along ties. The cross-correlation between the voltage ¯uctuations
the mid-plane of the cylinder and across the wake. Both was determined by collecting long time traces through an
streamwise and transverse components of velocity were Advantest spectrum analyser, followed by a numerical
measured. Time±mean velocity as well as ¯uctuations were integration of the product of the two signals (over 10 s).
recorded in the experiments. The scope of the present The frequency of vortex shedding was measured by
work is to map the near wake of the cylinder and compare the spectrum analyser using the FFT algorithm. The
it with the numerical results of Wang et al. (1996) and the time-averaged drag acting on the cylinder was deter-
experiments of Lyn et al. (1995). mined by the wake survey method. Momentum loss
The present experiments were carried out in an open calculation was carried out at the mid-plane of the cyl-
circuit wind tunnel with a nominal cross-section of inder at a dimensionless distance of x = 15. A consid-
40 ´ 40 cm. A 3-m-long test section of the wind tunnel was erable variation in the static pressure was seen within
available for experiments. Since the divergence angle of the the wake and so the measurements were directed indi-
test section was 4°, a fair constancy of pressure was vidually at the total and static pressures. Some variation
observed. The wind tunnel fan was electronically controlled in the tunnel static over a distance of 15 units was also
and stable velocities in the range 5±20 m/s were realizable in accommodated for. The conversion of voltages to ve-
the experiments. The speed control was formally rated at locities was accomplished through an explicit non-
‹0.5% but over a 3-h period, no noticeable change was ob- real-time method developed by Chew and Simpson
served in the approach velocity. Free stream turbulence in (1988). This technique has the advantage of determining
the approach ¯ow was damped by an array of mesh screens. the mean and the ¯uctuating components of velocity as
In all the experiments, turbulence was found to be less than well as their cross-correlations with very few assump-
the background noise of the anemometer, being equivalent tions regarding probe parity and turbulence level. Details
to 0.01% at an approach velocity of 10 m/s. The inlet velocity of the apparatus and instrumentation have been reported
pro®le was uniform to within ‹0.5% over 90% of the tunnel elsewhere (Saha 1999).
Velocity and velocity ¯uctuations in the wake of the
square cylinder were measured beyond a dimensionless
distance of 1.5 units. In view of the insensitivity of the hot-
wire probe to reversed ¯ow, measurements in the recir-
culation region of the cylinder are questionable. Earlier
studies show that the size of the recirculation region at
high Reynolds numbers is in the range 1.3±1.4 (DuraÄo
et al. 1988; Lyn et al. 1995). The recirculation zone was
established in these studies through alternative techniques
such as the LDV and the ¯ying hot wire. Hence, the choice
of the ®rst measurement station for the time-averaged
555
velocity as 1.5 can be taken as an acceptable starting point.
In experiments with low as well as high velocities, the
velocity ¯uctuations close to the cylinder can be expected Fig. 1. Comparison of measured Strouhal number with published
to be large. Under these conditions, the hot-wire mea- data
surements using the ®xed operating point approach is
questionable. In the present experiments, the ®rst mea- Reynolds number is shown in Fig. 2. The measured values
surement location was at x = 1.5, and a de®nite positive follow the pattern predicted by the numerical computa-
value of the mean velocity was recorded. Hence, on no tion. The drag coef®cient shows initially an increasing
occasion was the calibration curve utilized close to zero trend with Reynolds number and later becomes almost
velocity. Secondly, the instantaneous voltage signal of the asymptotic. It was not possible to measure drag below a
anemometer was converted to pointwise velocity and Reynolds number of 400 by the wake survey method,
subsequently, the time-averaged values were computed. owing to diffusive and wall effects. Measured values of
The difference between this strategy as against the con- drag at low Reynolds numbers have not been reported in
ventional operating point approach was found to be neg- the literature.
ligible, particularly in the high Reynolds number The experimental results of the present study are re-
experiments. This observation also ®nds support in the ported for Re = 8700 and 17,625. The blockage ratio based
work of Swaminathan et al. (1986). These authors have on the wind tunnel height is 6.25% and the length-to-width
shown a turbulence correction factor of 10% at a turbu- ratio of the cylinder is 16. To a ®rst approximation, this
lence intensity as high as 45% with respect to the mean represents the cylinder in an in®nite medium; blockage
¯ow. The corresponding probability of occurrence of re- and aspect ratio effect are thus expected to be of secondary
versed ¯ow was found to be less than 1%. The above importance (West and Apelt 1982).
discussion improves the degree of con®dence that can be At the Reynolds numbers studied, the drag coef®cients
placed on the hot-wire data. were found to be 2.13 and 2.2, respectively. These values
The uncertainty in the present set of experiments was are comparable with the numerical estimate of 2.07±2.67 at
found to be unexpectedly low. This observation was drawn a Reynolds number of 21,400 by Wang et al. (1996). It is in
from the fact that the dimensionless time-averaged ve- closer agreement with the experiments of Cheng et al.
locity pro®les at Reynolds numbers of 8700 and 17,625 (1992), who have presented the drag coef®cient as 2.1.
were practically identical. The clear peak in the near-wake Based on the vortex-shedding frequencies, the respec-
spectra ensured negligible uncertainty in the Strouhal tive Strouhal numbers have been found in the present
number. The uncertainty in the drag coef®cient assessed experiments to be 0.144 and 0.142 at Re = 8700 and
using the pilot tube and hot-wire data was found to be 17,625. These values differ from those reported by DuraÄo
within ‹5% with 95% con®dence. et al. (1988) as well as Lyn et al. (1995), who present the
Strouhal number as 0.138 and 0.132 at Reynolds numbers
2 of 14,000 and 21,400, respectively. The reasons for this
Results and discussion
The results of the present work have been organized as
follows: (1) drag coef®cient and Strouhal number; (2)
time-averaged velocity and velocity ¯uctuations; (3) Effect
of inlet turbulence; and (4) turbulence spectra.

2.1
Drag coefficient and Strouhal number
To benchmark the measurement technique against pub-
lished data, selected experiments were carried out at low
Reynolds numbers. Figure 1 depicts such a comparison for
the Strouhal number as a function of Reynolds number.
Data available in the published literature are also included.
The data points of the present work lie between the nu-
merical results of Davis et al. (1984) and the experiments
of Okajima (1982). The drag coef®cient as a function of Fig. 2. Comparison of measured drag with published data
discrepancy is possibly a high inlet turbulence level of This is also an indication of a low uncertainty level in the
2±6% in the referred experiments. The numerical simula- high Reynolds number experiments.
tion of Wang et al. (1996) shows a Strouhal number of 0.13 Figure 4 is a comparable plot for the transverse com-
at a Reynolds number of 21,400. The discrepancy here is ponent of velocity. Both Figs. 3 and 4 reveal good sym-
probably related to a smaller blockage in the simulation of metry (and antisymmetry) in velocities with respect to the
Wang et al. (1996). wake axis. The measurement of a negative transverse ve-
locity is a check on the correctness of the data-reduction
procedure. The agreement between experimental and nu-
2.2
merical data is uniformly satisfactory.  
Time-averaged velocity and velocity fluctuations
Figures 5 and 6 compare
 the streamwise u02 †1=2 and
A detailed comparison of the measured pro®les of velocity
556 and velocity ¯uctuations with the LDV measurements of transverse v02 †1=2 velocity ¯uctuations between experi-
Lyn et al. (1995) and the LES calculations of Wang et al. ments and numerical simulation. The agreement here is
(1996) is now presented. Time-averaged velocities and less satisfactory. Signi®cant differences can be seen at the
velocity ¯uctuations were non-dimensionalized using uav ®rst measurement station. Both experiments as well as
the average incoming velocity. The turbulent shear stress calculations reveal a double-peaked pro®le in u02 , while
was non-dimensionalized using uav2. single peak is seen in v02 with a maximum occurring along
The numerical calculations of Wang et al. (1996) em- the axis. The differences between the experiments and
ploy a Reynolds number of 21,400 as stated above. Besides simulations tend to diminish at larger streamwise locations.
this, the cylinder length is p, the blockage ratio is 0.05 and Figure 7 compares the turbulent shear stress from the
periodicity boundary conditions have been applied in the present experiments with the LDV measurements and the
spanwise direction. In the LDV experiments, the Reynolds numerical computation. The comparison is seen to be
number is 21,400, the cylinder length is 9.75 and the closer in these ®gures compared with v02 . The change of
blockage ratio is 7%. The inlet turbulence level in the sign in u0 v0 † about the wake centreline is clearly revealed.
present experiments was 0.01% at 10 m/s (Re = 17,500) Once again, the agreement improves in the downstream
and 0.05% at 5 m/s (Re = 8750). In the water tunnel ex- direction.
periments of Lyn et al. (1995), the inlet turbulence level The centreline recovery of the time-averaged stream-
was around 4% at Re = 21,400. In the numerical simula- wise velocity is shown in Fig. 8a and b for the Reynolds
tion of Wang et al. (1996), the inlet turbulence level was numbers under discussion. The present measurements
not speci®ed; one can expect it to be negligible. In view of show that the recovery is faster compared with the ex-
these differences, one cannot expect a complete match periments of Lyn et al. (1995) and LES results of Wang
among LDV measurements, numerical calculations and et al. (1996).
the present hot-wire measurements. In summary, the numerical results match experiments
Figure 3a and b shows the time-averaged streamwise at high Reynolds number when compared in terms of the
velocity pro®les at three locations downstream of the time-averaged velocities, the streamwise velocity ¯uctua-
cylinder at Reynolds numbers of 8700 and 17,625, re- tions and the turbulent shear stress. Numerical predictions
spectively. The numerical results are represented on the of the transverse velocity ¯uctuations are signi®cantly
same graph by solid lines. A remarkable aspect to be no- higher than the measured values. The possible reasons for
ticed in Fig. 3 is the invariance of the dimensionless ve- the discrepancy are the short spanwise dimension (=p) in
locity distribution with respect to both Reynolds numbers. computations as against a value of 16 used in experiments,

Fig. 3a, b. Comparison of


measured time-averaged
streamwise velocity pro®les: a
Re = 8700; b Re = 17,625
557

Fig. 4a, b. Comparison of


measured time-averaged trans-
verse velocity pro®les: a
Re = 8700; b Re = 17,625

Fig. 5a, b. Comparison of


measured time-averaged pro-
®les of streamwise velocity
¯uctuations: a Re = 8700; b
Re = 17,625

Fig. 6a, b. Comparison of


measured time-averaged pro-
®les of transverse velocity
¯uctuations: a Re = 8700; b
Re = 17,625
558

Fig. 7a, b. Comparison of


measured time-averaged pro-
®les of turbulent shear stress: a
Re = 8700; b Re = 17,625

stream. A minimum in this velocity is attained close to the


wake centreline. As seen in Fig. 3, the minimum velocity
increases in the downstream direction. Simultaneously,
there is an increase in the wake size. The two processes can
happen only when ¯uid is entrained from the edges of the
wake on each side of the centreline. A measure of the
entrainment of the ¯uid can be derived from the trans-
verse velocity pro®les in Fig. 4. It can now be seen that
entrainment, and hence the velocity recovery, are both
very high for small distances behind the cylinder. Both of
them asymptotically reach their respective limits for
higher downstream distances.
The variation in the turbulent ¯uctuations in the wake
requires greater elaboration. The mechanism involved is
not only dissipation but also the substantial energy
transfer from the core to the surrounding ¯uid. Hence, the
¯uctuations decay rapidly in the near wake, and their
transverse variation become progressively uniform.

2.3
Effect of inlet turbulence on the wake
Experiments with hot wire were in partial agreement with
LDV measurements and LES. Signi®cant differences were
seen in the pro®les of transverse velocity ¯uctuations. One
factor that has not been matched among the LDV exper-
iments, numerical simulation and the present experiments
is the turbulence level in the incoming ¯ow. The present
Fig. 8a, b. Comparison of recovery of measured time-averaged section is concerned with the study of inlet turbulence on
streamwise velocity along the wake centreline: a Re = 8700; the velocity ¯uctuations in the wake.
b Re = 17,625 The effect of inlet turbulence has been studied by
placing a circular cylinder centrally at a distance of 150
diameters (of the circular) upstream of the square cylin-
differences in the blockage ratio and the upstream tur- der. Consequently, the approximate turbulence level
bulence level. measured just upstream of the square cylinder (y = )2.0
A physical discussion on the results obtained can be to 2.0) was about 3.5%. The measurements at two different
conducted in the following manner. The separation of ¯ow streamwise locations, namely x = 1.5 and 2.5 were con-
over a bluff object causes a pressure drop across its sur- ducted at a Reynolds number of 12,000. The time-averaged
face, thus leading to a non-zero form drag. Equivalently, velocity pro®les at x = 1.5 (Fig. 9a) show the in¯uence of
this is manifested as loss of momentum of the ¯uid in the inlet turbulence on the recovery rate. The increase in the
wake. Thus, the time-averaged streamwise velocity is turbulence level leads to a slower recovery of the velocity
smaller within the wake compared with its value in the free de®cit. Simultaneously, the measured pro®les approach
559

Fig. 9a±c. Effect of upstream turbulence on time-averaged pro- Fig. 10a±c. Effect of upstream turbulence of time-averaged pro-
®les of: a streamwise velocity; b transverse velocity; and ®les of: a streamwise velocity; b transverse velocity; and c
c streamwise velocity ¯uctuations at x = 1.5 for a Reynolds streamwise velocity ¯uctuations at x = 2.5 for a Reynolds number
number of 12,000 of 12,000

those of Lyn et al. (1995). The effect of inlet turbulence on (Huot et al. 1986). When the entrainment becomes very
the transverse velocity as depicted in Fig. 9b is seen to be large, the free shear layers eventually reattach to the sides
small. The time-averaged streamwise ¯uctuations at of the obstacle. Consequently, separation is reinitiated at
x = 1.5 are presented in Fig. 9c. The ®gure shows that the rear corners. The entrainment of the ¯uid from the
there is a decrease in velocity ¯uctuations due to an in- freestream to the wake is then slow and the recovery rate is
crease in inlet turbulence. Figure 10 shows the effect of small. Figure 10b shows that the transverse entrainment in
inlet turbulence on the u, v and u02 †1=2 at a location the present measurements is marginally higher at a loca-
x = 2.5 over one half of the wake. These pro®les are greatly tion x = 2.5 compared with Lyn et al. (1995). The higher
in¯uenced and approach the data of Lyn et al. (1995). entrainment in the present experiments with low upstream
Thus an increase in the upstream turbulence in the present turbulence is possibly the cause for a higher recovery rate
experiments shows only a minor in¯uence at x = 1.5 but compared with the published results.
an improved comparison with the published measure-
ments at a distance of x = 2.5. 2.4
The role of upstream turbulence in modifying the wake Turbulence spectra
can be explained as follows. External turbulence increases Since hot-wire experiments reveal local pointwise infor-
the entrainment of the ¯uid into the free shear layer from mation, it is usually cumbersome to map the ¯ow ®eld in
the recirculation zone and hence increases their curvature all respects. Instead, the power spectra of x and y com-
ponents of velocity determined at selected locations in the consistent units. (Note: At low Reynolds numbers, the
wake can shed light on the wake dynamics. With these near-wake spectra showed dominant peaks. The far-wake
objectives, the u and the v spectra were recorded both spectra were not recorded owing to the potential in¯uence
along the centreline (y = 0) and an offset position of the bounding walls).
(y = 1.0) at a Reynolds number of 17,625. Figure 11 shows Figure 11 clearly brings out a dominant peak at the
the u spectra at locations x = 2.5, 5.0, 10.0 and 30.0 for two shedding frequency, particularly when the probe is away
y planes in the cylinder wake. Figure 12 is a comparable from the centreline. The spectra along the centreline do
plot for the v spectra. In these ®gures, the abscissa is the not reveal a clear trend. At y = 0, the second harmonic is
dimensionless frequency, while the ordinate is plotted in seen to be excited at x = 2.5, signifying the in¯uence of

560

Fig. 11. u Spectra at various streamwise locations


561

Fig. 12. v Spectra at various streamwise locations

vortices shed from both halves of the cylinder on the ¯ow for x- locations greater than 10.0. At x = 30.0, a
probe. At higher x locations the u spectra tend to become broadband spectrum in u is seen to be established.
broader, with a peak developing close to the origin. Thus, In contrast to the u velocity, clear peaks are visible in
¯ow near the centreline shows an established cascade with the v spectra at y = 0 as well as at y = 1.0. The peak is
energy transfer from the time-averaged streamwise veloc- observable all the way upto x = 30.0, though it is stronger
ity to the smaller ¯uctuations. at the offset location. The broadening of the v spectra at
The peak in the u- spectra seen at the offset location x = 30.0 indicates the loss of coherence in the ¯ow ®eld
diminishes in strength as compared with that for the mean and an evolution towards a fully developed state.
An interesting result can be discerned by comparing the
peak values of u and v spectra between x = 2.5 and 5.0
(y = 1.0). One can see an increase in the peak values in the
downstream direction, followed by decay at farther dis-
tances. This result suggests an intensi®cation of the vor-
tices in the near wake, followed by their decay.
Two new results that emerge from the spectral plots are:
1. The shift in the peak of the u spectra towards the origin
of the frequency axis at x = 2.5 con®rms energy transfer
from the mean ¯ow to the streamwise velocity ¯uctua-
562 tion close to the cylinder. The redistribution of kinetic
energy between u and v occurs over a much longer
length, and is not complete even at x = 30.0.
2. The initial growth of the spectral peaks followed by a
decay indicates a strengthening of the vortices outside the
vortex formation length, namely 1.0±1.5. The underlying
mechanism is the transport of transverse momentum
from the core of the wake to the free shear layer.
Fig. 13a, b. Vorticity variation in the near wake of a square
2.5 cylinder: a Re = 8700; b Re = 17,625
Identifiable physical mechanisms
A physical interpretation of the results presented above is
Figure 13 shows that the positions of the vorticity
taken up in this section. The discussion is in terms of the
maxima and minima upto x = 4 are essentially unchanged.
spanwise vorticity in the time-averaged ¯ow ®eld and the
This is proof that the vortices move parallel to the mid-
turbulence kinetic energy. During the measurement pro-
plane in the near wake. The authors have computed ¯ow
gram, the traversing mechanism was set up to resolve the
past a square cylinder at high Reynold numbers using the
vertical (y) direction transverse to the wake. The stream-
k±e model (Saha 1999). Their numerical prediction also
wise (x) direction was resolved to a lesser extent. From the
shows vortices originating on the top and bottom side of
de®nition
the square cylinder. Farther downstream, the vortex cen-
ov ou ters are closer to the midplane, as opposed to the vortex

ox oy shedding patterns for a circular cylinder.
the spanwise vorticity was hence approximated as The turbulence kinetic energy pro®les in the wake for the
two Reynolds numbers are shown in Fig. 14. A comparison
ou with Fig. 7 shows that k is similar to the turbulence shear

oy stress u¢v¢ at all x locations, except that the latter changes
This is not a serious limitation, since v as well as its sign. The change of sign is quite prominent at the ®rst
streamwise derivative are relatively small outside the base x location where turbulence is produced by the shear in the
region of the cylinder. The turbulent kinetic energy (k) was mean velocity pro®le. At subsequent locations, turbulence is
computed using the formula
3 02
kˆ u ‡ v02 †
4
where the approximation
w02 ˆ 0:5 u02 ‡ v02 †
was used (Hadid et al. 1992).
The vorticity plots using the de®nition above are shown
in Fig. 13 for both Reynolds numbers of the present study.
A comparison of this data at x = 1.5 with Figs. 5 and 6
shows that large streamwise normal stresses are correlated
with the vortex centres the vorticity reaches a maximum.
This correspondence is not seen either in the transverse
normal stress component or at other x locations. This
result can be explained as follows: At the ®rst x location,
turbulence production (primarily the streamwise normal
stress) originates in the shear present in the time-averaged
¯ow and thus correlates with the vorticity distribution. At
other downstream points, turbulence is governed by the
transport processes in the wake and its association with Fig. 14a, b. Kinetic energy variation in the near wake of a square
the time-averaged velocity gradient is weakened. cylinder: a Re = 8700; b Re = 17,625
dominated by transport (to and from a point) including References
transport from the core to the edge of the wake. Conse- Cheng CM; Lu PC; Chen RH (1992) Wind loads on square
quently, the k and u¢v¢ pro®les are much ¯atter. cylinder in homogeneous turbulent ¯ows. J Wind Eng Ind
Aerodyn 41: 739±749
3 Chew YT; Simpson RL (1988) An explicit non-real time data
reduction method of triple sensors hot-wire anemometer in
Conclusions three-dimensional ¯ow. J Fluids Eng 110: 110±119
Hot-wire measurements in the wake of a square cylinder Davis RW; Moore EF; Purtell LP (1984) A numerical-experi-
were reported at Reynolds numbers of 8700 and 17,625. mental study of con®ned ¯ow around rectangular cylinders.
The high Reynolds number experiments were compared Phys Fluids 23: 46±59
with LDV measurements and LES. The primary conclu- DuraÄo DFG; Heitor MV; Pereira JCF (1988) Measurements of
sions of this study are as follows: turbulent and periodic ¯ows around a square cross-section
cylinder. Exp Fluids 6: 298±304 563
1. The agreement among the measured and simulated Hadid AH; Sindir MM; Issa RI (1992) Numerical study of two
data is satisfactory in terms of the time-averaged velocity dimensional vortex shedding from rectangular cylinders. Comp
components, streamwise velocity and the turbulent shear Fluid Dyn J 2: 207±214
stress. The numerical prediction of transverse velocity Huot JP; Rey C; Arbey H (1986) Experimental analysis of the
pressure ®eld induced on a square cylinder by a turbulent ¯ow.
¯uctuations is substantially higher than the measured J Fluid Mech 162: 283±298
values. This discrepancy can be attributed to (1) a small Lyn DA; Einav S; Rodi W; Park J-H (1995) A laser-Doppler ve-
spanwise length in the simulation in comparison to the locimetry study of ensemble-averaged characteristics of tur-
experiments, (2) differences in the blockage ratio, and (3) bulent near wake of a square cylinder. J Fluid Mech 304:
a low upstream turbulence level in the present experi- 285±319
ments. Mukhopadhyay A; Biswas G; Sundararajan S (1992) Numerical
2. A fresh set of experiments with a higher turbulence investigation of con®ned wakes behind a square cylinder in a
channel. Int J Numer Meth Fluids 14: 1473±1484
level in the approach ¯ow has shown that the comparison Okajima A (1982) Strouhal numbers of rectangular cylinders.
improves with the published experimental data. J Fluid Mech 123: 379±398
3. The spectra of the streamwise velocity component at Saha AK (1999) Dynamical characteristics of the wake of a square
the near wake con®rms that the transfer of energy takes cylinder at low and high Reynolds numbers. Ph.D. Thesis,
place from the mean motion of the streamwise velocity Indian Institute of Technology Kanpur, India
¯uctuation and the redistribution of kinetic energy be- Swaminathan MK; Rankin GW; Sridhar K (1986) Evaluation of
tween the two components of velocity occurs over a longer the basic systems of equation for turbulence measurements
using the Monte Carlo technique. J Fluid Mech 170: 1±19
streamwise direction downstream. The spectra also show Tamura T; Ohta I; Kuwahara K (1990) On the reliability of two-
the transport of transverse momentum from the core of dimensional simulation for unsteady ¯ows around a cylinder-
the wake to the free shear layer. type structure. J Wind Eng Ind Aerodyn 35: 275±298
4. At the ®rst measurement station x = 1.5, there was a Wang G; Choudhuri PG; Bhandarkar MA; Vanka SP (1996)
good association between the locations of the vortex Report, Large eddy simulations of bluff body wakes on parallel
maxima and the turbulence shear stress. The correspon- computers, CFD 96-03, University of Illinois at Urbana ±
dence was lost at other downstream points. The turbulence Champaign, Contract No. 0014-92-J-1334 of Of®ce of Naval
Research
kinetic energy and the turbulence shear stress showed West GS; Apelt CJ (1982) The effect of tunnel blockage and as-
similar trends in the wake. However, the shear stress pect ratio on the mean ¯ow past a circular cylinder with
changed sign at the ®rst measurement point, revealing that Reynolds numbers between 104 and 105. J Fluid Mech 114:
turbulence production at this site was shear-dominated. 361±377

You might also like