Sie sind auf Seite 1von 14

International Journal of Heat and Fluid Flow 46 (2014) 2942

Contents lists available at ScienceDirect

International Journal of Heat and Fluid Flow


journal homepage: www.elsevier.com/locate/ijhff

The effect of mixed convection on the structure of channel ow at low


Reynolds numbers
Ahmed Elatar, Kamran Siddiqui
Department of Mechanical and Materials Engineering, University of Western Ontario, London, Ontario, Canada

a r t i c l e i n f o a b s t r a c t

Article history: An experimental study was conducted to investigate the effect of bottom wall heating on the ow struc-
Received 29 August 2012 ture inside a horizontal square channel at low Reynolds numbers (Re) and high Grashof numbers (Gr). The
Received in revised form 4 December 2013 ow eld was found to be complex and three-dimensional due to the interactions of buoyancy-induced
Accepted 15 December 2013
rising plumes of warm uid, falling parcels of cold uid and the shear ow. The mean streamwise velocity
Available online 21 January 2014
proles were altered by bottom wall heating; and back ow was induced in the upper half of the channel
when Gr/Re2 > 55. The bottom wall temperatures were found to have more signicant inuence on the
Keywords:
turbulent velocity magnitudes than the ow rate. The Reynolds stress became negative in the channel
Channel ow
Wall heating
core region indicating the momentum transfer from the turbulent velocity eld to the buoyancy eld.
Mixed convection The POD analysis revealed the presence of convective cells primarily in the lower half of the channel.
Low Reynolds number 2013 Elsevier Inc. All rights reserved.
Particle image velocimetry
Proper orthogonal decomposition

1. Introduction laminar regime when exposed to heat, generates turbulence and


hence, the investigation of turbulent ow behavior is crucial to
Mixed convection heat transfer where both forced as well as understand the underlying physical processes associated with the
free convection modes exist, can be found in several industrial mixed convection. To the best of authors knowledge, no previously
applications. Two main factors that control the heat transfer mech- reported studies had conducted a detailed investigation of the tur-
anism and consequently the ow behavior are Grashof number bulent ow behavior in low Reynolds number channel ows dur-
(Gr) and Reynolds number (Re). Grashof number is the ratio be- ing mixed convection.
tween buoyancy and viscous forces, and the Reynolds number is Gajusingh and Siddiqui (2008) experimentally studied the effect
the ratio between inertial and viscous forces (Incropera et al., of wall heating on the ow characteristics in the near wall region
2006). Mixed convection at low Reynolds numbers is important inside a square channel. Their main focus was on the region imme-
in electronics cooling, food process industry, chemical and nuclear diately adjacent to the bottom heated surface. They studied how
reactors, and biomedical applications. Recently, a new application heat transfer would affect ow dynamics in the near wall region
of low Reynolds number mixed convection is emerging in the for originally laminar and turbulent ows. They found that the tur-
green energy sector where the solar thermal systems that convert bulence was generated due to buoyancy for originally laminar ow
solar energy into heat operate at low Reynolds numbers and high while for originally turbulent ow, buoyancy dampened turbu-
Grashof numbers. lence. They argued that in originally turbulent ow, the turbulence
Several studies investigated the low Reynolds number mixed dampens due to working against buoyancy. They quantied the
convection inside channels. Different channel geometries, orienta- instability due to stratication using Richardson number. They ar-
tions and boundary conditions were examined. Generally, the main gued that for originally laminar ow, the instability produced by
focus of these previous studies was on quantifying the bulk prop- heating enhances turbulence while for originally turbulent ow,
erties such as Nusselt number or coefcient of friction and investi- instability due to heating would reduce turbulence magnitude.
gating its variation along the channel heating section (e.g. Several studies investigated the effect of bottom wall heating on
Mahaney et al., 1987; Maughan and Incropera, 1987). Some studies the Nusselt number in a horizontal rectangular channel (Lin and
were focused on ow visualization and identication of different Lin, 1996; Mahaney et al., 1987; Maughan and Incropera, 1987; Os-
ow patterns emerged due to convection (e.g. Lin and Lin, 1996 borne and Incropera, 1985a, 1985b; Ozsunar et al., 2002) and a
and Wang et al., 1996). The low Reynolds number ow in the horizontal tube (Choi and Choi, 1994) at low Reynolds numbers.
They all observed enhancement of Nusselt number for mixed con-
Corresponding author. Tel.: +1 519 661 2111x88234; fax: +1 519 661 3020. vection compared to forced convection along the channel length
E-mail address: ksiddiqui@eng.uwo.ca (K. Siddiqui).
and attributed this to the secondary ow enhancement which

0142-727X/$ - see front matter 2013 Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.ijheatuidow.2013.12.005
30 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

disrupts the thermal boundary layer. Decay in Nusselt number was important as they provide a better insight into the underlying
observed at the channel entrance region (Lin and Lin, 1996; Mau- physical processes that occurs in such ows. The present study
ghan and Incropera, 1987; Ozsunar et al., 2002) where the forced is focused on experimentally investigating the impact of bottom
convection was the dominant mode as the buoyancy-driven sec- wall heating on the ow structure inside a horizontal square
ondary ow was not developed yet. Osborne and Incropera channel at low Reynolds and high Grashof numbers. A detailed
(1985a,b) experimentally investigated the effect of buoyancy on qualitative as well as quantitative analysis of both mean and
convection heat transfer inside horizontal channels with top and turbulent ow elds have been conducted to obtain better
bottom heated walls for laminar, transient and turbulent ow re- understanding of the fundamental ow processes associated with
gimes. Their main focus was to quantify Nusselt number adjacent the mixed convection channel ow.
to the top and bottom walls. At the top wall, forced convection
was dominant in all ow regimes and the Nusselt number values 2. Experimental setup
were lower than that for the bottom wall. They argued that for
the transient regime at the top wall, the ow laminarization A 7 cm  7 cm square channel was built for the experiments.
due to stably stratied temperature distribution is responsible The channel consists of three sections as shown in Fig. 1(a). The in-
for the decrease in the Nusselt number values. For laminar ow re- let section has a inch diameter inlet followed by a divergent sec-
gime, forced convection was dominant in the top wall region and a tion that transitions into the 7 cm  7 cm square cross section. The
thermally stable boundary layer was formed preventing ascending length of the inlet section was 70 cm. A honeycomb was placed in-
plumes from the bottom wall to penetrate this region. They pro- side the square section to straighten the ow and damp any distur-
posed a correlation to quantify Nusselt number. They concluded bance before entering the test section. The inlet section was made
that the magnitude of heat ux at one wall has no inuence on of aluminum and contains a bleed valve to remove any air trapped
the convection heat transfer at the other wall. It has been reported inside the channel and a pressure gauge to monitor the inlet uid
in the previous studies that an increase in the Grashof number pressure. The inlet section was connected to the test section using
accelerates the onset of the secondary ow and consequently the two aluminum anges. The test section was 150 cm long and has a
mixed convection, while an increase in the Reynolds number de- 7 cm  7 cm square cross section. The top and side walls were
lays the onset of secondary ow and consequently the mixed con- made of inch non-tempered glass for visual access to the channel
vection (Maughan and Incropera, 1987; Osborne and Incropera, and the bottom wall was made of inch aluminum plate. The alu-
1985b; Ozsunar et al., 2002). minum bottom surface was colored in black with a marker to elim-
Flow visualization provides an insight into the nature of the sec- inate any light reection during the experiments.
ondary ow induced by the wall heating in low Reynolds number The test section was supported by two 5 cm high and 1.3 cm
mixed convection. The reported ow visualization studies were fo- thick aluminum plates. Two strip heaters (1500 W 250 V) 1.3 m
cused on the visualization along the channel length (Osborne and in length were installed in parallel directly underneath the bottom
Incropera, 1985a,b; Sakamoto et al., 1999; Toriyama and Ichimiya, aluminum surface 10 cm downstream of the test section entrance.
2010; Wang et al., 1996) and in the cross stream direction (Koiz- The temperature of the bottom wall was controlled by a tempera-
umi and Hosokawa, 1993; Lin and Lin, 1996; Sakamoto et al., ture controller (ZESTA-ZCP513) through a feedback loop from a
1999). Wang et al. (1996) identied four different ow patterns thermocouple embedded in the bottom wall close to the measure-
for mixed convection along the heated test section of a horizontal ment location. For a given controlled temperature, the variation of
square channel with bottom heated wall using shadowgraph tech- the surface temperature along the entire test section length was
nique. Grashof number ranged from 2.8  106 to 2.5  107, Rey- checked and was found to be normally within 1 C i.e. 23% of
nolds ranged from 100 to 1000. They found that the ow the wall temperature, which is reasonable to consider wall temper-
patterns changed with the Reynolds number and Grashof number. ature uniformity.
Based on the ow patterns they argued that the ow passes The end section was 30 cm in length and connected to the
through four different ow regimes along the channel heated sec- downstream end of the test section by two aluminum anges.
tion: laminar forced convection, laminar mixed convection, tran- The end section contains a bleed valve to remove any trapped air
sient mixed convection and turbulent free convection. Lin and in the channel and a pressure gauge to monitor the exit uid pres-
Lin (1996) experimentally investigated the unsteady mixed con- sure. It has a convergent end section with a 1/2 inch diameter exit.
vection for air in a bottom heated horizontal rectangular channel Clean tap water was used as the working uid. As the water was
in the cross stream direction using smoke tracer. The Reynolds being continuously heated through the channel in a closed loop,
numbers in the range from 9 to 186 and Grashof numbers up to the water temperature tended to build up with time. Four barrels,
5  106 were considered. They found that increasing Grashof num- 200 L each, were coupled together in series and used as a water
ber and/or decreasing Reynolds number alter the ow structure reservoir; This reservoir can supply water at room temperature
from periodic into quasiperiodic and even chaotic. throughout the experiment without a need to recycle i.e. water
Nandakumar et al. (1985) investigated the ow structure in the was circulated one time only during a given set of experiments.
cross stream direction for different horizontal channel geometries Due to a constant room temperature, the inlet water temperature
heated from below for Grashof numbers up to 5  105. Longitudi- maintained a constant value of around 24.5 C. Air bubbles present
nal vortex patterns of two or four vortices were observed and the in the tap water had to be removed to obtain good quality results.
bifurcation of the vortices was found to depend on the Grashof Therefore, water was stored in the barrels for 2 days with periodic
number and the channel aspect ratio. Huang and Lin (1994) stirring to remove air bubbles. A magnetic pump (Little Giant, 5
numerically investigated laminar mixed convection in a horizontal MD) installed downstream of the barrels was used to circulate
rectangular duct heated from below. Their main focus was on water through the loop. A ow meter with a control valve
studying the effect of buoyancy-inertia ratio on the cross-stream (FL4205, Omega Engineering) was installed between the pump
ow behavior. They found that with an increase of Gr/Re2, the and the channel to control the water ow rate (see Fig. 1(a)).
cross-stream ow behavior shifts from a steady vortex ow at Four mass ow rates 0.0210, 0.0315, 0.0420 and 0.0525 kg/s
Gr/Re2 < 4 into a chaotic ow at the channel exit at Gr/Re2 > 25. were used in the experimental runs. The Reynolds numbers corre-
Despite several studies, there is a scarcity of detailed investi- spond to these ow rates in the absence of heating are 300, 450,
gation of the turbulent ow structure in mixed convection 600 and 750 for reference. At each ow rate, experiments were
channel ows at low Reynolds numbers. Such studies are conducted at different bottom wall temperatures which were 30,
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 31

Fig. 1. Schematic of (a) the experimental setup, (b) PIV system components and setup, (c) position of thermocouples in the rake; thermocouples 14 were 1 mm apart,
thermocouples 57 were 2 mm apart, thermocouples 78 were 7.5 mm apart, and thermocouples 89 were 17.5 mm apart.

35, 40, 45, 50 and 55 C. The corresponding Grashof numbers ran- horizontal and 1752 pixel in vertical. An image acquisition system
ged from 6.37  106 to 3.86  107. For a given set of experiments, (DVR Express CORE, I.O. industries) connected to a PC was used to
the measurements were taken 30 min after adjusting the ow rate record images. A four-channel pulse generator (555-4C, Berkeley
and the heater, to allow the ow to reach steady state. Nucleonics Corporation) was used to control the laser pulses tim-
Particle image velocimetry (PIV) technique was used for mea- ing and synchronizing them with the camera frames. Water was
suring two-dimensional velocity elds. Fig. 1(b) shows the sche- seeded with silver-coated glass spheres, with the mean diameter
matic of the PIV setup. The measurements were taken 130 cm of 15 lm. These glass spheres were used as the tracer particles
downstream of the test section in the mid vertical plane of the for the PIV measurements. At each experimental run, 3000 images
channel (z/Dh = 0.5). The PIV system comprised of a 120 mJ Nd:YAG were captured at a sampling rate of 30 Hz. This resulted in 1500
laser (SoloPIV 120XT 532 nm) to illuminate the measurement instantaneous velocity elds at a rate of 15 Hz.
plane, and a CCD camera (VA-4M32, Vieworks) with the resolution PIV measuring technique computes velocity vectors by cross
of 2336  1752 pixels to capture the images in the measurement correlating two consecutive images (i.e. an image pair). Interroga-
plane. The camera was horizontally positioned i.e., 2336 pixel in tion window in the rst image is being correlated with the search
32 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

region in the second image of the image pair. In the present study, 5  107 < Ra < 5  108, indicating the presence of turbulent convec-
the interrogation windows size i.e. the region over which each tion. In the present conditions, there are two sources of turbulence
velocity vector was calculated was set as 32  32 pixels and the generation; the buoyancy and the shear ow. The contribution of
search windows was set as 64  64 pixels. The nominal resolution each source to the turbulence generation depends on their relative
of the velocity eld was increased to 16  16 pixels strengths, which is typically quantied in terms of the Richardson
(0.68  0.68 mm) by using a 50% overlap of interrogation windows. number (Ri) which is the ratio between the turbulence generation
Spurious velocity vectors were identied and then corrected by due to buoyancy and the turbulence generation by the shear (Kun-
using a scheme based on the local median test proposed by Siddiq- du and Cohen, 2002). The gradient Richardson number was com-
ui et al. (2001). The spurious vectors detected and corrected were puted using the measured temperature and velocity data as,
well below 1%.
  2
The uncertainty of the PIV velocity measurements was calcu- @q @U
lated at the highest ow rate and highest bottom wall temperature Ri g q 1
@y @y
as the largest velocity gradients occur at these conditions. The
uncertainty was estimated based on the criteria and data from where @@yq is the vertical density gradient and @U
@y
is the vertical gradi-
Cowen and Monismith (1997) and Prasad et al. (1992). The maxi- ent of the mean streamwise velocity (Turner, 1973). A negative va-
mum error in velocity measurements was estimated to be lue of the Richardson number implies the presence of unstable
0.071 cm/s which was less than 6.7% of the bulk ow velocity. stratication (Kundu and Cohen, 2002). The values of Richardson
Temperatures were measured in the same plane and under the number at the lowest and highest ow rates and wall temperatures
same conditions as for the velocity measurements in a separate set are presented in Table 1. The results show that for all cases, the ow
of experiments. A rake of nine T-type thermocouples with accuracy was unstably stratied (Ri < 0) as expected. For a given bottom wall
of 0.5 C was positioned vertically 135 cm downstream of the test temperature, the magnitude of the Richardson number decreased
section entrance. The spacing of the thermocouples was set in a with an increase in the ow rate, while for a given ow rate, the
way that most of the thermocouples were clustered near the bot- magnitude of Richardson number increased with the wall
tom wall to resolve the thermal boundary layer. The exact posi- temperature.
tioning of each thermocouple is show in Fig. 1(c). Inlet and outlet To illustrate the dynamical and complex ow structure due to
water temperatures were also measured using T-type thermocou- the interaction of shear and buoyancy-driven ows, turbulent
ples located at the upstream and downstream ends of the channel, velocity elds at different ow rates and heating conditions are
respectively. A 12 channel data acquisition module (National plotted in Fig. 2. Turbulent velocity elds were computed by sub-
Instruments NI 9211) was used to acquire the temperature data tracting the time-averaged velocity from the instantaneous veloc-
via LabVIEW data acquisition software. The data were recorded ity elds at each grid point. Fig. 2(a) shows the turbulent velocity
for 5 min at a rate of 3 Hz. The sampling rate of thermocouples eld for m _ 0:0315 kg/s at the lowest heated wall temperature
was set based on the response time of the thermocouples. Thermo- (T = 30 C). The plot shows the rising plumes of warm less dense
couples were calibrated using a mercury-bulb thermometer. uid parcels near the heated wall. The falling parcels of cooler
The repeatability of the measurements was checked in different dense uid are also visible in the plot in the mid channel region.
ways. In the rst approach, at a given bottom wall temperature, the The plot shows the interaction of rising and falling parcels of uid
ow rate was changed in the ascending order and then descending which formed complex ow structure. The plot also shows that
order. The results at the same ow rate were found comparable. multiple vortices are formed due to this interaction with clockwise
Furthermore, few experimental runs were repeated on another as well as counter clockwise orientations. It is also observed that at
day and the results were also comparable. this condition, the magnitude of turbulent velocities is very weak
in the upper half of the channel. This is likely due to the reason that
the thermal plumes are relatively weak due to the low heating con-
3. Results dition. As a result, most of the turbulent kinetic energy of these
plumes is utilized in the interactions discussed above and they be-
In mixed convection where both forced and natural convection come very weak as they reach the upper region of the channel.
modes are present, one parameter that quanties the relative con- Fig. 2(b) (m_ 0:0315 kg/s at T = 40 C) shows that as the wall
tributions of buoyancy versus inertial forces is Gr/Re2. The natural heating increases, the turbulence becomes stronger in magnitude
convection mode is considered to be dominant if Gr/Re2  1, while and spreads throughout the channel domain. An interesting obser-
the forced convection mode is dominant if Gr/Re2  1 (Incropera vation in the plot is a prominent source-like ow at the mid chan-
et al., 2006). In the present study, the Grashof number was calcu- nel height i.e., ow is emerging perpendicularly and diffusing into
lated based on the surface and bulk uid temperature difference the measurement plane. It implies the presence of a strong cross-
and using the hydraulic diameter of the channel (Dh) as the charac- stream component of turbulent velocity (perpendicular to the
teristic length scale, while for Reynolds number calculation, the measurement plane). This behavior manifests the existence of
average channel velocity and channel hydraulic diameter were three-dimensional turbulent ow within the channel. The plot also
used as velocity and length scales, respectively (Mahaney et al., shows that the magnitude of cross-stream velocity component is
1987). The results show that the Gr/Re2 ranged from about 10 at strong enough to diverge the falling parcels. The turbulent velocity
the highest ow rate and lowest wall temperature to over 200 at eld in this plot appears to be dominated by the falling parcels
the lowest ow rate and highest wall temperature, indicating that
in the present study, natural convection mode was dominant at all
cases. Table 1
Various parameters for the maximum and minimum wall temperatures and ow
The bottom wall heating induces buoyancy-driven secondary
rates.
ow which drives natural convection. This ow is in the form of
rising plumes of warm uid from the bottom wall and falling par- Mass ow rate (kg/s) 0.0210 0.0525 0.0210 0.0525

cels of cooler uid from the top unheated wall. For natural convec- Wall temperature (C) 30 30 55 55
tion mode, the ow becomes turbulent when Rayleigh number (Ra) Grashof number 6.37  106 6.61  106 3.86  107 3.45  107
Gr/Re2 53 9 206 37
becomes greater than 5  104 (Incropera et al., 2006; Turner,
Richardson number 0.43 0.08 0.84 0.13
1973). For the present study, the Rayleigh number is in the range
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 33

the vortex formation at multiple heights as the wall temperature


(a) 1 0.5 cm/s
increases.
The plots in Fig. 2 show that the wall heating induces strong
0.75
buoyancy-driven ow and the interaction of rising and falling uid
y/D h

parcels forms a complex and dynamic three-dimensional ow


0.5
structure. The plots also show that the bulk movement of the rising
and falling uid parcels is not organized, however, their interac-
0.25
tions locally generate organized patterns such as vortices. It is also
0
observed that at relatively low wall heating conditions, the strong
0 2 4 6 8 10 turbulent motions are mainly restricted near the heated wall, how-
x (cm) ever, as the wall heat ux increases, the turbulence becomes stron-
ger and its inuence reaches the upper unheated wall. The uid
(b) 1 0.5 cm/s parcels lose heat as they approach the upper wall, become denser
and fall towards the bottom wall. The results also indicate that the
0.75 buoyancy-driven ow also induced three-dimensional ow pat-
terns within the domain.
y/D h

0.5 The mean velocity eld for each case was obtained by time-
averaging the instantaneous velocities at each grid point. That is,
0.25 the mean velocity eld at each grid point (r, c) for both velocity
components is computed as,
0
0 2 4 6 8 10 PN
i1 ur;c;i
x (cm) U r;c 2
N
(c) 1 0.5 cm/s
where U r;c is the mean velocity (streamwise or vertical) at the grid
point (r, c), r is the row and c is the column, ur;c;i is the instanta-
0.75
neous velocity in each velocity eld at the same grid point, and N
is the total number of instantaneous velocity elds.
y/D h

To illustrate the overall structure of the mean velocity within


0.5
the measurement domain, the contours of the mean streamwise
velocity eld are presented in Fig. 3 at the lowest and highest ow
0.25
rates and wall temperatures. The white regions appear in the plots
correspond to the regions of bad data due to high noise in the PIV
0
0 2 4 6 8 10 images. The data in those regions were assigned as NaNs (Not-a-
x (cm) number) and hence excluded from the analysis as they could bias
the results. The gure shows that at the lowest ow rate and low-
(d) 1 0.5 cm/s est wall temperature (Fig. 3(a)), the mean streamwise velocity
magnitude increased sharply with the distance and then gradually
0.75 decreased to zero at the upper wall. However, as the wall temper-
ature increased to the maximum value (T = 55 C), the maximum
h

mean velocity remained almost at the same location but the mean
y/D

0.5
velocity became negative in the upper quarter of the channel i.e.
0.25 the mean ow direction in this region was opposite to the bulk
channel ow (Fig. 3(b)). The mean velocity magnitude in the lower
0 section of the channel was higher at the highest wall temperature.
0 2 4 6 8 10 At the maximum ow rate, the trends and magnitudes of the mean
x (cm) velocity are quite similar at both low and high wall temperatures
_ 0:0315 kg/s,
(Fig. 3(c and d)). Comparison of the plots at the lowest and highest
Fig. 2. Snapshots of turbulent velocity vector elds at, (a) m
T = 30 C, (b) m_ 0:0315 kg/s, T = 40 C, (c) m
_ 0:042 kg/s, T = 45 C, (d) ow rates shows that at the lowest wall temperature, the ow
m_ 0:0315 kg/s, T = 55 C. structure is quite similar at both ow rates although the velocity
magnitude is signicantly higher at the highest ow rate, as ex-
pected. However, the ow structure at the highest temperature is
within the measurement domain, however, a rising plume is ob-
quite different. It is observed that the negative velocity trends
served at the bottom left side of the velocity eld. The generation
present at the lowest ow rate diminished and the mean velocity
of a vortex due to the interaction of falling and rising uid parcels
became unidirectional at the highest ow rate.
is also visible.
The mean velocity proles were computed by spatially averag-
_ 0:042 kg/s at T = 45 C) shows a
Velocity eld in Fig. 2(c) (m
ing the time-averaged velocities at each height. That is, U r;c was
different ow structure. The plot shows the formation of multiple
averaged in space at each height (i.e. row, r) as follows,
vortices immediately above the bottom heated wall due to the
interaction of rising and falling parcels of uid. A strong rising PN c
c1 U r;c
plume in the middle section is prominent in the plot. The plume Ur 3
Nc
reached the top wall although it was diverted in the mid-height re-
gion by the strong cross-plane ow dispersing in the measurement where U r is the mean velocity at a given height, and Nc is the num-
plane. Strong sweeping ow is also visible in the upper right and ber of columns in the mean velocity eld. The mean streamwise
bottom left sections of the plot. The turbulent velocity eld in velocity proles normalized by the domain-averaged streamwise
Fig. 2(d) (m_ 0:0315 kg/s at T = 55 C) further highlights the com- velocity are presented in Fig. 4 for all ow rates and wall
plex interaction of ow throughout the measurement domain and temperature conditions. In the classical Poiseuille ow, the mean
34 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

(a) (a)

(b)
(b)

(c)
(c)

(d)
1
(d)
T = 30 C
y/Dh

T = 35 C
0.5 T = 40 C
T = 45 C
T = 50 C
T = 55 C

0
Fig. 3. Contours of mean streamwise velocity at (a) m _ 0:021 kg/s, T = 30 C (b) 0 0.5 1 1.5
m_ 0:021 kg/s, T = 55 C (c) m
_ 0:052 kg/s, T = 30 C (d) m
_ 0:052 kg/s, T = 55 C.
U/Uc
The colorbar is in cm/s.

Fig. 4. Normalized mean streamwise velocity proles for various wall temperatures
at (a) m_ 0:021 kg/s (b) m
_ 0:0315 kg/s (c) m
_ 0:042 kg/s (d) m
_ 0:0525 kg/s.

streamwise velocity exhibits a parabolic prole. However, the plots


in Fig. 3 and 4 clearly demonstrate that bottom heating altered the
the top unheated wall (Choi and Choi, 1994; Toriyama and Ich-
conventional parabolic velocity proles for all cases. At the three
imiya, 2010; Wang et al., 1994). Wang et al. (1994) reported a
higher ow rates, the mean velocity proles became asymmetric
mean velocity prole similar to that observed in Fig. 4 in the pres-
and skewed towards the bottom heating wall. The location of the
ence of the back ow. The issue of the back ow will be discussed
maximum mean velocity was around y/D  0.1 and was consistent
in detail in the discussion section.
for all cases.
Please note that in a channel ow in the absence of heating, the
The mean streamwise velocity trends at the lowest ow rate
mean ow is in the streamwise direction only. However, in the
(Fig. 4(a)) are signicantly different from that at the higher ow
presence of heating, the buoyancy-driven secondary ow intro-
rates particularly at the higher heating conditions. The plot shows
duces bulk movement in the other directions as well. That is, un-
that at the lowest wall temperature (T = 30 C), the mean stream-
like the channel ow in the absence of heating, the channel ow
wise velocity behavior is similar to that at the higher ow rates.
in the presence of heating contains mean velocities in the vertical
However, as the wall temperature increased, reverse ow observed
and spanwise directions as well. The magnitude of these velocities
near the top (unheated) wall region. The magnitude and extent of
changes with the heat ux at a given ow rate. Therefore, some
the reverse ow increased with an increase in the wall tempera-
variations observed in the mean streamwise velocity proles at
ture. Some previous studies have also reported back ow near
the same ow rate but different heating conditions are due to
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 35

the redistribution of mean resultant velocity into different


(a) 1
components.
The turbulent properties are typically normalized by the char- 0.75
0.4

acteristic velocity scale which however is dened differently for


0.3

y/D h
forced and natural convection. For forced convection, friction 0.5
velocity (u) is used as the characteristic velocity scale which is de- 0.2
ned as, 0.25
s
 0.1
dU  0
u m  4 0 2 4 6 8 10
dy y0
x (cm)
where v is the kinematic viscosity and dU is the mean streamwise
dy
vertical velocity gradient at the wall (Pope, 2000). For natural con-
(b) 1
vection, a convective velocity scale, w is used by Deardorff (1970) 0.4
and Adrian et al. (1986), which is dened as, 0.75

0.3

y/D h
q00s 0.5
w bgQ o y 1=3 ; Qo 5
qc p 0.2
0.25
where b is the thermal coefcient of expansion, g is the gravita- 0.1
tional acceleration, Qo is the kinematic heat ux, y is the length 0
scale equal to the channel height, q00s is the surface heat ux, q is 0 2 4 6 8 10
the density and cp is the specic heat (Adrian et al., 1986). x (cm)
None of these velocity scales can be used in the present study as
both forced and natural convection co-exist and their relative con- (c) 1

tribution changed with the change in the ow rate or the wall tem- 0.75
0.4
perature. That is, the forced convection contribution increased y/D h 0.3
with the ow rate while the natural convection contribution in- 0.5
creased with the wall temperature. As both forced and natural con- 0.2
vection modes are present in the mixed convection, the velocity 0.25
scale should account for both of these modes. Velocity scaling 0.1
accounting the contributions of both w and u has been proposed 0
0 2 4 6 8 10
in the eld of atmospheric sciences for mixed convection (Zeman
x (cm)
and Tennekes (1977), Driedonks (1982) and Moeng and Sullivan
(1994). The basic form of this velocity scale was used which is de-
1
ned as, (d)
q 0.4
0.75
v w2 u2 6
0.3
y/D h

0.5
The values of v for all cases are presented in Table 2. The results
0.2
show that the inuence of the wall temperature was more pro-
0.25
found on v compared to the ow rate. 0.1
The contours of Root-Mean-Square (RMS) streamwise turbulent
0
velocity at the lowest and highest ow rates and wall temperatures 0 2 4 6 8 10
are shown in Fig. 5. The results show that at the lowest ow rate x (cm)
and wall temperature, the streamwise turbulent velocity magni-
Fig. 5. Contours of RMS streamwise turbulent velocity at (a) m _ 0:021 kg/s,
tude is low (Fig. 5(a)). Relatively strong magnitudes are observed _ 0:021 kg/s at T = 55 C (c) m _ 0:052 kg/s at T = 30 C (d)
T = 30 C (b) m
in the middle of the channel and immediately above the heated m_ 0:052 kg/s at T = 55 C. The colorbar is in cm/s.
wall. The velocity magnitude decreased towards the upper un-
heated wall. As the ow rate increased while the wall temperature plots show that the wall temperature has a signicant impact on
remained the same, an overall increase in the streamwise turbu- the streamwise turbulent intensity in the channel. The peak turbu-
lent velocity was observed (Fig. 5(c)). The largest magnitudes of lent velocity in the channel increased by almost 100% and 70% at
turbulent velocity were found adjacent to the bottom heated wall the lowest and highest ow rates as the wall temperature in-
and the upper region of the channel core. The velocity magnitudes creased from 30 C to 55 C. The overall magnitude of streamwise
decreased sharply when approaching the upper unheated wall. turbulent velocity was increased by about 130% and 80%, respec-
Fig. 5(b and d) show the velocity contours at the highest wall tem- tively. Although the turbulent velocity magnitude increased with
perature at both lowest and highest ow rates, respectively. Both the wall temperature, the overall pattern of the streamwise turbu-
lent velocity distribution remained almost the same for all cases.
Table 2
That is, very strong streamwise turbulent velocity magnitude
Values of v in cm/s for all cases. was observed immediately above the heated bottom wall, which
decreased over a shorter distance and then increased to a maxi-
Mass ow rate (kg/s) Wall temperature (C)
mum magnitude in the channel core and then decreased sharply
30 35 40 45 50 55 to an almost negligible magnitude at the upper unheated wall.
0.021 0.47 0.61 0.67 0.74 0.8 0.83 The comparison also shows that at a given wall temperature, an in-
0.0315 0.49 0.58 0.71 0.76 0.83 0.89 crease in the ow rate did not signicantly inuence the turbulent
0.0420 0.0495 0.6 0.73 0.8 0.85 0.94
velocity magnitude but contributed to the diffusion of strong tur-
0.0525 0.51 0.65 0.72 0.785 0.86 0.92
bulence over a wider domain.
36 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

The RMS streamwise turbulent velocity proles normalized by channel core, which is due to the interaction of the buoyancy-dri-
v are presented in Fig. 6 for all ow rates and wall temperature ven ow with itself (i.e. the interaction of the rising plumes and
conditions. Figure shows that the pattern of streamwise turbulent falling uid parcels as shown in Fig. 2) and with the streamwise
velocity at all conditions are in general similar. That is, strong channel shear ow.
streamwise turbulent intensities in the region immediately above The contour plots of the vertical turbulent velocity elds are
the heated wall and in the channel core and almost negligible mag- plotted in Fig. 7 for the four extreme cases (maximum and mini-
nitudes at the upper unheated wall. The results however show that mum ow rates and wall temperatures). The structure of the ver-
the location of the enhanced turbulence in the channel core shifted tical turbulent velocity is quite similar for all these cases, i.e., the
with the ow rate. At the lowest ow rate, the enhanced turbu- vertical turbulent velocity magnitude increased sharply at the bot-
lence region was in the middle of the channel which shifted to- tom heated wall to a peak value and then gradually decreased to
wards the upper wall with an increase in the ow rate. In zero at the upper unheated wall. The strong magnitudes of vertical
classical channel ows in the absence of heating, the streamwise turbulent velocity are observed in the region 0.2 < y/Dh < 0.7, indi-
turbulent velocity magnitude peaks near the wall and then de- cating that the vertical turbulent velocities are signicant over al-
creases towards the channel core. Present results show that in most half of the channel height. The plots also show that the
the presence of wall heating, strong turbulence is observed in the inuence of the bottom wall temperature on the vertical turbulent
velocity magnitudes is more profound than the effect of the ow
rate. That is, an increase in the bottom wall temperature from
(a) 1 T = 30 C
30 C to 55 C enhanced the vertical turbulent velocity magnitude
T = 35 C
T = 40 C
T = 45 C
(a) 1 0.5
y/D h

T = 50 C
0.5 T = 55 C
0.4
0.75

0.3

y/D h
0.5
0.2
0
0.25
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0.1
urms /v*
0
0 2 4 6 8 10
x (cm)
(b) 1 T = 30 C
T = 35 C
T = 40 C (b) 1 0.5
T = 45 C
0.4
y/D h

T = 50 C
0.75
0.5 T = 55 C

0.3
y/D h

0.5
0.2
0.25
0 0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0
0 2 4 6 8 10
(c) 1 T = 30 C
x (cm)

T = 35 C
T = 40 C
T = 45 C
(c) 1 0.5

0.4
y/D h

T = 50 C
0.75
0.5 T = 55 C

0.3
y/D h

0.5
0.2
0.25
0 0.1
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
0
urms /v* 0 2 4 6 8 10
x (cm)

(d) 1 T = 30 C
T = 35 C
(d) 1 0.5

T = 40 C
0.4
T = 45 C 0.75
y/D h

T = 50 C
0.5 T = 55 C 0.3
y/D h

0.5
0.2
0.25
0.1
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0 2 4 6 8 10
urms /v* x (cm)

Fig. 6. Normalized RMS streamwise turbulent velocity proles for various wall Fig. 7. Contours of RMS vertical turbulent velocity at (a) m _ 0:021 kg/s, T = 30 C
_ 0:021 kg/s (b) m
temperatures at (a) m _ 0:0315 kg/s (c) m
_ 0:042 kg/s (d) (b) m_ 0:021 kg/s at T = 55 C (c) m
_ 0:052 kg/s at T = 30 C (d) m
_ 0:052 kg/s at
m_ 0:0525 kg/s. T = 55 C. The colorbar is in cm/s.
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 37

by a factor of approximately 5 and 4 for the lowest and highest adjacent to both heated and unheated walls which became nega-
ow rates, respectively. Whereas, an increase in the ow rate from tive in the middle region of the channel. The comparison of Rey-
0.021 to 0.0525 kg/s, increased the vertical turbulent velocity by nolds stress magnitude near both walls shows that the
approximately 60% and 25% at the lowest and highest bottom wall magnitude of positive Reynolds stress is relatively strong near
temperatures, respectively. It is also observed that the strongest the heated wall. It is also found that the relative strength of Rey-
vertical turbulent velocities are present in the lower half of the nolds stress near the heated wall compared to the unheated wall
channel. The normalized RMS vertical turbulent velocity proles increased with an increase in the wall temperature as well as the
are presented in Fig. 8 for all ow rates and wall temperature con- ow rate, however, the enhancement is more profound due to wall
ditions. The behavior for all ow rates and heating conditions is heating. It is also observed that at both ow rates, the magnitudes
relatively comparable where strong asymmetry is evident in all of positive and negative Reynolds stresses are quite comparable at
velocity proles. The velocity magnitude increased sharply from the low wall temperature while, the negative Reynolds stress mag-
the bottom surface up to a height of y/Dh  0.2 which is followed nitude was almost three times larger than that of the positive Rey-
by a decrease towards the upper wall. nolds stress at the high wall temperature. The normalized
The Reynolds stress (u0 v 0 ) contours are plotted in Fig. 9 for the Reynolds stress proles are presented in Fig. 10 for all ow rates
four extreme cases. The plots show positive Reynolds stress u0 v 0 and wall temperature conditions. The plots show similar overall
trend for all cases however, as the ow rate increased, the negative
Reynolds stress tends to become more uniform in the channel core.
(a) 1
T = 30 C
T = 35 C
T = 40 C
T = 45 C
T = 50 C (a)
y/D h

0.5 T = 55 C

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
vrms /v*

(b) 1
T = 30 C
T = 35 C
T = 40 C
T = 45 C (b)
T = 50 C
y/D h

0.5 T = 55 C

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
vrms /v*

(c) 1
T = 30 C
T = 35 C
T = 40 C (c)
T = 45 C
T = 50 C
y/D h

0.5 T = 55 C

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
vrms /v*

1
(d) T = 30 C
(d)
T = 35 C
T = 40 C
T = 45 C
T = 50 C
y/D h

0.5 T = 55 C

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
vrms /v*

Fig. 8. Normalized RMS vertical turbulent velocity proles for various wall Fig. 9. Contours of Reynolds stress at (a) m_ 0:021 kg/s at T = 30 C (b) m
_ 0:021
_ 0:021 kg/s (b) m
temperatures at (a) m _ 0:0315 kg/s (c) m
_ 0:042 kg/s (d) _ 0:052 kg/s at T = 30 C (d) m
kg/s at T = 55 C (c) m _ 0:052 kg/s at T = 55 C. The
m_ 0:0525 kg/s. colorbar is in cm2/s2.
38 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

Overall the region of negative Reynolds stress remained almost the signicantly from the classical Poiseuille trend, indicating that at
same at different ow rates and wall temperatures. Gajusingh and low Reynolds numbers and high Grashof numbers, the bottom wall
Siddiqui (2008) also reported negative Reynolds stress away from heating signicantly alters the mean velocity structure of the chan-
the heated wall in a channel ow at low ow rates that fall in nel ow. The mean velocity peak shifted towards the bottom
the laminar regime in the absence of heating. At higher ow rates heated wall at all cases. One plausible explanation for the increase
in the turbulent regime, they observed classical Reynolds stress in the mean velocity near the heated wall could be the decrease in
trends. viscosity of water due to higher temperatures near the heated wall.
This decrease in viscosity causes a reduction in the ow resistance
and hence, higher velocity. This increase in ow near the bottom
4. Discussion
wall automatically causes a decrease in the mean velocity towards
the upper unheated wall to satisfy mass conservation.
The results presented in the preceding section provide a quali-
The results also show a bulk back ow in the upper region of the
tative and quantitative insight into a channel ow subjected to
channel. This back ow is observed at the lowest ow rate, and its
heating from the bottom wall at low ow rates that falls within
magnitude increased with an increase in the wall temperature, and
the laminar regime in the absence of heating. The mean velocity
also at some higher ow rates at higher wall temperatures. It is ob-
results show that the mean streamwise velocity proles deviated
served that the back ow occurred when the Gr/Re2 was greater
than about 55 and the magnitude of the back ow increased in
(a) general with an increase in Gr/Re2. The contour plots in Fig. 3(b
and d) show almost horizontal contour lines associated with the
back ow indicating that the mean back ow observed at the mea-
surement location is not a local phenomenon. As a part of a sepa-
rate study in the same channel, we measured the ow elds at
different locations along the test section at different ow rates
and bottom wall temperatures and observed that in cases where
the back ow phenomenon is present, the back ow extended over
the entire heated section of the channel. In other words, the for-
ward ow is found along the bottom heated wall while the back
ow is found along the top unheated wall. This ow behavior indi-
cated the manifestation of a large convective cell that covered the
(b) entire heated section of the channel. The intensity of this channel-
scale cell increased with an increase in the bottom wall tempera-
ture. A thorough analysis of the PIV images and the corresponding
instantaneous velocity elds from the present dataset showed that
the local instabilities in the form of rising plumes and falling uid
parcels continuously interacted with the cell and the energy trans-
fer due to this interaction is likely the source that provided suste-
nance to this convective cell. Fig. 11 shows the snapshots of
instantaneous velocity elds depicting these interactions.
Reverse ow was reported in many studies that investigated
laminar mixed convection channel ows (Choi and Choi, 1994;
Sakamoto et al., 1999; Toriyama and Ichimiya, 2010; Wang et al.,
(c) 1994). Wang et al. (1994) numerically studied mixed convection
in a horizontal pipe for air at Re = 100 and Gr = 2  104. They ob-
served reverse ow near the top wall at the entrance region, which
distorted the mean velocity behavior. However, the reverse ow
diminished with the distance from the entrance region. They ar-
gued that the secondary ow development and the axial conduc-
tion are responsible for the reverse ow. Toriyama and Ichimiya
(2010) numerically and experimentally investigated laminar
mixed convection in a horizontal square channel for two condi-
tions considering isothermal walls. For the rst condition, wall
temperature was higher than the inlet water temperature while
for the second case, wall temperature was lower than the inlet
(d) water temperature. A reverse ow appeared at the top wall for
the rst case while for the second case, the reverse ow appeared
at the bottom wall. As shown above several studies reported the
backow in the channel, however, only few of them provided a
plausible explanation of this trend and attributed it to the positive
streamwise pressure gradient caused by the vertical motion of the
rising plumes and falling uid parcels (Choi and Choi, 1994;
Sakamoto et al., 1999).
The streamwise and vertical turbulent velocity results show
trends which are signicantly different from the conventional un-
heated channel ows. It is observed that both turbulent velocity
Fig. 10. Normalized Reynolds stress proles for various wall temperatures at (a) components have large amplitudes away from the wall. The
m_ 0:021 kg/s (b) m
_ 0:0315 kg/s (c) m
_ 0:042 kg/s (d) m
_ 0:0525 kg/s. proles of the vertical turbulent velocities were found to be more
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 39

layer similar to the conventional channel ows. However, unlike


(a) 1 0.5 cm/s
conventional channel ows, the Reynolds stress become negative
in the channel core and again become positive as approaching
0.75
the upper unheated wall. The region of the negative Reynolds
y/D h

stress approximately coincides with the region of strong turbulent


0.5
velocity magnitudes. This indicates that in this region, the interac-
0.25 tion of buoyancy induced rising plumes and falling parcels of cool-
er uid with the shear ow resulted in the strong turbulent
0 motions but at the same time these interactions caused a rapid en-
0 2 4 6 8 10
ergy and hence the momentum transfer from the energetic turbu-
x (cm)
lent eld.
The positive Reynolds stress is a measure of the momentum
(b) 1 0.5 cm/s
transfer by the turbulent velocity eld (Pope, 2000). Turner
(1973) argued that in the presence of buoyancy, the turbulent en-
0.75 ergy transfer is more rapid due to the stronger interaction of turbu-
lent and buoyancy elds and proposed a buoyancy subrange
y/D h

0.5 within the inertial subrange, where the energy transfer rate is fas-
ter (3 slope) than that in the classical inertial subrange (5/3
0.25
slope). The spectral analysis of our data (not shown here) indicated
that the spectra become steeper than 5/3 in the region of the neg-
0
0 2 4 6 8 10 ative Reynolds stress. Hence, it can be argued that the negative
x (cm) Reynolds stress implies the momentum transfer from the turbulent
velocity eld to the buoyancy eld. The plots in Fig. 9 also show
_ 0:021 kg/s and (a)
Fig. 11. Snapshots of instantaneous velocity vector elds at m
that the magnitude of negative Reynolds stress increased with
T = 50 C (b) T = 55 C.
the wall temperature but was almost independent of the ow rate.
This further conrms that the presence of the negative Reynolds
stress is due to the buoyancy eld.
organized with the peak magnitudes located in the region The results presented and discussed earlier have shown that the
0.2 < y/Dh < 0.4 for all cases, while, the streamwise turbulent structure of the turbulent velocity elds is very complex due to the
velocity proles were less organized. The streamwise turbulent three-dimensional interactions of buoyancy-induced secondary
velocities showed two peaks; the strongest peak located immedi- ow among themselves and with the shear ow. The vector plots
ately above the heated wall, whereas the secondary peak located in Fig. 2 and various contour maps of turbulent properties provide
in the middle of the channel at the low ow rate which shifted a general overview of this structure. However, in order to get a dee-
to the upper half of the channel with an increase in the ow rate. per insight into the underlying physical mechanisms that lead to
The streamwise turbulent velocity was in general found to be this turbulent ow, we performed the proper orthogonal decompo-
relatively uniform in the channel core. sition (POD) analysis of the turbulent velocity elds.
The rising plumes are more intense close to the bottom heated
wall and start to weaken with height due to decrease in the density 4.1. Proper orthogonal decomposition
uctuations and an increase in viscosity (Turner, 1973). Similarly,
the falling parcels of cooler uid accelerate as a result of the grav- The spatial distribution of ow energy content at different
itational force. As depicted in Fig. 2, the interaction between the orthogonal modes can be obtained by POD analysis. It allows deter-
rising plumes and the falling parcels is most intense in the mid mining the spatial scales and the associated energy of dominant
and lower sections of the channel. This is the region where the ver- turbulent ow structures. The orthogonal modes or basis functions
tical turbulent velocity magnitudes are relatively large. As dis- /n 
x are derived by decomposing the turbulent velocity elds.
cussed earlier, a shear ow is also superimposed on the These modes individually or in combination describe different tur-
buoyancy-driven secondary ow (plumes and falling parcels). This bulent ow energy patterns. The snapshot method proposed by
interaction of the shear ow with the secondary ow is expected to Sirovich (1987) was used in the current work with the planar PIV
manifest more profoundly in the streamwise turbulent velocity. As turbulent velocity data using the algorithm developed by Dod-
the shear ow is dominant in the bulk of the channel, it interacts dipatla (2010). In this algorithm, the velocity is expanded into a
with the rising plumes and the falling uid parcels throughout this sum of spatial and temporal components as (Holmes et al., 1996):
domain, Furthermore, this shear ow distorts the vertical motions,
i.e. the mean shear sweeps the secondary ow. This sweeping X
N
 x; t
u an t/n x 7
behavior occurred in the channel core region away from the bot- n1
tom and top walls. These interactions would likely be the reason
for strong and relatively uniform distribution of the streamwise where N is the number of snapshots equal to the number of turbu-
turbulent velocities in this domain as seen in Fig. 6. The results also lent velocity elds, and an(t) is the temporal coefcient.
show that the contribution of the top unheated wall to the stream- The basis functions /n 
x are obtained in such a way that the
wise turbulent velocity is almost negligible as compared to that minimum error between the original velocity and the recon-
contributed by the bottom wall heating. structed velocity eld square is minimal (Cizmas et al., 2003) as
The contours and proles of Reynolds stress in general show a follows,
different behavior than that in conventional unheated turbulent
channel ows. In conventional turbulent channel ows, the Rey- X
N

nolds stress increases sharply to a peak value in the inner layer  x; t 
jju an t/n xjj2 ! min 8
n1
and then decreases to zero in the channel core. The present results
show that in the region immediately adjacent to the heated wall, To achieve the minimum error, Eq. (8) is reduced into an eigen-
the Reynolds stress increases sharply to a peak value in the inner value problem (Holmes et al., 1996):
40 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

Z
hu   x0 ; ti  /n x0 dx0 kn /n x
 x; tu 9

where kn is the eigenvalue represent the energy associated with


mode n, hu   x0 ; ti is the spatial autocorrelation matrix at zero
 x; tu
time lag, and denoted the complex conjugate. The eigenvectors are
then used to determine different POD modes by projecting the
velocity eld onto them.
The POD modes are orthogonal and the temporal coefcients
an t are obtained using the POD modes by projecting the velocity
eld onto the POD modes (Cizmas et al., 2003; Maurel et al., 2001;
Holmes et al., 1996; Smith et al., 2005):
Z
 x; t  /n xdx Fig. 13. POD fractional energy (kn/E) distribution across all modes.
an t u 10

In order to validate the algorithm used for the POD analysis, a Fig. 13). At the lowest wall temperature and the ow rate, about
comparison was conducted between the original turbulent velocity 85% of the fractional mode energy resides in the rst 25 modes,
elds obtained from PIV data and the turbulent velocity elds while, at the highest wall temperature and the ow rate, 75% of
reconstructed by combining all POD modes with their temporal the fractional mode energy resides in the rst 25 modes. These re-
coefcients. Fig. 12 shows this comparison. As can be seen from sults indicate that an increase in both the buoyancy-induced as
the plots, both velocity elds are almost identical qualitatively well as shear ows redistributes the ow energy towards the high-
and quantitatively, thus, validating the POD scheme used in this er modes (see Fig. 13).
study. The detailed analysis of the streamwise and vertical turbulent
The fractional energy distribution over the POD modes is pre- velocities showed that the ow structure for all cases is quite sim-
sented in Fig. 13 for the highest and lowest bottom wall tempera- ilar at different modes. This indicates that the underlying mecha-
tures and ow rates. The gure shows that the fractional energy for nisms associated with the generation of the secondary ow and
all cases decreases with an increase in the POD mode, as expected. its interactions are similar however, the energy associated with
The plot also shows that up to mode 25, the fractional mode energy these mechanisms changes with the Grashof and Reynolds num-
remains approximately the same over the given range of ow rates bers. For example, it is found that the overall ow energy at all
and wall temperatures. However, the mode energy starts to vary modes increases with an increase in the wall temperature at the gi-
with the ow rate and wall temperature beyond mode 25. It is ob- ven ow rate, and increases with an increase in the ow rate at the
served that beyond mode 25, at a given ow rate, the fractional given wall temperature.
mode energy increases with an increase in the bottom wall tem- The detailed analysis of the vertical turbulent velocity at differ-
perature; and at a given wall temperature, the fractional mode en- ent modes revealed some intriguing features which provides a bet-
ergy increases with the ow rate. The increase in the magnitude of ter understanding of the overall ow dynamics. The contour plots
the fractional mode energy with temperature at a given ow rate is of the vertical turbulent ow energy at various modes are pre-
possibly due to the increase in the strength of buoyancy-driven sented in Fig. 14 for the ow rate of 0.042 kg/s and wall tempera-
secondary ow, whereas, the increase in the magnitude of frac- ture of 45 C. Regert et al. (2005) found that there is a physical link
tional mode energy with the ow rate at a given temperature is between the POD structures at certain modes and the real ow
likely due to the increase in the shear produced turbulence (see Ta- structure. They further argued that the mathematical structures
ble 1). The low fractional energy at higher modes implies that most within the POD modes that do not exist in the real ow are associ-
of the ow energy resides in lower modes (see modes 814 in ated with sudden drop in the mode fractional energy. Hence, the
ow structure presented in Fig. 14 can be related to the real phys-
ical behavior of the ow. The results at mode 2, which represents
(a) 1 0.5 cm/s
almost an overall ow behavior, showed two alternate positive
0.75 and negative contours. The positive contours represent the upward
motion while the negative contours represent the downward mo-
h
y/D

0.5 tion. This alternate upward and downward motion is hereinafter


referred to as a convective cell. Thus, the result at mode 2 showed
0.25 the presence of a large convective cell, which brings the warm
lighter uid from the bottom heated wall towards the upper un-
0
0 2 4 6 8 10 heated wall and brings the cool and dense uid from the upper
x (cm) end towards the bottom. As the mode number increased, the cells
become smaller in size and strength. As the turbulent ow is com-
prised of the superposition of all modes, the results in Fig. 14 show
(b) 1 0.5 cm/s
that buoyancy-driven ow induced these convective cells of differ-
0.75
ent scales and intensities. In general, these cells are more intense
near the heated wall and hence their inuence on the uid trans-
y/D h

0.5 portation is more dominant in the lower half of the channel. The
presence of these intense convective cells near the heated wall is
0.25 likely the reason for the enhanced vertical turbulent velocity in this
region as observed in Fig. 8. The contour plots of the streamwise
0 turbulent velocity (not shown here) do not show similar cell-type
0 2 4 6 8 10
features but rather a relatively uniform distribution of positive and
x (cm)
negative contours in the bulk of the channel domain. This could be
_ 0:0315 kg/s and T = 45 C, (a) original
Fig. 12. Turbulent velocity vector eld at m the reason for the relatively uniform magnitudes of streamwise
(b) reconstructed POD energy vector eld. turbulent velocity in the bulk channel domain as shown in Fig. 5.
A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942 41

(a) (b)

(c) (d)

(e) (f)

_ 0:042 kg/s and T = 45 C for (a) mode 2 (b) mode 4 (c) mode 9 (d) mode 18 (e) mode 25 (f) mode 30.
Fig. 14. Contours of vertical turbulent velocity at m

The results presented in this paper provided a deeper insight its magnitude increased with an increase in the wall temperature.
into the mean and turbulent ow structure in a horizontal channel The proles of RMS streamwise turbulent velocity showed peak
heated from below at low ow rates. Based on these results, the magnitude immediately above the heated wall and relatively high
overall physical processes associated with this type of ow can magnitudes in the channel core. It was found that increasing the
be described as follows. When heat is added from the bottom wall, shear ow diffuses the streamwise turbulence intensity in the
the mean streamwise velocity structure is modied and the veloc- channel core region. The vertical turbulent velocity structure was
ity peak shifts towards the bottom wall due to the lower ow resis- found to be similar for all cases and their magnitudes were rela-
tance. As the wall heating increases and the ow rate decreases, a tively large in the mid and lower sections of the channel. The bot-
channel-scale convective cell is induced which is sustained primar- tom wall temperature showed a stronger inuence than the ow
ily by the local convective cells. The interactions of the rising rate on the magnitudes of both turbulent velocity components.
plumes of warm uid and the falling parcels of the cooler dense Reynolds stress increased sharply to a peak value in the inner layer
uid generate turbulent velocity elds. The complexity further in- in the region adjacent to the bottom heated wall before becoming
creases due to their interactions with the mean shear ow. These negative in the channel core region. It was argued that the negative
interactions caused the momentum transfer from the turbulent Reynolds stress indicates the momentum transfer from the turbu-
eld to the buoyancy eld, i.e. the turbulence is utilized in working lent velocity eld to the buoyancy eld. It was found that the neg-
against buoyancy forces. The underlying mechanisms of these ative Reynolds stress magnitude increases with the wall
interactions are similar but their intensity changes with the wall temperature and was almost independent of the ow rate. The
heating and the ow rate. POD analysis revealed the presence of convective cells primarily
in the lower half of the channel. The analysis also showed that
the underlying mechanisms associated with the generation of sec-
5. Conclusions ondary ow and its interactions are similar however, the energy
associated with these mechanisms changes with the Grashof and
The impact of bottom heating on the ow behavior inside a hor- Reynolds numbers.
izontal square channel at low Reynolds numbers and high Grashof
numbers has been investigated experimentally. The results Acknowledgments
showed that free convection is dominant over forced convection
for all cases. Buoyancy-driven secondary ow was generated as a The authors would like to acknowledge Natural Sciences and
result of bottom heating in the form of rising plumes of warm uid Engineering Research Council of Canada (NSERC) and the Univer-
and falling parcels of cooler uid. Turbulence was generated in the sity of Western Ontario for providing the support.
ow due to the interaction between the buoyancy-driven second-
ary ow among themself and with the shear ow. These interac- References
tions created complex three-dimensional ow. The mean
streamwise velocity proles were altered as a result of bottom Adrian, R.J., Ferreira, R.T.D.S., Boberg, T., 1986. Turbulent thermal convection in
heating where the maximum streamwise velocity magnitude was wide horizontal uid layer. Exp. Fluids 4, 121141.
Choi, D.K., Choi, D.H., 1994. Developing mixed convection ow in a horizontal tube
shifted in the lower half of the channel. A back ow along the under circumferentially non-uniform heating. Int. J. Heat Mass Transfer 37,
upper unheated heated wall was induced when Gr/Re2 > 55 and 18991913.
42 A. Elatar, K. Siddiqui / International Journal of Heat and Fluid Flow 46 (2014) 2942

Cowen, E.A., Monismith, S.G., 1997. A hybrid digital particle tracking velocimetry Osborne, D.G., Incropera, F.P., 1985a. Experimental study of mixed convection heat
technique. Exp. Fluids 22, 199211. transfer for transitional and turbulent ow between horizontal, parallel plates.
Cizmas, P.G., Palacios, A., Brien, T.O., Syamlal, M., 2003. Proper-orthogonal Int. J. Heat Mass Transfer 28, 13371344.
decomposition of spatio-temporal patterns in uidized beds. J Chem. Eng. Sci. Osborne, D.G., Incropera, F.P., 1985b. Laminar, mixed convection heat transfer for
58, 44174427. ow between horizontal parallel plates with asymmetric heating. Int. J. Heat
Deardorff, J.W., 1970. Convective velocity and temperature scales for the unstable Mass Transfer 28, 207217.
planetary boundary layer and for Rayleigh convection. J. Atmos. Sci. 27, 1211 Ozsunar, A., Baskaya, S., Sivrioglu, M., 2002. Experimental investigation of mixed
1213. convection heat transfer in a horizontal and inclined rectangular channel. Heat
Doddipatla, L.S., 2010. Wake dynamics and passive ow control of a blunt trailing Mass Transf. 38, 271278.
edge proled body. PhD Thesis. Department of Civil and Environmental Kundu, P., Cohen, I., 2002. Fluid Mechanics. Academic Press.
Engineering, University of Western Ontario. Prasad, A.K., Adrian, R.J., Landreth, C.C., Offutt, P.W., 1992. Effect of resolution on the
Driedonks, A.G.M., 1982. Models and observations of the growth of the atmospheric speed and accuracy of particle image velocimetry interrogation. Exp. Fluids 13,
boundary layer. Bound. Layer Meteorol. 23, 283306. 105116.
Gajusingh, S.T., Siddiqui, M.H.K., 2008. The inuence of wall heating on the ow Pope, S.B., 2000. Turbulent Flows. Cambridge University Press.
structure in the near-wall region. Int. J. Heat Fluid Flow 29, 903915. Regert, T., Rambaud, P., Riethmuller, M. L., 2005. Investigation of the link between
Huang, C.C., Lin, T.F., 1994. Buoyancy induced ow transition in mixed convective physics and POD modes. RTO Applied Vehicle Technology Panel (AVT) meeting,
ow of air through a bottom heating horizontal rectangular duct. Int. J. Heat RTO-MP-AVT-124, Budapest, Hungary.
Mass Transfer 37, 12351255. Sirovich, L., 1987. Turbulence and the Dynamics of Coherent Structures. Quart. Appl.
Holmes, P., Lumley, J.L., Berkooz, G., 1996. Turbulence, Coherent Structures, Math. 45, 561571.
Dynamical System and Symmetry. Cambridge University Press, 1st ed. Sakamoto, Y., Kunugi, T., Ichimiya, K., 1999. Experimental and Numerical ow
Incropera, F. P., Dewitt, D. P., Bergman, T. L., Lavine, A. S., 2006. Fundamentals of visualization of mixed convection with ow reversal in a horizontal isothermal
Heat and Mass Transfer. John Wiley & Sons. channel. J. Flow Visual. Image Process. 6, 4150.
Koizumi, H., Hosokawa, I., 1993. Unsteady behaviour and mass transfer Siddiqui, M.H.K., Leowen, M.R., Richardson, C., Asher, W.E., Jessup, A.T., 2001.
performance of the combined convection ow in a horizontal rectangular Simultaneous particle image velocimetry and infrared imagery of micro-scale
duct heated from below. Int. J. Heat Mass Transfer 36, 39373947. breaking waves. Phys. Fluids 13, 18911903.
Lin, W.L., Lin, T.F., 1996. Experimental study of unstable mixed convection of air in a Smith, T.R., Moehlis, J., Holmes, P., 2005. Low-dimensional modeling of turbulence
bottom heated horizontal rectangular duct. Int. J. Heat Mass Transfer 39, 1649 using the proper orthogonal decomposition: a tutorial. Non-linear Dynam. 41,
1663. 275307.
Mahaney, H.V., Incropera, F.P., Ramadhyani, S., 1987. Development of laminar Turner, J.S., 1973. Buoyancy Effects in Fluids. Cambridge University Press,
mixed convection ow in a horizontal rectangular duct with uniform bottom Cambridge.
heating. Numer. Heat Transfer 12, 137155. Toriyama, K., Ichimiya, K., 2010. Effect of reverse ow on three-dimensional mixed
Maughan, J.R., Incropera, F.P., 1987. Experiments on mixed convection heat transfer convection in a horizontal square duct the case of three heated or cooled walls
for airow in a horizontal and inclined channel. Int. J. Heat Mass Transfer 30, of the duct. J. Flow Visual. Image Process. 17 (1), 6984.
13071318. Wang, L.W., Hou, K.H., Lu, I.G., Hsu, C.F., 1996. Flow patterns of mixed convection in
Moeng, C., Sullivan, P.P., 1994. A comparison of shear-driven and buoyancy-driven a horizontal square channel ow. Exp. Heat Transfer 9, 257265.
planetary layer ows. J. Atmos. Sci. 51, 9991022. Wang, M., Tsuji, T., Nagano, Y., 1994. Mixed convection with ow reversal in the
Maurel, S., Boree, J., Lumley, J.L., 2001. Extended proper orthogonal decomposition: thermal entrance region of horizontal and vertical pipes. Int. J. Heat Mass
application to jet/vortex interaction. Flow Turb. Combus. 67, 125136. Transfer 37, 23052319.
Nandakumar, K., Masliyah, J.H., Law, H., 1985. Bifurcation in steady laminar mixed Zeman, O., Tennekes, H., 1977. Parameterization of the turbulent energy budget at
convection ow in horizontal ducts. J. Fluid Mech. 152, 145161. the top of daytime atmospheric boundary layer. J. Atmos. Sci. 34, 111123.

Das könnte Ihnen auch gefallen