Accepted Manuscript: Applied Thermal Engineering

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

Accepted Manuscript

Off-design Performance of a Louvered Fin-Tube Heater Core in an Automotive


Climate Control System

T.I. Bhaiyat, S. Schekman, H.Y. Lim, Y.H. Jeon, T. Kim

PII: S1359-4311(19)30118-8
DOI: https://doi.org/10.1016/j.applthermaleng.2019.114155
Article Number: 114155
Reference: ATE 114155

To appear in: Applied Thermal Engineering

Received Date: 16 January 2019


Revised Date: 9 July 2019
Accepted Date: 21 July 2019

Please cite this article as: T.I. Bhaiyat, S. Schekman, H.Y. Lim, Y.H. Jeon, T. Kim, Off-design Performance of a
Louvered Fin-Tube Heater Core in an Automotive Climate Control System, Applied Thermal Engineering (2019),
doi: https://doi.org/10.1016/j.applthermaleng.2019.114155

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Off-design Performance of a Louvered Fin-Tube Heater Core in an
Automotive Climate Control System
T.I. Bhaiyata, S. Schekmana, H.Y. Limb, Y.H. Jeonb, T. Kimc,∗
aSchool of Mechanical, Industrial, and Aeronautical Engineering, University of the Witwatersrand, Johannesburg,
South Africa
bResearch Center, Hanon Systems, 95, Sinilseo-ro, Daedeok-gu, Daejeon, Korea
cCollege of Aerospace Engineering, Nanjing University of Aeronautics and Astronautics, Nanjing, Jiangsu Province, China

Abstract
Louvered fin-and-tube heat exchanger cores, operating inside of an automotive climate control system (ACCS), are
used to control temperature and humidity in a passenger cabin. The thermal performance of these cores is typically
sensitive to flow features inside the core, which have been well studied for uniform flow fields as well as some non-
uniform and inclined flow-fields upstream of the core. The compact geometry of an ACCS unit results in the flow-fields
upstream of the core being different from those used in previous studies. When measuring the performance of a louvered
fin-and-tube heater core inside a commercially manufactured ACCS unit, up to a 60% reduction in the core’s thermal
performance compared to the performance when exposed to a uniform upstream flow is detected. Such a drastic
performance reduction has not been reported in the literature. As such a systematic breakdown is made to elaborate how
some key and specific ACCS unit features designed for the sake of compactness could play a detrimental role in reducing
the overall thermal performance of a given louvered fin-and-tube heater core. By mimicking the actual ACCS features the
performance deterioration in the commercial unit could be largely replicated. The individual contributing effects of the
various fluidic features such as flow separation and recirculation are determined. Only 50% of the measured 60%
deterioration, however, could be accounted for, suggesting that other ACCS unit properties not assessed in detail in this
study may also adversely affect the core’s performance.

Keywords: Automotive climate control system, Design performance, Flow separation, Heater, Louvered fin-tube
core, Off-design performance

Nomenclature

Aa Total air-side heat transfer surface area m2


Aw Total water-side heat transfer surface area m2
At Tube wall conduction area m2
b Fin height mm
C Heat capacity rate W/K
Cr Heat capacity rate ratio -
cp Specific heat capacity J/kg-K
δ Fin thickness mm
D Ideal flow length mm
Dt Tube hydraulic diameter mm
d Turbine meter diameter mm
η Flow efficiency -
η0 Surface efficiency -
ηf Fin efficiency -
ε Effectiveness -
h Heat transfer coefficient W/m2-K
H Core height mm
Ht Tube height mm
j Colburn j-factor -
k Thermal conductivity W/m-K
lp Louver pitch mm
lh Louver height mm
1
ll Louver length mm
Lt Tube depth mm
Lf Fin depth mm
ll Louver length mm
𝑚 Mass flow rate kg/s
N Actual flow length mm
NTU Number of Transfer Units -
Nu Nusselt number -
Pr Prandtl number -
Pt Tube pitch mm
Pf Fin pitch mm
𝑞 Heat transfer rate W
Q Flow rate L/s
ρ Density kg/m3
Re Reynolds number -
Ri Tube edge radius mm
R Thermal performance ratio -
r Tube rib thickness mm
St Stanton number -
S1 Deflection louver length mm
S2 Turnaround louver length mm
θ Louver angle o

t Tube wall thickness mm


T Temperature oC

u Local fluid velocity m/s


um Mean fluid velocity m/s
U Overall heat transfer coefficient W/m2-K
W Core width mm
Wt Tube width mm

Subscripts
a Air
ACCS Automotive Climate Control System unit conditions
e Equivalent-ACCS unit conditions
h Half-ACCS unit conditions
in Inlet
min Minimum
max Maximum
out Outlet
Ref Reference conditions
t Tube
w Water

1. Introduction
Automobiles are equipped with an automotive climate control system (ACCS), consisting of heat exchanger cores as
illustrated in Fig. 1(a), which ensures the passengers’ comfort in the vehicle cabin. In this system, an evaporator core and
a heater core are sequentially arranged with a centrifugal fan placed upstream of the two cores blowing air at ambient
conditions through the cores. The cores typically consist of a corrugated louvered fin-and-tube configuration (Fig. 1(b),
(c)), due in part to it being lightweight and having well-established manufacturing process [1]. To heat up the cabin, hot
water from an engine block enters the inlet header of the heater core, where it is distributed to flat tubes within the core,
before exiting the outlet header without undergoing any phase-change. The temperature of the air-stream that passes over the
heater core (through the fins) is then increased while the water temperature inside the tubes is decreased due to convective
cooling by the air-stream, separated by the tube walls. The louvers (which are angled slats cut through the fins)
allow for heat transfer augmentation, as large temperature gradients can be maintained between the air-stream and the fin
2
surface due to periodic destruction and growth of the boundary layer [2]. The extent of such boundary layer behavior, and
thus the thermal enhancement capacity of the louvers, is dictated by the degree to which the louver-directed path (shown
in Fig. 1(c)) is followed; duct-directed flows dramatically reduce the thermal performance of the core [3]. The thermo-
fluidic performance of louvered-fins is generally well established for uniform, unidirectional air flows perpendicular to
the core and without any disturbances or flow area changes upstream and downstream of the core as shown in Fig. 2(a)
(referred to in this study as “design conditions”). Several correlations predicting the performance of louvered fins under
these conditions have been proposed in the literature [4-9].

(a)

(b)

(c)
Figure 1. Automotive climate control system: (a) Overall flow routes for internal and external working fluids in an
ACCS [18], (b) Typical heater core configuration [19], (c) Nomenclature of a louvered fin in a plan view.
3
Various off-design conditions are nevertheless encountered in typical air-conditioning (AC) systems, such as ACCS
units, mainly due to space constraints during their integration into the systems. The heater core positioned in a typical
ACCS unit is shown schematically in Fig. 2(b) where the geometrical features inside the unit substantially deviate from
the typical design setup of Fig. 2(a). Upstream of the heater core is a wedge obstruction to the flow, while downstream of
the core is a vertically oriented duct. As a result, the assumption of design conditions is not always expected to be valid
for performance predictions made inside such an ACCS unit and, therefore, an off-design study is required.
One of the common off-design conditions encountered in AC systems is that of geometry-induced macro-scale air-
flow non-uniformity at the inlet of the heat exchanger [10, 11]. The effects of various air-flow distributions on thermal
performance of fin-tube cores have previously been studied by blocking the air-flow path through the core, as shown by
Figs. 3(a,b), which results in a non-uniform velocity profile entering the core, as shown schematically in Fig. 3(c). T’Joen
et al. [12] found that linear and quadratic velocity distributions resulted in 8.2% and 3.4% reductions in the overall heat
transfer coefficient, respectively, for an inclined fin-tube heat exchanger. Datta et al. [13] considered four types of
blockages on a louvered fin-tube automotive condenser core and concluded that the side blockage results in the most
severe (up to 8.16%) reduction in cooling capacity. Blecich [14] found up to 30% reduction in thermal effectiveness for
extremely non-uniform coolant stream profiles incident on a fin-tube heat exchanger with single phase flow inside the
tubes, and 5-10% thermal performance reduction for moderate non-uniformities encountered in typical air-conditioning
systems. While the foregoing studies have successfully shown thermal performance degradation due to macro-scale air-
flow non-uniformity, the artificially created non-uniformities considered were essentially one-dimensional (i.e., non-
uniformity was introduced in only a single velocity component), downstream flow changes were not accounted for, and
the bulk flow direction remained perpendicular to the core, as shown in Fig. 3(c).
Another common off-design condition is encountered in several central AC systems where, due to space limitations,
either two cores are arranged with an “A” (or “V”) configuration or a single core is inclined relative to the oncoming air-
stream as shown in Fig. 3(d). Since either of these configurations may result in inlet coolant streams with velocities non-
perpendicular to the core, Kim et al. [15] tested the effect of inclining a full-scale louvered fin-tube core up to ±60o, from
the design condition, on heat transfer performance in the range of incoming velocities 0.6 - 2 m/s and found an insignificant
effect regardless of whether ducting was present upstream or downstream of the core. On the other hand, Henriksson [16]
carried out similar experiments for a higher range of incoming velocities (1.9 – 9.8 m/s) and found that the heat transfer
performance of a louvered fin-tube core increased (by as much as 36%) when the inclination relative to the design
condition increased (up to 60o). At low incoming velocities, an insignificant effect of inclination angle on heat transfer
was observed [16]. Both studies, however, did not include the influence of an upstream obstruction or air-flow non-
uniformity, and the downstream ducting was always aligned parallel to the bulk flow direction whereas the ducting in the
ACCS unit is aligned perpendicular to the bulk flow direction (Fig. 2(b)), and is expected to be similar to flow through a
90o bend [1, 17].
Certain off-design conditions typically encountered in engineering practice (such as non-uniform incident flows,
external coolant stream’s inclination, and downstream ducting) have thus been studied, through temperature
measurements, individually. Since each individual off-design condition has been reported to either increase or decrease
thermal performance, it is expected that a combination of off-design conditions will result in a superimposed thermal
performance deviation. However, to the best of the authors’ knowledge, there is no systematic study assessing these
combined off-design conditions, which exist inside the ACCS unit, altering the thermal performance of louvered fin-tube
heater cores. It is nonetheless important to make accurate performance predictions which influence the heater core design.
To this end, the present study firstly characterizes the differences in thermal performance between a louvered fin-tube
heater core in a uniform external coolant stream (i.e., at design conditions) and the core placed inside a selected ACCS
unit (i.e., at full off-design conditions). Thereafter, several systematic fluidic and temperature measurements are carried
out using wind-tunnels which are progressively modified to mimic conditions inside the ACCS unit.

(a)

4
(b)
Figure 2. Schematic representation for the air path upstream and downstream of a heater core: (a) At design conditions,
(b) At a combination off-design conditions, inside a typical ACCS unit.

(a)

(b)

(c)

(d)

Figure 3. Typical off-design conditions previously studied (a) Non-uniform air-flow distribution using a blockage screen,
(b) typical blockage screens studied by Datta et al. [13] and Blechich [14], (c) effect of a blockage screen on the velocity
profile of air entering a core, (d) typical central AC system configuration with “A” type and inclined one-slab fin-tube
cores.
5
2. Experimental Details
2.1. The heater core and hot water supply system
A commercially manufactured louvered fin-tube heater core with a width W = 225 mm and height H = 160 mm (Fig.
4) has been used in the current study. Detailed dimensions of the fins and flattened tubes are listed in Table 1 (based on
the parameters illustrated in Fig. 1(c)).
To measure the thermal performance of the selected heater core, a closed loop water heating system was constructed
as shown in Fig. 4. A 12 V brushless DC pump circulates water through electrical heaters and the louvered fin-tube core
at a constant volume flowrate of Qw = 0.11925 L/s. Each electrical heater (with maximum power rating indicated in Fig.
4) was connected to an independent voltage regulator to adjust the electrical power input. Similar to the method adopted
by Shinde [20], a 2 kW urn heater was installed to provide a backup heating system during testing and to minimize
variations of the inlet water temperature, maintained at Tw,in = 45 °C. Using this approach, the experimental setup was
capable of temperature variations less than 1 °C within the range of air volume flowrates, 0 ≤ Qa ≤ 160 L/s.

Figure 4. Schematic showing supply of hot water through the header tubes of a louvered fin-tube heater core with water
temperature and flow measurement instruments.

Table 1: Dimensions of the fins and tubes for the louvered fin-tube heater core test sample used in this study.
Component Parameter Value (+-0.01 mm / +-0.5o)
Lt 27.0 mm
Pt 7.62 mm
Tube Ht 1.28 mm
t 0.21 mm
Ri 0.64 mm
Lf 27.0 mm
b 6.15 mm
Pf 0.80 mm
S1 1.85 mm
Fin S2 1.85 mm
lp 0.93 mm
ll 5.09 mm
θ 30°
δ 0.04 mm

6
2.2. Test rig for off-design performance in an automotive climate control system (ACCS) unit
2.2.1 An ACCS unit configuration
For the off-design thermal performance characterization, the louvered fin-tube heater core furnished with the hot water
supply system was placed inside a complete climate control system unit as shown in Fig. 5. A centrifugal fan blows air
through a diffuser towards a heat transfer section containing a louvered fin-tube evaporator and heater core. A T-type
thermocouple was placed between the outlet of the centrifugal fan and the heat transfer section to measure the ambient air
temperature Ta,in entering the heater core.

(a)

(b)
Figure 5. Typical automotive climate control system (ACCS) unit for a passenger car: (a) Frontal view, (b) Side view.

2.2.2. Flowrate measurements


To measure the volume flowrate of the air-stream Qa, a pre-calibrated digital rotating vane anemometer from
AirflowTM, UK (LCA501) was installed at the inlet of the centrifugal fan. There was a fabric filter between the rotating
vane anemometer and the fan’s intake as a default setup in the ACCS unit. The centrifugal fan was connected to a DC
power supply unit (RPE-4323) capable of providing an output power range of 0 – 32V and 0 – 6A. For an output voltage
from the power supply unit fixed at 12V, the input current was regulated by the fan speed control circuit’s set electrical
resistances, which provides step settings for the fan speeds as (i) Low, (ii) Medium, (iii) Medium-High, and (iv) High.
Corresponding power inputs to the fan have been measured and summarized in Table 2. Three different modes have been
tested. First, the downstream heat transfer section consisting of the evaporator and heater cores was removed to
characterize the original pumping capability of the centrifugal fan without any downstream pressure drop. Second, the
heat transfer section was attached but the flow passage through the heater core was blocked so that the flow passes through
the evaporator core only and then exits directly to the cabin of the vehicle. This corresponds to the ‘cooling mode’ of the
ACCS unit. Third, a dividing lever was flapped to direct the air-stream through both the evaporator and heater cores. In
this configuration, the evaporator core is typically inactive, and the ACCS unit is said to be in the ‘heating mode.’
Volume flowrate measurement results plotted in Fig. 6 show that the air volume flowrate was reduced in the presence
of the two cores, due to the increase in flow resistance when air passes through each core. Figure 6 thus provides a
representative range of air volume flowrates (60 L/s ≤ Qa ≤ 120 L/s) for use in the current study to characterize the thermal
performance of a core in the heating mode.

7
Table 2. Volumetric flowrate based on speed settings, measured in the present study.
Configuration Speed Setting Flowrate [L/s]
Low 57.54
Without downstream unit Medium 79.66
Medium-High 109.62
Low 49.31
Through evaporator only (cooling mode) Medium 68.48
Medium-High 99.69
Low 38.71
Through evaporator and heaters (heating Medium 55.04
mode) Medium-High 77.76
High 99.36

160
No downstream unit
140 Through evaporator only
Through evaporator and heaters
120

100
Qa [l/s]

80

60

40

20

0
0 10 20 30 40 50 60 70 80 90 100
Input power at 12V [W]
Figure 6. Volume flowrate supplied by the centrifugal fan at four step input power settings; Low, Medium, Medium-high
and High.

Figure 7. A blowdown rectangular wind-tunnel with a louvered fin-tube heater core in the test section for measuring
design performance.

2.3. Test rig for design performance in a uniform air-stream


For thermal performance characterization at design conditions, the louvered fin-tube heater core connected to a hot
water supply system was placed in a uniform air-stream as shown in Fig. 7.
A 0.9 kW centrifugal fan capable of providing volume flowrates 0 ≤ Qa ≤ 200 L/s blows air at ambient temperature
𝑇𝑎,𝑖𝑛 through a circular tube that is connected to a turbine flowmeter AirflowTM, UK (LCA501). Downstream of which, a
long rectangular flow developing section (width, W = 225 mm and height, H = 160 mm) with its length of 10 times the
turbine’s rotor diameter (d), was attached. Inside the channel, a honeycomb layer was positioned at 2d downstream from
the inlet of the flow developing section. The heater core, of width (W) and height (H), was removed from the ACCS unit
and was attached to the end of a rectangular flow channel of the same dimensions (W × H). At the outlet side of the core,
8
another short rectangular channel was placed to avoid a sudden expansion of the air-stream that leaves the heater core.
The air-stream temperature (Ta,in) was measured using a T-type thermocouple placed 4 times the tube depth (i.e., 4Lt)
upstream of the heater core. All thermocouple signals (Ta,in, Tw,in, and Tw,out) were recorded at steady-state by a data
acquisition system (Agilent 34830A).
To measure the level of turbulence of the air-stream in the wind-tunnel, a hot-wire anemometer (a straight single-
sensor probe, 55P11, Dantec Inc.) of thickness 5 μm was positioned at the mid-channel height and width and about 4
times tube depth (i.e., 4Lt) upstream, acquiring signals at a sampling frequency of 10 kHz for a duration of 3.3 seconds.
Calibration has been performed by the probe against a Pitot tube in the main-stream. The turbulence intensity has been
measured to be 2.5% at the air-side Reynolds number of Relp = 200.

2.4. Wind-tunnel modifications to mimic off- design conditions


In investigating the off-design configuration (Fig. 2(b)), the test setup shown in Fig. 7 was modified to replicate flow
conditions in the ACCS unit. Two primary modifications were made: placement of a wedge obstruction upstream of the
core and a sharp redirection of the downstream flow by ducting. To replicate the upstream conditions, a right-angled
wedge with base 1.85Lt, height 0.625H , and depth W was fabricated and inserted immediately upstream of a heater core
as shown in Fig. 8(a). Downstream of the heater core, there were no further obstructions. To replicate ducting downstream
of the core, a plate was inserted downstream of the core, which closes off the air escape path behind it. Air vents drilled
into the windtunnel ceiling direct the air-stream upward, as shown schematically in Fig. 8(b). The downstream
modification could be removed so as to test the effect of the wedge in isolation. The effect of any flow area contractions
or expansions upstream of the heater core, as present in the ACCS (Fig. 2(b)) were understood to have a negligible effect
on the performance and not considered in this study [15].

(a)

(b)
Figure 8. Schematic of a wind tunnel modified: (a) With a wedge obstruction placed immediately upstream of the heater
core and no downstream ducting, (b) With an upstream wedge obstruction and upward ducting downstream of the heater
core.

2.5. Data reduction process


The thermal performance of a compact heat exchanger core is commonly reported in terms of the Colburn j-factor (j),
which first requires determination of the air-side heat transfer coefficient (ha) and corresponding Nusselt number (Nulp).
To obtain ha, the measured variables (Tw,in, Tw,out, Ta,in, Qa, and Qw) were processed using the 𝜖-NTU method as carried out
by several researchers [7-9, 21]. The heat transferred from the water-stream, 𝑞 is given by:
𝑞 = 𝑚𝑤𝑐𝑝,𝑤(𝑇𝑤,𝑖𝑛 ‒ 𝑇𝑤,𝑜𝑢𝑡) (1)

where 𝑚𝑤 is the mass flowrate of the water, cp,w is the specific heat capacity, Tw,in and Tw,out respectively represent the
water temperatures at the inlet and outlet of the header tubes. The maximum theoretically achievable heat rate is:
𝑞𝑚𝑎𝑥 = 𝐶𝑚𝑖𝑛(𝑇𝑤,𝑖𝑛 ‒ 𝑇𝑎,𝑖𝑛) (2)

where Cmin = min(Ca, Cw) and Ta,in is the temperature of the air entering the core. Ca and Cw are the specific heat rates of
the air and water-streams, respectively defined as:

9
𝐶𝑎 = 𝑚𝑎𝑐𝑝,𝑎 = 𝜌𝑎𝑄𝑎𝑐𝑝,𝑎 (3)

𝐶𝑤 = 𝑚𝑤𝑐𝑝,𝑤 = 𝜌𝑤𝑄𝑤𝑐𝑝,𝑤 (4)

where 𝑚, cp, ρ, and Q respectively represent the mass flow rate, specific heat capacity, density, and volumetric flow rate;
the subscripts a and w refer to the air and water-streams respectively. The effectiveness (𝜖) follows as:
𝑞
𝜖=𝑞 (5)
𝑚𝑎𝑥

from which the number of transfer units (NTU) for single pass cross flow conditions with both fluids unmixed can be
solved iteratively using the explicit relation:
𝑁𝑇𝑈0.22
𝜖 = 1 ‒ 𝑒𝑥𝑝 (
{𝑒𝑥𝑝 [ ‒ 𝐶 (𝑁𝑇𝑈)0.78] ‒ 1}
𝐶𝑟 𝑟 ) (6)

where
𝐶𝑚𝑖𝑛 min (Ca, Cw) (7)
𝐶𝑟 = =
𝐶𝑚𝑎𝑥 max (Ca, Cw)

From the definition of NTU, one may calculate the overall heat transfer coefficient U based on the total air-side heat
transfer surface area Aa, as:
𝐶𝑚𝑖𝑛𝑁𝑇𝑈 (8)
𝑈= 𝐴𝑎

The air-side heat transfer coefficient ha was then calculated using a one-dimensional steady-state surface energy balance,
assuming negligible fouling resistance and constant thermo-physical properties:
1 1 𝐴𝑎 𝑡 𝐴𝑎 ‒ 1 (9)
0
{
ℎ𝑎 = 𝜂 𝑈 ‒ 𝐴 𝑘 ‒ 𝐴 ℎ
𝑡 𝑡
}
𝑤 𝑤

where Aw is the tube inner surface area, t is the tube thickness, kt is the tube thermal conductivity, hw is the water-side heat
transfer coefficient, and the surface efficiency ηo is given by
𝐴𝑓 (10)
𝜂 = 1 ‒ (1 ‒ 𝜂 )
0 𝐴𝑎 𝑓

where Af is the finned surface area. The fin efficiency ηf is:


tanh (𝑚𝑙)
𝜂𝑓 = 𝑚𝑙
(11)

where
2ℎ𝑎 (12)
𝑚= 𝑘𝑎𝛿 (1 + ) 𝛿
𝐿𝑓

and
( 𝑏2 + 𝑃2𝑓 ‒ 𝛿) (13)
𝑙= 2

where δ is the fin thickness, Lf is the fin length in the flow direction, b is the fin height, and Pf is the fin pitch.
For the water-side Nusselt number Nuw, an analytical correlation which assumes laminar fully developed conditions,
constant axial wall heat flux, and constant peripheral wall temperatures prevailing throughout the tube length was used
[11]:
2 3
𝑁𝑢𝑤 = 8.235(1 ‒ 2.0421𝛼 ∗ + 3.0853𝛼 ∗ ‒ 2.4765𝛼 ∗ (14)
4 5
+ 1.0578𝛼 ∗ ‒ 0.1861𝛼 ∗ )

where

10
𝐻𝑡 ‒ 2𝑡 (15)
𝛼∗ = 1
2(𝑊𝑡 ‒ 2𝑡 ‒ 𝑟)

Ht is the tube height, Wt the tube width, and r the center rib thickness. The water-side heat transfer coefficient hw can then
be extracted using the definition of the Nusselt number:
𝑘𝑤 (16)
ℎ = 𝑁𝑢
𝑤 𝐷𝑡 𝑤

where kw is the thermal conductivity of water and Dt is the hydraulic diameter of the tube.
Equations (9)-(12) were solved iteratively, starting with ηo = 1, and refining the heat transfer coefficient ha until
convergence of ηo was achieved to within 10 ‒ 6. Thereafter, the air-side Colburn j-factor was computed using the
definition:
𝑁𝑢𝑙𝑝 (17)
𝑗= 1/3
𝑅𝑒𝑙𝑝 𝑃𝑟

where Nulp is the air-side Nusselt number Nulp = ha lp/ka based on louver pitch (lp), Relp is the louver pitch-based Reynolds
number, and Pr is the Prandtl number for air.

2.6 Measurement uncertainty


If y = f (x1, x2,…, xn) then the uncertainty propagated by 𝑥𝑖 in the variable 𝑦 is given by:

2 2 2 (18)
Δ𝑦 = ( ∂𝑦 ∂𝑦
) ( ) ∂𝑦
(
∂𝑥1Δ𝑥1 + ∂𝑥2Δ𝑥2 + … + ∂𝑥𝑛Δ𝑥𝑛 )
where Δxi is the absolute uncertainty in xi. Eq. 18 has been used to estimate the overall experimental uncertainties
associated with the steady-state temperature and flowrate measurements. For the range of Reynolds numbers considered
in this study, the maximum relative uncertainty estimate for the Colburn j-factor (Δj) was 15.2%; consistent with the range
of uncertainties commonly encountered by researchers employing a steady-state measurement technique [20].

3. Discussion of Results
3.1. Design performance (reference)
A reference case, to determine the credibility of the test setup, was established by placing a louvered fin-tube heater
core in a uniform, unidirectional air-stream without any upstream or downstream obstructions (Fig. 7). The setup
conformed to the experimental conditions adopted by previous researchers [5, 8, 9]. Under these reference conditions,
steady-state temperature measurements were taken for the range of air volume flowrates, 36.6 L/s ≤ Qa ≤ 225 L/s, which
encompasses the typical range of flowrates encountered inside the ACCS unit (i.e., 60 ≤ Qa ≤ 120 L/s). The corresponding
Reynolds number range based on louver pitch is 58.8 ≤ Relp ≤ 361, for which the variation of Nusselt number (Nulp) over
this range is given in Fig. 9(a).
Fig. 9(b) compares the Colburn j-factors (j) obtained in the current study with correlations developed in previous
studies [5, 6, 8, 9]. In the range of Reynolds numbers 130 ≤ Relp ≤ 361, Achaichia & Cowell’s (1988) prediction is in line
with the current data, with the maximum deviation being 13.2% while the rest of the points lie within the reported
correlation precision of 10%. The other correlation valid in this range (Chang & Wang) has significant deviation from the
current data (up to 26%). The discrepancy could be due to the fact that the data bank used for Chang & Wang’s correlation
encompasses a variety of different louver fins including corrugated fins with triangular and rectangular channels,
corrugated fins with splitter plates, as well as plate fins. Nevertheless, important trends in the data are consistent with the
previous studies, and it is assumed that comparisons between off-design and design conditions can reliably be made
against these results.

11
15

10

Nulp
5

0
0 100 200 300 400 500
Relp
(a)
1
10
Current study
Achaichia & Cowell (1988)
0 Chang & Wang (1997)
10
Dong et al. (2007)
Shinde & Lin (2017)

10-1
j

-2
10

-3
10 1 2 3 4
10 10 10 10
Relp
(b)
Figure 9. Design performance of a louvered fin and tube heater core: (a) Nusselt number (Nulp) variation with Reynolds
number (Relp); (b) Colburn j-factor varying with Reynolds number (Relp) relative to common correlations.

3.2. Off-design performance in an automotive climate control system unit


By placing the test core inside a complete automotive climate control system (ACCS) unit (Fig. 5), the thermal
performance of the core was characterized under the full combination of off-design conditions, i.e., with non-uniform,
multidirectional, obstructed flow upstream of the core and ducting (which imposes flow redirection and area change)
downstream of the core. This configuration is referred to as the “full off-design” case.
The Nusselt numbers obtained in the full off-design case (Nulp(ACCS)), for the range of air volume flowrates produced
by the unit’s centrifugal fan at its four step settings (Low, Medium, Medium-High, and High) in the heating mode,
corresponding to the Reynolds number range 121 ≤ Relp ≤ 180 are shown in conjunction with those of the reference case
(Nulp(Ref)) in Fig. 10(a). It is clear that the full combination of off-design conditions results in a significant reduction in
thermal performance. In ascertaining the effect of Reynolds number on this reduction a performance ratio RACCS,
comparing the Nusselt number of the off-design setup and design setup, acting as a reference value, was defined as:
𝑁𝑢𝑙𝑝(𝐴𝐶𝐶𝑆) (19)
𝑅𝐴𝐶𝐶𝑆 =
𝑁𝑢𝑙𝑝(𝑅𝑒𝑓)

The variation of RACCS with Relp is plotted in Fig. 10(b) from which it can be seen that at lower Reynolds numbers,
corresponding to low and medium fan speeds of the ACCS, the difference between design and off design performance is
60% while at higher fan speeds the performance deviation is 50%. It is understood that a smaller Reynolds number (Relp)
results in a thicker viscous boundary layer over the louver surface. It follows that main-stream air flow through a louver
gap is blocked off when the Reynolds number is small. Consequently, louver-directed flow is more likely to occur at a
larger Reynolds number, and this is accompanied by enhanced heat transfer [22], explaining the noted difference in Fig.
12
10(b) between the lowest and highest fan speed settings. Such a severe drop in overall thermal performance, 50% ~ 60%,
cannot be attributed to any single off-design condition reported in the literature. Moreover, simply adding the effects of
possible off-design configurations may be misleading since some configurations (such as non-uniform external coolant
stream) are detrimental to heat transfer while other configurations (such as external coolant stream’s inclination) may
enhance heat transfer.
There is merit, however, in identifying the causes of the performance reduction in the ACCS such that future design
iterations may better account for deviation and allow for more efficient operation. The performance deviation was to be
investigated by modifying the wind tunnel, used for the reference case, to incrementally mimic the setup within the ACCS
unit (Fig. 2(b)). In this way, the individual contributions of certain key features of the ACCS configuration, to the
performance deviation could be determined.

15
Design performance (Ref)
Off-design performance in an ACCS unit

10
Nulp

Nulp(Ref)

5
Nulp(ACCS)

ACCS
0
0 100 200 300 400
Relp
(a)
0.6

0.5
R ACCS

0.4

Operating range
in an ACCS unit
0.3
100 120 140 160 180 200
Relp
(b)
Figure 10. Comparison of thermal performance of a louvered fin-tube heater core under design and off-design conditions:
(a) Nusselt numbers measured at design conditions (Nulp(Ref)) and off-design conditions (Nulp (ACCS)) varying with
louver pitch-based Reynolds number (Relp), (b) Normalized Nulp(ACCS) variation with Relp.

3.3. Coolant flow upstream of the core


In modifying the wind tunnel setup to mimic that of the ACCS unit, a wedge obstruction was placed upstream of the
core (Fig. 8(a)). In this way the setup was seen to mimic only the upstream conditions in the ACCS unit, i.e., without any
downstream ducting, as shown in Fig. 12. This setup is referred to as the “half-ACCS unit” conditions. Nusselt numbers
obtained under half-ACCS unit conditions (Nulp(h)), for the same range 121 ≤ Relp ≤ 180, were normalized by Nulp(Ref).
This yielded the performance ratio Rh = Nulp(h)/Nulp(Ref) which is plotted for comparison with RACCS and the
reference/design performance in Fig. 13, where a thermal performance reduction of up to 32% relative to the reference
case is noted.

13
Figure 12. Schematic representation of air flow inside an ACCS unit without ducting downstream of the core.

Design performance
1
0.9
0.32
0.8
0.7
Rh
0.6
0.60
R

0.5 RACCS
0.4
0.3
0.2
ACCS unit
0.1 Main stream flow separation due to the wedge
0
0 100 200 300
Relp
Figure 13. Influence of half-ACCS unit conditions on the thermal performance of the heater core.
1

B
0.8
A

0.6
y/H

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 1.8
u/um
(a)

(b)
Figure 14. (a) Velocity profiles for Relp = 180, measured at A = 0.3H upstream from the wedge and at B = 0.2H
downstream from the heater core, (b) Infrared thermographs of the heater core showing the effect of non-uniform incident
coolant flow on the temperature distribution of the core.
14
Figure 15. Flow visualization image showing separated flow above the wedge.

Based on the results of T’Joen et al. [12], Datta et al. [13], Blecich [14], it was suspected that the wedge obstruction
resulted in the external coolant stream’s non-uniformity upstream of the core, which may have been responsible for the
thermal performance deterioration. This effect was quantified by measuring the air-stream velocity profiles upstream and
downstream of the wedge-core combination, as shown in Fig. 14(a). The upstream flow (as denoted as “A” in Fig. 14(a))
is skewed slightly due to the presence of the wedge; the flow being channelled between the apex of the wedge and the
channel upper wall. The downstream flow (as denoted as “B” in Fig. 14(a)), however, is distributed with a concentration
of the air-stream being towards the upper region (0.7 ≤ y/H ≤ 1.0), indicating that the core experiences a large flux of cool
air-stream in the upper region while the majority of the core experiences a much lower flux. The result is that the upper
fraction of the core’s surface is predominantly used to transfer heat. To supplement this reasoning, infrared temperature
maps of the core operating both under reference conditions and under half-ACCS unit conditions were obtained. In both
cases, the inlet water temperature and air volume flowrate were kept constant at Tw,in = 45 °C and Qa = 90 L/s (or Relp =
144), respectively. The resulting thermographs are shown in Fig. 14(b), where the temperature distributions of the
reference case and the half-ACCS unit case can be compared when a uniform and a non-uniform flow approaches the
heater core. Hot water enters at the lower end (y/H = 0) and flows upward (in the positive y-direction) as the air-stream,
flowing perpendicular to the image (in the positive x-direction), cools down the core. When the flow is uniform, a small
region near y/H = 0 is at the inlet water temperature Tw,in, after which the water-stream is gradually cooled as it traverses
the span of the core (H). In contrast, with a wedge upstream, gradual cooling of the water-stream is only evident towards
the upper core region (y/H > 0.5) due to higher momentum flow in the region 0.7 ≤ y/H ≤ 1.
On the other hand, it can be argued that the severity of the thermal performance deterioration (~30%) cannot solely be
due to non-uniform coolant stream distribution. In fact, Blecich [14] reported a 30% thermal performance deterioration
only for extremely non-uniform air-flows, where part of the core (particularly, 0 ≤ y/H ≤ 0.33) was completely depleted
of the external coolant stream. For all other cases of non-uniform stream where all parts of the core received at least some
coolant stream, the thermal effectiveness deterioration did not exceed 18%. In this connection, end-wall flow visualization
was carried out to discern any possible flow mechanism that could be responsible for a further reduction in thermal
performance. As seen in Fig. 15, separation (marked by a bright region of accumulated dye particles) occurs at the wedge
apex and in close proximity of the core. The separation is expected given the geometry of the obstruction, and this might
affect the ability of the coolant stream to follow the louver-directed flow path as indicated in Fig. 1(c), further decreasing
the performance of the core in the lower region.
15
Due to separation, the direction of the oncoming coolant stream, at least at the end-wall, is not expected to be
perpendicular to the core in the lower region. According to Kim et al. [15] and Henriksson [16], however, the inclination
angle of a coolant stream upstream of the core does not significantly affect heat transfer performance in the range of
incoming velocities considered here (~2.8 m/s). Therefore, it can be concluded that cumulative effects of both non-uniform
and separated coolant streams account for approximately half of the total performance reduction noted for the ACCS unit
(i.e., under the full off-design conditions).

3.4. Coolant flow downstream of the core


Vertical ducting downstream of the heater core was expected at further contributing towards the performance
reduction. To test this hypothesis, both upstream and downstream conditions were mimicked in the wind tunnel as shown
in Fig. 8(b). This would be testing the performance of the heater core in “equivalent-ACCS unit” conditions. Since velocity
and infrared measurements downstream of the core with a narrow duct in place were found to be impractical, only steady-
state temperature and flowrate measurements were carried out for the equivalent-ACCS unit. The average Nusselt
numbers obtained under equivalent-ACCS unit conditions Nulp(e) were normalized by the reference case Nusselt numbers,
leading to the performance ratio Re = Nulp(e)/Nulp(Ref) which is plotted along with Rh and RACCS in Fig. 16.

Design performance
1
0.9
0.8 0.32
0.7 Rh
0.6 0.52 XX Re
R

0.5 X RACCS
X
0.4
0.3
0.2 ACCS unit
0.1 Main stream flow separation due to the wedge
X Separation due to the wedge + duct
0
0 100 200 300
Relp
Figure 16. Combined influence of the wedge obstruction and upward turning of the air-stream downstream of the heater
core.

Due to the upward turning of the air-stream downstream of the core, a further reduction in thermal performance from
the half-ACCS unit conditions can be clearly seen with the thermal performance for the equivalent-ACCS unit conditions
decreasing by a further 20%. The equivalent unit performance approaches that of the actual off-design conditions to within
10%. The decrease of 20% is likely due to the flow exiting the lower region of the core becoming partially trapped. As
per the velocity profile shown in Fig. 14(a) at position B, the flow exiting the core is concentrated near the core upper
region. Given that the exiting flow is then redirected upwards it is theorized that the flow exiting the lower region of the
core, already having a lower flux than the upper region, becomes partially trapped, as shown in Fig. 17. This trapped,
recirculating, flow then reaches its saturation point for extracting heat from the core and the performance of the core
decreases.
Inclined, non-uniform, and obstructed flows tested previously and here in conjunction have been reasoned to
substantially affect the thermal performance of the heater core. However, there is still an unaccounted for 10% of thermal
performance deterioration, which could arise from effects produced by other configurational deviations inside the ACCS
unit. Examples include: unknown flow properties inside the diffuser up-stream of the heat transfer section, the effect of
an evaporator core upstream of the heater core, other practically irreproducible geometrical deviations, and unknown
turbulence intensity inside the ACCS unit. These aspects are suggested for further investigations, and once again stresses
the point that various combinations of off-design conditions are present in ACCS units, and the exact operating
environment (i.e., full off-design conditions) of the core should be considered when making a performance prediction.

3.5. Summary
Due to size constraints on the design, configurational features inside an ACCS unit have been shown to produce a
16
highly disturbed air-stream interacting with the louvered fin-tube heater core as shown in Fig. 17. This results in a
significant reduction of 50% to 60% in the thermal performance of the heater core compared to the design conditions,
which represents a larger performance reduction relative to off-design studies carried out previously. Approximately half
of this reduction can be attributed to the flow separating from an upstream wedge obstruction, as well as the air-flow non-
uniformity. This has been determined to be due to an uneven distribution of fluid momentum between the upper and lower
regions of the core resulting in a portion of the core exposed to a low momentum flow, consequently reducing the effective
heat transfer surface area of the heater core. A further performance reduction of 20% was attributed to sharp redirection
of the flow immediately downstream of the core. Thus, a louvered fin-tube heater core substantially underperforms, even
relative to off-design studies carried out previously, when exposed to the disturbed flows produced by contrivances inside
a typical ACCS unit configuration. In optimizing the design of a louvered fin-tube heater core, it is therefore insufficient
to only consider the geometry of the core, or individual off-design conditions in isolation, but to further consider the entire
set of geometrical features in which the core is installed on both the macro-scale flow features around the core as well as
the micro-scale features.

Figure 17. Schematic representation of the separated flow regions inside the ACCS unit.

4. Conclusions
This study firstly acknowledges that the spatially constrained design of the duct system in a typical automotive climate
control system (ACCS) unit results in a more drastic reduction, in thermal performance of louvered fin-tube heater cores,
than is expected based on the reports of previous off-design studies. The fluidic features caused by the contrivances of a
commercial ACCS unit duct system, which resulted in a thermal performance deterioration of up to 60%, have been
identified. The percentage contribution of each fluidic feature to the total thermal performance deterioration of the core
of a “particular” ACCS unit have been quantified as follows:
 A wedge obstruction upstream of the core contributes 30%. The wedge causes a non-uniform air-flow distribution
which may contribute 5~10%, and a separated flow region whose interaction with the core may contribute 20~25%.
 Sharp redirection of the flow field immediately downstream of the wedge-core combination caused a region of
trapped air, and contributed to 20%.
 Other deviations from the design conditions, such as the presence of an evaporator core upstream of the tested heater
core, unknown turbulence intensity, or unknown flow properties inside the diffuser may be responsible for the
unaccounted 10%.
To the best of the authors’ knowledge this study has, therefore, reported for the first time in open literature on the
thermal performance of a louvered fin-tube core operating inside a commercially manufactured ACCS unit, and has shown
that combined fluidic features of the external coolant stream both upstream and downstream of the core can be responsible
for thermal performance deviations. Hence, it is insufficient to only consider individual off-design features for
performance predictions of louvered fin-tube cores; the exact operating environment should be considered and the data
herein serves as a stepping-stone for the design of new robust heater cores applicable to ACCS units.

5. References

[1] A. Fraas and M. Ozisik, Heat Exchanger Design, New York : John Wiley & Sons, Inc, 1965.
[2] R. Webb and P. Trauger, "Flow Structure in the Louvered Fin Heat Exchanger Geometry," Experimental Thermal
and Fluid Science, vol. 4, pp. 205-217, 1991.
17
[3] T. Cowell, M. Heikal and A. Achaichia, "Flow and Heat Transfer in Compact Louvered Fin Surfaces,"
Experimental Thermal and Fluid Science , vol. 10, pp. 192-199, 1995.
[4] C. Davenport, "Correlations for Heat Transfer and Flow Friction Characteristics of Louvred," AIChE Symposium
Series, vol. 79, pp. 19-27, 1983.
[5] A. Achaichia and T. A. Cowell, "Heat Transfer and Pressure Drop Characteristics of Flat Tube and Louvered Plate
Fin Surfaces," Experimental Thermal and Fluid Science, vol. 1, pp. 147-157, 1988.
[6] Y. Chang and C. Wang, "A generalized heat transfer correlation for louver fin geometry," International Journal of
Heat and Mass Transfer, vol. 40, no. 3, pp. 533-544, 1997.
[7] M. Kim and C. Bullard, "Air-side thermal hydraulic performance of multi-louvered fin aluminium heat
exchangers," International Journal of Refrigeration , vol. 25, pp. 390-400, 2002.
[8] J. Dong, J. Chen, Z. Chen, W. Zhang and Y. Zhou, "Heat tranfer and pressure frop correlations for the multi-
louvered fin compact heat exchangers," Energy Conversion and Management , vol. 48, pp. 1506-1515, 2007.
[9] P. Shinde and C. Lin, "A heat transfer and friction factor correlation for low air-side Reynolds number applications
of compact heat exchangers," Science and Technology for the Built Environment , vol. 23, pp. 192-210, 2017.
[10] J. Kitto and J. Robertson, "Effects of maldistribution of flow on heat transfer equipment performance," Heat
Transfer Engineering, vol. 10, no. 1, pp. 18-25, 1989.
[11] R. Shah and D. Sekulic, Fundamentals of Heat Exchanger Design, Hoboken, New Jersey: John Wiley & Sons Inc,
2003.
[12] C. T'Joen, M. De Paepe and F. Vanhee, "Heat Exchanger Behaviour in Non Uniform Flow," Experimental Heat
Transfer , vol. 19, no. 4, pp. 281-296, 2006.
[13] S. Datta, P. Das and S. Mukhopadhyay, "Obstructed airflow through the condenser of an automotive air
conditioner - Effects on the condenser and overall performance of the system," Applied Thermal Engineering , vol.
70, pp. 925-934, 2014.
[14] P. Blecich, "Experimental investigation of the effects of airflow nonuniformity on performance of a fin-and-tube
heat exchanger," International Journal of Refrigeration , vol. 59, pp. 65-74, 2015.
[15] M. Kim, B. Youn and C. Bullard, "Effect of inclination angle on the air-side performance of a brazed aluminium
heat exchanger under dry and wet conditions," International Journal of Heat and Mass Transfer , vol. 44, pp.
4613-4623, 2001.
[16] L. Henriksson, "Performance of Compact Heat Exchanger in Non-Perpendicular Cooling Airflows," Chalmers
University of Technology, Gothenburg, 2015.
[17] P. Dutta, S. Saha, N. Nandi and N. Pal, "Numerical study on flow separation in 90 degree pipe bend under high
Reynolds number by k-e modelling," Engineering Science and Technology, and International Journal, vol. 19, pp.
904-910, 2016.
[18] "What is CLimate Control?," Hanon Systems , [Online]. Available:
http://www.hanonsystems.com/En/Technology/WhatisClimate#what_tit4.. [Accessed 5 November 2018].
[19] J. L. R. C. Parts, "JAGUAR XJ SERIES X308 PARTS - RADIATOR-3.2/4.0 LITRE-V8," [Online]. Available:
https://www.jaguarclassicparts.com/uk/jaguar-xj-x308-parts/engine-cooling-system/radiator-and-hoses/radiator-3-
2-4-0-litre-v8. [Accessed 5 November 2018].
[20] P. Shinde, "Investigation of Low Reynolds Number Flow and Heat Transfer of Louvered Surfaces," FIU
Electronic Theses and Dissertations, Miami, 2016.
[21] Y. Chang and C. Wang, "Air side performance of brazed aluminium heat exchangers," Journal of Enhanced Heat
Transfer, vol. 3, no. 1, pp. 15-28, 1996.
[22] C. Davenport, "Heat Transfer and Flow Friction Characteristics of Louvred Heat Exchanger surfaces," in Heat
Exchangers: Theory and Practice, New York, Hemisphere/McGraw-Hill, 1983, pp. 397-412.

18
Highlights

 A commercial ACCS results in unique flow features interacting with the heater core
 Previous studies have not accurately modelled or replicated the unique flow features
 Thermal performance of a heater core in an ACCS is up to 60% lower than expected
 Several off-design features act in combination to deteriorate thermal performance
 Only 50% of the deterioration in performance could accounted for

19

You might also like