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

International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Enhancement of convective heat transfer in an air-cooled heat exchanger


using interdigitated impeller blades
Jon M. Allison, Wayne L. Staats ⇑, Matthew McCarthy 1, David Jenicek, Ayaboe K. Edoh, Jeffrey H. Lang,
Evelyn N. Wang, J.G. Brisson
Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA

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

Article history: The enhancement of convective heat transfer through a finned heat sink using interdigitated impeller
Received 11 January 2011 blades is presented. The experimentally investigated heat sink is a subcomponent of an unconventional
Received in revised form 9 June 2011 heat exchanger with an integrated fan, designed to meet the challenges of thermal management in com-
Accepted 9 June 2011
pact electronic systems. The close integration of impeller blades with heat transfer surfaces results in a
Available online 7 July 2011
decreased thermal resistance per unit pumping power. The performance of the parallel plate air-cooled
heat sink was experimentally characterized and empirically modeled in terms of nondimensional param-
Keywords:
eters. Dimensionless heat fluxes as high as 48 were measured, which was shown to be about twice the
Heat exchanger
Heat sink
heat transfer rate of a traditional heat sink design using pressure-driven air flow at the same mass flow
Fan rate.
Nondimensionalization Ó 2011 Elsevier Ltd. All rights reserved.
Thermal management

1. Introduction mounted on top and drives the rotors on a single shaft running
through the condenser plates. Air enters the top through an axial
Thermal management is a critical bottleneck for high-power intake and is forced radially outward between the condenser plates
systems, such as phased-array radar and microwave and digital by the interdigitated impeller blades. The working fluid (water)
electronics, where performance and reliability are limited by the evaporates in the primary evaporator wick; the vapor travels
ability to dissipate heat efficiently. Fluidic-based cooling solutions through vertical pipes to the condenser layers where it is convec-
have commonly been incorporated using traditional large-scale tively cooled to the liquid phase and then wicked back into the
air-cooled fin-fan arrays and pumped liquid-based cooling [1–3]. evaporator. The high thermal conductance of the loop heat pipe re-
The current work focuses on the convective heat transfer in an sults in an isothermal condenser surface for heat removal; the ra-
alternative air-cooled heat exchanger comprised of a loop heat dial outflow fan convectively cools the stack of condenser layers.
pipe with an integrated fan. In contrast to a conventional pumped More conventional finned heat sinks have been studied by
liquid system, it is self-contained and has no external fluidic many investigators. Tuckerman and Pease [5] studied a microchan-
connections. This device (Fig. 1) is 10 cm  10 cm  10 cm and is nel heat sink etched on silicon and minimized thermal resistance
targeted to consume less than 33 W of electrical power while dis- by choosing the appropriate channel width, fin width and aspect
sipating over 1000 W of heat with a thermal resistance of less than ratio subject to constraints on the geometry and pressure drop.
0.05 K/W [4]. Knight et al. [6] developed an analytical method to minimize ther-
The integrated design utilizes a stack of rotating blades inter- mal resistance of a heat sink with pressure driven flow in a closed
digitated between thermal stator plates, each of which is a con- finned channel by varying the geometry. Teertstra et al. [7] devel-
denser chamber of a loop heat pipe. A single evaporator layer is oped an analytical model to calculate the Nusselt number for a
located at the bottom of the unit, which is in contact with the heat plate fin heat sink as a function of geometry and flow properties.
source and connected to the thermal stator plates via vertical Culham and Muzychka [8] presented a method of optimizing the
pipes. Fig. 2 shows a cross-sectional schematic of the loop heat geometry of plate fin heat sinks by minimizing the entropy gener-
pipe. A low-profile radial-flux permanent magnet motor is ation rate while incorporating fan performance into their model.
Several recent studies have shown that finless designs can provide
improved performance in small heat sinks. Egan et al. [9] looked at
⇑ Corresponding author. Tel.: +1 617 253 2237; fax: +1 617 258 7754.
a miniature, low profile heat sink with and without fins and used
E-mail address: wstaats@mit.edu (W.L. Staats).
1
Current address: Drexel University, Department of Mechanical Engineering and particle image velocimetry to detail flow structures and heat
Mechanics, 3141 Chestnut Street, Philadelphia, PA 19104, USA. transfer. Stafford et al. [10] studied forced convection cooling on

0017-9310/$ - see front matter Ó 2011 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ijheatmasstransfer.2011.06.023
4550 J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

Nomenclature

DTLM log mean temperature difference v air velocity


A area W_ pumping power
B blade aspect ratio x coordinate along exit plane
Cf flow coefficient y coordinate along shaft axis
cp specific heat capacity of air z coordinate normal to exit plane
Dh hydraulic diameter
G flow channel aspect ratio Greek symbols
h convective heat transfer coefficient based on wall-to- g efficiency
ambient temperature difference l dynamic viscosity of air
hb height of blade x rotational frequency
hc height of flow channel Um dimensionless heat flux
hLMTD convective heat transfer coefficient based on log mean q density of air
temperature difference r slip factor
I inlet size ratio e exchanger effectiveness
Ie electrical current
k thermal conductivity of air Subscripts
Kb minor loss coefficient, bend amb ambient
Kc minor loss coefficient, contraction b blade, bearing/brush
m_ mass flow rate c air flow channel
Nurf Nusselt number for radial outflow e electrical
P pressure fan fan in pressure driven heat sink
Pr Prandtl number f fluid (air)
Q_ thermal power sunk by heat sink i inlet
q00m mean heat flux from wall to air max maximum
r radius mech mechanical
R convection thermal resistance m motor
Rarm motor armature resistance o outlet
Rex rotational Reynolds number rf radial flow
Rehc flow Reynolds number t blade tip
T temperature w wall
To bulk outlet air temperature h azimuthal direction
Uave log mean velocity 1, 2, 3, 4 Points 1, 2, 3 or 4 in Fig. 4
V voltage

low profile heat sinks with and without fins and showed that heat
transfer rates of the finless designs were better than their finned
counterparts.
Typical air-cooled heat exchanger systems have separate fans
ducted to the heat sink; the interdigitated impeller design offers
increased heat transfer efficiency through several mechanisms.

Motor Shaft Liquid Flow


Vapor Flow
Impeller Blades

Condenser

Air Flow

Fig. 1. Schematic of the 10  10  10 cm integrated fan-heat sink device. The rotors


Evaporator
are driven by a common shaft attached to the motor. Air enters axially at the top
and is blown radially outward by the rotors. Air flows across the condensers, Heated Surface
cooling and condensing the heat pipe fluid inside the condensers. The condensed
fluid then travels down the vertical fluid connectors to the evaporator where it is Fig. 2. A simplified cross-sectional diagram of the integrated fan-heat sink showing
heated and evaporates, returning to the condensers. the motor shaft, impeller blades, and loop heat pipe.
J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559 4551

The large heat transfer coefficients associated with the developing ature to the ambient temperature as a function of mechanical
and high-shear flows slipping past the blades in the fan section are power input to the fluid ðWÞ _ required to drive the flow.
located directly on the fin surfaces. Additionally, the rotating The independent control variables governing the thermal per-
blades shear off boundary layers and mix the air, while the rotary formance of the layer shown in Fig. 3 are the geometric parameters
nature of the flow path increases the residence time of the air with- (hc, hb, ri, rt), the rotational speed of the impeller, x, and the fluid
in the parallel plate geometry. These mechanisms are derived di- properties of the coolant air. In this work, the number of blades
rectly from the integration of the fan and heat sink structures. is five and is not varied. Using the tip radius, rt, as the characteristic
The combination of these components not only achieves high effi- length for system, three nondimensional geometric parameters are
ciency and a compact design, but also directly enhances the con- defined as:
vective heat transfer between the air and heat sink. Accordingly,
hb
this work focuses on the experimental characterization of the con- B¼ ð1Þ
rt
vective heat transfer rates associated with impeller-driven flows
through representative parallel plate geometries. Rather than hc
G¼ ð2Þ
studying the entire stack, a single layer was examined for experi- rt
mental simplicity. Empirically derived correlations for the relevant
ri
I¼ ð3Þ
nondimensional parameters are established focusing on thermal rt
performance as a function of the interdigitated geometry. Addi- B and G are the aspect ratios of the impeller blade and flow channel,
tionally, the analysis and derivation of characteristic design charts respectively, and I is defined as the ratio of the flow inlet radius to
for implementing such devices is reported. the impeller blade tip radius. Accordingly, the ratio of B/G is the fill
factor of the parallel plate gap height. A value of B/G = 1 corresponds
2. Modeling to an impeller blade completely filling the flow passage, while B/
G = 0 represents a vanishingly thin blade as compared to the flow
A diagram of a single layer of the interdigitated design is shown passage height. Both of these cases, however, are not physical pos-
in Fig. 3 along with the variables and geometric parameters affect- sibilities due to manufacturing constraints.
ing convective heat transfer from the plates. Ambient air at a tem- The impeller rotational speed and the fluid properties of air are
perature of Tamb is drawn in from the top of the stack through an characterized nondimensionally using the Prandtl number and the
_ Flow is driven outward
inlet of radius of ri at a mass flow rate of m. rotational Reynolds number (as in Beretta and Malfa [11]), defined
through the parallel plates, separated by a distance of hc, by an respectively as
impeller blade with a thickness of hb and tip radius of rt. The impel-
cp l
ler spins at a rotational speed of x and is connected to a shaft (not Pr ¼ ð4Þ
_ to the air. k
shown). The impeller delivers a mechanical power of W
qx r 2t
The parallel plates are at a uniform temperature of Tw and transfer Rex ¼ ð5Þ
a thermal power of Q_ into the air between the plates, which exits l
the periphery of the plates at a mass-averaged temperature of T o .
where cp, l, k, and q are the specific heat capacity, viscosity, thermal
conductivity, and density of the ambient air. The relevant parame-
2.1. Nondimensionalization _ mechanical
ters for the heat exchanger are the mass flow rate ðmÞ
shaft power ðWÞ _ heat flow rate into the air (Q_ Þ wall temperature
The complexities of the flow field and the resulting thermal per-
(Tw) and the mass-averaged exit flow temperature ðT o Þ. These terms
formance of the heat exchanger geometry shown in Fig. 3 cannot
have been nondimensionalized as follows. The flow coefficient (Cf)
be easily modeled using closed form analytic techniques. Accord-
is a turbomachinery parameter [12] relating the net flow rate
ingly, this work focuses on the experimental characterization of
through the plates, the tip speed (xrt) and the area swept out by
the net heat transfer, fluid flow, and mechanical shaft power of
the impeller blades (which is proportional to r 2t ). It is given by
the system for various geometries. These experimental results
are then used to establish empirical correlations for the relevant _
m
Cf ¼ ð6Þ
nondimensional parameters governing such systems. In addition, qxr3t
performance characteristics and design charts have been devel-
The heat exchanger effectiveness (e) is defined as the ratio of
oped for the resulting thermal resistance (R) from the wall temper-
the actual convective heat transfer rate to the maximum possible
heat transfer rate. The maximum heat transfer rate to the air oc-
Inlet Air - m, T amb curs when the air entering the channel at Tamb is heated to the wall
temperature of the heat exchanger, or

Rotating Impeller - ω, W Q_ max ¼ mc


_ p ðT w  T amb Þ ð7Þ
Heated Wall - Q, T w
rt ri The effectiveness e is then
hc
Q_
e¼ _ ð8Þ
mcp ðT w  T amb Þ
Using the mass-averaged temperature of the exiting flow ðT o Þ; e
can be rewritten as the ratio of the outlet flow temperature to the
y x wall temperature taken relative to the ambient inlet conditions.
Impeller
z Since the specific heat does not significantly vary over the expected
hb Outlet Air - m , T o
temperature range, Eq. (8) simplifies to
Fig. 3. Schematic representation of a single-layer of the heat exchanger device T o  T amb
including a cutaway section showing the interdigitated impeller blade as well as the e¼ ð9Þ
relevant variables and geometric parameters.
T w  T amb
4552 J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

The slip factor (r) is another turbomachinery parameter [13] gi- for designing heat sinks and evaluating the trade-offs between
ven by the effective thermal resistance from the heat load to the ambient
_ air and the mechanical power necessary to operate the device. The
W total thermal resistance between the isothermal wall and the
r¼ _ 2 2 ð10Þ
mx r t ambient air is given by
A slip factor of one implies the fluid streamlines follow the contours
T w  T amb
of the rotor, which can be seen by substituting the Euler turbine R¼ ð16Þ
equation (with no pre-swirl) [14] Q_
_ ¼m
W _ xr t v h ðr t Þ ð11Þ Inserting Eqs. (4)–(9) into Eq. (16) and rearranging yields

into Eq. (10). The Euler turbine equation relates the power required R ¼ ðeC f Rex Prkrt Þ1 ð17Þ
to turn the impeller rotor with the mass-averaged tangential veloc-
Similarly, the mechanical power delivered to the fluid can be eval-
ity at the rotor tip (vh(rt)). Accordingly, Eq. (10) characterizes the
uated by combining Eqs. (6) and (10) and rearranging terms, result-
amount of flow slipping past the impeller blades as they rotate. It
ing in
is the ratio of the tangential velocity of the air at the blade tip to
the velocity of the blade tip itself that describes how well the flow _ ¼ C f rqx3 r 5
W ð18Þ
t
is driven by the impeller. In the limit of no slip, the tangential veloc-
ity at the tip (vh(rt)) approaches the blade tip velocity (xrt) and as a
result the slip factor approaches unity. 2.2. Radial flow model
The primary nondimensional parameter of interest for this de-
vice is the dimensionless heat flux (Um), defined by Shah and Lon- For comparison purposes, the performance of a similar heat ex-
don [15] as changer geometry without the interdigitated impeller blade is
evaluated here. The resulting performance of such a system can
q00m Dh be confidently modeled using closed form analysis and existing
Um ¼ ð12Þ
kðT w  T amb Þ correlations for convective heat transfer in radially outward fluid
Here, k is the thermal conductivity of the air, the hydraulic diameter flow through a parallel plate gap. This model is used to compare
is defined as Dh = 2hc, and q00m is the mean heat flux from the walls the performance of the integrated fan-heat sink design to a tradi-
into the air. The dimensionless heat flux (Um) can be thought of tional fin-fan design, where pressure-driven flow is used to con-
as a Nusselt number whose temperature scale is based on the wall vectively cool the parallel plate heat sink. In this scenario, a fan
temperature and the inlet temperature (which is equal to the Tamb provides power to drive the flow between the parallel plates in a
in this heat sink). radially outward direction from the center inlet section to the out-
An average convective heat transfer coefficient (h) for the device er periphery. The effect of the four corner regions (where r > rt) on
can be defined by Newton’s Law of Cooling as the pressure drop is neglected.
Suryanarayana et al. [16] performed an experimental study on
q00m Q_ the heat transfer in radial outflow between parallel discs held at
h¼ ¼ ð13Þ
ðT w  T amb Þ AðT w  T amb Þ uniform temperature, both stationary and rotating, with an inlet
radius ri and outer radius rt separated by a gap of hc. The geome-
where Q_ is the net heat transfer rate to the air. The area over which
tries studied in the present work (in particular I and G) are within
the heat is transferred is the inner surface of both plates, accounting
Suryanarayana’s experimentally characterized range. Accordingly,
for the inlet hole, and is given by
the average Nusselt number for the radial flow can be calculated
A ¼ 2ð4r 2t  pr 2i Þ ð14Þ using his correlation for stationary plates as

The average heat transfer coefficient (h) is defined using the differ- hLMTD hc
ence between the isothermal wall temperature and the ambient in- Nurf ¼ ¼ 0:0332Re0:782
hc ð19Þ
k
let flow temperature. By doing so, the heat transfer rate through the
system is quantified relative to the variables pertinent to real world where hLMTD is the average heat transfer coefficient based on the log
implementation, namely the temperature of the device and the sur- mean temperature difference between the heated plate and the air
rounding ambient environment. By combining Eqs. (12)–(14), along and Rehc is the Reynolds number based on the plate spacing (hc) and
with the definition of the hydraulic diameter, the dimensionless the log mean velocity. The Nusselt number in the Suryanarayana
heat flux can be rewritten in terms of the previously defined nondi- correlation, Eq. (19), uses the plate spacing as the length scale,
mensional parameters from Eqs. (2)–(9) as which is important to note when comparing this to the dimension-
  less heat flux in this work, which uses Dh = 2hc as the length scale,
PrG
Um ¼ eC f Rex ð15Þ as defined in Eq. (12). Furthermore, in contrast to Eq. (13) which de-
4  pI2 fines the heat transfer coefficient using the temperature difference
The independent variables in Eq. (15) (G, I, Rex, Pr) are deter- between the wall and the ambient, Suryanarayana’s heat transfer
mined by the device geometry, coolant fluid, and the rotational coefficient in Eq. (19) references the log mean temperature
speed of the impeller. However, the heat exchanger effectiveness difference
(e) and flow coefficient (Cf) are dependent parameters varying with
Q_
device geometry, impeller speed and fluid properties. Accordingly, hLMTD ¼ ð20Þ
ADT LM
correlations for the dimensionless heat flux are developed by
empirically characterizing the effectiveness and the flow coeffi- The log mean temperature difference DTLM for this heat sink is de-
cient as functions of geometry and operating speed. fined as
While the dimensionless heat flux can be used as a measure of  
the effective heat transfer within a system, it is also useful to eval- ðT w  T amb Þ  T w  T o
DT LM ¼   ð21Þ
uate the resulting thermal resistance and the input mechanical
ln TTw TTamb
power required. These two parameters are of particular interest w o
J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559 4553

The Reynolds number in Suryanarayana’s correlation (Eq. (19)), inserting the dimensionless variables described above, the pressure
Rehc, is given by difference between points 3 and 4 in Fig. 4 becomes
"
Rehc ¼
qU ave hc
ð22Þ
_2
m 16 2  2 
P3  P4 ¼ G I 1
l 2q
4
hc 63p 2

 0:25  #
where Uave, the log mean velocity, can be calculated as in [16] as m_ 
0:75 0:75
þ    0:007089  G I  1:682  2
_
m ln ðrt =r i Þ
lr t
U ave ¼ ð23Þ
2pqhc r t  r i ð27Þ

The average heat transfer coefficient based on the temperature dif- For the geometries investigated in this work, the flow from point 3
ference between the plate and the inlet air, as defined in Eq. (13), is to point 4 experiences a pressure increase due to the increasing
related to the heat transfer coefficient in Eq. (19) (hLMTD) by the con- cross sectional area. Since the flow exits the parallel plate gap into
servation of energy as atmospheric conditions as a jet, the pressure at the exit (P4) is atmo-
 spheric. The pressure upstream of the parallel plate gap entry (that
_ p
mc hLMTD A is, point 2 in Fig. 4) is determined using the head loss equation:
h¼ 1  exp ð24Þ
A _ p
mc

P2 v 22 P3 v 23 v 23
where A is the total wetted area as given by Eq. (14). By determining þ  þ ¼ ðK b þ K c Þ ð28Þ
q 2 q 2 2
the average heat transfer coefficient (h) using Eqs. (19) and (22)–
(24), the total convective thermal resistance can be determined as where v2 and v3 are the velocities at points 2 and 3, and Kb and Kc
are the minor losses associated with the right angle bend and the
1
R¼ ð25Þ contraction from point 2 to point 3 respectively. The turning loss
hA
is assumed to be Kb = 1, which is typical of a sharp right angle bend
where A is once again the total area over which convective heat [18]. The contraction loss is given by Fay [18] as
transfer between the parallel plates and the air occurs. 
A3
The power required to drive air radially outward through the K c ¼ 0:4 1  ð29Þ
A2
parallel plates can determined in several steps. Fig. 4 shows a dia-
gram of the pressure driven radial flow model. In essence, a fan is where A2 and A3 are the flow areas at points 2 and 3, given by
used to bring stagnant, atmospheric pressure air to a pressure and
velocity sufficient to drive the flow through a channel between two A2 ¼ pr2i
parallel discs; the work requirement of the fan can be determined A3 ¼ 2pr i hc ð30Þ
for a specified mass flow rate.
First, the pressure difference between points 3 and 4 in Fig. 4 is A3 must be smaller than A2 in Eq. (29). Rearranging Eq. (28) gives
calculated using an expression derived by Moller [17]. Moller the pressure difference from point 2 to point 3 as
determined an approximate solution for the pressure distribution qv 23 qv 22
of a turbulent radial outflow in a parallel plate gap by using an P2  P3 ¼ ð1 þ K b þ K c Þ  ð31Þ
2 2
integral solution to the momentum equation. Moller also derived
a laminar solution; however, the error introduced by using the tur- Mass conservation can be used to simplify Eq. (31) by eliminating v2
bulent solution compared to a composite laminar-turbulent solu- and v3, giving the pressure difference as a function of the mass flow
tion was found to be small for the cases considered in this work. rate as
Therefore, for simplicity, the turbulent solution was used; this " #
_ 2 ð1 þ K b þ K c Þ 1
m
has the form P2  P3 ¼  ð32Þ
" 2q A23 A22
 2  2
_2
m 16 hc rt
PðrÞ  Pðrt Þ ¼  1 Finally, the ideal work required to bring stagnant air at atmospheric
2qhc
4 63p2 r t r
!# pressure (point 1 in Fig. 4) to the pressure and velocity at point 2 in
hc l0:25 1 1:682 Fig. 4, determined using the conservation of energy, is
þ    0:007089  ð26Þ
_ 0:25 r 0:75 ð2r t Þ0:75
m 
_ f;fan ¼ m P1  P 2 v 22
W _  ð33Þ
Applying Eq. (26) between the entrance to the parallel plate gap q 2
(point 3 in Fig. 4, r = ri) and the exit (point 4 in Fig. 4, r = rt) and where W_ f;fan is the useful work delivered to the fluid (air) by the fan.
Eliminating v2 and reformulating in terms of the mass flow rate
using continuity, Eq. (33) becomes
1
Welec,fan Patm
_ f;fan ¼ m_3 _ 1  P2 Þ
mðP
Adiabatic Exterior W  ð34Þ
Surface 2q A2
2 2 q
Fan
hc
ri 2
Patm Of course, only a fraction of the electrical power input to the fan is
rt 3 Air Flow 4 available as useful fluid power. To determine how much mechanical
work input would be required by the fan, 9 commercially available
fans in the 1–8 W fluid power range were surveyed (Delta AH-
CL Adiabatic Exterior B1448EH/SH/VH, Delta AFC0948DE-TP20, Delta FFB1012EHE, Sunon
Heated Walls
Surface PMD4809PMB1-A/2-A/3-A, Delta FFB0912SH). Fan curves (static
Fig. 4. Schematic of the radial flow model. A fan forces air axially into a parallel
pressure rise vs. volume flow) from the manufacturer’s datasheets
plate gap where it then flows radially outward, providing convective cooling to the were used to produce fluid power vs. volume flow curves. The point
heated walls. of maximum fluid power was determined from these curves; these
4554 J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

powers were divided by the specified electrical power input to Fig. 5(a) is a schematic representation of the entire assembly, while
determine an overall fan efficiency Figs. 5(b) and (c) show a cross section of the flow test section and
the impeller design, respectively.
_
W
goverall;fan ¼ _ f;fan ð35Þ Two 6 mm thick, polished square copper plates with an edge
W e;fan length of 100 mm are used to simulate the parallel condenser
plates. The top plate has a 40 mm diameter hole cut through it to
Of the fans surveyed, the highest overall efficiency observed was
function as the air inlet. Thermal power is supplied to the parallel
0.33 (Delta AHB1548EH). Therefore, in this analysis, an overall fan
plates with polyimide film insulated heaters (Omega KHLV-102/
efficiency of 0.35 was used as an optimistic estimate for a well de-
(10)-P) while the wall temperature is monitored using embedded
signed fan of this size. Ultimately, the mechanical power require-
thermocouples (Omega 5TC-TT-J-30-36). The backsides of the plate
ment of the fan must be determined to give a fair comparison to
assembly are insulated with balsa wood to reduce the heat lost to
the impeller driven flow. The efficiency of the motor must be con-
the environment. The top plate is mounted above the bottom using
sidered to determine the mechanical power input to the fan.
annular shims in the corners to define the channel height (hc). The
_ mech;fan
W plates, insulation, and shims, have clearance holes and are fastened
gmotor;fan ¼ _ e;fan
ð36Þ with machine screws to an aluminum stand. Multiple impeller ro-
W
tors were laser cut out of acrylic sheets (with various thicknesses)
All of the fans in the survey described above are driven by brushless and separately mounted and aligned within the parallel plate flow
DC motors. An efficiency of 0.85 is typical for a well designed motor section using a drive shaft and bearing assembly adjusted with a
of this type and size, based on manufacturers’ datasheets (e.g. Max- positioning stage. The rotor blades are shaped as the trailing end
on 339285). of a NACA 0011 airfoil, as shown in Fig. 5(c). The rotor blades used
Eq. (34) can be rewritten in terms of the mechanical fan power for this work are simple extrusions of the two-dimensional airfoil
by using Eqs. (35) and (36) as geometry with no complex out-of-plane contours. The blade pro-
! file begins radially at r = 0 and transitions at a constant rate to a
_ mech;fan ¼ _3
m _ 1  P2 Þ
mðP gmotor;fan sweptback angle of 45° at the trailing edge, which is rounded off
W  ð37Þ
2q2 A2 2 q goverall;fan with a circular arc of radius 1.3 mm. The shaft is supported by
two self-aligning, double-row, unshielded, oiled ball bearings
Eqs. (37), (32) and (27) can be solved to determine the mechanical (SKF 126TN9). The shaft is connected to a commercially available
power input required by the fan to yield an arbitrary mass flow rate. DC motor (Maxon 110930) using a flexible coupling.
This power estimate represents a well designed commercial motor- The experiments were carried out for a fixed impeller blade pro-
fan combination based on an assessment of fans and motors the file as well as inlet radius, ri = 2 cm, and blade tip radius, rt = 5 cm.
power range expected in the radial flow comparison. These fixed geometries were chosen as optimal solutions from a 3D
finite element simulation using ANSYS FLUENT [4,19], while taking
3. Testing into consideration realistic manufacturing and size constraints.
These numerical results showed that the dominant parameters
Using the relations developed in Section 2.1, the flow coefficient affecting performance of the layer were the channel and blade
(Cf), slip factor (r), and the heat exchanger effectiveness (e), were thicknesses. Accordingly, the experiments have been conducted
evaluated experimentally as a function of rotational Reynolds for a variety of channel heights and impeller blade thicknesses cre-
number (Rex), and geometry. Using these relations the dimension- ated by adjusting the spacers within the assembly and varying the
less heat flux as well as the total thermal resistance and mechani- impeller blades used. During testing, electrical power is delivered
cal power required to drive the flow have been correlated to the to the heaters while monitoring the wall temperature and calculat-
independent variables. Fig. 5 shows the setup used to experimen- ing the total heat transfer rate to the fluid accounting for the heat
tally characterize the various thermal and fluidic parameters rele- lost to the environment [19]. Similarly, the rotor is controlled by
vant for developing an empirical model of device performance. supplying electrical power to the DC motor and monitoring the

(a) (b)
Thermal
Insulation Machine Shaft Impeller
Heated Screw Machine
Copper Plate Screw
Film
Heater Thermal
Heated Insulation
y
Shim Copper Plates
z x Film
Shaft Impeller
Heaters
Aluminum
Bearing Table

5 DOF
Motor Stage (c)

Fig. 5. Schematic representation of the experimental set-up showing (a) the exploded assembly, (b) a cross section of the flow section, and (c) the impeller blade geometry.
J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559 4555

rotational speed. The mechanical power delivered to the flow is


(a) vz / vz,max
calculated accounting for the motor, shaft, and bearing losses.
The exit air flow rate and average temperature are measured using 2.5 0.5
hot-wire anemometry at the periphery of the assembled stack. 0.9

0.1
0.5
2

y (mm)

0.1
1.5

0.9
_ f 1

0.5
3.1. Mechanical power characterization – W
0.5 0.5
0
The electric power supplied to the DC motor ðW _ e Þ is converted 0 20 40 60 80 100
into three components: motor losses ðW _ m Þ, mechanical losses x (mm)
including bearing and brush losses ðW_ b Þ, and the power delivered
to the fluid ðW_ f Þ according to (b) To – Tamb (°C)
2.5

30
W _ mþW
_ e¼W _ bþW
_f ð38Þ 25

15
2

y (mm)

20
30

25
30
The electrical power input to the motor and the motor losses are gi- 1.5

25
ven by 1
0.5 30
30
_ e ¼ Ie V
W ð39Þ
0 20 40 60 80 100
_ m ¼ I2 Rarm
W ð40Þ x (mm)
e

where Ie is the current, V is the voltage, and Rarm is the armature Fig. 6. (a) Normalized air velocity perpendicular to the exit plane and (b)
resistance of the DC motor. Inserting these relations into Eq. (38) temperature rise above ambient of air on the exit plane, where the blades and
flow are rotating from right to left. ðQ_ ¼ 90W; hb ¼ 1:65 mm; hc ¼ 2:80 mm;
and rearranging yields
x ¼ 2  60  3000 rpm; m_ ¼ 3:69g=s; T o  T amb ¼ 28:5  CÞ.
_ f þW
W _ b ¼ Ie V  I2 Rarm ð41Þ
e

The left hand side of Eq. (41) is the mechanical shaft power, which is non-uniform over the exit plane, where the effects of the corner
the power delivered to the fluid by the impeller blades (W _ f ) plus the supports are evident by the reduced speeds and temperatures at
mechanical power lost in the bearings and brushes (W _ b ). The the rightmost sections.
mechanical losses can be calibrated as a function of rotational speed Using contours similar to the ones shown in Fig. 6, the maxi-
by rotating the shaft without an impeller attached (so that W _ f ¼ 0) mum and mass-averaged flow temperature and velocity can be
and evaluating Eq. (41) for the unloaded condition W _ ¼ found for each geometry and at each operating condition. It was
b
W _ f ¼ 0Þ. Using this calibration the mechanical power delivered
_ b ðW found that the location of the maximum and average values for
to the flow can be written as both temperature and velocity did not vary substantially with
_ f ðxÞ ¼ Ie V  I2 Rarm  W
_  ðxÞ the impeller speed or heat input. Therefore, for testing simplicity,
W e b ð42Þ
the contours shown in Fig. 6 were used to determine the placement
Using the calibration for the mechanical losses, the mechanical of the anemometer and thermocouple probe such that the maxi-
power delivered to the fluid is evaluated at every test condition mum velocity and average temperature were recorded in subse-
based on the electrical excitation of the motor and the mechanical quent tests. This greatly reduced the complexity and length of
losses at the resulting rotational speed. the testing procedure with little effect on the resulting data. The
experimental techniques used for acquiring mass flow rates and
_ To
3.2. Thermofluidic characterization – m; temperatures via scanned hot-wire anemometry and thermocou-
ple probes, as well as corrections for flow entrainment, are dis-
The heat transfer rate to the flow ðQ_ Þ is measured by monitoring cussed in detail elsewhere [19]. Using these techniques, the mass
the electrical input to the heaters and adjusting for heat lost to the averaged exit temperature ðT o Þ and total mass flow rate ðmÞ_ were
ambient through natural convection [19]. The temperature at the experimentally determined.
wall surface is simultaneously measured using several thermocou-
ples embedded within the copper plates. In the actual heat exchan-
4. Results and discussion
ger device shown in Figs. 1 and 2, the walls of the flow passages are
parallel condensers of a loop heat pipe and will operate at near iso-
The experimental measurements were carried out using air as
thermal conditions. Accordingly, the experimental apparatus con-
the working fluid (Pr = 0.72) for a fixed impeller tip radius (rt = 5
sidered in this work uses 6 mm thick copper plates to ensure
cm) and inlet flow radius (ri = 2 cm), resulting in a value of
isothermal conditions across each plate. In addition to these values,
I = 0.4. Using the measurement scheme outlined in Sections 3,
measurements of the average exit flow conditions are necessary to
3.1, 3.2, data was collected for ten independent geometries corre-
evaluate the nondimensional parameters in Eqs. (6)–(15).
sponding to channel heights of hc = 1.6  3.4 mm, blade thick-
The mass flow rate ðmÞ _ and mass-averaged flow temperature
nesses of hb = 0.5  2.8 mm, and for rotational speeds of
ðT o Þ are measured at the exit of the single layer’s periphery. By
3000  7000 rpm. Nondimensionally, this corresponds to blade as-
scanning a hot-wire anemometer (TSI 1210-60) and a thermocou-
pect ratios of 0.01 < B < 0.49, flow passage aspect ratios of
ple probe (Omega 5TC-TT-J-30-36) across the exit plane, contour
0.032 < G < 0.068, and rotational Reynolds numbers of
plots of exit velocity (vz) and flow temperature (To) have been gen-
4.7  104 < Rex < 1.1  105.
erated for each geometry. The anemometer and thermocouple
probe were continuously scanned vertically (y direction) at multi-
ple discrete horizontal (x direction) locations. Using these data sets 4.1. Empirically-derived correlations
velocity and temperature contours at the flow exit were created.
Fig. 6 shows contours for normalized flow velocity perpendicular The major dependent parameters to be empirically correlated to
to the exit plane as well as the exit flow temperature relative to the geometry and the rotational speed are the net mass flow rate
ambient conditions, where the impeller blades (and therefore flow) through the system ðmÞ, _ the net heat flow rate from the wall to
rotate from right to left. Fig. 6 shows that the flow conditions are the air ðQ_ Þ, and the net mechanical power delivered to the fluid
4556 J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

ðW_ f Þ. Accordingly, the nondimensional versions of these parame- independent parameters. Using a least squares fit applied to Eq.
ters have been correlated to the geometry and the rotational Rey- (43) for each experimentally measured parameter yields
nolds number using a functional dependence of the form
B
B C f ¼ 0:32G þ 2:4B  0:035 ð44Þ
f ðG; B; Rex Þ ¼ C 1 þ C 2 G þ C 3 B þ C 4 þ C 5 Rex ð43Þ G
G B
e ¼ 1  7:2G þ 4:5B  0:55 ð45Þ
This correlation used to relate the flow coefficient (Cf), the heat ex- G
changer effectiveness (e), and the slip factor (r), to the relevant B
r ¼ 0:11  0:96G  3:5B þ 0:54 ð46Þ
G

Fig. 7 shows the correlation results where the measured data is


(a) plotted relative to the calculated values from Eqs. (44)–(46). While
Eq. (43) was found to accurately correlate the geometric effects to
first order, it also showed little to no dependence of Cf, e, or r on
the rotational Reynolds number (Rex).
Using the correlations for Cf, e, and r in Eqs. (44)–(46) the
dimensionless heat flux can be expressed in terms of nondimen-
sional geometric parameters and the rotational Reynolds number
using Eq. (15). Fig. 8 shows the experimentally measured dimen-
sionless heat flux plotted against the developed correlation.
Dimensionless heat fluxes as high as 48 have been measured at a
rotational Reynolds number of 1.0  105, corresponding to an aver-
age convective heat transfer coefficient of h = 197 W/m2 K at a
rotational speed of 7000 rpm.

4.2. Characteristic design charts

The empirical correlations in Eqs. (44)–(46) can be substituted


into Eqs. (17) and (18) to evaluate the thermal resistance and the
(b) mechanical power required to drive the flow. Using these relations,
design charts of the form shown in Fig. 9 can be generated for a gi-
ven characteristic device size operating at a fixed speed.
Fig. 9 shows contour plots of the device thermal resistance and
required pumping power as a function of the flow channel geomet-
ric parameters, B and G. Alternatively, Fig. 10 shows the thermal
resistance plotted directly as a function of pumping power for
eight experimentally characterized flow geometries.
Multilayer heat exchangers with interdigitated impeller blades
can be designed and optimized using the empirical correlations
developed in this work along with characteristic design charts sim-
ilar to those presented in Figs. 9 and 10. The analysis and modeling
presented here can be used to balance the design trade-offs of a
single layer geometry against the total number of layers permissi-
ble in a given application. This is exemplified by examining Fig. 10,
which shows increased performance for designs with a larger
parallel plate aspect ratio, G. For a multilayer stack with negligible

(c)

Fig. 7. Comparison of experimentally measured values and empirically derived


correlations for (a) flow coefficient, (b) heat exchanger effectiveness, and (c) slip Fig. 8. Experimentally measured dimensionless heat flux plotted relative to the
factor as a function of nondimensional geometric parameters B and G. developed empirical model.
J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559 4557

R - Thermal Resistance (K/W) 1.4


G = hc / rt B = hb / rt
.032 .010

(a) 1.2
.032
.050
.016
.010
.050 .023

R - Thermal Resistance (K/W)


.050 .036
.067 .010
1 .067 .029
0.04 .067 .049
0.4
0.8
B = h b / rt

0.6
0.03
0.5

0.45

0.4
0.5
5

0.6 0.2
0.02 0 0.5 1 1.5 2 2.5
0.55
0.6
5 Wf - Pumping Power (W)
0.6
0.7
0.
8

0.8 0.65 Fig. 10. Calculated thermal resistance of the single layer as a function of pumping
5 0.75 power (Eqs. (17), (18) and (44)–(46)) for various geometries. The inset table shows
0.7
the values of B and G corresponding to each symbol in the main plot. The markers
0.04 0.05 0.06 indicate linearly spaced rotational speeds ranging from 2000 to 8000 rpm in
increments of 500 rpm.
G = hc / rt

W f - Pumping Power (W) 4.3. Comparison with traditional approaches

(b) 1. 6 The novelty of the proposed heat exchanger design is the direct
integration of the fan and heat sink components. Traditionally, a
1.4 separate and independent fan is used to force air over a high-
surface-area heat sink, convectively cooling the structure. Using
0.04
1.2 the correlations developed for this work, the integrated fan-heat
sink design can be compared to traditional separate fin-fan
1 approaches where pressure-driven flow is used to cool the heat
B = hb / rt

sink. In this comparison, the traditional pressure-driven design


0.03 has geometry identical to the parallel plate structure studied in
0.8 the current work but without an impeller, taking the configuration
described in Section 2.2. The empirical correlations from Section
2.2 were used to determine the convective heat transfer and pres-
0.6
sure drop in the traditional pressure-driven design’s simulta-
0.02 neously developing radial flow through parallel plates. Fig. 11
0.4 shows a direct comparison of the two designs for mass flow rate,
heat transfer coefficient, thermal resistance, and required pumping
power at two different plate spacings. In both of the impeller-dri-
ven cases, the impeller occupies 75 percent of the plate spacing;
that is, B/G = 0.75. This B/G ratio represents a readily manufactured
0.04 0.05 0.06 design at these channel spacings. The integrated impeller-driven
G = hc / rt design is presented over the range of operating conditions directly
measured and empirically modeled in this work alongside the cor-
Fig. 9. Contour plots of (a) the convective thermal resistance between the wall and responding performance of the traditional pressure-driven design.
ambient air, and (b) power required to drive the flow for a single layer based on the Fig. 11(a) shows the average heat transfer coefficient over the
developed correlations with rt = 5 cm and 5000 rpm (Rex = 7.8  104). The dashed
lines represents a limit of manufacturable geometries; above this line the blade
internal surfaces of the parallel plates as a function of mass flow
thickness becomes too close to the gap thickness. rate for both approaches, showing higher heat transfer for impel-
ler-driven flow at a fixed mass flow rate. This enhancement in
the heat transfer coefficient is thought to be a consequence of in-
inlet resistance, the total required pumping power will increase creased mixing of the flow, higher velocity magnitudes due to
linearly with the number of layers, while the total thermal resis- the spiral-shaped path the air takes through the gap, and the devel-
tance will be inversely proportional to the number of layers. If a oping boundary layer being repeatedly sheared off with each pass-
multilayer device is to be designed to fit inside of a cube of a given ing blade. As would be expected, smaller channel heights lead to
size, a design with a larger flow passage will have a lower thermal higher heat transfer coefficients for both approaches.
resistance per layer but fit a fewer number of total layers. This Fig. 11(b) shows the net mass flow rate pumped through the
model allows a designer to determine the heat transfer for a given plates as a function of required pumping power for both
geometry, which can be considered along with manufacturing approaches. It can be seen that the impeller-driven flow is more
limitations to optimize application-specific performance efficiently pumped through geometry as compared to the pres-
requirements. sure-driven flow and, as would be expected, larger plate spacings
lead to more flow per unit power. This demonstrates how the
4558 J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559

(a)
h − Heat Transfer Coeff. (W / m 2K)
Predicted performance of an impeller-driven
200 heat sink, based on empirical correlations
developed in the current work
hc = 3 mm
150 hb = 0.75 · hc
hc = 4 mm

100
Predicted performance of a pressure-driven,
radial flow heat sink, based on empirical
50 correlations [Eqs. (19) and (33)]
hc = 3 mm
0 hc = 4 mm
0 2 4 6 8 10 12 14
m − Mass Flow Rate (g/s)

1.5
12
(b) (c)

R − Thermal Resistance (K / W)
m − Mass Flow Rate (g / s)

9
1

0.5
3

0 0
0 0.5 1 1.5 2 2.5 3 3.5 4 0 0.5 1 1.5 2 2.5 3 3.5 4
W − Pumping Power (W) W − Pumping Power (W)

Fig. 11. Performance of the integrated impeller-driven device as compared to a pressure-driven radial flow heat sink: (a) the average heat transfer coefficient as a function of
mass flow rate, (b) the mass flow rate as a function pumping power, and (c) the resulting thermal resistance from the heat sink to the air as a function of pumping power. The
uncertainties in the impeller-driven device performance are based on the propagation of uncertainties in the underlying empirical correlations (Eqs. (44)–(46)). The
uncertainties in these correlations are based on the standard deviations of the errors when compared to experimental data. The markers in the impeller-driven series
represent equally spaced rotational speeds ranging from 2000 to 6500 rpm.

integration of the fan directly within the heat sink improves the tally characterized. The parameters dictating the performance of
fluidic performance. The pumping power in the pressure-driven the system have been correlated to independent control variables
design is an underestimate in this analysis because entry losses using a set of nondimensionalized correlations. These correlations
to the fan section were neglected and generous estimates for both relate the parallel plate geometry and rotational speed to the
fan efficiency and motor efficiency were employed. resulting thermal resistance and the power required to drive the
The total thermal resistance between the structure and the flow. It has been shown that the use of impeller blades integrated
ambient air as a function of pumping power is the most relevant directly within the heat exchanger results in a factor of two
metric of heat exchanger performance and is plotted in Fig. 11(c). increase in performance over traditional approaches for millime-
A single layer thermal resistance as low as 0.4 K/W has been ter-scale plate spacings. The developed correlations provide a
achieved for the impeller-driven design with a plate spacing of comprehensive set of guidelines for designing such devices to meet
hc = 4 mm at a pumping power of 2 W. This corresponds to an in- next-generation cooling requirements using ambient air as the
crease in device performance of over a factor of two as compared coolant. The results of this work will be used to model and design
to the pressure-driven approach. Interestingly, the larger plate an interdigitated heat exchanger and fan design with multiple lay-
spacing (hc = 4 mm) performs better than the smaller plate spacing ers driven using an integrated permanent magnet motor.
(hc = 3 mm) for impeller-driven flow, while the opposite is seen for
pressure-driven flow. This can be explained by examining Figs.
Acknowledgements
11(a) and (b), where the heat transfer coefficient is greater for
smaller spacings (at a fixed mass flow rate), but the power required
This work was supported by the DARPA Microtechnologies for
to drive the flow is larger. The combination of these two effects re-
Air-Cooled Exchangers (MACE) program, grant # W31P4Q-09-1-
sults in an increase in performance for a single layer with larger
0007, under program manager Dr. Thomas Kenny. The authors
plate spacings (hc). Realistically, this effect is balanced by the num-
thank Dr. Barbara Hughey for her assistance with hot-wire
ber of layers that can be included in a design with a fixed total
anemometry.
height; designs with larger spacing can accommodate fewer layers.
As would be expected, this is not seen for pressure-driven flow due
to the weaker dependence of heat transfer coefficient on mass flow References
rate as seen in Fig. 11(a).
[1] I. Mudawar, Assessment of high-heat-flux thermal management schemes, IEEE
Trans. Compon. Pack. Technol. 24 (2001) 122–141, doi:10.1109/6144.926375.
5. Conclusions [2] S. Garimella, V. Singhal, D. Liu, On-chip thermal management with
microchannel heat sinks and integrated micropumps, Proc. IEEE 94 (2006)
1534–1548, doi:10.1109/JPROC.2006.879801.
A single layer of an air-cooled parallel plate heat exchanger sys- [3] S.S. Anandan, V. Ramalingam, Thermal management of electronics: a review of
tem driven by interdigitated impeller blades has been experimen- literature, Thermal Sci. 12 (2008) 5–26.
J.M. Allison et al. / International Journal of Heat and Mass Transfer 54 (2011) 4549–4559 4559

[4] M. McCarthy, T. Peters, J. Allison, A. Espinosa, D. Jenicek, A. Kariya, C. Koveal, J. [10] J. Stafford, E. Walsh, V. Egan, P. Walsh, Y.S. Muzychka, A novel approach to low
Brisson, J. Lang, E. Wang, Design and analysis of high-performance air-cooled profile heat sink design, J. Heat Transfer 132 (2010) 091401.
heat exchanger with an integrated capillary-pumped loop heat pipe, in: 12th [11] G.P. Beretta, E. Malfa, Flow and heat transfer in cavities between rotor and
IEEE Intersociety Conference on Thermal and Thermomechanical Pheno- stator disks, Int. J. Heat Mass Transfer 46 (2003) 2715–2726.
mena in Electronic Systems (ITherm), 2010, pp. 1–8. doi:10.1109/ITHERM. [12] F.M. White, Fluid Mechanics, fifth ed., McGraw-Hill, 2003. pp. 760-761.
2010.5501375. [13] S.L. Dixon, Fluid Mechanics and Thermodynamics of Turbomachinery, fifth ed.,
[5] D. Tuckerman, R. Pease, High-performance heat sinking for vlsi, IEEE Electr. Elsevier, Butterworth-Heinemann, 2005. p. 227.
Device Lett. 2 (1981) 126–129, doi:10.1109/EDL.1981.25367. [14] R. Lewis, Turbomachinery Performance Analysis, Wiley, New York, 1996.
[6] R.W. Knight, J.S. Goodling, D.J. Hall, Optimal thermal design of forced [15] R.K. Shah, A.L. London, Laminar Flow Forced Convection in Ducts: A Source
convection heat sinks-analytical, J. Electron. Pack. 113 (1991) 313–321. Book for Compact Heat Exchanger Analytical Data, Academic Press Inc., 1978.
[7] P. Teertstra, M. Yovanovich, J. Culham, T. Lemczyk, Analytical forced [16] N.V. Suryanarayana, T. Scofield, R.E. Kleiss, Heat transfer to a fluid in radial,
convection modeling of plate fin heat sinks, in: Fifteenth Annual IEEE outward flow between two coaxial stationary or corotating disks, J. Heat
Semiconductor Thermal Measurement and Management Symposium, 1999, Transfer 105 (1983) 519–526.
pp. 34–41. doi:10.1109/STHERM.1999.762426. [17] P. Moller, Radial flow without swirl between parallel discs, Aeronaut. Quart. 14
[8] J. Culham, Y. Muzychka, Optimization of plate fin heat sinks using entropy (1963) 163–186.
generation minimization, IEEE Trans. Compon. Pack. Technol. (2001) 159–165, [18] J. Fay, Introduction to Fluid Mechanics, MIT Press, Cambridge USA, 1994.
doi:10.1109/6144.926378 24. [19] J. Allison, Air flow in a high aspect ratio heat sink, Master’s thesis,
[9] V. Egan, P. Walsh, E. Walsh, R. Grimes, Thermal analysis of miniature low Massachusetts Institute of Technology, 2010.
profile heat sinks with and without fins, J. Electr. Pack. 131 (2009).

You might also like