Sie sind auf Seite 1von 12

Building and Environment 46 (2011) 657e668

Contents lists available at ScienceDirect

Building and Environment


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

Numerical study of a M-cycle cross-ow heat exchanger for indirect evaporative


cooling
Changhong Zhan a, b, Xudong Zhao c, *, Stefan Smith c, S.B. Riffat a
a

Department of the Built Environment, University of Nottingham, University Park, Nottingham NG7 2RD, UK
School of Civil Engineering, Northeast Forestry University, Harbin 150040, China
c
Institute of Energy and Sustainable Development, De Montfort University, The Gateway, Leicester LE1 9BH, UK
b

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 16 June 2010
Received in revised form
15 September 2010
Accepted 21 September 2010

In this paper, numerical analyses of the thermal performance of an indirect evaporative air cooler
incorporating a M-cycle cross-ow heat exchanger has been carried out. The numerical model was
established from solving the coupled governing equations for heat and mass transfer between the
product and working air, using the nite-element method. The model was developed using the EES
(Engineering Equation Solver) environment and validated by published experimental data. Correlation
between the cooling (wet-bulb) effectiveness, system COP and a number of air ow/exchanger parameters was developed. It is found that lower channel air velocity, lower inlet air relative humidity, and
higher working-to-product air ratio yielded higher cooling effectiveness. The recommended average air
velocities in dry and wet channels should not be greater than 1.77 m/s and 0.7 m/s, respectively. The
optimum ow ratio of working-to-product air for this cooler is 50%. The channel geometric sizes, i.e.
channel length and height, also impose signicant impact to system performance. Longer channel length
and smaller channel height contribute to increase of the system cooling effectiveness but lead to reduced
system COP. The recommend channel height is 4 mm and the dimensionless channel length, i.e., ratio of
the channel length to height, should be in the range 100 to 300. Numerical study results indicated that
this new type of M-cycle heat and mass exchanger can achieve 16.7% higher cooling effectiveness
compared with the conventional cross-ow heat and mass exchanger for the indirect evaporative cooler.
The model of this kind is new and not yet reported in literatures. The results of the study help with
design and performance analyses of such a new type of indirect evaporative air cooler, and in further,
help increasing market rating of the technology within building air conditioning sector, which is
currently dominated by the conventional compression refrigeration technology.
2010 Elsevier Ltd. All rights reserved.

Keywords:
Evaporative cooling
Cross-ow
Heat and mass transfer
Numerical simulation

1. Introduction
Air conditioning of buildings is currently dominated by
conventional compression refrigeration system, which takes over
95% of the market share in this sector. This kind of system is highly
energy intensive due to extensive use of electricity for operation of
the compressor, and therefore, is neither sustainable nor environmentally friendly. The use of indirect evaporative cooling has a high
potential for meeting air conditioning needs at low energy costs.
This, however, is dependent on the capacity of additional water
vapour that can be held by the cooling air stream. Whilst more
commonly applied in hot, arid climatic regions such as the Middle
East, part of the Far East, North/South America and Europe, there is
* Corresponding author. Tel.: 44 116 257 7971; fax: 44 116 257 7981.
E-mail address: xzhao@dmu.ac.uk (X. Zhao).
0360-1323/$ e see front matter 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.buildenv.2010.09.011

an increasing trend for such systems to be applied in low energy


building designs in less suited climatic regions such as in the UK.
Recent research associated with projected future climate in the UK
shows at least a probable increased potential for evaporative
cooling in this region, particularly when being jointly operated
with desiccant dehumidication [1,2].
Indirect evaporative cooling systems have the advantage of
being able to lower the air temperature without increasing
humidity of the conditioned space and avoid potential health issues
from contaminated water droplets entering occupied spaces (as
associated with direct evaporative cooling systems). These systems
usually require much less electric power that mechanical vapour
compression uses for air conditioning [3]. Therefore, such systems
will help reduce electricity consumption, and thus contribute to
reducing greenhouse gas emissions. It has widely been used as
a low energy consuming device for various cooling and air

658

C. Zhan et al. / Building and Environment 46 (2011) 657e668

Nomenclature
A
cp
COP
d
h
hm
i
L
Le
m
Nu
P
Pr
Q
Re
t
u

heat transfer area, m2


specic heat of air, J/kg  C
energy efciency of the IEC (Indirect Evaporative
Cooler)
equivalent diameter of the air passage, m
convective heat transfer coefcient, W/m2  C
mass transfer coefcient, m/s
specic enthalpy of air, J/kg
length, m
Lewis number
air mass ow rate, kg/s
Nusselt number
theoretical fan power, W
Prandtl number
heat ux, W/m2
Reynolds number
temperature,  C
velocity, m/s

conditioning applications in industrial, agricultural and residential


sectors [4e7] for providing low temperature uids (e.g. air, water).
Indirect components can also be combined with mechanical vapour
compression air conditioning systems to achieve very high efciencies while delivering comfort cooling that is equal to conventional air conditioning.
In an indirect evaporative cooler (IEC), a primary (also called
product) air stream is cooled by simultaneous heat and mass
transfer between a secondary (also called working) air stream and
a wet wall surface. The latent heat transport, in connection with the
vaporization of the liquid lm, plays an important role in the heat
transfer process [8,9]. Most commercially available IECs are
equipped with standard cross-ow heat exchangers that have
a stacked structure of heat and mass transfer plates as shown in
Fig. 1. In principle, the structure allows the product air to ow over
the dry side of a plate and the working air to ow perpendicular to
the product ow direction over the opposite wet side of the plate.
The wet side absorbs heat from the dry side by evaporating water
and therefore cooling the dry side, while the latent heat of vaporizing water is given to the wet side air. In an ideal operation, the
product air temperature on the dry side of the plate will reach the
wet-bulb temperature of the incoming working air, and temperature of the working air on the wet side of the plate will increase to
the incoming product air dry-bulb temperature and will reach 100%
saturation. However, practical systems are far from ideal. It has
been suggested that only 50e60% of the incoming working air wetbulb temperature can be achieved for a typical indirect evaporative
cooling device [10], while in most systems the working and product
air come from the same source (i.e. ambient air) and therefore have
the same temperature level. This type of exchanger has been
comprehensively studied and developed, as suggested in the literature [8e13], with no great potential to further improve the cooling
effectiveness (efciency) of the exchanger.
In recent years, a new type of heat and mass exchanger (Fig. 2a)
utilizing the benets of the Maisotsenko cycle [14] has been
developed commercially [15e17]. In this type of exchanger, part of
the surface on the dry side is designed for the working air to pass
through and the rest is allocated to the product air. Both the
product and working air are guided to ow over the dry side along
parallel ow channels. There are numerous holes distributed
regularly on the area where the working air is retained and each of
these allows a certain percentage of air to pass through and enter

V
w
Dp

g
3
h
r
F0

air volume ow rate, m3/s


humidity ratio of moist air, kg/kg dry air
pressure loss, Pa
latent heat of water evaporation, J/kg
effectiveness, %
dynamic viscosity , Pa s
density, kg/m3
cooling capacity, W

Subscripts
1
dry side
2
wet side
a,f
air ow
db
dry bulb
in
inlet
l
latent
su
supply air
w
wall
wb
wet-bulb
wk
working air

the wet side of the sheet. The air is gradually delivered to the wet
side as it ows along the dry side, thus forming an even distribution
of airstreams over the wet surface. This arrangement allows the
working air to be pre-cooled before entering the wet side of the
sheet by losing heat to the opposite wet surface. The pre-cooled air
delivered to the wet side ows over the wet surface along channels
arranged at right angles to the dry side channels, absorbing heat
from the working and product air. As a result, the product air is
cooled before being delivered to spaces where cooling is required,
and the working air is humidied, heated and discharged to the
atmosphere. Owing to effect of pre-cooling, the working air in the
wet side (working air wet channel) has a much lower temperature
and therefore, is able to absorb more heat from its two adjacent
sides, i.e. the dry working air ow side and the dry product air ow
side. As a result, the cooling (wet-bulb) effectiveness of the new
structure would be higher than that in the traditional cross-ow
exchanger (Fig.1). The cooling process is shown on a psychometric
chart in Fig. 2b. The manufacturers data has indicated that the
exchanger, namely M-cycle heat exchanger, could obtain a wetbulb effectiveness of 110% to 122%. [16,17]
Although signicant progression has been achieved in industrial
and manufacturing exercise of such a new type of M-cycle
exchanger, to the authors knowledge there is no numerical study of
the new design being so far reported. To overcome the shortfall in
the theoretical study of the exchanger and to further enable

Fig. 1. Schematic of the traditional cross-ow heat and mass exchanger for indirect
evaporative cooling.

C. Zhan et al. / Building and Environment 46 (2011) 657e668

659

Fig. 2. Air ow and heat/mass transfer associated with the new heat and mass exchanger. (a) Air ow prole. (b) air treatment process (psychrometric indication).

optimization of the exchanger performance, a numerical model has


been developed to enable solving the coupled governing equations
of the heat and mass transfer between two adjacent airstreams
using EES software [18]. Based on this development, the effect of
various exchanger operating parameters to the system performance have been investigated. This work is of signicant importance to optimization of system conguration and development of
the solutions towards the better performance of the system operation. The work is expected to achieve high level of impact in terms
of increasing energy efciency of the indirect evaporative cooling
systems, extending its market share in building air conditioning
sector, and thus contributing to achieve the global targets in energy
saving and carbon reduction measures.
2. Description of the cooler with new type of heat and mass
exchanger
Fig. 3a presents the structure of the M-cycle exchanger in an
ISAW [17] indirect evaporative cooler e tac-150. This type of
exchanger consists of numerous sheets of a bre designed to wick
uids evenly. The sheets are stacked together, separated by channel
guides located on one side of the sheet. One side of each sheet is also
coated with polyethylene to avoid penetration of water. The guides
are fabricated with a plastic material, and run along the length of one

sheet, and the width of the next sheet to form a cross-ow within the
exchanger. There are numerous regularly distributed holes made
along the dry air ow paths, which are located at the working air
ow area. This conguration gradually diverts air from the dry
channel to the wet channels e the air ow is perpendicular to the
wet channels and has an even velocity distribution. With heat and
moisture exchange this warmer and more highly saturated air is
discharged to the atmosphere. In the meantime, the product air is
being cooled along its ow path. The pre-cooling of the working air
provides a greater temperature difference between the dry and wet
channel air, so improving the cooling effectiveness of the system. In
this studied case, the bre of the exchanger is 0.24 mm in thickness;
and the whole package incorporates a total of 35 dry passages and 34
wet passages, each of 4 mm in height.
Air ow distribution across the channels is shown schematically
in Fig. 3b, with heat and mass transfer taking place between the dry
and wet air channels. All the incoming air is initially led into the dry
passages (from Nos. 3 to 6 for product and Nos. 1 to 2 for working
air), with the working air being gradually diverted into the wet
channels, via the dedicate-designed holes. This channel layout
contributes to even distribution of the air ow across the wet
channels without imposed ow adjustment at the outlets of the
supply and exhaust air. Heat and mass transfer will take place
between the dry and wet channel air.

Fig. 3. Schematic of the heat and mass exchanger in ISAW TAC-150. (a) Structure view. (b) Air ow distribution in the dry/wet channels.

660

C. Zhan et al. / Building and Environment 46 (2011) 657e668

Fig. 4. Cell element applied for numerical simulation. (a) Cell element for simulation. (b) Differential illustration.

3. Simulation approach

dQl dQ1 dQ2

(2)

3.1. Heat and mass transfer mechanisms e mathematical indication


(3) The energy balance in dry passages
The cell element selected for numerical analyses is shown in
Fig. 4. The element consists of half the height of the dry channel, the
plate wall and half the height of the wet channel. Energy balance
equations were applied to each single element, with consideration
of a pre-set boundary condition. This allowed the temperature and
humidity distribution across the dry and wet channel sections to be
established.
To simplify the modelling process and mathematical analysis,
the following assumptions were made:
1. The heat and mass transfer is in steady state. The IEC enclosure
is considered as the system boundary.
2. The wet surface of the bre sheet is completely saturated. The
water vapour is distributed uniformly within the wet channel.
3. A temperature gradient for the channel cross-section was set to
zero. Heat transfer in the separating plate is considered in the
vertical direction only. Within the working uid, the crossstream convective heat transfer is considered as the dominant
mechanism of heat transfer.
4. Each element has a uniform wall surface temperature. An
analysis carried out by Zhao et al. [9] showed that the thermal
conductivity of the plate wall has little impact on the magnitude of the heat and mass transfer rates, owing to its small
thickness (0.24 mm). The temperature difference between dry
and wet sides of the wall can be ignored.
5. Air is treated as an incompressible gas.
By applying principles of mass and energy conservation [19] into
the differential element shown in Fig. 4, the heat and mass transfer
processes in an IEC can be described with the following set of
differential equations.

Dry passage air involves the forced convective heat transfer,


leading to change of the enthalpy of the air. Energy balance in a dry
passage could be written as,





ma;f 1
dia;f1
dQ1 h1 ta;f1  tw dA
2
(4) The energy balance in wet passages

Wet passage air involves the forced heat and mass exchange,
which leads to a change of enthalpy of the air within the passages.
The energy balance within the passages can be written as,


dQl  dQ2


ma;f2
dia;f2
2



dQ2 h2 ta;f2  tw dA

(5)



dQl hm rw;a2  ra;f2 g dA

(6)

The air ow within the pipes remains in a laminar ow state


when ReD < 2300 and becomes turbulent ow when ReD > 4000.
Due to the passage size and air velocity, the air ow within the

The level of moisture in the working air could be calculated as


follows:




ma;f2
dwa;f2 hm rw;a2  ra;f2 dA
2

(1)

(2) The general energy balance within the element in Fig. 4 can be
expressed as:

(4)

where, for the forced convective heat and mass transfer occurring
in the wet passages,

(1) The mass balance in the wet channel

(3)

Fig. 5. The calculating grids/meshes.

C. Zhan et al. / Building and Environment 46 (2011) 657e668

661

Fig. 6. Experimental validation e supply air temperature. (a) Case 1. (b) Case 2.

passage is considered to be laminar. In this case, the thermal entry


length for laminar ow can be calculated as follows [20]:

The theoretical energy efciency of the system can be dened as


the ratio of cooling capacity to fan power consumption:

L
0:05Re  Pr
d

COP

(7)

For both entry region and fully developed ow conditions, the


Nusselt number can be calculated using the following equation:

!0:14


Re  Pr 1=3 ha;f
Nu 1:86
hw;a
L=d

(8)

The thermal entrance Nusselt numbers are higher than those for
the fully developed case. For the developing ow conditions in the
entry region, the Nusselt number can be calculated as below.

 

0:0688  Re  Pr dL
Nu 3:66
h
 i2=3
1 0:04 Re  Pr dL

(9)

The mass transfer coefcient between wet passage air ow and


the wet surface of the wall may be calculated using the following
equation:

h
rcp Le2=3
hm

(10)

The mathematical expression for wet-bulb effectiveness can be


written as follows

3wb

tdb;wk;in  tdb;su
tdb;wk;in  twb;wk;in

(11)

It should be stressed that the wet-bulb effectiveness is the major


parameter for evaluating the performance of the exchanger and
cooler, which represents the extent of the outlet air temperature to
approach its relative wet-bulb of the inlet air.

f0

(12)

It should be stressed that cost of the water consumed in the


system was neglected owing to its minor value compared to the
cost of the electricity.
Cooling capacity, 40, can be expressed as:

f0 mpt iwk;in  ipt

(13)

The theoretical fan power, P, can be written as:

P Dpwk Vwk Dppt Vpt

(14)

It should be emphasized that the energy efciency obtained


from the simulation is an ideal value, which involves use of the
theoretical fan power. Actual fan power will be 120e170% of the
ideal value, leading to a drop in the calculated efciency by 60e80%
[21]. It should be noted that in this paper all the subsequent gures
related to COP are ideal rather than practical values.
By solving the above coupled differential equations, values of
temperature and moisture content of the air at each single element
can be obtained, which results in a solution of the wet-bulb effectiveness. A computer model incorporating the above equations was
developed in EES, by employing the nite-element approach. Fig. 5
presents the air ow prole across a plate wall; the upper side of
the plate is arranged with wet passages, and the underside the dry
passages. In terms of a single element, the following assumptions
were made: (1) each element has a uniform wall surface temperature; (2) at the inlet and outlet of the dry or wet channel, the air
has a uniform temperature and moisture content.
The trial computation results showed that under the specied
conditions a change of 0.02  C (0.2%) in the supply air temperature
resulted from increasing the mesh grid from 12  12 to 24  24. The

Fig. 7. Experimental validation e wet-bulb effectiveness. (a) Case 1. (b) Case 2.

662

C. Zhan et al. / Building and Environment 46 (2011) 657e668

Fig. 8. Experimental validation e supply air moisture content. (a) Case 1. (b) Case 2.

Table 1
Operational conditions of the TAC-150 air cooler.
Inlet air dry-bulb
temperature ( C)

30

Inlet air relative


humidity (%)

50

Inlet air wet-bulb


temperature ( C)

22.0

Supply air
ow rate
(m3/h)

(kg/s)

150

0.0475

signicant increase in computing time for the 24  24 grid was not


considered a rational modelling burden for the relatively small
temperature change. The mesh grid of 12  12 was considered to
provide sufcient accuracy for engineering applications and was,
therefore, adopted in the model set up. The Newton Iterative
method was used to solve a set of 6,357 equations in relation to
uid ow and heat and mass transfer within the passages of the
heat exchanger.
4. Validation of the model accuracy using the existing
experimental data

Modelling and experimental data regarding the supply air


temperature, wet-bulb effectiveness and air moisture are shown
respectively in Figs. 6e8. For case 1, the difference between
experimental and simulated supply air temperature is
0.69e1.28  C, but for case 2 much closer agreement is shown with
a difference of as small as 0.02e0.05  C. For the two cases, the
highest deviation in simulated to experimental supply air
temperature is 3.4%. Case 2 shows greater agreement between
experimental and simulated wet-bulb effectiveness, with case 1
showing a difference in a range of 7.3%e9.4% and case 2 a difference
of 0.2%e0.4%.
In theory, the moisture content of supply air should be equal to
that of the inlet air. Fig. 7 shows good agreement to this in the
experimental data for case 2, and shows supply air moisture to be
clearly higher than inlet air moisture for case 1. With the reported
unsteady behaviour in the experiment of [22] and the indication
that the experimental data of case 2 is more accurate, the level of
agreement shown between the experimental results and simulation is considered to offer sufcient condence in the modelling
process for the IECs air ow, heat and mass transfer.
5. Simulation results and analyses

The model was set to the same operating conditions as for


experimental cases 1 & 2 (i.e. the same inlet air parameters and
ow rates). Comparison between EES modelling results and testing
data obtained from [22] was carried out, and differences between
the results were analysed. This analysis established the accuracy of
the model in predicting the performance of the real system.
The experimental cases (1 and 2) selected for validation refer to
the testing to a ISAW TAC-150 cooler, as shown in Fig. 3, which was
carried out by Qiu [22] at the standard environmental chamber
conditions. Case 1 was carried out at the controlled inlet air
conditions of 35% RH, 25 to 40  C dry bulb and 130 m3/h air ow
rate; whereas case 2 was at the condition of 50% RH, 25 to 40  C dry
bulb and 130 m3/h air ow rate. During the testing, T-type thermocouple probes and PT100 humidity sensors were installed to
measure the temperatures and relative humidity of air ow,
whereas the Testo 425 type handheld hotwire anemometer used to
measure the air velocity which resulted in calculation of the air
ow rate. All these measurement sensors were linked to a DT500
data logger and a computer for data recording and analyses. A
programme was established in the Datatakers software to control
the Datataker to scan signals and report them at 5 second intervals.
The data was also saved at the same intervals as well.

5.1. Start-up operation and system performance


After being validated with the experimental data, the model was
utilized to investigate the effect of various operational factors to
system performance. The recommended operating conditions of
the TAC-150 unit (see Table 1) were used to set the initial conditions
of the model [17]. The simulation indicated that the system can
achieve a cooling capacity of 200 W. Further information resulting
from the simulation is presented in Table 2.
For this operating condition, the temperature proles of dry air,
wet air and the exchanging wall are presented in Fig. 9aec, the heat
ux in Fig. 10aec, and the moisture content prole of the dry and
wet air in Fig. 11a, b.
In Fig. 9a, it can be seen that the temperature of supply air in dry
channels (Nos. 3e6, referring to Fig. 3b) decreases along its direction of ow, and the temperature of working air in the dry channels
(Nos. 1 and 2, referring to Fig. 3b) has a bigger drop because of the
reduction of air mass ow along the way. Wet air temperature
shows a different trend. As shown in Fig. 9b, the temperature of
working air in the wet channels of Nos. 1e3 are lower than those of
Nos. 4e6. Moreover, both of them initially fall before rising again

Table 2
Results of simulation.
Supply air ow
rate (m3/h)

Average air speed in


dry channel (m/s)

Wet-bulb
effectiveness (%)

Supply air average


temperature ( C)

Cooling
capacity (W)

Total inlet air


ow rate (m3/h)

Exhaust air
ow rate (m3/h)

150

1.77

51.1

25.8

200.4

228

78

C. Zhan et al. / Building and Environment 46 (2011) 657e668

663

Fig. 9. Temperature distribution across the exchanger plate. (a) Dry side (dry passages). (b) Wet side (wet passages). (c) Wall.

along the wet air ow direction. The temperature of working air in


wet channels Nos. 4e6 has a smaller increase to the end of the
channels - relative to the other three wet channels. Shown in
Fig. 9c, the general trend of wall temperature is that it decreases
along the dry-airs direction of ow and increases along the wetairs ow direction.
As shown in Fig. 10a, the convective heat transfer decreases
along the ow path of dry air as a result of the observed (see Fig. 9a
and c) decrease in the temperature difference between the dry
channel air and the wall. The heat transfer rate in dry channels Nos.
3e6 is higher than that of Nos. 1 and 2 as they have different air
mass ow rates (see Table 2).
Referring to Fig. 10b, the wet air is not initially saturated and has
a higher temperature than the wet wall close to the entrance of the
wet channels (comparison of Fig. 9b and c). This results in heat being
transferred to the water reserved on the wet side of the wall e leading
to the evaporation of the water. After travelling to a critical point, the
temperature of wet air is lower than that of the wet wall, so the
convective heat ux has become negative (as shown in Fig. 10b),
which means that the wet air picks up both sensible and latent heat
from the wall.
From Fig. 10c, the working air in wet channels 1e3 is pre-cooled
over a very short distance in dry channels 1e2 to a temperature
lower than that of the wall surface. Below this temperature the
working air absorbs heat from the wet wall.
The moisture content of the air in dry channels keeps constant,
shown in Fig. 11a. As shown in Fig. 11b, the moisture content of the
air in each wet channel increases along the main direction of ow of
the wet air. This is due to the continuous addition of moisture to the
air along the direction of ow. This results in a reduction in the
difference of moisture concentration between the wall surface and
the air, ultimately leading to a smaller driving force for evaporation
and a smaller associated heat ux (Fig. 10c). This behaviour is

consistent in each wet channel; the heat ux in all wet channels


decreases along the main direction of air ow. The air in each wet
channel does not reach saturation; instead the maximum relative
humidity is about 90%.
The results from the simulation allow determination of wetbulb effectiveness. Changing the values of air ow rate, ratio of
working-to-product air ow rates, temperature, and moisture
content allows different sets of simulation results to be obtained.
Further analyses of the results will allow the impact of these variables on cooling effectiveness to be determined.
5.2. Inlet air temperature impact
Varying the inlet air temperature between 20  C and 40  C while
all other parameters remain unchanged, the simulation was carried
out using the above established computer model and the results are
presented in Fig. 12a and b. A trend in increasing supply air
temperature, cooling capacity, wet-bulb effectiveness and COP of
the evaporative cooler coincides with increasing inlet air temperature. This is due to a higher inlet air temperature resulting in
a larger temperature difference between the inlet air and the water.
Although the temperature of inlet air has doubled (from 20  C to
40  C), the wet-bulb effectiveness increased from 46.5% to 56.3%,
and COP increased from 230 to 440. This shows the IEC is more
efcient at higher temperatures, suggesting it is more suited to
a high-temperature environment.
5.3. Air relative humidity impact
Keeping other parameters unchanged, the impact of relative
humidity of inlet air can be seen in Fig. 13a and b. When the relative
humidity of inlet air increased from 0.1 to 0.9, accordingly the wetbulb effectiveness increased from 48.3% to 54.1%, but COP and the

Fig. 10. Heat transfer rate across the exchanging plate. (a) Dry side. (b) Wet side. (c) On the wall.

664

C. Zhan et al. / Building and Environment 46 (2011) 657e668

Fig. 11. Moisture content distribution across the exchanging plate. (a) Dry side. (b) Wet side.

5.4. Impact of air speed

400

35

350

Temperature
Cooling capacity

30

300

25

250

20

200

15

150

10

100

50

20

24

28

32

36

40

5.5. Impact of ratio of supply-to-total inlet air mass ow rate


When the resistance of ow varies in either the dry or wet
channels, the resulting variation in air velocity between the dry and
wet channels will inuence the performance of the cooler. Under
this situation, the resistance variation is hard to determine, so the
inuence of the ratio of supply-to-total inlet air ow rate on COP
cant be given. For discrete total inlet air ow rates, the inuence of
different average velocity in wet channels and dry channels on both
wet-bulb effectiveness and cooling capacity were investigated,
through changing the ratio of supply air ow rate to total inlet air
ow rate from 0.1 to 0.9 by interval of 0.1, shown in Fig. 15a and b.
Fig. 15a shows wet-bulb effectiveness decreases as the considered ow rate ratio increases, but with different paths. The nature of
these paths are such that above 228 m3/h (at which point the supply
air ow rate reaches the specied maximum of 150 m3/h) the curve
becomes increasingly more convex, and below 200 m3/h it becomes
increasingly more concave. The wet-bulb effectiveness can even
achieve a value of 144% when the total inlet air ow is 50 m3/h with
a ratio of 0.1, but the corresponding cooling capacity is 38.8 W, which
apparently is not applicable and economic for a practical application.
At or below an inlet ow rate of 228 m3/h, Fig. 15b shows cooling
capacity reaches a maximum in the range of 0.5 to 0.6 for the ratio
of supply air to inlet air (decreasing towards the limits of the
considered ratio). Above 228 m3/h the curves display two maxima,
but with differing measures of cooling capacity. The maximum
cooling capacities for ow rates of 250 m3/h and 300 m3/h are

0.7

500
450

0.65

400

0.6

350

0.55

300

COP

40

Cooling capacity [W]

a
Supply air temperature [C]

Simulations were carried out to investigate the effect of air


speed on the performance of the cooler. When the total inlet air
ow rate increases, the working air ow rate and product air ow
rate will increase in proportion, so does the air speed in dry
channels or wet channels. Varying the total inlet air ow rate from
50 to 500 m3/h while keeping other parameters unchanged, the
simulation results are shown in Fig. 14a and b. Fig. 14a shows the
cooling capacity (W) to increase with increased ow rate (i.e.
increase in the air speed). The sensitivity of cooling capacity,
however, reduces at higher inlet air ow rates. Whilst the cooling
capacity increases (from 83 W to 270 W) the supply air temperature also increases, which may be a limiting factor to achieving
desired internal environmental conditions.
Fig. 14b shows a steep decline in both the wet-bulb effectiveness
and the COP (more so for COP) as inlet air ow rate increases. The
pressure drop across the cooler (due to increased ow rate) negatively impacts the COP e dropping from 3700 to 114 over the
considered range of inlet ow rate. If the average supply air

temperature is limited to a value of 26  C, the recommended


average air velocities in dry and wet channels should be less than
1.77 m/s and 0.7 m/s, respectively.

0.5

250
200

0.45
COP

150

0.4

Wet-bulb effectiveness

100

0.35

50
0

20

Inlet air temperature [C]

24

28

32

Inlet air temperature [C]


Fig. 12. Impact of inlet air temperature.

36

40

0.3

Wet-bulb effectiveness

cooling capacity dropped from 640 to 60 and from 388.2 W to


38.17 W, respectively. At RH 0.9 the temperature drop across the IEC
cooler was only 0.76  C, so showing the evaporative cooler to be
redundant in a humid climate. Therefore, the inlet air humidity should
not be higher than 65%, corresponding to a supply air temperature of
26  C, which is advised in the European standard of [23]. As the wetbulb effectiveness shows similar trends in both increasing air
temperature and increasing RH, but with different outcome to the IEC
performance, the wet-bulb effectiveness cannot be considered to
independently characterize the performance of the IEC.

C. Zhan et al. / Building and Environment 46 (2011) 657e668

b 900
800

15
10

Temperature
Cooling capacity

Wet-bulb effectiveness

Wet-bulb effectivenes

20

COP

700
600
500

COP

25

Cooling capacity [W]

Supply air temperature [C]

a 30

665

400
300
200

100
0
0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0
0

0.1

0.2

0.3

Inlet air Relative Humidity

0.4

0.5

0.6

0.7

0.8

0.9

Inlet air Relative Humidity


Fig. 13. Impact of inlet air relative humidity.

5.6. Impact of exchanger geometry e channel height, and length


Simulations were carried out to investigate effect of channel size
(height and length) on the COP, wet-bulb effectiveness, supply air
temperature and cooling capacity.
Varying the height from 2 to 20 mm while keeping other
parameters unchanged (as shown in Table 1), different sets of
simulation results were obtained. As shown in Fig. 16a and b, it can
be seen that both the wet-bulb effectiveness and cooling capacity

3
Total inlet air flow rate (m /h)

decrease with increasing channel height. However, a small channel


height results in increased ow resistance and decreased energy
efciency. The COP reaches its maximum of 578 when the channel
height is 12 mm, but the wet-bulb effectiveness and cooling
capacity are respectively 20% and 78.4 W, which are quite low. At
the given channel lengths as shown in Table 1, if the channel height
is greater than 4 mm, the supply air temperature will exceed the
indoor thermal comfort temperature of 26  C. A compromise
among the cooling effectiveness and COP suggests that the channel
height should not be greater than 4 mm.
Varying the length of the dry channel from 0.2 m to 2.4 m,
leading to change of the dimensionless length, i.e., ratio of length to
height, from 50 to 600, while keeping all other parameters
constant, simulation was carried out to investigate the impact of
channel length to cooling performance. As shown in Fig. 17a and b,
it can be seen that wet-bulb effectiveness and cooling capacity
increase with increasing dry channel length, whereas the COP and
supply air temperature present the adverse trend under this variation. When the dimensionless length exceeds 300, the variation
rates of the above parameters tend to slow down. Considering the
factors of material use and cooling performance, it is suggested that
the dimensionless length should be controlled to between 100 and
300. For the exchanger of 4 mm channel height, the length of the
dry channel should be in the range 0.4e1.2 m.
Similar simulation work to wet channel was carried out and the
results present similar trend of variation, as shown in Fig. 18a and b.
The wet channel dimensionless length should be in the range
100e300. For the exchange of 4 mm channel height, the length of
the wet channel should be in the range 0.4e1.2 m.

20
15
Temperature
Cooling capacity

3500

0.9
0.8
COP
effectiveness Wetbulb

2500

0.7
0.6

2000

0.5
0.4

1500

0.3

1000

0.2

500

0
0

4000

3000

COP

25

Cooling capacity [W]

Supply air temperature [C]

30

10

3
Total inlet air flow rate (m /h)

0.1

0
0.5

1.5

2.5

3.5

0.5

0.17

0.33

0.50

0.66

0.82

0.98

1.14

1.30

1.5

2.5

3.5

Air speed in Dry channels (m/s)

Air speed in Dry channels (m/s)


1.46

1.62

0.17

Air speed in Wet channels (m/s)

0.33

0.50

0.66

0.82

0.98

1.14

Air speed in Wet channels (m/s)

Fig. 14. Impact of air velocity.

1.30 1.46

1.62

Wet-bulb effectiveness

within the lower half of the considered ratio range, for 400 m3/h
and 500 m3/h the maximum cooling capacities move towards the
upper limit of the considered ratio range.
Based on Fig. 15b, the cooler has an optimal (theoretical) ow
rate ratio for different inlet ow rates to achieve maximum cooling
capacity. As total inlet air ow increases (up to 250 m3/h), the ratio
of supply-to-inlet air ow rate decreases to maintain maximum
cooling capacity.
At the specied conditions (total inlet air ow rate is 228 m3/h,
shown in Table 2), the practical ratio of the supply-to-inlet air ow
rate is 0.657, which result in a deviation of 7% for cooling capacity
from the best ratio value of 0.5 (Fig. 14b). Moreover, the wet-bulb
effectiveness is decreasing from 70.52% to 50.4% when the
considered ratio rises from 0.5 to 0.657, which means that the
supply air temperature is higher according to Eq. (11). Therefore,
within the limits of the considered test conditions, the IEC unit can
be said to be operating within 93% of the maximum cooling
capacity, and within 71% of the wet-bulb effectiveness corresponding to the maximum cooling capacity.

666

C. Zhan et al. / Building and Environment 46 (2011) 657e668

Fig. 15. Impact of the supply-to-total air ratio.

450

26

400

25

350
Temperature

24

Cooling capacity

300

23

250

22

200

21
0

150

100

200

300

400

500

600

Dimensionless dry channel length (L/H)

500

1.1

450

400

0.9

350

COP
Wet-bulb effectiveness

300

COP

27

Cooling capacity [W]

Supply air temperature [C]

250

0.7
0.6

200

0.5

150

0.4

100
50
0

0.8

Wet-bulb effectiveness

Fig. 16. Impact of air passage height.

100

200

300

400

500

600

0.3

Dimensionless dry channel length (L/H)

Fig. 17. Impact of dry channels length.

5.7. Comparison between the new (M-cycle) exchanger and


conventional (cross-ow) exchanger

(2) The same inlet air parameters (shown in Table 1).


(3) The same working/product air ow rates.

Based on the optimised geometrical sizes of the exchanger, i.e.,


4 mm channel height and 1.2 m channel length, comparison was
carried out to examine the difference in the performance of the
new type of M-cycle exchanger and conventional cross-ow
exchanger. To enable the comparison, it is assumed:

The simulation results are listed in Table 3. In general, the new


type of exchanger achieved much higher cooling performance
than the conventional cross-ow exchanger. Under the given
conditions, the supply air temperature of the new exchanger is
1.4  C lower than that in the conventional exchanger; the wetbulb effectiveness is 15.7% higher than that in the conventional
exchanger. As a result, the system cooling capacity can be
increased by 62 W which is 16% higher than the conventional
exchanger.

(1) Both exchangers have the same effective heat/mass transfer


area. However, the new type of exchanger has extra working air
pre-cooling space.

26

400

25

350
Temperature

24

Cooling capacity

300

23

250

22

200

21
0

150

100

200

300

400

500

600

667

b 500

1.1

450

Dimensionless wet channel length (L/H)

400

0.9

350

0.8

COP

300

Wet-bulb effectiveness 0.7

250

0.6

200

0.5

150

0.4

100
50
0

Wet-bulb effectiveness

450

COP

27

Cooling capacity [W]

a
Supply air temperature [C]

C. Zhan et al. / Building and Environment 46 (2011) 657e668

100

200

300

400

500

600

0.3

Dimensionless wet channel length (L/H)

Fig. 18. Impact of wet channels length.

Table 3
Comparison between improved tac-150 (M-Cycle) and conventional (cross-ow)
exchanger.

New exchanger
Conventional exchanger

Supply air
temperature ( C)

Wet-bulb
effectiveness

Cooling
capacity (W)

20.7
22.1

116.4%
99.7%

456.2
394.2

6. Conclusions
The research has established a computer model able to simulate
the thermal performance of a M-cycle cross-ow heat exchanger.
By using the model, detailed analyses into relation between the
cooling (wet-bulb) effectiveness, system COP and air ow/
exchanger operational parameters were undertaken. These led to
suggestion to the most favourite operating conditions including air
velocity, inlet air temperature and humidity and ratio of workingto-product air, and optimised exchanger conguration e.g. channel
length and height etc. The model was also validated by the published experimental data which indicated that the sufcient accuracy in simulation could be obtained. The model is therefore
suitable for use in design of the indirect evaporative cooling system
and prediction of the system operational performance. This work
will help with enhancing the energy efciency of this kind of
system, exploring its market share in building air conditioning
sector, and thus contribute to achieve the global targets in energy
saving and carbon reduction measures. Furthermore, extension of
the model could be used in simulating the performance of other
types of exchanger for indirect evaporative cooling, e.g., counter
ow exchanger with no divisional holes along the air ow paths,
which is part of the follow-up project to develop a commercial
exchanger and will be detailed in a separate paper.
It is indicated that (1) the new type of M-cycle heat and mass
exchanger is able to achieve 16.7% higher cooling effectiveness
compared with the conventional cross-ow heat and mass
exchanger for the indirect evaporative cooler; (2) a higher channel
air velocity in the new exchanger results in a relatively lower wetbulb effectiveness and system COP, though the smaller system size
is considered to be spatially and economically benecial to
potential users. The recommended average air velocities in dry
and wet channels should be less than 1.77 m/s and 0.7 m/s,
respectively; (3) at the specied conditions, the optimum ratio (in
terms of cooling capacity) between the exhaust and supply air
ow rate is 1:1; (4) reducing the channel height led to an increase
in cooling capacity or wet-bulb effectiveness and decrease of the
system COP. A compromise among these performance indices

suggests that the channel height should be set to no more than


4 mm; (5) increasing the channel (both dry and wet) length led to
improved cooling effectiveness but reduced the system COP. It is
suggested that the dimensionless channel length should be
controlled to between 100 and 300. For the exchanger of 4 mm
channel height, both dry and wet channel lengths should be in the
range 0.4e1.2 m; and (6) the system performance is highly
dependent on the climatic conditions where it is applied. For the
given inlet air condition, the system can achieve 4.22 kW of
cooling capacity as per kg/s of supply air, which is 50% of the
maximum capacity the system can achieve.
It should also be stressed that the above analysis is based on
a small size unit (TAC-150) for domestic use. This type of exchanger
can be extended to the large-scale central air handle unit which will
result in signicantly higher energy saving and carbon reduction
potential.
Acknowledgement
The authors would like to acknowledge the nancial support
provided for this research by the EU FP7 Marie Curie International
Incoming Fellowship (PIIF-GA-2008-220079).
References
[1] CIBSE knowledge series: sustainable low energy cooling: an overview. Plymouth PL6 7PY, UK: Latimer Trend & Co. Ltd; Sept 2005.
[2] Zhao X, Duan Z, Zhan C, Riffat SB. Dynamic performance of a novel dew point
air conditioning system for the UK climate. International Journal of Low
Carbon Technology 2009;4(1):27e35.
[3] Cerci Y. A new ideal evaporative freezing cycle. International Journal of Heat
and Mass Transfer 2003;46:2967e74.
[4] Costelloe B, Finn D. Indirect evaporative cooling potential in air e water
systems in temperate climates. Energy and Buildings 2003;35:573e91.
[5] Sethi VP, Sharma SK. Survey of cooling technologies for worldwide agricultural greenhouse applications. Solar Energy 2007;81:1447e59.
[6] Goshayshi HR, Missenden JF, Tozer R. Cooling tower e an energy conservation
resource. Applied Thermal Engineering 1999;19:1223e35.
[7] Maheshwari GP, Al-Ragom F, Suri RK. Energy saving potential of an indirect
evaporative cooler. Applied Energy 2001;69:69e76.
[8] Madhawa Hettiarachchi HD, Golubovic Mihajlo, Worek WM. The effect of
longitudinal heat conduction in cross ow indirect evaporative air coolers.
Applied Thermal Engineering 2007;27:1841e8.
[9] Zhao X, Liu Shuli, Riffat SB. Comparative study of heat and mass exchanging
materials for indirect evaporative cooling systems. Building and Environment
2008;43:1902e11.
[10] Stoitchkov NJ, Dimitrov GI. Effectiveness of crossow plate heat exchanger for
indirect evaporative cooling. International Journal of Refrigeration 1998;21
(6):463e71.
[11] Ren Chengqin, Yang Hongxing. An analytical model for the heat and mass transfer
processes in indirect evaporative cooling with parallel/counter ow congurations. International Journal of Heat and Mass Transfer 2006;49:617e27.
[12] Maclaine-cross IL, Banks PJ. A general theory of wet surface heat exchangers
and its application to regenerative cooling. ASME Journal of Heat Transfer
1983;103:579e85.
[13] Erens PJ, Dreyer AA. Modelling of indirect evaporative coolers. International
Journal of Heat Mass Transfer 1993;36(1):17e26.

668

C. Zhan et al. / Building and Environment 46 (2011) 657e668

[14] <http://www.idalex.com/technology/how_it_works_engineering_
perspective.htm>.
[15] Maisotsenko V. et al. Method and plate apparatus for dew point evaporative
cooler, United State Patent 6,581,402; June 24, 2003.
[16] Coolerado, CooleradoHMX (heat and mass exchanger) brochure. Arvada,
Colorado, USA: Coolerado Corporation; 2006.
[17] ISAW. Natural air conditioner (heat and mass exchanger) catalogues. Hangzhou, China: ISAW Corporation Ltd.; 2005.
[18] Klein SA. Engineering equation solver (EES) for Microsoft windows operating
systems, professional versions, Madison USA, WI: F-Chart Software. Available
from: http://www.fchart.com.

[19] Welty JR, Wicks CE, Wilson RE, Rorrer G. Fundamentals of momentum, heat,
and mass transfer. USA: John Wiley & Sons Inc.; 2000. 500e589288e326.
[20] John H. Lienhard V. A heat transfer textbook, Cambridge, Massachusetts:
Phlogiston Press, pp. 352.
[21] CIBSE guide B2, ventilation and air conditioning. Norwich, UK: Page Bros
(Norwich) Ltd.; 2001. 5e285e25.
[22] Guoquan Qiu. A novel evaporative/desiccant cooling system. Dissertation for
the degree of Doctor of Philosophy, The University of Nottingham; June 2007.
pp. 41, 60.
[23] CEN prEN15251, Criteria for the indoor environment including thermal,
indoor air quality, light and noise; 2005.

Das könnte Ihnen auch gefallen