You are on page 1of 10

Experimental Thermal and Fluid Science 31 (2007) 751–760

www.elsevier.com/locate/etfs

The influence of temperature gradient on the Strouhal–Reynolds


number relationship for water and air
a,*
T. Vı́t , M. Ren b, Z. Trávnı́ček c, F. Maršı́k c, C.C.M. Rindt b

a
Faculty of Mechanical Engineering, Technical University of Liberec, Hálkova 6, 461 17 Liberec, Czech Republic
b
Energy Technology Division, Department of Mechanical Engineering, Eindhoven University of Technology,
WH-3.127, P.O. Box 513, NL-5600 MB Eindhoven, Netherlands
c
Institute of Thermomechanics, Academy of Sciences of the Czech Republic, Dolejškova 5, 182 00 Prague 8, Czech Republic

Received 9 February 2006; received in revised form 27 July 2006; accepted 2 August 2006

Abstract

This paper focuses on the wake flow behind a heated circular cylinder in the laminar vortex shedding regime. The phenomenon of
vortex shedding from a bluff body is an interesting scientific and engineering problem. Acquisition of reliable experimental data is con-
sidered an indispensable step toward a deeper physical understanding of the topic.
An experimental study of the wake flow behind a heated cylinder in the forced convection regime is performed using water as the
working fluid. Firstly, qualitative visualization experiments were performed and the parallel vortex shedding mode was adjusted. Next,
hot-wire anemometry was used for St–Re data acquisition. Data analysis confirmed the so-called thermal effect in water: cylinder heating
increases the vortex shedding frequency and destabilizes the wake flow.
The effective temperature concept was used and the St–Re data were successfully transformed to the St–Reeff curve. Furthermore, a
comparison with air as the working fluid was discussed (cylinder heating decreases the vortex shedding frequency in air, thus stabilizing
the wake flow). The formula to determine the effective temperature in water was experimentally derived from the present data, while the
data and formula for air is already known. The relationship between the Strouhal number and the effective Reynolds number for water
and air is represented by the same, universal formula: St ¼ 0:2660  1:0160Re0:5
eff , where Reeff is calculated at the effective temperature.
Finally, the measurement results were compared to the thermodynamic St–Re equation derived by Maršı́k et al. [F. Maršı́k, Z. Tráv-
nı́ček, R.H. Yen., A.-B. Wang, Fluid dynamics concept for the critical Reynolds number of a heated/cooled cylinder in laminar cross-
flow, in preparation]. A satisfactory agreement between the derived equation and experimental data for both fluids (water and air) was
achieved.
 2006 Elsevier Inc. All rights reserved.

Keywords: Laminar flow; Vortex shedding; Heated circular cylinder; Effective temperature; Thermal effect

1. Introduction engineering point of view. The phenomenon of vortex


shedding from a bluff body has been studied by many
Fluid flow around a heated bluff body, namely a circular authors – e.g. hundreds of references can be found in the
cylinder, is of principal importance for fluid dynamics as comprehensive monograph by Zdravkovich [4]. This phe-
well as for heat transfer (see, e.g., [2,3]). The low Reynolds nomenon is of fundamental importance in the theoretical
number range, where laminar vortex shedding occurs, is study of hydrodynamic instability which includes many
considered very important from a scientific as well as an problems dealing with wake flow dynamics (e.g., [5]), such
as the onset of vortex shedding, the passing frequency of
*
Corresponding author. Tel.: + 420 485 353 405; fax: + 420 485 353
vortices, and the influence of geometrical and material
644. parameters. From an engineering point of view, the
E-mail address: tomas.vit@email.cz (T. Vı́t). phenomenon of vortex shedding is considered one of the

0894-1777/$ - see front matter  2006 Elsevier Inc. All rights reserved.
doi:10.1016/j.expthermflusci.2006.08.002
752 T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760

sources of flow-induced vibrations, noise, or even body col- A recent numerical study by Shi et al. [16] concluded
lapse. It influences drag as well as heat transfer in an exter- that the effective temperature defined by Eq. (2) [11] agrees
nal flow. well with their results [16].
This paper focuses on the wake flow behind a heated It is obvious that a quite opposite situation occurs in
bluff body, namely a circular cylinder in the laminar vortex water, where the kinematic viscosity decreases with temper-
shedding regime. The heated cylinder is studied in the ature. Therefore, cylinder heating destabilizes the wake
forced convection regime, as is explained in the following flow in water. This was confirmed experimentally by Lecor-
text. dier et al. [17]. However, a lack of adequate experimental
data for quantitative confirmation (or adaptation) of the
1.1. Thermal effects in air and water for forced convection Teff concept for fluids other than air is evident.
It is worth mentioning that the effective temperature is
The working fluid properties such as viscosity, density not ‘‘just an artificial value’’, like the well-known film tem-
and thermal conductivity are fundamentally important perature, which is defined as the arithmetic mean of the
for thermal effects. In the text below two of the most wall and free-stream temperatures: T1 + 0.5(Tw  T1).
common working fluids – air and water – are discussed. The effective temperature is close to the hot recirculation
In air, a heat input stabilizes the wake flow, thus laminar zone temperature according to Dumouchel et al. [15]; fur-
vortex shedding can be completely suppressed by heating thermore the maximum temperature in the wake measured
the cylinder. The onset of vortex shedding (the lower limit by Yahagi [18] was apparently very close to the effective
of the vortex shedding regime) for an unheated cylinder has temperature according to the calculation by Wang et al.
been studied many times. The commonly accepted critical [11]. However, no consistent and reliable experimental con-
Reynolds number (Re = dU/m, where d is the diameter of firmation of this idea has been published in the available
the cylinder, U the velocity of the undisturbed flow and m literature thus far, a fact that is one of the main motiva-
the kinematic viscosity at the temperature of the undis- tions of this study.
turbed flow) ranges from 40 to 49 (e.g., Rec = 40 by The dynamics of the wake behind a bluff body is com-
Kovasznay [6]; 44 by Collis and Williams [7]; 45.9 by Lange monly quantified by means of the Strouhal number (St,
et al. [8]; 47 by Fey et al. [9]; 49 by Williamson [10]). denoting a non-dimensional frequency: St = df/U, where
Cylinder heating suppresses this onset of instability, thus f is the flow frequency). For the St–Re relation of the
the Rec value increases with heating. This increase was isothermal case, several equations have been presented in
evaluated in the range of Rec = 47.7–70 when the cylinder the literature, e.g., [19,20,9]. The influence of heating on
temperature increases to nearly 290 C [11]. A possible the frequency of vortex shedding was studied recently in
explanation for this thermal effect in air is the increase in air, and it was concluded that the vortex shedding fre-
kinematic air viscosity with temperature, which causes a quency decreases with increasing cylinder temperature.
decrease in the local Reynolds number. Another known This frequency decrease was quantified by means of the
explanation of the thermal effect in air emphasizes the effective temperature concept. When Reeff is evaluated
reduction in fluid density with a temperature increase, at Teff defined above in Eq. (2), the derived relationship
and thus a reduction in absolute instability [12]. Another St–Reeff is found to be ‘‘universal’’, i.e. valid for both
approach, based on the analytical description of the vari- heated and unheated cylinders [11]:
able properties of fluids, was suggested by Herwig and 1:0160
Wickern [13]. St ¼ 0:2660  pffiffiffiffiffiffiffiffiffiffi ð3Þ
Reeff
If the idea of the so-called effective temperature is
applied, the onset of vortex shedding even for the case of A recent numerical study by Shi et al. [16] uses the results
a heated cylinder can be described by the critical effective of Wang et al. [11] as the reference experimental data for
Reynolds number Rec,eff, which is the same for both heated the heated cylinders. Shi et al. [16] concluded that the
and unheated cylinders. A value of Rec,eff = 47.5 ± 0.7 was experimental data [11] agree well with the numerical results
evaluated by Wang et al. [11]. [16], [21], and that their numerical results [16] confirm the
The idea of effective temperature was proposed origi- experimental findings of the effective temperature (Eq. (2)
nally by Lecordier et al. [14], and used later by Dumouchel [11]).
et al. [15], who worked out this concept and calculated the It can now be assumed that this (or a similar) relation is
effective kinematic viscosity meff from an effective tempera- valid for vortex shedding in water as well, where the cylin-
ture Teff that is defined by der heating (logically) increases the frequency of the shed-
ding of vortices. However, no studies have focused on this
T eff ¼ T 1 þ cðT w  T 1 Þ; ð1Þ problem.
where T1 and Tw are the free-stream and cylinder surface It is worth noting that a precise evaluation of the St–Re
temperatures, respectively. Recently, the effective tempera- curve is intrinsically complicated if the wake is under the
ture was derived by Wang et al. [11] in the following form influence of the so-called end effects caused by the end con-
ditions of the tested cylinder. Under these circumstances,
T eff ¼ T 1 þ 0:28ðT w  T 1 Þ: ð2Þ the vortex shedding from the cylinder is not parallel but
T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760 753

slanted to the cylinder axis and typical discontinuities in The measurement methods and experimental setup used
the St–Re curve exist [10]. There are a few different end- to define the St–Re relation for water are briefly described
manipulating methods to isolate the end effects and gener- in the following chapter, as are the construction of the tow-
ate parallel vortex shedding. These have been shown to be ing tank and heated cylinder. Possible sources of errors and
very effective by Williamson [20]; Eisenlohr and Eckelmann inaccuracies in the non-isothermal flow in water are also
[22]; Miller and Williamson [23]; Hammache and Gharib described here. One paragraph deals with the problems of
[24] and Wang et al. [11]. parallel vs. oblique vortex shedding. The results of the
experiments are presented in Section 3.
1.2. Criteria for forced convection
2. Experimental apparatus and techniques
The forced convection regime occurs at a low level of
heating. In contrast, at a higher level of heating buoyancy The experiments were carried out in a towing tank
effects are added to the viscous phenomena and a mixed installed at Eindhoven Technical University (a detailed
convection flow occurs. The vortex shedding frequency description of the structure is given by Kieft [31]). The
can be basically altered. Typically, the Strouhal number main dimensions and design of the towing tank are shown
increases with the Grashof number. However, the vortex schematically in Fig. 1. A towing mechanism enables the
shedding can also be suppressed, which is denoted as a cylinder to move within a speed range 0–2 cm/s through
‘‘breakdown of the von Kármán vortex street’’ [25,26]. the tank which corresponds to a Reynolds number interval
For not too high heat inputs the vortex street can be of Re = 0–90 using a cylinder of diameter d = 4.5 mm. The
deflected [27], and the 2D wake flow can be turned into a heated cylinder was designed from a copper tube with
3D structure [28]. Of course, all these mixed convection external diameter d = 4.5 mm, as shown schematically in
effects depend on the Grashof and Reynolds numbers Figs. 2 and 3. The cylinder was equipped with end cylinders
(Gr and Re respectively), as well as on the orientation of made of Plexiglas with a diameter of 10 mm on both ends.
the free-stream with respect to the buoyancy force (buoy- The cylinder together with the end cylinders were attached
ancy effects can be opposite or parallel or cross into the free between end plates (detail in Fig. 2) that were tightly
stream). Different criteria for the mixed convection region mounted to the solid frame of the towing device. The
to occur are suggested in the literature, all based on the presence of end cylinders and end plates, as proposed by
ratio Gr/Res. A typical value used for the exponent s is Eisenlohr and Eckelmann [22] (where the diameter of the
within the range of 1.8–3. The value s = 2 is very common end cylinders is recommended in the range of D = 1.8d–
where the ratio Gr/Re2 is called the Richardson number, Ri. 2.2d and the length of the end cylinders should be at least
The criterion for mixed convection to occur in the cross l = 5d) should ensure parallel vortex shedding during the
flow situation is Ri > 0.5, as used by Wang et al. [11] and experiment. The free length of the cylinder between these
by Wang and Trávnı́ček [29]. Maas et al. [28] found that two end cylinders was l = 250 mm (this corresponds to
for water at Ri = 0.3 the flow becomes 3D. Other criteria the aspect ratio k = l/d = 45.3).
are mentioned by Dumouchel et al. [15] and by Wang The construction of the heating element is depicted in
and Trávnı́ček [29]. In fact, the maximum Ri values in Fig. 3. Heat is provided by a manganese resistance wire
the experiments by Dumouchel et al. [15] and by Wang with a length of 250 mm and an electric resistance of
et al. [11] were much lower, around Ri 6 0.012 and 0.02, 21 X, which is placed in a ceramic capillary. The maximum
respectively, to avoid mixed convection in the laminar electrical load is approximately 80 W, and the range of the
vortex shedding regime. temperature differences DT = Tw  T1 was from 0 to
The value of the Richardson number in this study satis- 2.8 K for all experiments.
fies the relation of Ri 6 0.15 (Re1 = 56.3, Gr = 483), hence
the buoyancy effects can be neglected.

1.3. Outline of the paper

The main goals of the present study are: (1) to experi-


mentally identify and qualify the thermal effects in water
for forced convection on the Reynolds–Strouhal number
relation; (2) to compare the observed results for water to
the results for air; and (3) to demonstrate the effective
temperature concept by comparing it with (a) the St–Re
relationship for the isothermal case (e.g., the relationship
proposed by Williamson and Brown [30]) and the St–Reeff
relationship for the non-isothermal case [11], and with (b)
the relation derived theoretically from wake flow dynamics
by Maršı́k et al. [1]. Fig. 1. Towing tank.
754 T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760

mocouple (Th. 2) was placed on the cylinder surface.


During the calibration, Th. 2 was placed in the center of
the cylinder span; then Th. 2 was moved to the end of the
cylinder to avoid wake disturbances during the experiments
(see Fig. 3). A third thermocouple (Th. 3 in Fig. 2) was
placed 30 mm in front of the cylinder to measure the
free-stream temperature T1.

2.1. Temperature calibration

If we disregard the heat flow to the end plates and to the


end cylinders we can consider the value of heat flow from
the surface of the cylinder to the fluid to be directly propor-
tional to the current flow through the resistant wire. Dur-
ing the calibration for a known value of the Reynolds
number and a selected value of current I, the temperature
difference DT = Tw  T1 was measured. The calibration
characteristic will be deduced from the following overall
convective heat transfer equation:

Fig. 2. Construction of the end cylinders and the end plates. Position of
Q ¼ aADT ; ð4Þ
thermocouple Th. 3.
where a is the mean heat transfer coefficient and A is the
overall surface area of the exposed body.
The heat transfer for forced convection for moderate
temperature loading can be expressed by the following
non-dimensional form – see e.g. [32]:
Nu ¼ ðA þ cRen ÞPrm ; ð5Þ
where Nu is the Nusselt number (Nu = ad/k), Pr is the
Prandtl number, k is the heat conductivity, and A, c, n,
and m are experimentally determined constants.
Evidently, heat transfer varies predominantly with the
Reynolds and Prandtl numbers and many other parameters
Fig. 3. Construction of the heated cylinder and positions of thermocou- affect the process such as temperature loading, boundary
ples Th. 1 and Th. 2; 1, ceramic capillary; 2, copper cylinder; 3, end conditions, aspect ratio (end effects), blockage effects due
cylinder; 4, end plate; 5, to the power supply. to the wind tunnel and wakes, and free-stream turbulence.
It is worth noting here that the overall heat transfer in its
general form should include the so-called temperature
Experiments carried out in water are very sensitive to loading factor – see, e.g., the discussion by Wang and Tráv-
the temperature of the liquid. For this reason special atten- nı́ček [29]. However, as the temperature differences in the
tion was paid to measure this temperature precisely during present experiments are relatively small, the so-called tem-
the experiments. The form of heating of the cylinder corre- perature loading factor is neglected in Eq. (5).
sponds to a boundary condition of constant heat flux, and The material properties of water (e.g. by Gebhart [34]:
so it was necessary to carry out a temperature calibration heat conductivity, kinematic viscosity and Prandtl number)
before the actual experiment started. This means that we can be expressed as a function of temperature in the follow-
had to define the characteristic: DT = f(Re, I), where I is ing forms:
the value of amperage through the resistant wire. The pro-
 x1  x2  x3
cess of calibration used calibration equations and assumed T T T
error in temperature measurement are discussed in Section k ¼ k0 ; m ¼ m0 ; Pr ¼ Pr0 ð6Þ
T0 T0 T0
2.1. To measure the temperature at significant places, Alu-
mel–Chromel thermocouples with diameter 0.076 mm were where T0 is the reference temperature, T1 6 T0 6 Tw.
used. The first thermocouple for measuring the tempera- Taking Eqs. (4) and (5), Ohm’s law (Q = RI2, where R is
ture inside the cylinder was placed between the copper tube the resistance), and assuming that A  cRen in Eq. (5) (in
and the ceramic capillary (Th. 1 in Fig. 3). This thermocou- the investigated Re range), the calibration function can be
ple was used to define the steady state of the system. Exper- expressed as
iments were started only after the temperature at this 1=2
thermocouple settled on a constant value. A second ther- I ¼ ½ð1=RÞðk=dÞcRen Prm AðT w  T 1 Þ : ð7Þ
T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760 755

For a specific wire resistance (R), cylinder (d, A), and water show the parallelism of vortex shedding. The frequency
as the working fluid (Pr, k), and assuming that Eq. (5) is results extracted from the visualization experiments were
valid for material properties Pr and k corresponding to not considered in the final (quantitative) results.
T1 (this assumption can be made for the relatively small
temperature differences present in water), Eqs. (6) and (7) 2.3. Frequency measurement using CTA
yield
X 1=2 Although there were problems mainly due to the low
I ¼ CRen=2 ðT 1 =T 0 Þ ðT w  T 1 Þ ð8Þ
velocity of the water flow (approximately 10–20 mm/s)
where X = (x1 + mx3)/2. and transfiguration of the velocity field (parallel versus
The purpose of the calibration was to find values for the oblique vortex shedding), the constant temperature ane-
constants C, X, m, and n so that function (8) would corre- mometer (CTA) with a film probe proved suitable to find
spond to the values measured during calibration. St–Re dependence correctly even for non-isothermal flow.
For selected values of Re and DT, the value of heater A ‘‘hot film’’ probe (type P55R36), and a StreamWare
current I was set before each experiment. Real temperature system together with StreamLine software produced by
values T1,r(t) and Tw,r(t) were measured during the exper- DANTEC were used to perform the experiments. The
iment. The
R mean value of the real temperature difference parameters of the CTA circuit were adjusted as shown in
DT r ¼ 1t ðT w;r  T 1;r Þdt was calculated from these tem- Table 1. The probe was placed in the axis of the cylinder
peratures, where t is the duration time of the experiment at a downstream distance of 40–50 mm during the experi-
(typically up to 30 s). Next, the error of measurement ments as shown in Fig. 2.
e = DT  DTr was calculated. If this error was smaller than The relative velocity of the flow past the cylinder was set
0.1 K the results were used for further processing. The by the velocity of the towing mechanism. The calibration of
uncertainties of the thermocouples were 0.05 K. the film probe was also carried out in the towing tank. The
probe was fixed to the towing appliance and was towed
2.2. Visualization with known velocity through the tank. Fig. 5 shows a
typical time line record of the voltage over the CTA Bridge.
Ions of tin were used for visualization. A tin wire (actu-
ally four wires formed into a grid) with diameter 0.6 mm
Table 1
was used as an anode. More information about the method Parameters of CTA bridge
can be found in Maas et al. [28]. It was noticed during the
Overheat ratio 0.12 –
experiments that the grid formed by the wires as well as the Over temperature 40 C
tin wire itself influenced the experiments too much which Offset 2.766 –
led to a decrease in the frequency of vortex shedding. For LP filter 0.3 kHz
this reason the results of the visualization were used only Sampling rate 0.05 kHz
Number of samples 8192 –
to study the qualitative structure of the flow, mainly to

Fig. 4. Parallel and oblique regimes of vortex shedding at Re = 70, DT = 1.4 K. Pictures past (a) 7 s, i.e. 25d and (b) 63s, i.e. 225d from the beginning of
the experiment.
756 T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760

Fig. 5. Time record of the CTA signal. The change in the frequency during the experiment resulted from the change from parallel to oblique modes of
vortex shedding. Recorded at Re = 71, DT = 0 at the beginning (0–20 s: black line) and at the end (32–52 s: grey line) of the experiment.

2.4. Discussion of experimental uncertainties

2.4.1. Parallel vortex shedding


It is a well-known fact that vortex shedding from a cir-
cular cylinder is influenced by so-called end effects, which
are related to the transition to oblique vortex shedding
modes, and thus to discontinuities in the St–Re relation-
ship [20,22]. Parallel vortex shedding is related to the con-
tinuous St–Re curve.
Initially, after the start of cylinder motion in the towing
tank, the shedding pattern is parallel. However, if the end
effects change the mode of shedding at the cylinder ends,
these changes diffuse over the whole cylinder span. This
time development is demonstrated in Figs. 4 and 5.
Firstly, parallel vortex shedding was confirmed on the Fig. 6. Variation of St in spanwise direction y; Re = 87; Tw/T1 = 1.0072.
Corresponding visualization is shown in the background; flow is oriented
basis of visualization. Fig. 4(a) shows parallel vortex shed- bottom-up.
ding along almost the whole length of the cylinder. The
effect of the end cylinders is clearly visible behind them.
If the movement of the cylinder lasts longer, the influence Obviously, precise measurements required a well-stabilized
of the end effects appears and the flow is no longer parallel flow and temperature fields in the standing water of the
– see Fig. 4(b). To make a comparison, Fig. 4(a) and (b) towing tank. Fluid flow in the tank was practically stabi-
shows the velocity field after the cylinder was moved 25d lized after a long time when the experiment was carried
and 225d, respectively, from the beginning of the experi- out – it was necessary to wait approximately 2 hours
ment. Only the data obtained under parallel vortex shed- between experiments.
ding were collected for evaluation. In addition, the presence of large structures in the tank
Fig. 5 demonstrates a response of the parallel and non- causes a motion of the wake in the vertical direction. Fig. 7
parallel vortex shedding. The non-parallel (oblique) vortex demonstrates schematically how this effect can even
shedding is characterized by a decrease in the recorded fre- increase the obtained frequency. Because this effect is sig-
quency on the time line record. The distortion of the vortex nificant at low Re numbers near the critical Reynolds num-
street always appeared after a certain period during the ber, the critical Re number was difficult to detect in the
experiments and the regime of shedding could no longer towing tank. Therefore the present quantification of effec-
be considered parallel. tive temperature (c in Eq. (1)) could not be based on mea-
Parallel vortex shedding was also proved by CTA mea- surement of the critical Re number (similarly as for air by
surement in the spanwise direction. Typical results for the Wang et al. [11]), but it was based on the whole St–Re
CTA and visualization, obtained at the same parameters, curve – see the explanation in Section 3.2 below.
are shown in Fig. 6. This demonstrates that the frequency Another significant complication is the non-uniformity
along the span was practically constant except in the vicin- of the temperature field. To avoid this non-uniformity,
ity of the cylinder end (y = 100 mm), where a decrease in the tank was filled at least one day before the experiment
frequency was found. This end effect agrees very well with to ensure equalization of the water temperature with the
the spanwise measurement of the vortex shedding fre- laboratory temperature. However, temperature uniformity
quency, which is known from the literature (e.g., [10,33]). cannot be ideal – residual thermal drifts, which are partly
connected with the large eddy structures and partly with
2.4.2. Disturbances in the towing tank natural thermal stratification, remain permanently. The
In comparison with similar experiments carried out in maximum difference between the highest and lowest tem-
air, several problems complicated the experiments in water. perature of the water in the tank during one experiment
T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760 757

3. Results

3.1. St–Re relationship

Fig. 8 presents an example of the results: St–Re relation-


ships for an unheated cylinder (Tw/T1 = 1.0000) and for
several heated cylinders (Tw/T1 = 1.0024–1.0096). The
scale on the vertical axis is shifted by St = 0.01 for each
of the temperature differences to make the chart easier to
read.
For comparison purposes, the curve corresponding to
the unheated cylinder according to Williamson and Brown
Fig. 7. Schematic demonstration of error in frequency measurement due [30] is drawn to each of the data series. This curve is con-
to motion of the large structures.
sidered the reference curve for the isothermal case; its form
is [30]
of T1,max  T1,min = 0.2 K was measured during the
experiments. Inaccuracies caused by the non-uniformity 1:0175
St ¼ 0:2665  pffiffiffiffiffiffi ð9Þ
of the temperature field influenced the experimental results Re
by changing the thermo-physical properties of the fluid. It
is worth pointing out that in contrast with air (with rela- It is noteworthy here that the ‘‘universal’’ (i.e. valid for
tively weak temperature dependencies of its thermo-physi- both heated and unheated cylinders) St–Reeff relationship
cal properties) the differences in the kinematic viscosity of defined by Eq. (3) [11] gives practically the same results
water are much bigger. For example, a rather small temper- in the isothermal case, where Reeff = Re and the maximum
ature increase of 1 K from 290 K to 291 K decreases the deviation of Eqs. (3) and (9) is less than 0.23% – see Wang
kinematic viscosity by 2.5%. et al. [11].
The uncertainty of the CTA measurement, based on the Fig. 8 confirms the expected thermal effect in water:
calibration procedure and fitting of the calibration curve cylinder heating causes an increase in frequency. The
[35] was estimated at 1%. To illustrate this value, the differ- perturbation of the data is less than 1.5% (i.e. the stan-
ences ±1% from resultant Eq. (3) are plotted by the dotted dard deviation between data and a smooth curve fitting
lines in Fig. 9 – see the text below. them).

Fig. 8. St–Re relationship of the unheated and heated cylinder in the water flow. Comparison of measured data (points) with the reference unheated
cylinder by Eq. (9) – Williamson and Brown, [30].
758 T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760

Quantification of the thermal effect is discussed in T1 = 1.0072 and 1.0096 probably point to influences of
the following two sections by means of the effective free convection, which causes an increase in the frequency.
temperature concept [11] (in Section 3.2) and the thermo- The present value of c = 0.97 for water is different from
dynamic derivation of the St–Re relationship [1] (in Section the known c = 0.28 for air (Eq. (2) [11]). A reasonable
3.3). explanation for these differences for water and air seems
to correspond to the different Prandtl numbers and to the
3.2. Effective temperature and St-Reeff relationship different thicknesses of the momentum and temperature
boundary layers.
The idea of the effective temperature assumes a flow For comparison purposes, an expansion of the St–Reeff
similarity at the onset of vortex shedding. An original relationship for air [11] (i.e. Eqs. (2) and (3)) into the
construction of the effective Reynolds number assumes St–Re domain was made according to the derivation by
that critical effective Reynolds numbers are the same for Trávnı́ček et al. [36], which yielded the following formula:
both unheated and heated cases [14,15,11]. However, the  0:8887
1:0160 Tw
present Teff calculation could not use this approach. It St ¼ 0:2660  pffiffiffiffiffiffi 0:72 þ 0:28 ð10Þ
was impossible to find the critical Re exactly at different Re T1
DT’s because of the experimental limits mentioned above. Fig. 10 show the relationship of St–Re with T* as a
Therefore, the definition of the effective temperature was parameter. As the reference St–Re curve, the formula by
based on the whole St–Re curve instead of the critical Re Williamson and Brown [30] Eq. (9) for the isothermal case
number only. In other words, instead of the commonly is plotted there. The available data of the heated cylinder in
used approach [11,14,15], which postulated that the critical air by Wang et al. [11] are plotted there: thermal effect in air
effective Reynolds number is independent of temperature, means that cylinder heating causes a frequency decrease
we postulated that all the present data for different temper- and thus the wake flow is stabilized. A very good agree-
ature ratios (DT = 0–0.28) collapse to the ‘‘universal’’ ment between data and Eq. (10) can be concluded.
St–Reeff curve by using the Reeff concept, if Teff is computed Furthermore, the present data for the experiments in
from Eq. (1). Next, the constant c in Eq. (1) was water are also shown in Fig. 10 (the scale on the vertical
determined. axis is shifted by St = 0.01 for the experiments in water
It was found from the measured data that the best fit is to make the chart easier to read, similarly as in Fig. 8.
achieved at c = 0.97. The dependence of St–Reeff for Because the temperature and frequency differences in air
c = 0.97 is shown in Fig. 9. All frequency data for the are much bigger, the data for the experiments in air are
non-heated and heated cylinders collapse to the ‘‘univer- not shifted).
sal’’ St–Reeff curve of Eq. (3). The dotted lines in Fig. 9 cor- The different thermal effect in water can be elucidated in
respond to the differences ±1% from Eq. (3). Deviations Fig. 10: cylinder heating in water increases the vortex shed-
that appear at a lower Re and particularly at Tw/ ding frequency and thus destabilizes the wake flow.
It is noteworthy here that the effective temperature is not
‘‘just an artificial value’’; the Teff appears to be close to the
hot recirculation zone temperature [15]. Moreover, the
maximum temperature in the wake measured by Yahagi
[18] is apparently very close to the calculated Teff according
to the discussion by Wang et al. [11]. The rather high
present value of c = 0.97, which was found for water, indi-
cates that the maximum temperature of the water wake
appears to be quite close to the cylinder temperature Tw.
However, the measurement of the water wake temperature
lies outside the scope of this work.

3.3. Thermodynamic St–Re relationship

A second possible approach to interpreting the mea-


sured frequency is based on the thermodynamic derivation
of the St–Re relationship by Maršı́k et al. [1]. This
approach, contrary to the effective temperature concept,
is based on a detailed analysis of the velocity and temper-
ature fields. The derivation takes into account the physical
properties of various fluids. The resultant equation, whose
Fig. 9. St–Reeff relationship of the unheated and heated cylinder in the
water flow; Reeff was calculated from Eq. (1) with c = 0.97. The smooth constants 0.2665 and 1.0175 match Eq. (9) (i.e. the isother-
line represents correlation Eq. (3), dotted lines correspond to 1% error mal case by Williamson and Brown [30]) has the following
from this. form:
T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760 759

Fig. 10. St–Re relationship of the unheated and heated cylinder in air and water. Comparison of the experimental data (open symbols denote the present
experiments in water; full symbols denote experiments in air by Wang et al. [11]) with the reference unheated cylinder (Eq. (9)), with correlation based on
the experiment in air (Eq. (10)), and with the theory (Eq. (11)).

sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
1:0175 x 0:227ð1  T  Þ A comparison of the data for experiments in water and
StðRe1 ; T  ; Pr1 Þ ¼ 0:2665  pffiffiffiffiffiffiffiffiffi ðT  Þ 2 1 þ 1=3 air with Eq. (11) is shown in Fig. 10. A satisfactory agree-
Re1 Pr1 ð2T   1Þ
ment between the derived Eq. (11) and the experimental
ð11Þ data for both fluids is demonstrated.
where T* is the temperature ratio T* = Tw/T1, and x is the
exponent in the dependence of the viscosity on temperature 4. Conclusions
defined as
 x This paper is concerned with vortex shedding behind a
T
l ¼ l0 ð12Þ heated circular cylinder in water and air. The first experi-
T0
mental step related to an adjustment of the parallel vortex
where l and l0 are molecular (dynamic) viscosities at tem- shedding mode was performed using the flow visualization
peratures T and T0. in water. Next, hot-wire anemometry in water was used to
Eq. (11) is valid for various fluids, and their different obtain the St–Re data. It was confirmed that the frequency
material properties are embodied in their different x and of vortex shedding changes with the temperature gradient
Pr1. The exponent x for water and air was discussed by in the boundary layer. It was found that cylinder heating
Maršı́k et al. [1]: in water increases the frequency (thermal effect in water).
This increase was measurable even for a relatively small
• For water in the temperature range of 15–30 C, the temperature difference DT = Tw  T1.
exponent x was found from data by Gebhart [34] in The effective temperature (Teff) concept was used for
the temperature range T 2 (10; 30) C as x = 7.00. processing the St–Re data from the hot-wire measurement.
• The exponent for air was taken as x = 0.7774. The present evaluation of Teff was based on the fitting of all
St–Re data onto one St–Reeff curve (instead of the com-
The Prandtl number of water and air at atmospheric pres- monly known evaluation by means of the flow similarity
sure and temperature T1: at the onset of vortex shedding). The resultant effective
temperature in water was found to be Teff = T1 +
• The Prandtl number of water monotonically decreases 0.97(Tw  T1).
from Pr = 7.99 at 15 C to Pr = 5.39 at 30 C (by Geb- The present experimental data for water flow were sup-
hart [34]). The curves plotted in Fig. 10 were evaluated plemented by available data for a heated cylinder in airflow
at Pr1 = 7.053 (T1 = 20 C). [11]. The present results show that the concept of effective
• The Prandtl number of air monotonically decreases temperature, originally suggested for heated cylinders in
from 0.71 at 20 C to 0.68 at 300 C (e.g., by Hilsenrath air, can be used for heated cylinders in water. The relation-
and Touloukian [37]); the curves plotted in Fig. 10 were ship between the Strouhal number and the effective
obtained using the common approximation by the con- Reynolds number for water and air is represented by the
0:5
stant value of 0.7. same (universal) formula: St ¼ 0:2660  1:0160Reeff , where
760 T. Vı́t et al. / Experimental Thermal and Fluid Science 31 (2007) 751–760

Reeff is evaluated at Teff defined as Teff = T1 + c(Tw  T1), [16] J.-M. Shi, D. Gerlach, M. Breuer, G. Biswas, F. Durst, Heating effect
where c = 0.97 or 0.28 for water or air, respectively. on steady and unsteady horizontal laminar flow of air past a circular
cylinder, Phys. Fluids 16 (12) (2004) 4331–4345.
The measurement results were also compared to the [17] J.C. Lecordier, L.W.B. Browne, S. Le Masson, F. Dumouchel, P.
thermodynamic St–Re equation derived by Maršı́k et al. Paranthoën, Control of vortex shedding by thermal effect at low
[1], which has the form of St = f(Re1, T*, Pr1), and which Reynolds numbers, Exp. Therm. Fluid Sci. 21 (4) (2000) 227–237.
is valid for various fluids. A satisfactory agreement [18] Y. Yahagi, Structure of two-dimensional vortex behind a highly
between the derived Eq. (11) and the experimental data heated cylinder, Trans. JSME B 64 (1998) 209–215.
[19] A. Roshko, On the drag and shedding frequency of two-dimensional
for both fluids was achieved. bluff bodies, NACA TR 1191, 1954.
[20] C.H.K. Williamson, Oblique and parallel modes of vortex shedding in
Acknowledgements the wake of a circular cylinder at low Reynolds numbers, J. Fluid
Mech. 206 (1989) 579–627.
We gratefully acknowledge the support of the Eindhoven [21] M. Sabanca, F. Durst, Flow past a tiny circular cylinder at high
temperature ratios and slight compressible effects on the vortex
University of Technology and the Grant Agency AS CR shedding, Phys. Fluids 15 (7) (2003) 1821–2003.
(No. IAA200760504). The valuable editorial corrections [22] H. Eisenlohr, H. Eckelmann, Vortex splitting and its consequences in
by Mr. Dennis Ferner are also sincerely appreciated. the vortex street wake of cylinders at low Reynolds number, Phys.
Fluids A 1 (2) (1989) 189–192.
References [23] G.D. Miller, C.H.K. Williamson, Control of three-dimensional phase
dynamics in a cylinder wake, Exp. Fluids 18 (1994) 26–35.
[24] D. Hammache, M. Gharib, An experimental study of the parallel and
[1] F. Maršı́k, Z. Trávnı́ček, R.H. Yen, A.-B. Wang, Fluid dynamics
oblique vortex shedding from a circular cylinder, J. Fluid Mech. (232)
concept for the critical Reynolds number of a heated/cooled cylinder
(1991) 567–590.
in laminar crossflow, in preparation.
[25] N. Michaux-Leblond, M. Bélorgey, Near-wake behavior of a heated
[2] H. Schlichting, K. Gersten, Boundary-Layer Theory, eighth ed.,
circular cylinder: viscosity-buoyancy duality, Exp. Therm. Fluid Sci.
Springer-Verlag, Berlin, 2000.
15 (2) (1997) 91–100.
[3] F.P. Incropera, D.P. DeWitt, Introduction to Heat Transfer, third
[26] K.-S. Chang, J.-Y. Sa, The effect of buoyancy on vortex shedding in
ed., John Wiley & Sons, New York, 1996.
the near wake of a circular cylinder, J. Fluid Mech. 220 (1990) 253–
[4] M. Zdravkovich, Flow Around Circular Cylinders, vol. 1, Oxford
266.
University Press, 1997.
[27] R.N. Kieft, C.C.M. Rindt, A.A. van Steenhoven, The wake behav-
[5] P.A. Monkewitz, C.H.K. Williamson, G.D. Miller, Phase dynamics
iour behind a heated horizontal cylinder, Exp. Therm. Fluid Sci. 19
of Kármán vortices in cylinder wakes, Phys. Fluids 8 (1996) 91–96.
(4) (1999) 183–232.
[6] L.S.G. Kovasznay, Hot-wire investigation of the wake behind
[28] W.J.P.M. Maas, C.C.M. Rindt, A.A. van Steenhoven, The influence
cylinders at low Reynolds numbers, Proc. R. Soc. London, Ser. A
of heat on the 3D-transition of the von Karman vortex street, Int.
198 (1949) 174–190.
J. Heat Mass Transfer 46 (16) (2003) 3069–3081.
[7] D.C. Collis, M.J. Williams, Two-dimensional convection from heated
[29] A.-B. Wang, Z. Trávnı́ček, On the linear heat transfer correlation of a
wires at low Reynolds numbers, J. Fluid Mech. 6 (1959) 357–384.
heated circular cylinder in laminar cross flow by using a new
[8] C.F. Lange, F. Durst, M. Breuer, Momentum and heat transfer from
representative temperature concept, Int. J. Heat Mass Transfer 44
cylinders in laminar crossflow at 104 6 Re 6 200, Int. J. Heat Mass
(24) (2001) 4635–4647.
Transfer 41 (1998) 3409–3430. p
[30] C.H.K. Williamson, G.L. Brown, A series in 1/ Re to represent
[9] U. Fey, M. König, H. Eckelmann, A new Strouhal–Reynolds number
the Strouhal–Reynolds number relationship of the cylinder wake,
relationship for the circular cylinder in the range 47 < Re < 2 · 105,
J. Fluids Struct. 12 (8) (1998) 1073–1085.
Phys. Fluids 10 (1998) 1547–1549.
[31] R.N. Kieft, Mixed convection behind a heated cylinder, Ph.D. Thesis,
[10] C.H.K. Williamson, Vortex dynamics in the cylinder wake, Ann. Rev.
Eindhoven University of Technology, 2000.
Fluid. Mech. 28 (1996) 477–539.
[32] V.T. Morgan, The overall convective heat transfer from smooth
[11] A.-B. Wang, Z. Trávnı́ček, K.-C. Chia, On the relationship of
circular cylinders, Adv. Heat Transfer 11 (1975) 199–264.
effective Reynolds number and Strouhal number for the laminar
[33] M. König, H. Eisenlohr, H. Eckelmann, The fine structure in the
vortex shedding of a heated circular cylinder, Phys. Fluids 12 (6)
Strouhal–Reynolds number relationship of the laminar wake of a
(2000) 1401–1410.
circular cylinder, Phys. Fluids A 2 (9) (1990) 1607–1614.
[12] M.-H. Yu, P.A. Monkewitz, The effect of nonuniform density on the
[34] B. Gebhart, Heat Conduction and Mass Diffusion, McGraw-Hill,
absolute instability of two-dimensional inertial jets and wakes, Phys.
New York, 1993.
Fluids A 2 (7) (1990) 1175–1181.
[35] H.H. Bruun, Hot Wire Anemometry, Oxford University Press, 1995.
[13] H. Herwig, G. Wickern, The effect of variable properties on laminar
[36] Z. Trávnı́ček, A.-B. Wang, F. Maršı́k, Flow visualization of the
boundary layer, Wärme Stoffübertragung 20 (1986) 47–57.
laminar vortex shedding behind a cooled cylinder, in: W.-C.Wang
[14] J.C. Lecordier, L. Hamma, P. Paranthoën, The control of vortex
(Ed.), International Symposium on Experimental Mechanics ISEM
shedding behind heated circular cylinders at low Reynolds numbers,
(CD – proceedings), December 28–30, 2002, Taipei, ROC, C209.
Exp. Fluids 10 (4) (1991) 224–229.
[37] J. Hilsenrath, Y.S. Touloukian, The viscosity, thermal conductivity,
[15] F. Dumouchel, J.C. Lecordier, P. Paranthoën, The effective Reynolds
and Prandtl number for air, O2, N2, NO, H2, CO, CO2, H2O, He, and
number of a heated cylinder, Int. J. Heat Mass Transfer 41 (12) (1998)
A, Trans. ASME 76 (1954) 967–985.
1787–1794.

You might also like