Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

Applied Thermal Engineering 147 (2019) 781–788

Contents lists available at ScienceDirect

Applied Thermal Engineering


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

Research Paper

Development of an additive manufacturing-enabled compact manifold T


microchannel heat exchanger
Ratnesh Tiwari, Rohit S. Andhare, Amir Shooshtari , Michael Ohadi

Advanced Heat Exchangers and Process Intensification Laboratory, Department of Mechanical Engineering, University of Maryland, College Park, MD 20742, USA

HIGHLIGHTS

• AFlow
novel heat exchanger was designed using fin tubes as microchannels.
• Shell-side
distribution in microchannels was provided by a 3D printed manifold.
• The heat exchanger 2
heat transfer coefficients up to 45,000 W/m K were obtained.
• can be scaled up in a multitube bundle for larger applications.

ARTICLE INFO ABSTRACT

Keywords: This paper discusses the design and performance characterization of a compact tubular manifold microchannel
Additively manufactured heat exchanger heat exchanger. The purpose of this study was to explore the role of more precise flow distribution in the heat
Manifold microchannel heat exchanger exchanger utilizing an additively manufactured manifold for single-phase flow under low to moderate heat flux
3D printed heat exchangers conditions. The heat exchanger uses a commercially available enhanced tube having a fin structure on its outer
Compact heat exchanger
surface and helical pattern of grooves (rifling structure) on the inner bore of the tube. A 3D printed manifold
made of ABS plastic was used to properly distribute the flow on the shell-side of the heat exchanger. Water was
used as the working fluid for both shell and tube-sides. Single-phase experimental tests showed an overall heat
transfer coefficient of 22,000 W/(m2 K) and shell-side heat transfer coefficient of 45,000 W/(m2 K) for the
shell-side and tube-side water flow rates of 82 g/s and 806 g/s , respectively. The shell-side heat transfer coef-
ficient was found to be an order of magnitude higher than that found in typical shell and tube and plate-type heat
exchangers.

1. Introduction electronics cooling industries. Major reasons for this are (1) the high
pressure drop associated with narrow flow channels; (2) flow mal-
Heat exchangers play a vital role in the energy efficiency of any distribution issues; and (3) high manufacturing cost. In our approach
industrial chemical process. With the increased push to make such we utilize a commercially available and low cost fin tube as micro-
processes more energy efficient, there is demand for a new generation channels and a manifold system for flow distribution over the micro
of heat exchangers that can deliver superior performance while keeping channels to reduce the pressure drops. The manifold system allows the
the cost and size of the heat exchanger to minimum. One way to im- flow length to be divided into several small sections, as shown in Fig. 1.
prove the heat transfer performance of heat exchangers is by increasing The short flow lengths help reduce the pressure drop. This is further
the surface area to volume ratio using mini and microchannels, with the explained in the following sections of this paper.
net effect of substantial volume/weight savings for the same thermal One of the earliest works on manifold-microchannel heat sinks for
performance or increased thermal performance for the same size/ electronics cooling was reported by Harpole and Eninger [1] in 1991.
weight and pumping power consumptions. Following this work, Copeland et al. [2] developed a silicon-base mani-
Extensive research has been done on microchannel heat transfer in fold-microchannel heat sink that used fluorocarbon (FX3250) as the heat
the past couple of decades. However, microchannel heat exchangers exchanging medium. Since then, many studies have been reported in the
have not realized their full potential yet and their market penetration literature that investigate the physics of the flow, geometrical optimiza-
has been rather limited to applications in the HVAC, automotive, and tion, manufacturing aspects, performance characterization, and expansion


Corresponding author.
E-mail address: amir@umd.edu (A. Shooshtari).

https://doi.org/10.1016/j.applthermaleng.2018.10.122
Received 23 April 2018; Received in revised form 19 October 2018; Accepted 27 October 2018
Available online 28 October 2018
1359-4311/ © 2018 Elsevier Ltd. All rights reserved.
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

Nomenclature x+ dimensionless axial distance for hydrodynamic entry re-


gion [-]
a cross-sectional area [m2 ] x dimensionless axial distance for thermal entry region [-]
A surface area [m2 ]
B1 constant [-] Greek symbols
B2 constant [-]
Cp specific heat [J/(kg K) ] α area enhancement ratio [-]
D diameter [m ] difference
f Darcy’s friction factor [-] density [kg/m3]
g gravitational acceleration constant [m/s2 ] uncertainty
h heat transfer coefficient [W/(m2 K) ] viscosity [kg/(ms) ]
H height [m ]
k thermal conductivity [W/(mK) ] Subscripts
LMTD log mean temperature difference [K ]
l length [m ] actual actual
m mass flow rate [kg/s ] avg average
n number ch microchannel
Nu Nusselt number [-] diff difference
NuH Nusselt number for constant wall heat flux boundary fin fin
condition [-] flow cross-sectional
NuT Nusselt number for constant wall temperature boundary fpl fins per length
condition [-] h hydraulic
P pressure drop [Pa ] hot , cold hot side and cold side
P perimeter [m ] in inlet
Pr Prandtl number [-] i inner side
Q heat transfer [W ] q, r summation indices
R arbitrary variable manifold manifold side
Re Reynolds number [-] min , max minimum or maximum
STC Sieder-Tate coefficient [-] out outlet
T temperature [K ] o outer side
t thickness [m ] p pass
U overall heat transfer coefficient [W/(m2 K) ] ridge ridge of the rifling structure
v velocity [m/s ] s shell side
W width [m ] t tube side
X independent variable wall wall
x axial distance in the flow direction [m ] water water side

of applications [3–9]. With careful design and optimization of micro- et al. [7,10] achieved heat flux more than 1000 W/cm2 in two-phase re-
channels as well as the manifold geometry, very high heat fluxes can be gime for high heat flux electronics applications.
obtained using manifold microchannel heat sinks. For example, Mandel While these early studies focused on electronics cooling applica-
tions, some recent works have focused on larger-scale applications of
manifold heat exchangers such as HVAC. Jha et al. [8,11] designed and
tested a tubular manifold microchannel heat exchanger using a high fin-
density (100 µm fin width) microgrooved tube for absorption re-
frigeration systems. Both studies achieved significant enhancement in
overall heat transfer coefficient up to 10,000 W/(m2 K) using water as
the working fluid. Andhare et al. [12] tested a manifold microchannel
flat plate heat exchanger for single-phase heat transfer with equal water
flow rates on both sides of the heat exchanger. An overall heat transfer
coefficient of up to 22,000 W/(m2 K) was reported, which is re-
markably high for single-phase heat transfer. A review of high tem-
perature heat exchangers including the manifold microchannel heat
exchangers is provided by Xiang et al. [13].
Manifold designs may become too complex to fabricate using tra-
ditional manufacturing processes and thus may require the use of ad-
ditive manufacturing. Recent advances in additive manufacturing
technology has helped fabricate complex heat exchanger geometries
that would otherwise be impossible to fabricate. Several heat ex-
changers are being manufactured using additive manufacturing for
variety of applications including polymer as well as metallic heat ex-
changers [9]. Due to the ability to fabricate the geometries best suited
for heat transfer, such heat exchangers are compact and lightweight and
often have superior heat transfer. A titanium-based air-water manifold-
Fig. 1. Comparison of flow path in a traditional microchannel (a) with that in a microchannel heat exchanger printed using the 3D process of direct
manifold microchannel (b). laser metal sintering (DLMS) showed a 15–50% increase in heat transfer

782
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

compared to the state-of-the-art wavy fin heat exchangers [14].


Apart from the manufacturing challenges, manifold-microchannel
heat exchangers also need to address the pressure drop occurring inside
the manifold itself. It is desirable to estimate the pressure drop inside
the manifold and microchannels separately and optimize the geometry
to achieve maximum heat transfer with the lowest pressure drop.
Several efforts have been made to model the manifold microchannel
using CFD and optimize the design to achieve the most efficient heat
exchanger designs. Sarangi et al. [15] conducted a CFD analysis of a
unit cell of the manifold microchannel heat sink. They also performed
an optimization study to understand the effect of geometric parameters
on performance. Arie et al. [16] developed a multi-objective, approx-
imation-based optimization that sought to maximize heat transfer Fig. 3. 3-D printed manifold used for the manifold microchannel heat ex-
density and coefficient of performance. Arie et al. [17,18] expanded the changer.
study to optimize a single element of the manifold microchannel flat
plate heat exchanger. They used a flat plate manifold microchannel
heat exchanger geometry for the optimization study. 4350028) [19]. The enhanced surface copper tube is a ¾ inch outer
Although microchannel heat exchangers have been a topic of in- diameter tube consisting of outer circumferential fins and internal ri-
terest for various studies, they have not been commercially viable in fling (Fig. 2(a) and (b)). The outer circumferential fins are further
most single-phase liquid applications due to their higher cost and dif- notched to provide enhancement as discussed by Thors et al.[20].
ficulty in scaling up. Some of the printed circuit microchannel heat The internal rifling provides a circumferential component to the velo-
exchangers are slowly appearing in the market; however, the high cost city of the flow. This disturbs the temperature gradients that develop
remains an issue. The higher cost is mainly due to the use of expensive near the surface of the tube, and thus improves the heat transfer.
fabrication methods such as chemical etching as well as diffusion A critical component of the heat exchanger is the manifold. Fluid
bonding. Scaling up of these heat exchangers also becomes expensive as enters the manifold and is driven into the microchannels. In a manifold
the heat exchanger headers become very large and hence expensive. microchannel heat exchanger, it is important that the manifold be de-
Thus, there is a need to develop heat exchangers which use inexpensive signed to fit tightly on the microchannel tube. A loose fit will allow
microchannel fabrication techniques and can be fabricated and scaled water to bypass the microchannels and flow between the manifold and
up cost effectively. the microchannel surface, negating the benefits the manifold. To ad-
The current work aims to develop manifold-microchannel heat ex- dress these issues, a 3-D printed tubular manifold was utilized (Fig. 3).
changers which can address these issues. First, the design presented Table 1 provides a summary of the geometrical and material spe-
herein utilizes commercially available fin tubes as the microchannel cifications of each component of the tubular manifold microchannel
geometry to reduce the cost of microchannel manufacturing. Secondly, heat exchanger.
the use of the patented 3D printed polymer manifold design helps flow The manifold has a multi-stage arrangement wherein the adjacent
distribution and thus facilitates designing longer heat exchangers for stages are staggered to guide the fluid in a zigzag path from the inlet to
ease of scale up. The fabrication of the heat exchanger is similar to that the outlet of the heat exchanger. This staggered arrangement helps the
currently used for shell and tube heat exchangers. Thus, the combina- fluid move from one stage to the adjacent stage via the microchannels
tion of high thermal performance along with the use of mass-manu- in the circumferential direction of the tube. As the fluid enters the
factured fin tubes helps reduce the cost of the heat exchanger. manifold’s open channels, it travels longitudinally until it is prevented
from going further, as the manifold channels are closed at the end of
2. Design of the test section every stage (Fig. 4(a)). Thus, the fluid is forced to move into the mi-
crochannels (fin gaps on the enhanced tube) during which it exchanges
The prototype tubular heat exchanger was developed using com- heat. As the fluid moves through the microchannels, it is guided into the
mercially available enhanced fin tube, Turbo CIII (catalog number 95- adjacent manifold channels, which are the path of least resistance for

Fig. 2. (a) Enhanced tube used in the study [19] (b) detailed dimensions of the fin tube.

783
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

Table 1
Specifications of the tubular manifold microchannel heat exchanger.
General component Parameter Value

Tube (catalog number: Material Copper


95-4350028) Nominal tube ID, Di (mm ) 16.21
Finished fin OD, Do (mm ) 18.80
Effective tube length, l (mm ) 152.4
Fins per inch 43
Actual outside surface area, A actual (m2) 0.088
Area enhancement ratio, α = A actual / Ao 2.95
Average fin width, Wfin (mm ) 0.16
Average channel width, Wch (mm ) 0.43
Average fin height, Hch (mm ) 0.61
Ridge height, Hridge (mm ) 0.48

Manifold Material ABS


ID (mm ) 18.8
OD (mm ) 24.4
Pass length, lp (mm ) 25.4
Rib Width, x (mm) 5
Number of manifold openings, np 6

Outer shell Material HDPE


ID, (mm ) 24.4
OD, (mm ) 25.4

the fluid. Thus, the fluid moves across the second (neighboring)
manifold stage, and then is forced to move into further sets of micro-
channels and so on. This process of the fluid moving alternatively be-
tween the manifold and microchannels continues until the fluid exits
the heat exchanger.
A prototype of a tubular manifold microchannel heat exchanger was
fabricated by inserting a 3-D printed manifold over the enhanced tube,
depicted in Fig. 4(b). The manifold and fin tube subassembly was then
inserted inside a transparent polymer tube. Two tee connections were
used as the inlet and outlet for the shell-side (manifold-side) of the heat
exchanger. Temperature and differential pressure probes were con-
nected at the inlet and outlet as shown. The shell-side was connected to
the cold fluid supply, while the tube-side was connected to the hot fluid
supply.

3. Experimental setup, testing, and data reduction


Fig. 4. (a) Cut section schematic showing the position of the manifold on top of
Experimental test setup schematic is shown in Fig. 4(c). T-type the fin tube. (b) Prototype tubular manifold microchannel heat exchanger (c)
thermocouples were used to measure the tube-side and shell-side fluid Experimental test setup schematic diagram.
temperatures, whereas the temperature difference of the inlet and
outlet temperatures for each side was measured using the differential were estimated using the correlations provided by the manufacturer.
thermocouples. The tube-side fluid flow rate was varied from 189 g/s to Dracy’s friction factor for the tube-side is calculated using the following
806 g/s , and the shell-side flow rate was varied from 9 g/s to 100 g/s . equation:
Inlet temperatures of tube side and shell side fluids were fixed at 15 °C
f = B1 (Ret ) B2
(1)
and 30 °C respectively. The shell-side pressure drop was measured using
a differential pressure transducer, whereas the tube-side pressure drop The tube-side pressure drop can be calculated using the following
was calculated using manufacturer provided correlation. To estimate equation:
the pressure drop inside the inlet and outlet tees, the pressure drop was
separately measured for each flow rate condition without connecting l vt 2
Pt = f
the heat exchanger section. The pressure drop in the inlet and outlet Di 2 (2)
tees was then subtracted from the total pressure drop recorded by the Tube-side heat transfer coefficient is calculated using the following
pressure transducers to reflect the pressure drop on the shell-side. Mass equation:
flow rates were measured using a coriolis flow meter. The in-
0.14
strumentation data were recorded once per second using a data ac- k 1 µ
ht = (STC )(Re t )0.8Pr ( 3 )
quisition system. Di µ wall (3)
Fluid on both shell and tube sides were supplied by two separate
chillers to maintain the constant inlet temperatures to the heat ex- For the selected tube, B1 is 1.773, B2 is 0.331 and STC (Sieder-Tate
changer. At the beginning of each test, the flow rates on both sides were coefficient) is 0.078, Di is the inner diameter of the tube (Fig. 2 (b)).
set to the desired value and heat exchanger was allowed to reach steady Tube-side Reynolds number is given as
state. 4mt
The calculations are broadly split into two parts: shell-side and tube- Re t =
Di µ (4)
side calculations. Tube-side pressure drop, and heat transfer coefficient

784
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

The above Eqs. (3) and (4) are used with the parameters provided in Qavg
U=
the tube datasheet [17] to calculate the heat transfer coefficient and Ao (LMTD ) (14)
pressure drop.
where Ao is the base area for the heat transfer, which is calculated
To obtain the shell-side heat transfer coefficient, the first step was to
based on the finished fin OD, Do . This was done to simplify comparison
calculate the mass flow rate of water inside the microchannels. As
of the heat exchanger with the conventional non-enhanced surface heat
discussed previously, the fluid enters the manifold’s openings and then
exchangers. With the overall heat transfer coefficient and the tube-side
is forced into the microgrooves present on the enhanced tube. As the
heat transfer coefficient known, the shell-side heat transfer coefficient
fluid enters the microgrooves, each flow stream is divided into two
is calculated using the following equation:
streams 180°to each other. Thus, if the number of manifold openings is
np (six in present design), and the number of channels per length is nfpl ,
which can be calculated from the fin density given in Table 1, the mass 1
=
1 1 ln ( Do 2Hch
Di )
flow into each channel is given by hs Ao UAo ht Ai 2 kl
(15)
ms
mch = where Ai is the tube surface area based on the nominal tube ID, Di , Hch
2np nfpl lp (5)
is the height of fins on the outer side of the tube, and Hridge is the height
where ms is the shell-side mass flow rate and lp is the length of each pass of the ridge of the internal rifling structure.
inside the manifold. Reynolds number into the microchannels is cal- Ao = Do l (16)
culated as
Ai = Di l (17)
mch Dh
Rech =
µach (6) where l is the length of the tube. Based on the shell-side heat transfer
coefficient, the microchannel Nusselt number is calculated using the
where hydraulic diameter, Dh , cross sectional area of the channel, ach , following equation:
and perimeter of the channel cross section, Pch , are represented as
(hs / ) Dh
4ach Nu ch =
Dh = k water (18)
Pch (7)
where hs / represents the shell side heat transfer coefficient based on
ach = Wch Hch (8) the actual outside surface area of the tube.

Pch = 2(Wch + Hch) (9)


Uncertainty analysis
where Wch is the width and Hch is the height of the microchannel. Shell-
side pressure drop can be divided into two components: pressure drop Uncertainty analysis was performed using the law of propagation of
inside the manifold and pressure drop inside the microchannels. It is uncertainty where uncertainty of the quantity R , which is a function of
desired to minimize the frictional pressure drop inside the manifold, variable Xi , is given by:
since it doesn’t contribute to the heat transfer enhancement. Experi- n 2
mental characterization of pressure drops inside the manifold and in- R
R= dXq
side the microchannels is not straightforward. A numerical analysis was q
Xq (19)
performed to estimate the pressure drop inside the manifold for this
given geometry, and it was found that the pressure drop inside the where
manifold was less than 20% of the total pressure drop for all flow R = R (X1 , X2 , X3 , X 4 , XN ) (20)
conditions.
The total heat duty for the heat exchanger is calculated using the Different parameters and the estimated uncertainties are presented
following equations: in Table 2. Maximum uncertainty during the experiments was due to
the measurement of temperatures using thermocouples. Since the tube-
Qt = mt Cp (Tt , in Tt , out ) (10) side Nusselt number was calculated using the manufacturer’s correla-
tion, the uncertainty in Sieder-Tate coefficient in Eq. (3) is needed to
Qs = ms Cp (Ts, out Ts, in) (11) estimate the uncertainty on the tube-side and shell-side heat transfer
Due to uncertainties in measurements and heat losses, there may be coefficients. This uncertainty for Turbo CIII was not provided by the
a difference in the heat transfer rates calculated for the cold and the hot manufacturer; however, in an experimental study by Webb et al. [21]
sides of the heat exchanger. Therefore, the hot and cold-side heat eight different tubes, including Turbo CII (which has very similar in-
transfer rates are calculated separately, and the average of the two is ternal geometries and tube side performance to Turbo CIII) were tested
used to calculate the heat transfer rate, which is used to calculate the
Table 2
thermal performance. Ideally, the heat transfer rates on both sides
Parameters and estimated uncertainties.
should be equal; however, calculations showed that the difference was
found to be within ± 6% for all tests. Parameters Uncertainty

Qt + Qs Temperature (°C) ±0.3


Qavg =
2 (12) Temperature difference (°C) ±0.1
Tube diameters (mm ) ±0.025
The logarithmic temperature difference (LMTD ) was calculated Microchannel dimensions (μm) ±10
using the following equation: Tube-side mass flow rate (kg/s ) 0.1% (reading)
Shell-side mass flow rate (kg/s ) 0.1% (reading)
((Tt , in Ts, out ) (Tt , out Ts, in )) Pressure drop (Pa ) 0.25% (FS)
LMTD =
ln ( Tt , in Ts, out
Tt , out Ts, in ) (13)
Absolute pressure (Pa )
Overall heat transfer coefficient (W/(m2 K) )
0.25% (FS)
6.3%
Shell-side heat transfer coefficient W/(m2 K) 18%
Using this LMTD , the overall heat transfer coefficient is calculated Tube-side heat transfer coefficient. (W/(m2 K) ) 7.6%
using the following equation:

785
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

35 55,000

Shell-side HTC [W/m2K]


Tube-side Pressure Drop [kPa]

30 50,000
Experimental
45,000
25
Correlation 40,000
20
35,000
15
30,000
10
25,000
0.030 0.040 0.050 0.060 0.070 0.080 0.090 0.100
5
Shell-side water flow rate [kg/s]
0
0 0.2 0.4 0.6 0.8 1
(a)
Tube-side Mass Flow Rate [kg/s] 28,000

Fig. 5. Comparison of tube-side pressure drop with the manufacturer correla- 26,000
0.19 kg/s

Overall HTC [W/m2K]


tion. 24,000
0.32 kg/s
22,000
20,000 0.45 kg/s

using same tube-side fluid. That work estimated an uncertainty of 7% 18,000 0.63 kg/s
for Sieder-Tate coefficient. It is thus reasonable to assume a similar 16,000
0.81 kg/s
uncertainty for Sieder-Tate coefficient for Turbo CIII tube as well. 14,000
Considering a maximum uncertainty of 7%, shell-side heat transfer 12,000
coefficient uncertainty was calculated to be ± 18%.
10,000
0.030 0.040 0.050 0.060 0.070 0.080 0.090 0.100 0.110 0.120
4. Results and discussion
Shell-side water flow rate [kg/s]
Before conducting the final experiments, a five-foot Turbo CIII tube (b)
piece was first tested experimentally for tube-side pressure drop to
2,400
verify the validity of the pressure drop correlation provided in the
manufacturer datasheet (Eqs. (1–2)). Water flow rate into this tube was 2,200 0.19 kg/s

varied, and pressure drop was measured using a differential pressure 2,000
Heat Duty [W]

0.32 kg/s
transducer. The pressure drop obtained through the correlations was 1,800
compared with the experimental results shown in Fig. 5. Experimental 1,600
0.45 kg/s

results were within 10% of the investigated flow range. 1,400 0.63 kg/s
Since the purpose of the present study was to develop microchannel 1,200 0.81 kg/s
heat exchangers using commercially used fin tubes, it will be logical to
1,000
represent the heat transfer coefficient based on the nominal tube area.
800
This will help compare the heat transfer enhancement of the present 0.03 0.04 0.05 0.06 0.07 0.08 0.09 0.10 0.11 0.12
concept with the conventional shell and tube heat exchangers. Fig. 6(a)
Shell-side water flow rate [kg/s]
shows the variation of shell-side heat transfer coefficient based on the
nominal tube area with shell-side flow rate. Shell-side HTC increases (c)
monotonically from 34,000 W/(m2 K) to 47,000 W/(m2 K) for flow rate
Fig. 6. (a) Shell-side heat transfer coefficient variation for a fixed tube-side flow
variation from 35 g/s to 91 g/s . Overall heat transfer coefficient in-
rate of 0.81 kg/s; (b) overall heat transfer coefficient variation with shell-side
creases with increasing flow rates on both sides (Fig. 6(b)). As discussed
flow rate; (c) heat duty variation with shell-side flow rate.
previously, tube-side heat transfer coefficient was evaluated using the
heat transfer correlations provided by the tube manufacturer. It varied
between 13,000 W/(m2 K) and 42,000 W/(m2 K) for the tube-side flow
rates investigated. It should be noted that the tube-side heat transfer from the flow inlet (Table 1), Dh is the hydraulic diameter, and Re and
becomes limiting for low tube-side flow rate cases, and thus overall heat Pr are Reynolds and Prandtl numbers, respectively.
transfer does not change much for these cases even with the increase in The flow is assumed to be hydrodynamically and thermally devel-
shell-side flow rates, as shown in Fig. 6(b). oped in the channel if x + > 0.05 and x > 0.059 for the current rec-
A maximum heat duty of 2.1 kW was achieved at LMTD of 12.1°C tangular channel with aspect ratio of 1.41 [22,23]. x + and x values for
and flow rates of 0.81 kg/s and 0.09 kg/s on tube and shell sides, re- all the experimental flow cases showed that the flow was simulta-
spectively (Fig. 6(c)). Microchannel Nusselt number and Reynolds neously developing. The shell-side microchannel Reynolds number
number corresponding to this flow rate were 22.94 and 630. The (Rech ) for these cases varied between 250 and 630. Experimental
Reynolds number corresponding to the maximum flow rate indicated averaged Nusselt numbers in the microchannels, Nuch , are compared
that the flow was in the laminar region. with the averaged Nusselt number correlations by Muzychka and Yo-
In order to understand the heat transfer phenomena in the present vanovich [24] for constant heat flux as well constant temperature
concept, it is necessary to evaluate the heat transfer characteristics boundary conditions (Fig. 7). Experimental Nusselt values were in be-
inside the microchannels as well. Nusselt number increase with the tween the constant wall heat flux and constant wall temperature
increasing microchannel Reynolds number indicates a developing la- boundary conditions and followed a similar trend to that predicted by
minar flow. Dimensionless numbers x + and x , which respectively re- the correlations. This result was as expected, since the majority of the
present hydrodynamically and thermally developing flow lengths, are heat exchangers’ boundary conditions lie between these two extremes.
defined as x + = x / DhRech and x = x / DhRech Pr , where x is the distance However, the numbers were expected to be closer to the constant

786
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

18

Nusselt number inside microchanels (-)


Experimental data
16
Nu_H correlation
14 Nu_T correlation

12

10

4
0.0020 0.0025 0.0030 0.0035 0.0040 0.0045 0.0050 0.0055 0.0060
Dimensionless axial distance in flow direction, *
Fig. 7. Experimental Nusselt number compared with the correlations.

0.8

0.7
Shell-side Pressure Drop [bar]

0.6

0.5

0.4

0.3

0.2

0.1

0.0
0.00 0.02 0.04 0.06 0.08 0.10 0.12
Shell-side water flow rate [kg/s]
Fig. 8. Shell-side pressure drop versus shell-side water flow rate.

temperature boundary condition. This could be explained by the fol- Shell-side heat transfer coefficient was an order of magnitude higher
lowing: The fins (microchannels) in a fin tube run in the circumferential than that for a typical plate heat exchanger as reported in literature
direction, and temperature of the tube wall is not expected to vary [18]. Thus, the manifold microchannel heat exchanger has a potential
substantially across the circumference at any given axial distance of the to be developed as a compact heat exchanger for industrial applica-
tube. This, combined with the high microchannel fin efficiency results tions. Since the microchannel surface used is mass-manufactured, the
in a condition closer to the three-side heated constant temperature cost of such heat exchangers should not be prohibitive. Mass manu-
boundary condition. facturing the fin tubes using a rolling process is much cheaper than
The experimental results presented here, however, indicate that the other microchannel geometries manufactured using expensive pro-
actual condition is more towards the middle of the two extremes. This cesses such as chemical etching, EDM or micromachining. It should also
could be due to (1) uncertainty in the experimental Nusselt number, (2) be noted that due to the inherent design of the manifold, the length of
the prediction accuracy of the Nusselt number correlations themselves, the flow inside the smaller channels (fin gaps) is much shorter than the
as the correlation Nusselt number is expected to be accurate within ± length of the tube. A careful design selection of manifold and micro-
15%, and (3) the correlation used was for a typical four-sided heated channel sizes can result in a heat exchanger with high thermal perfor-
channel, whereas the present microchannels represent a case closer to a mance without heavy penalty on the pressure drop.
three-sided heated and one-sided adiabatic channel, as the manifold is
not expected to transfer significant heat. Typical Nusselt numbers for 5. Conclusions
one adiabatic side and constant temperature boundary condition for the
present aspect ratio are about 10% higher than all-side heated channels A novel tubular manifold microchannel heat exchanger was de-
[25]. Apart from this, the notched fin geometry of the test section is also signed and experimentally investigated for thermal and hydraulic per-
expected to enhance the heat transfer. formance. The heat exchanger concept utilizes a commercially available
Pressure drop inside the test section varied from 0.1 bar to 0.7 bar fin tube as the geometry for the microchannels and a 3D printed
for the test conditions (Fig. 8). As expected, pressure drop increases polymer manifold for flow distribution. The following conclusions can
with the increase in the mass flow rate. be drawn from this study:

787
R. Tiwari et al. Applied Thermal Engineering 147 (2019) 781–788

1) Shell-side heat transfer coefficient between 28,000 W/m2 K and [7] R. Mandel, A. Shooshtari, S. Dessiatoun, M. Ohadi, Streamline modeling of manifold
45,000 W/m2 K based on the nominal tube area was achieved with microchannels in thin film evaporation, Proc. ASME Summer Heat Transf. Conf. vol.
2, (2013).
water as the working fluid. [8] V. Jha, S. Dessiatoun, A. Shooshtari, E.S. Al-Hajri, M.M. Ohadi, Experimental
2) High heat transfer was attributed to the simultaneously developing characterization of a nickel alloy-based manifold-microgroove evaporator, Heat
laminar flow inside the microchannels. Transfer Eng. 36 (1) (2015) 33–42.
[9] X. Zhang, R. Tiwari, A.H. Shooshtari, M.M. Ohadi, An additively manufactured
3) The overall heat transfer coefficient for the heat exchanger was as metallic manifold-microchannel heat exchanger for high temperature applications,
high as 25,000 W/m2 K, or roughly an order of magnitude higher Appl. Therm. Eng. 143 (2018) 899–908.
than that of most conventional shell and tube heat exchangers. [10] R. Mandel, S. Dessiatoun, P. McCluskey, M. Ohadi, Embedded Two-Phase Cooling of
High Flux Electronics via Micro-Enabled Surfaces and Fluid Delivery Systems
4) Superior thermal performance of the present design concept could (FEEDS), ASME Paper No. IPACK2015-48496, 2015.
help in the development of a compact heat exchanger, whereas the [11] V. Jha, S. Dessiatoun, M. Ohadi, A. Shooshtari, E. Al-Hajri, High performance
use of mass-manufactured fin tubes as the microchannel heat micro-grooved evaporative heat transfer surface for low grade waste heat recovery
applications, in: ASME. International Electronic Packaging Technical Conference
transfer surface and the shell-and-tube type design makes the fab-
and Exhibition, ASME 2011 Pacific Rim Technical Conference and Exhibition on
rication cost effective. Packaging and Integration of Electronic and Photonic Systems, MEMS and NEMS,
vol. 2, 277–283. doi:10.1115/IPACK2011-52179.
Based on the above observations, it can be concluded that the pre- [12] R.S. Andhare, A. Shooshtari, S.V. Dessiatoun, M.M. Ohadi, Heat transfer and
pressure drop characteristics of a flat plate manifold microchannel heat exchanger
sent heat exchanger concept could be used in large-scale heat ex- in counter flow configuration, Appl. Therm. Eng. 96 (2016) 178–189.
changers. However, further research is needed to investigate the geo- [13] X. Zhang, H. Keramati, M. Arie, F. Singer, R. Tiwari, A. Shooshtari, M. Ohadi,
metries which could minimize the pressure drop. The effect of flow Recent developments in high temperature heat exchangers: a review, Front. Heat
MassTransf. (2018) 11, https://doi.org/10.5098/hmt.11.18.
distribution in a multitube bundle also needs to be investigated to de- [14] M. Arie, A. Shooshtari, S. Dessiatoun, M. Ohadi, Performance characterization of an
velop a shell and tube type multitube bundle heat exchanger. additively manufactured titanium (Ti64) heat exchanger for an air-water cooling
application, Proceedings of the ASME Heat Transfer, Fluids Engineering, &
Nanochannels, Microchannels, and Minichannels Conference HT/FE/ICNMM2016,
Acknowledgements Washington, DC, (2016), https://doi.org/10.1115/HT2016-1059.
[15] S. Sarangi, K.K. Bodla, S.V. Garimella, J.Y. Murthy, Manifold microchannel heat
The financial support of this research by The Petroleum Institute, sink design using optimization under uncertainty, Int. J. Heat Mass Transf. 69
(2014) 92–105.
Abu Dhabi, UAE, and the Advanced Heat Exchangers and Process [16] M. Arie, A. Shooshtari, S. Dessiatoun, M. Ohadi, E. Al Hajri, Simulation and thermal
Intensification consortium within the Center for Environmental Energy optimization of a manifold microchannel flat plate heat exchanger, in: Proc. ASME
Engineering at the University of Maryland is gratefully acknowledged. 2012 International Mechanical Engineering Congress and Exposition, American
Society of Mechanical Engineers, pp. 209–220. doi:10.1115/IMECE2012-88181.
[17] M.A. Arie, A. Shooshtari, S. Dessiatoun, M. Ohadi, Thermal optimization of an air-
Appendix A. Supplementary material cooling heat exchanger utilizing manifold-microchannels, in: Proc. Thermal and
Thermomechanical Phenomena in Electronic Systems (ITherm), 2014 IEEE
Supplementary data to this article can be found online at https:// Intersociety Conference on, IEEE, pp. 807–815. doi:10.1109/ITHERM.2014.
6892364.
doi.org/10.1016/j.applthermaleng.2018.10.122. [18] M. Arie, A. Shooshtari, S. Dessiatoun, E. Al-Hajri, M. Ohadi, Numerical modeling
and thermal optimization of a single-phase flow manifold-microchannel plate heat
References exchanger, Int. J. Heat Mass Transf. 81 (2015) 478–489.
[19] Wolverine Turbo-CIII ® Technical Datasheet, http://www.wlv.com/wp-content/
uploads/2014/03/TurboCIII.pdf.
[1] G.M. Harpole, J.E. Eninger, Micro-channel heat exchanger optimization, Proc. [20] P. Thors, N.R. Clevinger, B.J. Campbell, J.T. Tyler, Heat Transfer Tubes and
Semiconductor Thermal Measurement and Management Symposium. SEMI-THERM Methods of Fabrication Thereof, U.S. Petenet 5,697,430, 1997.
VII. Proceedings, Seventh Annual IEEE, IEEE, 1991, pp. 59–63. [21] R.L. Webb, R. Narayanamurthy, P. Thors, Heat transfer and friction characteristics
[2] D. Copeland, M. Behnia, W. Nakayama, Manifold microchannel heat sinks: iso- of internal helical-rib roughness, J. Heat Transf.-Trans. ASME 122 (1) (2000)
thermal analysis, IEEE Trans. Comp. Packag. Manuf. Technol. Part A 20 (2) (1997) 134–142.
96–102. [22] P.-S. Lee, S.V. Garimella, Thermally developing flow and heat transfer in rectan-
[3] Y.H. Kim, W.C. Chun, J.T. Kim, B.C. Pak, B.J. Baek, Forced air cooling by using gular microchannels of different aspect ratios, Int. J. Heat Mass Transf. 49 (17)
manifold microchannel heat sinks, KSME Int. J. 12 (4) (1998) 709–718. (2006) 3060–3067.
[4] Y. Wang, G.-F. Ding, Numerical analysis of heat transfer in a manifold microchannel [23] S. Kakaç, R.K. Shah, W. Aung, Handbook of Single-Phase Convective Heat Transfer,
heat sink with high efficient copper heat spreader, Microsyst. Technol. 14 (3) second ed., 1987.
(2008) 389–395. [24] Y. Muzychka, M. Yovanovich, Laminar forced convection heat transfer in the
[5] J. Ryu, D. Choi, S. Kim, Three-dimensional numerical optimization of a manifold combined entry region of non-circular ducts, J. Heat Transf. 126 (1) (2004) 54–61.
microchannel heat sink, Int. J. Heat Mass Transf. 46 (9) (2003) 1553–1562. [25] R.K. Shah, A.L. London, Laminar Flow Forced Convection in Ducts: A Source Book
[6] Y. Wang, G.-F. Ding, S. Fu, Highly efficient manifold microchannel heatsink, for Compact Heat Exchanger Analytical Data, Academic press, 1978.
Electron. Lett. 43 (18) (2007) 978–980.

788

You might also like