Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

The influence of windbreak wall orientation on the cooling performance


of small natural draft dry cooling towers
Yuanshen Lu ⇑, Hal Gurgenci, Zhiqiang Guan, Suoying He
Queensland Geothermal Energy Centre of Excellence, Qld 4072, Australia
School of Mechanical and Mining Engineering, The University of Queensland, Qld 4072, Australia

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

Article history: While crosswind has a negative influence on natural draft dry cooling towers (NDDCTs) of all sizes, the
Received 25 April 2014 influence may be fatal for short towers (height < 30 m) proposed for geothermal or solar thermal power
Received in revised form 15 July 2014 plants. In a previous paper, the authors demonstrated the potential for tri-blade-like windbreak walls not
Accepted 4 September 2014
to only maintain but significantly improve the short tower cooling performance. The effect of crosswind
Available online 26 September 2014
attack angle (windbreak walls orientation) was not examined in that paper. The present paper investi-
gates that effect for a 15 m-high small-size NDDCT with horizontally-arranged heat exchangers. 3D
Keywords:
CFD models with different wind attack angles (0°, 10°, 20°, 30°, 40°, 50°, and 60°) are set up and computed
Crosswind
Natural draft cooling tower
at different crosswind speeds. The results indicate that the way the cooling tower performance varies
Dry cooling with the crosswind speed is highly sensitive to the wind attack angles. At attack angles of 0° and 60°
Windbreak wall the cooling performance is improved by windbreaks over the entire crosswind speed range investigated.
CFD modelling Other attack angles lead to unfavourable effects at certain wind speeds. The differences are related to the
turbulent airflow field in the tower bottom. The results suggest that the tri-blade-like windbreaks place-
ments always with one symmetry axis alignment with the dominant crosswind direction. The findings
could be used to determine the windbreak installation angles with respect to the most frequent direc-
tion(s) of the ambient wind in a given district.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction geothermal power plant is a 15-m tall cylindrical steel tower with
a diameter of 12 m [5]. This tower is equipped with horizontally
Thermal power plants based on Rankine cycles, no matter what arranged finned-tube heat exchangers and has a heat rejection
heat sources they use, produce nearly 90% of electric power in capacity around 578 kW at the free convection air speed of less
today’s world [1]. In these power plants, the redundant heat needs than 0.5 m/s (i.e. the mean velocity of the hot air rising in the tower
to be removed through condensers or heat exchangers [2], which in still ambient air), predicted by the 1D model developed by the
usually work with cooling towers of various types. Natural draft authors.
dry cooling towers (NDDCTs) feature no water loss and no parasitic While crosswind is not considered in the traditional design fun-
power consumption and therefore are widely used in thermal damentals of NDDCTs [6], its effects on the cooling performance of
power plants in arid areas around the world [3]. Geothermal and conventional towers have been widely investigated experimentally
solar-thermal power plants, two types of renewable thermal power as well as numerically (CFD) in recent decades [7–15]. These stud-
plants, are more likely to be located in arid areas. The Queensland ies reported that crosswind had a negative influence on the
Geothermal Energy Centre of Excellence (QGECE) has been devel- NDDCTs with either horizontally or vertically arranged heat
oping small-scale geothermal and solar thermal power plants with exchanger bundles; for example, the approach temperature
net power generation up to a few megawatts for remote Australian increases by 4–7 °C [8,11] or the heat rejection rate decreases by
communities [4]. The NDDCTs proposed for these plants are con- 25–34% [7,10,15] when the crosswind speed is 10 m/s.
siderably shorter than towers designed for conventional fossil- Methods of mitigating the crosswind effect have been proposed
fired or nuclear power plants. One proposed NDDCT for a 100 kW using windbreak walls or wind shells. A cross-shape windbreak
wall installed underneath the horizontally arranged heat exchang-
ers in a 165 m-high NDDCT was proposed and investigated by Du
⇑ Corresponding author. Tel.: +61 7 336 542 48. Preez and Kröger [16,17]. The wall was porous and as high as the
E-mail address: y.lu7@uq.edu.au (Y. Lu). tower inlet and was able to decrease the approach by up to 8 °C

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.09.012
0017-9310/Ó 2014 Elsevier Ltd. All rights reserved.
1060 Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

Nomenclature

A area (m2) Vc numerical cell volume (m3)


Aa, Afr, Ar air-side area, front area and fin-root area of heat v velocity scalar (m s1)
exchangers, respectively (m2) x, y, z Cartesian co-ordinates
Ac surface area of numerical cell (m2)
a constant Greek letters
C inertial resistance factor a, a⁄ constants in generation terms of x
cp specific heat (J kg1 K1) b bulk thermal expansion coefficient (K1)
dr outer diameter of finned tube (m) C/ diffusion coefficient for variable quantity /
F source term for momentum equations c ⁄, c constants in dissipation terms of k and x, respectively
FT temperature correction factor d velocity ratio
H height, elevation (m) l, le, lt laminar, effective, and turbulent viscosity, respectively
h convective heat transfer coefficient (W m2 K1) (kg m1 s1)
Gk, Gkb generation term of k due to mean velocity gradients and q, q density and mean density (kg m3)
buoyancy, respectively rk, rx turbulent Prandtl number for k and x, respectively
Gx, Gxb generation term of x due to mean velocity gradients rq constant in generation terms of k
and buoyancy, respectively / scalar quantity (u, v, w, T, k, e. . .)
Kr pressure loss coefficient u permeability (m2)
K, Ke, Kt laminar, effective, and turbulent thermal conductivity, x turbulence energy specific dissipation rate (s1)
respectively (W m1 K1)
k turbulent kinetic energy (m2 s2)
Vectors
m mass flow rate (kg s1)
n iteration number in CFD calculation
v
~ velocity
P pressure (Pa)
Pr, Prt laminar and turbulent Prandtl number, respectively Subscripts
pt, pd fin pitch and tube diagonal pitch, respectively (m) a, l air side, liquid (water) side
Q heat transfer rate (W) cw cross wind
q heat flux (W m2) e effective
Ret turbulence Reynolds number hx heat exchanger
S modulus of the mean rate-of-strain tensor i, o inside or inlet and outside or outlet
S/ volumetric source term for variable quantity / r radiator
T temperature (K) t tower
DTlm logarithmic mean temperature difference (K) u overall
U, V, W velocity components in x-, y-, and z-direction (m s1) 0, ref reference value

at wind speeds below 18 m/s. This conclusion was verified by Al- v cw


d¼ ð1Þ
Waked et al. [18], who numerically studied the effect of this type v ao
of windbreak wall on the thermal performance of NDDCT and they where vcw is the crosswind speed at the tower height and vao stands
further reported that either porous walls or solid walls have similar for the net upward air speed inside the cooling tower.
favourable effects on cooling tower. Chen et al. [19] ran experi- Since tall towers provide high air draft speeds, the velocity
ments on a scaled wet cooling tower model installed with the same ratios were generally limited to below 10 in these past studies,
windbreak walls and found that improvement in the cooling per- while for shorter towers, d can easily exceed 10. In a previous
formance of the tower due to windbreak walls depended on the study, the present authors considered crosswind effects on a short
setting angles of the walls. Alternative wind shells on the periphery NDDCT at velocity ratios up to 47 (corresponding to a wind speed
of the tower base were investigated by Wang et al. [20] using a of 18 m/s) [21]. They found that the heat rejection performance of
scaled model tower in the laboratory. They found that the air flow
the short tower kept declining along with the increasing of the
rate and the cooling efficiency increased remarkably after the inlet velocity ratio until reaching the maximum reduction of 37% at
air was directed by the wind shells with various installation angles. the velocity ratio of around 13, which corresponded to an actual
Zhai et al. [15] proposed a similar but much simpler version of crosswind speed of 5 m/s, only a slightly annoying speed on most
outer shells—the placement of two walls at two opposite lateral large NDDCTs [21]. It was proposed that, by introducing tri-
sides of towers, which was found to improve the cooling efficiency blade-like windbreak walls in small NDDCTs, the negative effect
by about 50% by hindering the cross-airflow and forcing the air of the crosswind in a wide range of velocity ratios (up to 40) could
flowing into the towers. be effectively converted into a significant performance boost. But
All these past studies focused on natural draft cooling towers or that paper did not consider the performance differences of the
their prototypes with heights usually over 100 m under crosswind windbreak walls when the crosswind approached in different
speeds up to 20 m/s. Compared to these tall towers employed in angles.
conventional power plants, the effect of crosswind on the cooling Therefore in this follow-up study, investigations are made on
performance of short towers is much more complicated since small the effects of the wind attack angles on the cooling tower perfor-
towers are more sensitive to the ambient conditions. The reason is mance. 3D CFD models with different wind attack angles are built
related to the relative magnitude of the crosswind with the natural and computed at different crosswind speeds. The variation of the
air flow through the heat exchanger for a particular tower, which heat transfer rates of the heat exchanger are examined and
can be defined by a ratio of the crosswind speed to the free convec- explained by considering the vortices in the airflow. The results
tion air speed—the velocity ratio d [9] as Eq. (1) provide further assistance to designers who need to design
Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069 1061

relatively short natural draft dry cooling towers for renewable to the angle between the incoming crosswind direction and the
power plants. axis of the leeward wall, as shown in Fig. 1.
The whole computational domain is discretised by structured
prism meshes. A series of grid-independence tests have been done
2. Methodology of numerical simulations
in the cases with and without windbreaks at a crosswind speed of
4 m/s using different mesh sizes until the integral heat transfer
2.1. Governing equations and solver
rate over the whole radiator (i.e. Qr, as discussed below) is nearly
independent of the mesh refinement. Fig. 2 indicates that three
The physical flow problem in NDDCTs can be expressed as a ser-
million mesh cells would be fine enough for all the cases. The final
ies of unsteady, three-dimensional Navier–Stokes equations sup-
models of the cooling tower contain over 3.7 million cells in total.
plemented with some simplified mathematical component
Cells near the walls and the heat exchangers are further refined
models. These equations are solved to obtain the conserved solu-
through the inflation method so that the minimum thickness of
tions using a general-purpose CFD code, Fluent. The governing
the cell layers drops to 5 cm with the aspect ratios of 4–6 in these
equations can be expressed in the form of the following transport
areas. An average cell size of 12 cm is used inside the cooling
equation Eq. (2):
tower. Fig. 3 shows the final mesh of the tower. The tests showed
@ q/ that the final mesh size allows the capturing of most major fea-
v /Þ ¼ rðC/ r/Þ þ S/
þ rðq~ ð2Þ tures of the turbulent airflow at the length scales that the two-
@t
equation RANS model could resolve.
The generalised scalar /, diffusion coefficient C/ and source term S/
for each governing equation are defined in Table 1. 2.3. Boundary conditions
The air was assumed to be incompressible, and the Boussinesq’s
approximation was applied so that the air density was a function of The velocity inlet boundary condition is used on the windward
temperature only. The turbulent air flow was simulated by the half surface of the domain. The profile of the inlet x-velocity U
two-equation RANS model SST k–x in the comprehensive consid- obeys the power law defined by Eq. (3) while the other two veloc-
eration of accuracy, computing time and robustness [22]. ity components V and W equal to 0. A constant ambient tempera-
All numerical calculations of the governing equations were run ture is applied, and the turbulence quantities k and x at the
first using the pressure-based steady-state solver with SIMPLE seg- boundary are determined using Eqs. (4) and (5).
regated algorithms and second-order discretization [23]. The con- !a
vergence criterion in each calculation is that all the scaled y
U ¼ v cw ¼ mref ð3Þ
residuals of the dependent variables drop to the order of 104 yref
[24] and remain invariable. In addition, the integral variables mon-
3
itored remain stable when iterating. Based on these converged k¼ ðv cw IÞ2 ð4Þ
2
results, the transient solver was used further to solve the conserva-
0:09k
tion equations in a time-dependent manner. Then we compared x¼ ð5Þ
the steady-state solutions and the time-averaged transient results lt
carefully, finding that the difference between them was sufficiently where y is the height and vref is the reference velocity at the refer-
small. ence height yref = 10 m [25]. The crosswind speed in discussions
below refers to the wind speed at the height of the cooling tower
2.2. Model geometry and meshes outlet. The turbulence intensity I and the turbulent viscosity lt
are determined according to a preliminary CFD simulation for the
The natural draft dry cooling tower and the computational same cylindrical domain without any object inside.
domain are both modelled as regular cylinders, as shown in At the leeward half surface and the top surface of the domain,
Fig. 1. The windbreak walls consist of three solid radial walls the pressure outlet boundary is used. The static pressure gradients
arranged symmetrically with separating angles of 120°. The walls in x-, y-, and z-direction are defined in Eq. (6), while the velocity
are located under the heat exchangers at the same height as the (U, V, W), temperature (T) and turbulence quantities (k, x) are
tower inlet and divide the tower base into three identical sectors, solved in the governing equations by CFD codes.
denoted sector A, B, and C respectively. The walls themselves are @P @P @P
named after the sectors they separate, for example, wall A–B is ¼ ¼ 0; ¼ q0 g ð6Þ
@x @z @y
the wall between sectors A and B. The wind angle of attack refers

Table 1
Summary of governing equations.

Equation / C/ S/
Continuity 1 0 0
 @ 
x Momentum U le @x þ r  le @x  v þ F x
 @P ~
 
y Momentum V le
@y þ r  le @y  v  q0 bðT  T 0 Þg þ F y
 @P @ ~

 
z Momentum W le @z þ r  le @z  v þ F z
 @P @ ~

Energy T Ke 1 qAc
cp cp ð V c Þ
k k l þ rlkt Gk  c⁄qkx + Gkb
x x l þ rlxt Gx  cqx2 + Gxb

pffiffiffiffiffiffiffiffiffiffiffiffiffi  
cp lt l
Where le = l + lt, lt ¼ a qxk, Ke = K + Kt, K t ¼ Pr t
@U
, Gk = ltS2, S ¼ 2Sij Sij ;Sij ¼ 12 @xij þ @U@xj ,
i
Gkb ¼ bg rqt @T
@y , Gx ¼ a xk Gk ,
   
x 0:52 0:111þRet =2:95  0:024þRet =6 qk 1 ⁄
Gxb ¼ k ½ða þ 1Þ maxðGkb ; 0Þ  Gkb , a ¼ a 1þRet =2:95 , a ¼ 1þRet =6 , Ret ¼ lx, b ¼  T 0 , c = 0.075, c = 0.09, rk = 2.0, rq = 0.9,
rx = 2.0, Prt = 0.85, T0 = 293.15.
1062 Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

72 Top view of tower


The angle of attack
Pressure outlet

A B
Crosswind
Velocity inlet
Windbreak wall C
90 12
Wall Pressure outlet

15 Heat exchangers
3
Ground Radiator
0.8
Porous media zone
y
x Wall (windbreak)
z Unit: m

Fig. 1. The dimensions of the CFD model and the boundary conditions.

800 surface. The porous media represent the pressure loss within the
heat exchanger by adding an additional source term (as defined
Integral heat transfer rate (KW)

700
in Eq. (7)) in each momentum equation [26] in Table 1. The radia-
600 tor boundary condition only reflects the heat transfer between heat
500 exchangers and the air, calculating the heat flux qr using Eq. (8)
[27]:
400  
le 1
300 Fi ¼  v þ C qv 2i ð7Þ
u i 2
200 without windbreak wall
with windbreak wall (0°)
100 qr ¼ hr ðT r  T ao Þ ð8Þ
with windbreak wall (30°)
0 where Fi and vi are the source term and velocity for the ith (x, y, or z)
0 1 2 3 4 5 6 momentum equation. 1/u and C are resistance factors; Tr and Tao
Total amounts of mesh cells (Million)
are the radiator reference and air outlet temperatures respectively.
Fig. 2. The integral heat transfer rate of the radiator against the total number of The resistance factors 1/u and C and the convective heat transfer
cells in the CFD model with and without windbreak walls at a wind speed of 4 m/s coefficient hr are all functions of air velocity and the heat exchanger
and different wall attack angles. specifications, which are derived from correlations shown in Eq. (9)
[28] and Eq. (10) [29], respectively.
 0:927  0:515
pt pt
K r ¼ 37:86nr Re0:316
c ð9Þ
dr pd

hr ¼ 0:38Re0:6
c Pr
0:333
ðAa =Ar Þ0:15 K=dr ð10Þ
where Kr is the pressure loss coefficient of heat exchangers and Rec
is the air-side Reynolds number based on the minimum free flow
area of the finned tubes. Parameters dr, nr, Pt, Pd, Aa, and Ar are all
the specifications of the heat exchangers. K is the molecular thermal
conductivity of air. For the horizontal directions, i.e. the x- and
z-axis, the source Fi is set significantly larger than that in vertical
direction y-axis so that the horizontal air flow inside the heat
exchanger zone is prevented.
Non-slip and adiabatic conditions are applied to the tower wall
and the windbreaks [30]. The pressure drop due to the tower sup-
port structures is also simulated by a cylinder face with pressure
resistance coefficients in the tower model.

2.4. Model validations


Fig. 3. Refined meshes inside and outside of cooling towers.
The simulation outcomes under the normal conditions without
the windbreak walls are compared with the results obtained by
In Eq. (6) P is the relative static (gauge) pressure defined by using the analytical NDDCT design methods, and this is a prelimin-
P = Pa  P0, where Pa is the atmospheric absolute pressure and P0 ary validation of the CFD model. The validation was described in
is a reference pressure. Here P0 is equal to Pa at ground level. q0 greater detail in an earlier paper by the present authors [21]. There
is the mean ambient air density. is no research data on such a small-size NDDCT to compare it with
The heat exchangers are modelled by a cylindrical porous media in open publications. However, to offer more support for the valid-
zone associated with the radiator boundary condition on its upper ity of the above CFD modelling methodology, i.e. the settings of
Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069 1063

12 0 2 4 6 8 10 12 14 16
tower base 4
1a Crosswind speed—vcw (m/s)
10 2a windbreak 0° 10° 180
3.5

Ratio of air mass flow rate ma/ma0


3a wall 20° 30°
40° 50° 160
8
1b 3 60° No windbreak
Δ(Two - Tai ) (°C)

2b 140

Air mass flow rate—ma (kg/s)


wind
3b 2.5
6 120

2 100
4
80
1.5
60
2
1
40
0 0.5
20
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16
Crosswind speed (m/s) 0 0
0 5 10 15 20 25 30 35 40 45
Velocity ratio (δ)—vcw/vao
Fig. 4. The effect of crosswind on the approach difference of three large NDDCTs:
(1) 129 m high, 285 MW heat dumping, CFD results [11,18]; (2) current 120 m high
Fig. 5. The dimensionless air mass flow rate ma/ma0 as a function of the velocity
327 MW, CFD results (built for model validation only); (3) 165 m high, 650 MW
ratio d in all cases of angles of attack. The legend refers to the wind attack angle. The
heat dumping, experiment results [8,16], where letters a and b following the
secondary x-axis and y-axis show the corresponding dimensional values of wind
numbers denote the cases without and with the windbreak wall, respectively.
speed and air mass flow rate, respectively.

boundary conditions, turbulence model, solver, etc., a 120 m-high 0 2 4 6 8 10 12 14 16


2.25
3D NDDCT model has been built using the same aforementioned Crosswind speed—vcw (m/s)
1,245

Ratio of radiator heat transfer rate—Qr / Qr0


methodology. The analytical heat-dumping capacity of this big 2
0° 10°

Heat transfer rate of radiator—Qr (kW)


20° 30° 1,145
cooling tower under windless condition is around 327 MW [6]. In 40° 50°
1,045
1.75 60° No windbreak
addition, a cross-shape windbreak wall with the same porosity as
945
that in [17] was used underneath the heat exchangers inside the 1.5
845
tower base, as indicated in Fig. 4. The crosswind effect on this cool-
745
ing tower with and without the windbreak wall is assessed in the 1.25

form of approach temperature difference, where the approach 645


1
temperature is defined as the water outlet temperature minus 545

the air inlet temperature: Two  Tai. Fig. 4 compares the approach 0.75 445
temperature difference in this CFD result with those obtained in 345
0.5
previous studies. 245
In Fig. 4, the solid lines (curves 1a, 2a, and 3a) represent the 0.25 145
results without the windbreak wall while the dashed lines (curves 0 5 10 15 20 25 30 35 40 45
Velocity ratio (δ)—vcw/vao
1b, 2b, and 3b) are the results with the windbreak. Particularly,
curves 3a and 3b are obtained in a field measurement [8] and a Fig. 6. The dimensionless radiator heat transfer rate Qr/Qr0 as a function of the
model test [16], respectively. It is noticed that the results of our velocity ratio d in all cases of angles of attack. The legend refers to the wind attack
current large cooling tower model (curves 3a and 3b) are relatively angle. The secondary x-axis and y-axis show the corresponding dimensional values
closer to their counterparts reported by Al-Waked et al. [11] of the wind speed and the heat transfer rate, respectively.

(curves 2a and 2b), which is due to that all the cooling towers have
different heat dumping rates, and the wind effect tends to be Z
increasingly less along with the increase of heat dumping rates ma ¼ qv ao dA ð11Þ
A
of the cooling tower [8]. In spite of this, all these studies concluded
the same change trend in the cooling performance, which implies Z Z
that current modelling methods are reliable. Qr ¼ hr ðT r  T ao ÞdA ¼ qr dA ð12Þ
A A

3. Results and discussion All variables on the right-hand sides of both equations are solved
and conserved in the numerical iterations to enable reporting of
In the above three-dimensional CFD model of the 15 m-high the quantities on the left side after the computations are completed.
small NDDCT, the air flow and the heat transfer within the compu- When both ma and Qr are divided by their corresponding values
tational domain are calculated. The simulations are carried out at under no-crosswind condition ma0 and Qr0, respectively, they
different wind speeds (0, 1, 2, 3, 4, 6, 8, 10, 12, 14, and16 m/s) become dimensionless quantities. These two dimensionless quanti-
and different wind attack angles (0°, 10°, 20°, 30°, 40°, 50°, and ties are plotted against the velocity ratio d for different angles of
60°). attack using solid lines, as shown in Figs. 5 and 6, respectively.
The installation of windbreak walls in the 15 m-high tower are For comparison purposes, the simulated results without the wind-
found to have a strong effect on the air flow behaviour in the tower break walls [21] are plotted in dashes in Figs. 5 and 6 as well.
base, and this resulted in a large change in the heat dissipation For wind attack angles from 0° to 40°, the ratios of both air mass
capacity of the heat exchangers. The parameters of the integral flow and the heat transfer rate start to decline at low velocity ratios
net upward mass flow rate ma and the convective heat transfer rate until the velocity ratio reaches a critical value. Different attack
Qr at the radiator are introduced here to assess quantitatively the angles show different declinations. Above the critical value, the
overall thermal performance of the heat exchangers and the cool- trend is reversed, which indicates the benefits of the windbreak
ing tower. In this model, ma and Qr are computed using Eqs. (11) walls. The troughs of these curves, depending on the wind attack
and (12) respectively. angle, occur in the range of 2.5–10 of the velocity ratio. In this
1064 Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

Fig. 7. Side views of the time-averaged 3D streamlines passing through the tower bottom at a crosswind speed of 4 m/s and at different wind attack angles as indicated.

15 m-high NDDCT, an air velocity ratio of 10 corresponds to a The underlying reasons for the interesting trends of the numer-
crosswind speed of approximately 4 m/s. A comparison between ical results presented in Figs. 5 and 6 are investigated by detailed
the solid lines and the dashed lines shows the significant effective- examination of the air flow patterns around the heat exchangers
ness of the windbreak walls at high velocity ratio (d > 10). under two sets of conditions: the same crosswind speed at differ-
It is interesting to note the turn-around of heat transfer rate ent angles of attack; and the same angle of attack at different cross-
without windbreaks (i.e. the dashed curve in Fig. 6) at velocity wind speeds.
ratios above 13, which cannot be found in previous open published
results. Most of the research studies on natural draft cooling tow- 3.1. Flow and temperature fields at various attack angles
ers involved air velocity ratios of below 10. A previous study [21] of
the current authors found that the turn-around feature in the The air flow patterns in the tower interior and around the tower
dashed curve could be attributed to the reverse airflow in the base are visualised by the time-averaged 3D streamlines at differ-
windward part of the heat exchanger area caused by the suction ent attack angles at the wind speed of 4 m/s, as shown in Figs. 7
effect of crosswind underneath the heat exchanger. The inverse and 8.
flow occurs only at certain crosswind speeds. With this inverse With no windbreak walls, the air flows directly across the tower
flow, the total heat transfer Qr between the heat exchangers and base with significant vortices activity being observed inside the
the air can be dissipated in two paths at the same time—one tower. By contrast, towers with windbreak walls experience a
through the tower top whose mass flow rate is described by ma, smoother and more uniform air flow inside the tower. This
and the other via the tower bottom. Increasing wind speed improvement leads to a difference in the distribution of air tem-
depresses the former but boosts the latter surprisingly. The change perature. Fig. 9 compares the temperature contours at mid-xy
in the total Qr of the entire cooling tower is thus the result of both plane of the cooling tower for different wind attack angles at the
the negative and positive effects of crosswind. same crosswind speed. No hot air region is seen underneath the
While similar trends have been observed at the curves for all heat exchanger from the contours when the windbreaks exist.
wind attack angles, Figs. 5 and 6 show that the critical crosswind The windbreak walls enhance the convective heat transfer from
speed corresponding to the troughs of ma/ma0 or Qr/Qr0 curve for the heat exchangers by improving the airflow above the heat
attack angles from 0° to 40° are different at different angles of exchangers.
attack. It is noted that at the attack angles of 50° and 60°, the cool- However vortices are generated in the wake of the tower base
ing performance is almost unaffected at low velocity ratios (d < 10) caused by the separation of the air flow at the tips of windbreak
region, which implies a great advantage compared with other wind walls. At a wind speed of 4 m/s (d = 10.5), the Reynolds number
attack angles. Once the wind velocity ratio exceeds d > 10 (vcw > 4 - based on the tower base diameter is around 3.2  106. The wake
m/s), a significant advantage is observed for the attack angles of 0°, structures are complex at such a high Reynolds number and sensi-
10°, and 20°, where the wind direction is closer to one of the walls. tive to windbreak wall orientations as well. At attack angles of 0°
Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069 1065

Fig. 8. Top views of the time-averaged 3D streamlines passing through the tower bottom at a crosswind speed of 4 m/s and at different wind attack angles as indicated.

Fig. 9. Temperature contours at mid-xy plane for different wind attack angles as indicated when crosswind speed is 4 m/s.

and 60°, where the walls are arranged symmetrically about the biased streamline pattern behind the walls as seen in Fig. 8 at
wind direction, the time-averaged streamlines of air flow vortices attack angles in the range of 10–50°.
are symmetrical as expected. For other attack angles, the tips of The air flow through the tower base in one of the cases is pre-
the windbreak walls where flow is separated are not symmetrically sented in greater detail in Fig. 10, which shows the horizontal com-
positioned about the wind direction, resulting in the different vor- ponents of the time-averaged velocity vectors in the plane of
tex distributions on the two sides. This causes an asymmetric and y = 2.5 m at a non-symmetric attack angle of 30°. This height
1066 Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

corresponds to the layer just underneath the plane of the heat


exchangers, which are placed at a height of 3 m. Fig. 10 shows that
the air flow is separated at the tips of the windbreak walls forming
the wake. An imaginary boundary between the free stream zone
and the wake zone (free-wake boundary) can be seen as shown
in shaded lines. The flow regimes are distinctly different at the
two sides of this boundary.
The magnitude of vorticity, which is a measure of the local spin-
ning motions of air, is introduced to help in understanding the dis-
tribution of the vortices and the wake flow structure. Fig. 11 shows
the vorticity contours at the horizontal plane of y = 2.5 m at differ-
ent angles of attack under a crosswind speed of 4 m/s. The vorticity
in the wakes of windbreak walls is generally higher than those in
the ambient air, and it is especially strong along the free-wake
boundaries and closer to the wall tips, as in these areas, the veloc-
ity gradients are much higher, causing the shear stress to increase
dramatically. If the free-wake boundaries are located inside a sec-
tor, the strong shear force near the boundaries substantially dom-
inate the air flow field in this sector by inducing a large-scale
circulation of air movement in the wake of this leading wall. In
Fig. 10. The time-averaged horizontal velocity components at y = 2.5 m, an attack
angle of 30° and a crosswind speed of 4 m/s.

Fig. 11. The vorticity contour at plane y = 2.5 m for a crosswind speed of 4 m/s and at different wind attack angles as indicated.

Fig. 12. Pressure P contours at a surface 1 cm under heat exchangers at a crosswind speed of 4 m/s and at different wind attack angles as indicated.
Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069 1067

Fig. 13. Heat flux qr contours at heat exchanger surface at a crosswind speed of 4 m/s and at different wind attack angles as indicated.

Fig. 14. The vorticity contour at plane y = 2.5 m for a wind attack angle of 30° at different crosswind speeds as indicated.

Fig. 11, this phenomenon is clearly visible in sector C in cases of experience relatively low vorticity despite the sectors falling into
attack angles of 30–50°. the wake zones, such as sector B in Fig. 11. The vortices in the lead-
The air flow underneath the heat exchangers at y = 2.5 m is ing edge of windward sector A relate to a reverse suction effect of
influenced by not only the separation but also the lifting effect of the heat exchangers, as discussed in [21].
buoyance force. Unlike the shear stress, the lifting force tends to The larger magnitude of local vorticity indicates a more severe
regulate the air flow by drafting air moving upward through the spinning of the air nearby which consequently implies that lower
heat exchanger. Sectors without significant shear stresses air pressure occurs and vice versa. According to the working
1068 Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069

Fig. 15. Heat flux qr contours at the heat exchanger surface for a wind attack angle of 30° at different crosswind speeds as indicated.

principles of a cooling tower with horizontally arranged heat Concerning the local heat flux distributions (Fig. 15), increasing
exchangers [6], the variation of the inlet pressure in a given zone crosswind speeds boost the heat transfer rate in sector A, where the
of heat exchangers directly influences the air flow rate through upward air flow rate and pressure underneath the heat exchangers
the heat exchanger bundles and consequently the heat transfer are both enhanced. Meanwhile, the cooling performance is also
rate in that zone. Fig. 12 plots the contours of the air pressure, P, improved slightly in sector B, benefiting from the extension of the
in the surface under the heat exchangers at different angles of separated flow areas of both walls A–B and A–C. Because the low-
attack, while Fig. 13 shows the corresponding locally averaged velocity zone is enlarged, providing more intake air for this sector.
heat flux, qr, in the upper surface of the heat exchangers. The However, a low heat flux zone appears in sector C because of the
distributions of the net heat transfer rate Qr can then be calculated large local air circulation. The qr in this sector experiences a turn-
by applying Eq. (12). around process along with increasing crosswind speed.
Clearly, low-value zones for P and qr are observed at positions of The variation of the overall heat transfer rate of the cooling
high vorticity. The heat transfer performance in sectors facing or tower Qr at different crosswind speeds is therefore subject not only
mainly facing toward the oncoming wind (e.g. sector A in all cases) to increased heat transfer in sectors A and B but also the decrease
has been significantly enhanced compared to the same areas with in sector C. By integrating the local heat flux qr over the entire heat
no windbreak walls since both the air flow rate and pressure in this exchanger area using Eq. (12), it is found that the minimum Qr
area are increased, which can be understood from the streamlines occurs at the wind velocity ratio of 8 with a arrangement of the
(Fig. 7) and pressure distributions (Fig. 12). By contrast, the later- windbreaks at the attack angle of 30°. For d > 8, the total air flow
ally facing sectors (sector C in cases 20°, 30°, and 40°) suffer rate, ma, across the heat exchanger surface starts to increase,
reduced P and the qr for the opposite reason, causing a lower over- resulting in the enhancement of cooling performance. In fact, the
all heat transfer rate of the entire cooling tower. results at other attack angles follow the same trend, which
explains the existence of critical wind speeds for both the ma and
Qr curves in Figs. 5 and 6.
3.2. Flow and temperature fields at various wind speeds
4. Conclusions
For different crosswind speeds at the same wind attack angle,
the vortices at y = 2.5 m are shown in Fig. 14 for the attack angle The influence of crosswind speed on the heat transfer perfor-
of 30°. With the existence of wall B–C, the wake of wall A–C is con- mance of a small NDDCT equipped with tri-blade-like windbreak
fined within sector C at a wind speed of 1 m/s. As a result, the shed- walls in the bottom has been studied using CFD numerical model-
ding vortices accumulate in this region rather than dissipate ling. The overall heat rate Qr of the tower was found to be signifi-
downstream, forming a great local circulation of air flow. The cir- cantly enhanced compared to that of the same tower without a
culation reduces the local air pressure and therefore yields less windbreak wall when the velocity ratio of air d was over 10. For
air flow through the heat exchangers. As the crosswind speed the short tower examined in this study, this velocity ratio corre-
increases, the air circulation expands gradually until it covers the sponds to crosswind speeds larger than 4 m/s. The results confirm
whole area of sector C when the crosswind speed is 3 m/s. How- the benefits of using windbreak walls for cooling performance in
ever, a further increase in the wind speed causes the wake of the small NDDCTs with horizontally arranged heat exchangers. The
wall A–C to extend out of sector C so that vortices can dissipate variation of this benefit depends on the structure of the turbulent
far downstream of the cooling tower and the large local circulation airflow in the tower bottom, which in turn is sensitive to the orien-
shrinks. As a result, the air flow rate through the heat exchangers tations of the windbreak walls with respect to the crosswind veloc-
recovers. ity vector. The study also finds that:
Y. Lu et al. / International Journal of Heat and Mass Transfer 79 (2014) 1059–1069 1069

1. The flow separation causes the forming of vortices in the wake [7] M.D. Su, G.F. Tang, S. Fu, Numerical simulation of fluid flow and thermal
performance of a dry-cooling tower under cross wind condition, J. Wind Eng.
of the windbreak walls. The vortices with high magnitude of
Ind. Aerodyn. 79 (3) (1999) 289–306.
vorticities waste the majority of the kinetic energy of the air [8] A.F. Du Preez, D.G. Kröger, Effect of wind on performance of a dry-cooling
flow, resulting in reductions in the air pressure and the heat tower, Heat Recovery Syst. CHP 13 (2) (1993) 139–146.
transfer rate of heat exchangers in these regions. [9] Q.D. Wei, B.Y. Zhang, K.Q. Liu, X.D. Du, X.Z. Meng, A study of the unfavorable
effects of wind on the cooling efficiency of dry cooling-towers, J. Wind Eng. Ind.
2. The increase of crosswind speed could consistently enhance the Aerodyn. 54 (1995) 633–643.
air flow and the cooling performance in the windward and lee- [10] S. Fu, Z.Q. Zhai, Numerical investigation of the adverse effect of wind on the
ward sectors of the cooling tower heat exchangers. heat transfer performance of two natural draft cooling towers in tandem
arrangement, Acta Mech. Sin. 17 (1) (2001) 24–34.
3. When the windbreak walls are arranged with wind attack [11] R. Al-Waked, M. Behnia, The performance of natural draft dry cooling towers
angles of 0° and 10° with respect to the crosswind direction, under crosswind: CFD study, Int. J. Energy Res. 28 (2) (2004) 147–161.
the heat-dumping rate of the cooling tower is significantly [12] R. Al-Waked, M. Behnia, CFD simulation of wet cooling towers, Appl. Therm.
Eng. 26 (4) (2006) 382–395.
higher than other angles at air velocity ratios of over 10. At [13] M. Gao, F.-Z. Sun, K. Wang, Y.-T. Shi, Y.-B. Zhao, Experimental research of heat
the lower velocity ratios, the performance is slightly better for transfer performance on natural draft counter flow wet cooling tower under
the attack angle of 50° or 60° and then 0°. cross-wind conditions, Int. J. Therm. Sci. 47 (7) (2008) 935–941.
[14] R. Al-Waked, Crosswinds effect on the performance of natural draft wet
cooling towers, Int. J. Therm. Sci. 49 (1) (2010) 218–224.
In general, at wind attack angle of 0° the windbreak walls give [15] Z. Zhai, S. Fu, Improving cooling efficiency of dry-cooling towers under cross-
the most beneficial performance as they enhance the cooling rate wind conditions by using wind-break methods, Appl. Therm. Eng. 26 (10)
(2006) 1008–1017.
of the heat exchanger over the entire range of crosswind speed.
[16] A.F. Du Preez, D.G. Kröger, The effect of the heat exchanger arrangement and
This implies that the tri-blade-like walls should be placed with wind-break walls on the performance of natural draft dry-cooling towers
one wall, i.e. one symmetry axis, always aligned with the dominant subjected to cross-winds, J. Wind Eng. Ind. Aerodyn. 58 (3) (1995) 293–303.
direction of the crosswind. The most practical implication of this [17] A.F. Du Preez, D.G. KrÖGer, Effect of the shape of the tower supports and walls
on the performance of a dry-cooling tower subjected to cross winds, Heat
result is that the findings can be used to determine the windbreak Transfer Eng. 16 (2) (1995) 42–49.
installation angles with respect to the most frequent direction(s) of [18] R. Al-Waked, M. Behnia, The effect of windbreak walls on the thermal
the environmental crosswind in a given district. If there is no dom- performance of natural draft dry cooling towers, Heat Transfer Eng. 26 (8)
(2005) 50–62.
inant crosswind direction, the results can be used to quantify the [19] Y. Chen, F. Sun, H. Wang, N. Mu, M. Gao, Experimental research of the cross
benefits of designing a rotatable windbreak wall fitted at the tower walls effect on the thermal performance of wet cooling towers under
bottom under the heat exchangers. crosswind conditions, Appl. Therm. Eng. 31 (17) (2011) 4007–4013.
[20] K. Wang, F. Sun, Y. Zhao, M. Gao, L. Ruan, Experimental research of the guiding
channels effect on the thermal performance of wet cooling towers subjected to
Conflict of interest crosswinds – air guiding effect on cooling tower, Appl. Therm. Eng. 30 (5)
(2010) 533–538.
[21] Y. Lu, Z. Guan, H. Gurgenci, Z. Zou, Windbreak walls reverse the negative effect
None declared. of crosswind in short natural draft dry cooling towers into a performance
enhancement, Int. J. Heat Mass Transfer 63 (2013) 162–170.
Acknowledgements [22] Z. Zhang, Z.Q. Zhai, W. Zhang, Q.Y. Chen, Evaluation of various turbulence
models in predicting airflow and turbulence in enclosed environments by CFD:
Part 2-comparison with experimental data from literature, HVAC&R Res. 13 (6)
This research was performed as part of the Australian Solar (2007) 871–886.
Thermal Research Initiative (ASTRI), a project supported by the [23] ANSYS Inc., FLUENT User’s Guide, ANSYS Inc., New Hampshire, USA, 2011.
[24] M. Lucas, P.J. Martinez, J. Ruiz, A.S. Kaiser, A. Viedma, On the influence of
Australian Government, through the Australian Renewable Energy psychrometric ambient conditions on cooling tower drift deposition, Int. J.
Agency (ARENA). Heat Mass Transfer 53 (4) (2010) 594–604.
[25] L.J. Yang, X.Z. Du, Y.P. Yang, Wind effect on the thermo-flow performances and
its decay characteristics for air-cooled condensers in a power plant, Int. J.
References Therm. Sci. 53 (2012) 175–187.
[26] H.C.R. Reuter, D.G. Kröger, Computational fluid dynamics analysis of cooling
[1] W.H. Wiser, Energy Resources.: Occurrence, Production, Conversion, Use, first tower inlets, J. Fluids Eng. Trans. ASME 133 (8) (2011).
ed., Springer Verlag, New York, 2000. [27] L.J. Yang, L. Chen, X.Z. Du, Y.P. Yang, Effects of ambient winds on the thermo-
[2] B. Péter, G. Gyula, G. Iván, Efficiency and cost modelling of thermal power flow performances of indirect dry cooling system in a power plant, Int. J.
plants, Therm. Sci. 14 (3) (2010) 821–834. Therm. Sci. 64 (2013) 178–187.
[3] S. He, H. Gurgenci, Z. Guan, A.M. Alkhedhair, Pre-cooling with munters media [28] K.K. Robinsion, D.E. Briggs, Pressure drop of air flowing across triangular pitch
to improve the performance of natural draft dry cooling towers, Appl. Therm. banks of finned tubes, Chem. Eng. Prog. Symp. Ser. 62 (64) (1966) 177–184.
Eng. 53 (1) (2013) 67–77. [29] A. Ganguli, S.S. Tung, J. Taborek, Parametric study of air-cooled heat exchanger
[4] Z. Zou, Z. Guan, H. Gurgenci, Y. Lu, Solar enhanced natural draft dry cooling finned tube geometry, Am. Inst. Chem. Eng. Symp. Ser. C 81 (245) (1985) 122–
tower for geothermal power applications, Sol. Energy 86 (9) (2012) 2686– 128.
2694. [30] A.S. Kaiser, M. Lucas, A. Viedma, B. Zamora, Numerical model of evaporative
[5] Y. Lu, Z. Guan, H. Gurgenci, Development of Small Natural Draft Dry Cooling cooling processes in a new type of cooling tower, Int. J. Heat Mass Transfer 48
Towers for Geothermal Power Plants, in: The 15th IAHR Cooling Tower and (5) (2005) 986–999.
Air-cooled Heat Exchanger Conference, Beijing, China, 2011, pp. 308–316.
[6] D.G. Kröger, Air-cooled Heat Exchangers and Cooling Towers, vol. II, Pennwell
Corp, Tulsa, Okl, USA, 2004. pp. 38–91.

You might also like