Vous êtes sur la page 1sur 13

148

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

Comparison of Micro-Pin-Fin and Microchannel


Heat Sinks Considering Thermal-Hydraulic
Performance and Manufacturability
Benjamin A. Jasperson, Yongho Jeon, Kevin T. Turner, Frank E. Pfefferkorn, and Weilin Qu

Abstract This paper explores the potential of micro-pinfin heat sinks as an effective alternative to microchannel heat
sinks for dissipating high heat fluxes from small areas. The
overall goal is to compare microchannel and micro-pin-fin heat
sinks based on three metrics: thermal performance, hydraulic
performance, and cost of manufacturing. The channels and pins
of the microchannel and micro-pin-fin heat sinks, respectively,
have a width of 200 m and a height of 670 m. A comparison of
the thermal-hydraulic performance shows that the micro-pin-fin
heat sink has a lower convection thermal resistance at liquid flow
rates above approximately 60 g/min, though this is accompanied
by a higher pressure drop. Methods that could feasibly fabricate
the two heat sinks are reviewed, with references outlining current
capabilities and limitations. A case study on micro-end-milling
of the heat sinks is included. This paper includes equations that
separate the fabrication cost into the independent variables that
contribute to material cost, machining cost, and machining time.
It is concluded that, with micro-end-milling, the machining time
is the primary factor in determining cost, and, due to the
additional machining time required, the micro-pin-fin heat sinks
are roughly three times as expensive to make. It is also noted
that improvements in the fabrication process, including spindle
speed and tool coatings, will decrease the difference in cost.
Index Terms Micro heat sink, micro-manufacturing, micromachining, pin-fin heat sink.

N OMENCLATURE

nf
tchange
tcleaning
tmachining
tsetup
ttoolchange
dstraight
dstgpin

Total cost of heat sink [$].


Total cost of tools [$].
Total cost of materials [$].
Feedrate [mm/min].
Time to fabricate heat sink [min].

Manuscript received November 25, 2008; revised February 19, 2009. First
version published October 13, 2009; current version published March 10,
2010. This work was supported by the National Science Foundation, Grant
CBET-0729693 at the University of Wisconsin-Madison, and Grant CBET0730315 at the University of Hawaii at Manoa. Recommended for publication
by Associate Editor A. Bhattacharya upon evaluation of the reviewers
comments.
Y. Jeon was with the Department of Mechanical Engineering, University of
Wisconsin-Madison, Madison, WI 53706 USA. He is now with the Hyundai
Motors, Seoul, South Korea (e-mail: yjeon@hyundai-motor.com).
K. T. Turner, B. A. Jasperson, and F. E. Pfefferkorn are with the Department of Mechanical Engineering, University of Wisconsin-Madison, Madison, WI 53706 USA (e-mail: kturner@engr.wisc.edu; bajasperson@wisc.edu;
pfefferk@engr.wisc.edu).
W. Qu is with the Department of Mechanical Engineering, University of
Hawaii at Manoa, Honolulu, HI 96822 USA (e-mail: qu@hawaii.edu).
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TCAPT.2009.2023980

Chipload [mm].
Operator rate [$/hr].
Cost per tool [$].
Width of heat sink for case study [mm].
Length of heat sink for case study [mm].
Spindle speed [rpm].
Number of tools required for fabrication of one
heat sink.
Number of flutes.
Time to change one tool [min].
Time to clean up after machining [min].
Machining time [min].
Setup time before machining [min].
Total time to change tools [min].
Tool path to machine straight channel heat sink
[mm].
Tool path to machine staggered pin fin heat sink
[mm].

Thermal-Hydraulic Performance Variables


At
Aht
Aht, eff

Manufacturing Variables
Ctotal
CT
CM
fr
t

tc
R
CT
w
l
Ns
Ntools

h
Hfin
L hs
P
Pdh
Pfh

qeff
Rconv
Tf
Tw
Whs
Wch
Wfin
Wt

Area of heat sink base surface [m2 ].


Total heat transfer area of microscale enhancement
structure [m2 ].
Total effective heat transfer area of microscale
enhancement structure [m2 ].
Heat transfer coefficient [W/m2 C].
Height of fin [m].
Length of heat sink [m].
Pressure drop across heat sink [bar].
Pressure drop in developing region [bar].
Pressure drop in fully-developed region [bar].
Heat flux based on heat sink base area [W/cm2 ].
Average convection thermal resistance [C/m].
Water bulk temperature [C].
Fin base temperature [C].
Width of heat sink [m].
Width of flow channel [m].
Width of fin [m].
Mass flow rate [g/min].

Greek Symbols

Aspect ratio of microchannel.


Viscosity evaluated at coolant bulk temperature,
[Ns/m2 ].

1521-3331/$26.00 2010 IEEE

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

Whs = 1.0 cm, Lhs = 3.38 cm

149

Whs = 1.0 cm, Lhs = 3.38 cm

SL = 400 m

Hfin = 670 m

Hfin = 670 m
Lfin = 200 m

Wfin = 200 m

Wch = 200 m

Wfin = 200 m

Flow

Flow
(a)

Fig. 1.

Viscosity evaluated at fin base temperature, [N.s/m2 ].


Density, [kg/m3 ].

Average.
Liquid (water).
Inlet.
Microchannel heat sink.
Micro-pin-fin heat sink.
Outlet.
Wall.
I. I NTRODUCTION

(b)

Structure and dimension of (a) microchannel heat sink and (b) micro-pin-fin heat sink.

Subscripts
ave
f
in
mc
mpf
out
W

Wch = 200 m

HE CEASELESS pursuit of improved performance with


simultaneous reduction in volume leads to ever-increasing
dissipative heat loads in a wide range of modern microelectronic devices. It has been shown that the performance
and reliability of these devices are strongly affected by their
temperature as well as immediate thermal environment. As
a result, there is an increasing demand for highly efficient
thermal management techniques capable of dissipating high
heat fluxes from small areas.
Single-phase liquid-cooled miniature heat sinks, which
have internal heat transfer enhancement structures and flow
passages that are tens to hundreds of micrometers in size,
have emerged as one solution to the aforementioned thermal
management challenges. Among the large variety of possible
microscale enhancement structures, parallel-plate fins have
received the most attention so far [1][5]. These miniature

heat sinks consist of parallel channels aligned with the flow


[Fig. 1(a)]. Key technical merits of microchannel heat sinks,
as demonstrated by the previous studies, include low thermal
resistance to dissipative heat flux, high heat transfer area
to volume ratio, compact dimensions, and small coolant
inventory requirement [1][5].
Recent advances in microfabrication technologies, however,
allows more complex microscale geometries to be fabricated
directly into high-thermal-conductivity solid substrates (e.g.,
metals) at low cost. This makes it possible to explore more
complex and 3-D enhancement structures that may be more
effective in transferring heat than the aforementioned parallelplate fins. A possible configuration is staggered [Fig. 1(b)] or
aligned micro-pin-fin arrays [6][14].
Staggered micro-pin-fin heat sinks have the potential to
remove a high heat flux for a given volume of the heat
sink and flow rate of working fluid, and hence improve the
performance of the heat-generating component. Despite the
potential for improved heat transfer from micro-pin-fin heat
sinks, economics and realistic microfabrication options will
continue to play an important role in whether these devices
are a viable choice over the nearest alternative (e.g., straight
microchannel heat sinks). Unlike microchannel heat sinks,
whose thermal-hydraulic performance can be fairly accurately
described by conventional macrochannel analytical models [3],
[4], reliable analytical or numerical models for micro-pin-fin
heat sinks have not been developed yet due to the complex
nature of fluid flow and heat transfer. Existing studies on liquid
single-phase heat transfer and pressure drop in micro-pin-fin
arrays are mostly empirical. In these previous studies, specific

150

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

TABLE I
P OTENTIAL M ANUFACTURING M ETHODS FOR M ICRO H EAT S INKS
Mfg Method

Can Mfg

Mass Production
Suitability*

Prototyping
Suitability*

Cost Comparison
of Designs

EDM
Wire
Plunge

Channel
Both

Poor
Poor

Average
Average

NA
Pin Channel

Etching
LIGA/Electroforming

Both

Good

Good

Pin Channel

Casting

Both

Very Good

Poor

Pin Channel

Extrusion

Channel

Very Good

Poor

NA

Machining
End Mill
Slot/Form Mill

Both
Channel

Average
Good

Very Good
Very Good

Pin > Channel


NA

Both

Very Good

Poor

Pin Channel

Sintering

micro-pin-fin configurations were tested and new heat transfer


and pressure drop correlations proposed [6][14].
The objective of the present paper is to simultaneously
compare the thermo-hydraulic performance and manufacturability of the aforementioned two types of miniature heat
sinks. Material is presented in four sections: 1) a review of
manufacturing techniques that can be used to make these micro
heat sinks out of metals; 2) a thermal-hydraulic analysis of
single-phase water cooled copper heat sinks to explore whether
the micro-pin-fin design has the potential to outperform the
microchannel design; 3) a case study of micro-end-milling to
determine the difference in manufacturing cost of the two heat
sink designs; and 4) a discussion of the results.
II. R EVIEW OF M ANUFACTURING T ECHNIQUES
A. Scope of Analysis
Because it is a highly specialized and emerging area, there is
a need to review the different manufacturing methods that can
be used to fabricate microscale heat sinks. Due to the higher
thermal conductivity and mechanical performance of metal
alloys as compared to nonmetallic (i.e., silicon) materials,
this review focuses on the fabrication of micro heat sinks
out of metal alloys. This paper does not attempt to predict
which technique is best suited for making micro heat sinks,
because there are too many production variables that must
be considered when making that decision (material, design,
tolerances, quantity, existing equipment, etc.). Instead, the goal
is to critically review a variety of methods that may be well
suited for prototyping, low-volume production, or high-volume
production of heat sinks.
An excellent source on the fabrication of heat sinks with
features similar in size to those discussed in this paper is the
paper by Eugene et al. [15], which discusses the fabrication
of micro-meso heat sinks embedded in turbine blades. Eugene
et al. conclude that the three most viable candidates for
mass manufacturing microscale features inside turbine blades

are micro electrical discharge machining (micro-EDM), micro


laser machining, and micro casting.
B. Potential Fabrication Methods
Table I summarizes the potential fabrication methods
discussed below, their ability to make the two heat sink
geometries, their suitability for mass production and prototype
fabrication, and a comparison of the manufacturing cost for
each heat sink design.
1) Electrical Discharge Machining: EDM erodes/removes
material when a spark discharges between an electrode (tool)
and a workpiece. Material is removed from the workpiece
because of the rapid temperature rise and explosive phase
change resulting from the concentrated energy released by the
electric arcs [16]. The electrode does not experience the same
rate of material removal because its high thermal diffusivity
dissipates the heat more rapidly. Repeatedly discharging a
spark at high frequencies under controlled conditions allows
for bulk material removal around the tool. Hence, the cavity
that is created takes the inverse shape of the tool (or wire)
that is used as the electrode. Intricate and microscale designs
can be created in electrically conductive materials without
imparting large forces or a significant heat-affected zone [17].
The heat-affected zone is minimal because of the localized
nature of the repeated material removal events and the EDM
tool and workpiece are immersed in a dielectric fluid that
removes heat and debris while also controlling the arcs [16].
EDM uses either a thin wire or a shaped electrode as
the tool [17]. Wire EDM [Fig. 2(a)] uses wires down to a
diameter of 20 m and can create straight through-thickness
slots or cuts [18]. Electrodes are also commonly machined into
cylindrical or square cross-section bars and plunged straight
into the material to drill a hole or moved laterally to mill out
shapes [Fig. 2(b)]. Very complicated geometries that need to
be created repeatedly (high-volume production) often use a
shaped electrode that is plunged into the workpiece once in a
process called die sinking [Fig. 2(c)] [19]. Micro-EDM milling

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

151

Wire Direction

Workpiece
Direction
(a)

(b)

(c)

(d)

(e)

(f)

(a)

(b)

Fig. 3. Schematic of LIGA process: (a) deposit a conductive seed layer, (b)
spin on a thick layer of photoresist, (c) expose photoresist to high-energy
X-rays through a mask, (d) develop photoresist removing X-ray exposed
material, (e) deposit metal into photoresist mold, and (f) dissolve photoresist
mold.

(c)
Fig. 2. Illustrations of three electrodischarge machining techniques: (a) wire
EDM, (b) die sinking, and (c) EDM milling.

and drilling can utilize electrodes as small as 5 to 10 m in


diameter [16], [20], [21] and aspect ratios as large as 20 have
been obtained [22], [23]. These tools are more than adequate
for all features in the micro heat sinks being considered.
The cost of EDM is primarily a function of the time
required to shape each part due to the relatively low material
removal rate. Die sinking is the fastest method of creating the
heat sinks; however, it requires the most complicated (hence
expensive) die to be manufactured. Dies and all other types of
EDM electrodes wear out after repeated use [24]. Hence the
selection of an EDM method is a function of the part geometry
and volume. There are examples in the literature of heat sinks
that have been fabricated through these methods [25].
2) Photolithographic-Based Techniques: Photolithography
is a core fabrication technique utilized in the manufacturing
of integrated circuits and microelectromechanical systems.
Photolithography uses a transparent mask containing a desired
device pattern and an exposure source (e.g., a UV light source)
to transfer patterns onto a photodefinable polymer resist. The
patterned resist can be used as a mask for etching a substrate
or serve as a mold that can be filled with a metal [26]. Because
of the multiple process steps involved and significant overhead
associated with the facilities and equipment, photolithography

is best suited to batch production and most economical for


high-volume production.
Etching methods, such as deep reactive ion etching [27],
[28], can be employed to create heat sinks out of silicon,
but they cannot generally be used to create high-aspect-ratio
structures in metals and will not be discussed here. However,
lithography can be used to form metal heat sinks using
electrodeposition-based techniques such as the lithographie,
galvanoformung, und abformung (LIGA) process. The original
LIGA process has three main steps (Fig. 3): 1) A thick layer
of X-ray resist, typically poly(methyl methacrylate) (PMMA),
is deposited on a carrier substrate coated with a conductive
seed layer [Fig. 3(a) and (b)], 2) The resist is exposed to
high-energy X-rays through a mask [Fig. 3(c)] and then
developed [Fig. 3(d)], yielding a 3-D mold; 3) A method of
metal deposition, most commonly electroplating, is used to
fill the mold [Fig. 3(e)]; and 4) The resist mold is dissolved
(i.e., expendable), resulting in the final free-standing metal
component [Fig. 3(f)] [26]. Similar processes that use thick
photosensitive resists, such as SU-8 and PMMA, eliminate the
need for an X-ray source and provide the ability to produce
similar structures with reduced cost [29][31]. SU-8 processes
can also be incorporated to create positive molds, which then
can be used for subsequent metallic device creation [32].
LIGA processes are able to produce structures with aspect
ratios as large as 60:1 [33] with tolerances on the order of
micrometers. This method can make the smallest features of
any technique described in this paper; however, one must
sacrifice some resolution (i.e., tolerance) for increased aspect
ratio [34]. Common metals used in LIGA and LIGA-like
processes include nickel [30], copper [35], and gold [33].

152

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

Heating/Cooling Passages

(a)

(b)

(c)

Fig. 4. Schematic of die casting: (a) metal molds with runner and cooling passages, (b) molds pressed together with molten metal being inserted, and
(c) separation of molds and removal of part.

In addition, LIGA-like fabrication has been utilized to make


heat sinks or cooling plates in the past [36], [37]. The ability
to create complex 2-D shapes, as shown in [30], provides the
option for creating nonstandard pin shapes and optimizing the
heat sinks for thermal and hydraulic performance.
3) Casting: The process of casting, in its most basic form,
involves pouring a molten metal into a pre-fabricated mold,
allowing the metal to solidify, and then removing the part from
the mold [17]. Of the numerous casting methods, only the two
that produce the finest features and, hence, are most likely to
make microscale heat sinks will be described here: die casting
and investment casting.
Investment casting utilizes a wax (typically for macroscale)
or plastic (microscale) pattern that defines the shape of the
final part. Ceramic powder is poured around the pattern, dried,
and then sintered to increase the strength of the ceramic mold
and melt out the pattern. The mold is filled with molten metal
by vacuum die casting (evacuate the mold and pressurized gas
forces metal into it) or centrifugal casting (forces generated by
spinning are utilized), and after solidification the expendable
mold is removed [38].
Important considerations that determine the quality of
a microcasting include the preheating temperature of the
mold and the filling pressure of the mold. Baumeister et al.
[38] showed, for a particle-hardened gold-based alloy and
Albronze microcastings, that flowlength increases with an
increase in preheating temperature and filling pressure. Likewise, grain size increases with increasing preheating temperature due to the slower cooling rates.
In comparison to investment casting, metal-mold-based
microcasting (Fig. 4) offers the ability to reuse molds, increase
efficiency in production, and greater repeatability in part
production [39]. Aspect ratios of 8 or 9 can be achieved with
microcasting [40], and casting of features as small as 200 m
in size is feasible [38][42]. Die casting does have some
geometric limitations; notably, undercuts cannot be included

in the permanent die as it would be impossible to remove the


solidified part.
Cast parts only approach the theoretical density of the metal
and, hence, may have a slightly lower thermal conductivity
than the same part milled out of a forged or extruded billet.
Casting has been the mainstay of high-volume production of
complex metal parts. The cost of the permanent mold would
not make this method suitable for prototyping or low-volume
production.
4) Extrusion: Extrusion is a method of producing constant
cross-sectional area parts through the plastic deformation of
billets through a die (Fig. 5) [17]. Hence, this method could
make straight channel heat sinks [Fig. 1(a)] but not the
staggered micro-pin-fin design [Fig. 1(b)]. Most macroscale
metal heat sinks used for cooling computer chips are made
via extrusion. However, before extrusion can be applied to
the mass production of micro heat sinks, further research and
development is required. Microextrusion is an area of active
research [43], [44] with the promise of industrial application in
the not too distant future. Microextrusion processes encounter
two problems that are not found in their macroscale counterparts. Current process limitations include the precision of the
tools used in creation of the dies and the precision (i.e., backlash) present in forming machinery [45]. In addition, the size
of the final extruded part relative to the grain size of the billet
material has a significant effect on manufacturing. Krishnan
et al. [46] showed that 568 m diameter extruded pins with
grains 211 m in size tended to curl due to inhomogeneous
deformation, while pins with 32 m grain size did not.
5) Sintering: Sintering in the microfabrication realm may
take the form of micro powder injection molding (PIM). In
this process, a metal powder combined with a binder system
is injected into a mold of the final part shape (Fig. 6). After
injection, the binder is removed (through thermal means or
other methods) and the part is sintered [47]. Fu et al. [48]
demonstrated the ability to create 316L stainless steel pillars

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

153

Billet

Die
(a)
Fig. 5.

Schematic of extrusion process.

Fig. 7.

(a)

(b)

(c)

(d)

Fig. 6. Schematic of micro powder injection molding, (a) inject metalbinder mix into mold, (b) heat to remove binder, (c) sinter metal powder, and
(d) remove part.

through PIM that were 100 m in diameter, 200 m in


height, and had a 200 m pitch. A silicon master was created
to serve as the expendable mold. This process differs from
casting because the powderbinder mixture does not have to be
injected at elevated temperature, allowing for a wider variety
of mold materials to be used. The part can be sintered in
the mold or after it is removed from the mold. Similar to
casting, the near-net-shape part shrinks upon cooling, which
can induce distortion and stresses in the part, and a small
amount of porosity must be taken into account.
Other sintering fabrication methods include selective laser
sintering (SLS), where powder is deposited and then selectively sintered layer by layer to form a bulk part. Macroscale
systems may have a powder-feed cylinder which supplies
the powder to the machine, and a part-build cylinder, which
is incrementally lowered to create each layer. The powder
is typically transferred using a roller [17]. In micro-SLS,
a powder deposition device replaces the roller and selectively
places the powder for the micro features [49], [50]. Using this
concept, feature sizes as small as 100 m are reported [49].

(b)

Schematics of milling: (a) slot milling and (b) end milling.

6) Machining: a) Slot/form milling: Slot milling is a potential method of manufacturing the straight channel heat sinks.
A single circular cutter with teeth on the outer portion of the
bit or a cluster of cutters [Fig. 7(a)] can be used. Slotting saws
as thin as 150 m are commercially available, sufficient for the
feature sizes on the micro heat sinks being compared in this
paper.
b) End milling: Micro-end-milling [Fig. 7(b)] refers to an
end-milling process that uses cutting tools between 5 and
1000 m in diameter to create microscale features on micro-,
meso-, and macroscale parts [51]. It is a direct method of
creating true 3-D shapes in myriad materials, frequently in
a single process step. The fact that the geometry of interest
is created by a part program that controls the movement
of the end mill makes this method flexible. Therefore, it is
clearly suited for prototyping metal heat sinks and low-volume
production.
To maintain the same cutting speed as the diameter of
an end mill decreases, the spindle speed must be increased
proportionally. For example, to achieve the recommended
cutting speed for wrought aluminum alloys being end-milled
with a tungsten carbide tool (3.15 m/s [52]) a 200 m-diameter
end mill requires a spindle speed of 300 000 rpm. Currently,
there are 200 000 rpm spindles commercially available, and
ongoing research aims to develop spindles that can achieve
more than 1 million rpm [53]. However, most micro-endmilling is done with spindles between 50 000 and 100 000 rpm,
because it is not yet known if the cutting speeds that decades
of empirical data have shown to work well at the macroscale
are optimal for micromachining [54], [55]. Micro-end-mills
remove small amounts of material with each rotation, thus
high-speed spindles do not need to be powerful with costs
ranging from approximately $5,000 for a 50 000 rpm air-drive
spindle (fixed rpm) to $25 000 for a 200 000 rpm electric-drive
spindle with variable rpm.
As will be shown in the case study, the cost of machining
a heat sink is inversely proportional to the time it takes to
machine a part (productivity), which is mainly a function of the
feed rate (mm/min) at which a micro-end-mill can be moved
through the material. The feedrate fr is the product of the
chipload tc number of flutes (cutting edges) n f , and spindle
speed Ns
f r = tc n f N s .
(1)
Hence, doubling the spindle speed or number of flutes
will double the feedrate and cut the time to machine a

154

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

Chip Load
[ft/tooth]

Cutter
Rotation

Material
Feed
Fig. 8.

Schematic of chip load.

part in half. The third variable that influences the feedrate


(hence, productivity) is the chipload tc (Fig. 8): the depth of
engagement of a flute in the direction of travel. The magnitude of the chipload for a micro-end-mill is fundamentally
limited by the strength and flexibility of these small diameter
tools [51]. Decreasing the diameter of an end mill decreases
the flexural stiffness and the cutting forces it can withstand
without bending and/or breaking. The force acting on the tool
is a function of the chipload, depth of cut, and material being
machined. Decreasing the depth of cut will enable an increase
in the chipload but requires more passes of the tool to create
a feature of the desired depth.
Micro-end-milling is a viable option for prototyping and
low-volume production.
III. T HERMAL -H YDRAULIC P ERFORMANCE
Thermal-hydraulic performance of the micro-pin-fin heat
sink was determined experimentally with details provided in
previous papers [13], [14]. Only a brief overview is presented
here. Made of 110 copper, the micro-pin-fin heat sink had a
platform area of 1.0 cm in width (Whs ) by 3.38 cm in length
(L hs ). An array of 1950 staggered micro-pins with 200
200 m2 cross section and 670 m height were milled out of
the top surface (Fig. 9).
Thermal performance of the micro-pin-fin heat sink is
represented by an average convection thermal resistance Rconv
Tw,ave Tf ,ave
Rconv =
(2)
 A
qeff
t
where At is the total base area of the heat sink
At = Whs L hs = 1.0 3.38cm2 .

qeff

(3)

is the effective input heat flux, Tw,ave is the average


pin-fin base (wall) temperature, and Tf ,ave is the average water
(fluid) bulk temperature. Hydraulic performance is represented
by the measured pressure drop across the heat sink P.
 , T
Details on how to determine qeff
w,ave , Tf ,ave , and P
for the micro-pin-fin heat sink can be found in previous
papers [13], [14].
As heat transfer and pressure drop in microchannel heat
sinks can be adequately described by available analytical models developed for macrochannels [3], [4], a pseudo microchannel heat sink is proposed. Thermal-hydraulic performance of

the microchannel heat sink is determined using the heat transfer and pressure drop models summarized in Table II [56],
[57].
The performance of the two micro heat sink geometries
are compared assuming identical heat sink substrate material,
single-phase liquid coolant, overall dimensions, microscale
structure dimensions, and operating conditions. Fig. 1(a) illustrates the structure and key characteristic dimensions of the
microchannel heat sink along side those of the micro-pin-fin
heat sink. In particular, the (Wfin , Wch , Hfin ) combination for
the microchannel heat sink is chosen to be (200 m, 200 m,
670 m), which is the same as that for the micro-pin-fin heat
sink. Average convection thermal resistance for the microchannel heat sink is similarly evaluated from (2) using the heat
transfer models provided in Table II. Pressure drop across the
microchannel heat sink P is the sum of the pressure drop
across the upstream hydrodynamically developing entrance
region Pdh and the pressure drop across the downstream
fully developed region Pfh . Analytical models for evaluating
the two pressure drop components are provided in Table II.
Fig. 10(a) and (b) compare the average convection thermal
resistance for the micro-pin-fin heat sink and microchannel
heat sink for Tin = 30 and 60 C, respectively. The solid
line and dashed line in these figures are power-law curves
to best-fit the micro-pin-fin heat sink and microchannel heat
sink data, respectively, and are used to indicate the overall
data trend. It can be seen from Fig. 10(a) and (b) that Rconv
for the microchannel heat sink is fairly constant throughout the
total flow rate Wt range, while Rconv for the micro-pin-fin heat
sink is more sensitive to Wt and decreases significantly with
increasing Wt . In the low Wt range, Rconv for the micro-pin-fin
heat sink is higher than that for the microchannel heat sink,
but becomes lower at a higher Wt . The comparison indicates
a better micro-pin-fin heat sink thermal performance at an
elevated cooling water flow rate.
Fig. 11(a) and (b) compare the pressure drop across the
micro-pin-fin heat sink and microchannel heat sink for Tin =
30 and 60 C, respectively. It can be seen from Fig. 11(a)
and (b) that P in the micro-pin-fin heat sink is significantly
higher than that in the microchannel heat sink at all flow rates
tested.
IV. C ASE S TUDY: M ICRO -E ND -M ILLING
The authors have the most experience with micro-endmilling [Fig. 7(b)] and have used it to manufacture copper
micro heat sinks (Fig. 9). In this section, the cost of micro-endmilling two different heat sink geometries, pin-fin and straight
channel (Fig. 1), is compared. Only relative differences will
be highlighted since any productivity improvements that would
be applied to machining of one design would also be applied
to the other. The goal of this case study is to determine which
heat sink is more expensive to manufacture by micro-endmilling and why.
In order to compare the geometries directly, the same base
material (110 copper), and hence material cost and overall
heat sink geometry (width = 1.0 cm, length = 3.38 cm) are
assumed. Likewise, the pin width and gap (200 m) is consistent between heat sinks, and hence the same tool diameter

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

155

(a)
Fig. 9.

(b)

Photographs of (a) copper heat sink and (b) pin fin geometry created by micro-end-milling.
TABLE II
A NALYTICAL M ODELS FOR M ICROCHANNEL H EAT S INK [16], [17]
Heat transfer coefficient
For L 0.2 (thermally fully-developed flow),



k
w 0.14
h = N u3 d f

N u 3 = 8.235(1 1.883 + 3.767 2 5.814 3 + 5.361 4 2.0 5 ); =


h

ST Wfin
Hfin

L = Red z Pr
h
f
For L 0.2 (thermally developing flow),

  N u 3   k f   w 0.14

h = N u 4 + 8.68(103 L )0.506 exp (9.9776 ln () 26.379) L

Nu
d
4

N u 4 = 8.235(1 2.042 + 3.085 2 2.477 3 + 1.058 4 0.186 5 )


Pressure loss components

Pdh

2 f app,dh f u 2f L dh
Pdh =
; L dh = (0.06 + 0.07 0.04 2 )Rein dh
dh






+ 0.5


Re3.44
L
+
f

0.5 K () 4L +

fh
dh
dh
w 0.58
1 3.44 L +
f app,dh = Re
+

2

b
dh
+
1+C L dh

f fh Re = 24(1 1.355 + 1.947 2 1.701 3 + 0.956 4 0.254 5 )


2
3
dh
L+
dh = Redh ; K () = 0.6740 + 1.2501 + 0.3417 0.8358
L

C = (0.1811 + 4.3488 1.6027 2 ) 104


  0.58 
2
Pfh

Pfh =

2 f fh w b
dh

f u f L fh

; L fh = L L dh

is used for both geometries. Assuming the same tool material


results in a constant tool cost. Furthermore, if the same feed
rate is used in both cases, the tool life should be the same.
Machining both parts on the same machine with the same
operator running means that the tool change time is constant.
Since the heat sinks have the same geometric envelope and
similar features sizes, it is assumed that setup and cleaning
times are the same for both designs. The remaining variables
that determine the cost difference between the two heat sink

geometries are tool path, which dictates the machining time t,


and cost rate R.
The overall cost of the heat sink can be broken down into
the cost of tools C T , the cost of materials C M , as well as
the product of manufacturing time t and cost rate R (4). The
change in Ctotal between the heat sinks can be calculated
by taking the difference of each subcomponent of the total
cost
Ctotal = C T + C M + t R.

(4)

156

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

0.16

0.16
Micro-pin-fin heat sink
Microchannel heat sink

0.14

0.14
0.12
Rconv[C/W]

Rconv[C/W]

0.12
0.10
0.08

0.02
30

0.10
0.08
0.06

0.06
0.04

Micro-pin-fin heat sink


Microchannel heat sink

0.04
Water
Tin = 30 C
40

50

60
wt[g/min]

70

80

0.02
30

90

Water
Tin = 60 C
40

50

(a)
Fig. 10.

60
wt[g/min]

70

90

(b)

Comparison of micro-pin-fin heat sink and microchannel heat sink average convection thermal resistance for (a) Tin = 30 C and (b) Tin = 60 C.

0.12

0.12

Micro-pin-fin heat sink


Microchannel heat sink

0.10

0.10

0.08

0.08
P [bar]

P [bar]

Micro-pin-fin heat sink


Microchannel heat sink

0.06

0.06

0.04

0.04

0.02

0.02

Water
Tin = 60 C

Water
Tin = 30 C
0.00
30

40

50

60
wt[g/min]

70

80

90

(a)
Fig. 11.

80

0.00
30

40

50

60
wt[g/min]

70

80

90

(b)

Comparison of micro-pin-fin heat sink and microchannel heat sink pressure drop for (a) Tin = 30 C and (b) Tin = 60 C.

The cost of the tools C T is a function of the number of tools


required Ntools and the cost per tool C T . The number of tools
required to manufacture one heat sink is a function of tool life
and final part geometry, because it dictates the total tool path
and, along with feed rate, determines the machining time. In
this analysis, we are assuming that only one tool design is
used. However, in reality tools with shorter flute lengths will
be used where possible because they last longer
C T = Ntools C T.

(5)

Since the cost per tool does not change between heat sinks, the
total tool cost only varies with the number of tools. The tool
life and tool geometry are assumed constant between heat
sinks, meaning that the number of tools varies solely with
the final part geometry of the heat sinks.
The material costs C M are dependent on the volume of
material required and the per unit cost. The heat sinks

have the same overall dimensions and are made out of the
same material, so C M does not vary between the two heat
sinks.
The final term in (4) is the cost of the processing time.
The cost rate R includes the capital cost of machinery, any
overhead and utilities required for operation, additional training required for machining/setup of process, labor, etc. These
factors do not change, regardless of fabrication geometry, and
as such can be excluded from the current comparison.
The total time t required to manufacture the heat sink is a
function of the machining time, the amount of time required
for tool changes, setup time, and cleaning time (6). The
amount of time spent machining is a function of the feedrate
and geometry of the final design. More complex designs
require more passes with the tool, resulting in longer machining time (7). Feedrate is a complex function of multiple parameters, including material properties, tool strength/geometry,

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

157

wfin

wfin

2(wfin)
2(wfin)
(a)
Fig. 12.

(b)

Illustration of tool path for milling channel heat sink.

tool coating, and metal working fluids (8)


t = tmachining + ttoolchange + tsetup + tcleaning
(6)
tmachining = f (feed rate, part geometry, tool geometry) (7)
Feed rate = f (material, tool strength, tool geometry,
tool coating, metalworking fluid).

(8)

None of the variables that feed rate is a function of changes


between heat sinks, meaning that the feed rate can be assumed
to be constant. Likewise, the tools used to machine both
geometries can be 180 m in diameter (allowing for runout)
because both designs have the same characteristic dimension
(spacing between pin/wall). Therefore, machining time is
based solely on geometry.
The time required for tool changes is also related to the work
piece material. Shaping a material that is harder to machine
results in shorter tool life, more tool changes, and longer
machining time due to the increased tool changes. The time
per tool change is a function of the machine being used, and,
when automated, takes less than 1 min.
Since setup time, cleaning time, feed rate, machine used,
and operator are the same, the time to manufacture the heat
sinks varies only with the geometry of the part. The cost
equation (4) for comparing the two geometries simplifies to (9)
Ctotal = Ntools C T + (tmachining + ttoolchange ) R.

(9)

The remainder of this section will focus on the geometric


differences between straight channel and staggered-pin-fin heat
sinks and how to calculate the tool path length. The method
for machining a straight channel heat sink in a piece of copper
with length l and width w is shown in Fig. 12. For analysis
purposes, assume that w and l are multiples of wfin , which
is the width of one fin. The machining length dstraight that is
required to fabricate one layer for a straight channel for this
geometry is given by (10)
w wfin
(l) + w 3(wfin ).
dstraight =
(10)
2(wfin )
The tool path for machining one layer of a staggered pin-fin
heat sink dstgpin is more complicated (Fig. 13) and longer since
more material must be removed in order to create the extra
surface area that benefits heat transfer (11)



l wfin
w wfin
2w + 6wfin + 2wfin
. (11)
dstgpin =
2wfin
2wfin

(c)

(d)

(e)

(f)

Fig. 13. Illustration of tool path for milling staggered pin heat sink: (a) first
pass, illustrating the effect of tool radius on the corners of the pins, (b) second
pass, which finishes the first column of pins and makes the first cut on the
second column, and (c) through (f) repeating the process to make multiple
columns of pins.

Substituting typical numerical values (l = 3 cm, w = 1 cm,


w f in = 200 m) into (10) and (11) and comparing them
show that the tool distance for the staggered-pin heat sink
design used in this paper is approximately three times greater
than for the straight channel. Experience has shown that for a
200 m diameter tool, a depth of cut of approximately 50 m
is appropriate. Therefore, each heat sink would require 12
layers, each of length d, to machine the 600 m-deep pins.
Fig. 14 shows the total machining distance as a function of
the pin/wall width.
V. D ISCUSSION
When comparing the manufacturability of pin-fin versus
straight channel heat sinks, the geometries shown in Fig. 1
are assumed. The width of and distance between the pins
or channel walls [Fig. 1(a)] are 200 m and the aspect ratio
is 3:1, making the height of the pins/walls 600 m. For the
staggered pins [Fig. 1(b)], the distance between rows of pins is
200 m with alternating patterns. This geometry was also used
in the thermal-hydraulic performance analysis as well as the
micro-end-milling case study, to allow for a direct comparison
between all the different aspects of the study.

158

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

Machining Distance [m]

140
120
Staggered Pins
Straigth Pins

100
80
60
40
20
0

100

200
300
400
500
600
Pin/Wall Width, wfin [microns]

Fig. 14. Total machining distance (tool path as a function of pin/wall width
for a 1 cm 3.38 cm area).

The tool path for one layer of a straight channel and


staggered-pin-fin heat sink with 200 m feature size (wfin )
is approximately 830 and 2600 mm, respectively. The total
machining distances for micro-pin-fin and microchannel heat
sinks of this size (12 layers) are approximately 10 and 31.25 m,
respectively. With a feedrate of approximately 100 mm/min
for 200 m diameter end mills shaping pure copper, total
machining times of 100 and 312.5 min result. If the tool life
is approximately 2 meters, and each tool change takes 1 min,
then the straight channel can be made with 5 tools and the pin
fin requires 16. As a first-order estimation, the cost of each
tool is approximately the same as the hourly rate. Examining
these numbers in (9) shows that tmachining R is an order
of magnitude greater than any other costs that differ between
the two heat sink designs being considered. Therefore, the
cost of machining the heat sinks will scale with the machining distance (Fig. 14). Hence, the pin-fin heat sink will be
approximately three times as expensive to make if the material
costs, setup time, and cleaning time are much smaller than the
machining time. As improvements in productivity (i.e., feed
rate) are made by developing new tool coatings and using
higher speed spindles, the percentage of the part cost related to
machining time will decrease. This also means that the difference in cost between the two heat sink designs will decrease.
An appealing option that would minimize the difference
in manufacturing cost between the two miniature heat sinks
would be to fabricate a mold with micro-end-milling to
facilitate casting of the final part. If casting were used for
high-volume production of micro heat sinks, the difference in
unit cost between microchannel and micro-pin-fin heat sinks
would be small. Approximately the same amount of material
is used, and the tolerances and dimensions are similar. The
only significant cost difference would be in the mold, which
would be more complex for the micro-pin-fin heat sink.
Returning to the discussion on the thermal-hydraulic
performance, the average convection thermal resistance Rconv
for the two heat sinks can be written as
1
Rconv =
(12)
h ave Aht, eff
where h ave represents the average heat transfer coefficient, and
Aht,eff represents the total effective heat transfer area of the
microscale enhancement structures. Equation (12) indicates

that an improved thermal performance (low Rconv ) can be


achieved by either enhancing heat transfer (increasing h ave )
or increasing total effective heat transfer area Aht,eff .
Total heat transfer area Aht for the micro-pin-fin heat sink
is approximately 13.3 cm2 , and for the microchannel heat sink
13.0 cm2 . Total effective heat transfer area Aht,eff assumes
values lower than those of Aht due to the fin effect. Aht,eff
ranges from 12.0 to 12.9 cm2 for the micro-pin-fin heat sink,
and is 12.5 cm2 for the microchannel heat sink. This shows
that the pin-fin geometry that was chosen does not have a
significant area advantage over the chosen microchannel heat
sink geometry.
The better thermal performance for the micro-pin-fin heat
sink at high flow rate can therefore be attributed to enhanced
heat transfer. The data trend as shown in Fig. 10(a) and (b)
may be explained by the nature of water flow in the micropin-fin array. At low flow rate, flow in the micro-pin-fin array
is dominated by laminar flow, and vortices in the wake are
relatively weak. As a result, the downstream faces as well as
a substantial portion of the side faces of the square micropin-fins are not exposed to the main flow, which leads to a
less efficient use of the total heat transfer area. As flow rate
increases, flow in the micro-pin-fin array is more tortuous,
and vortices in the wake become stronger, which enhances
heat transfer through reducing boundary layer thickness and
activating a larger portion of pin-fin surface areas.
The higher pressure drop across the micro-pin-fin heat sink
as shown in Fig. 11(a) and (b) is a result of the drag force
presented by each and every pin-fin. Because a staggered
micro-pin-fin configuration was used, every pin-fin sees a flow
impinging on its upstream face. The pressure drop may be
decreased, while maintaining the same surface area, by using
micro-pin-fins with airfoil cross sections [6].
VI. C ONCLUSION
After comparing copper microchannel and micro-pin-fin
heat sinks (same characteristic dimensions; single-phase water
flow) using thermal performance, hydraulic performance, and
cost of manufacturing as metrics, it is concluded that neither
design is better for all applications.
The average convection thermal resistance decreases with
increasing flow rate for the micro-pin-fin, but it does not vary
significantly with flow rate for the microchannel heat sink.
Below a flow rate of approximately 60 g/min, the micro-pin-fin
heat sink has a higher thermal resistance than the microchannel
heat sink. Above 60 g/min, the micro-pin-fin heat sink has a
lower thermal resistance. This variation in thermal resistance
is attributed to the more tortuous flow and strong vortices in
the wake at high flow rate. Therefore, the micro-pin-fin heat
sink would be chosen for its better thermal performance at
flow rates above 60 g/min.
The pressure drop across the micro-pin-fin heat sink is
approximately twice as large as that across the microchannel
heat sink at low rates. The difference in pressure drop increases
with increasing flow rate for the range of flow rates evaluated
in this paper. Therefore, the improved thermal performance at
high flow rates comes with a significant increase in pressure

JASPERSON et al.: COMPARISON OF MICRO-PIN-FIN AND MICROCHANNEL HEAT SINKS

drop. Other pin designs (diamond, circular, airfoil, etc.) that


are being studied have the potential to provide the same thermal performance without the same increase in pressure drop.
Multiple manufacturing methods exist for creating heat
sinks out of metal. Casting and extrusion are the most economical choices for mass production; micro-end-milling and
micro-EDM are ideal for prototyping. If casting a significant
volume of heat sinks, the cost per unit would be similar for
both heat sink designs. The tool path needed to end-mill the
micro-pin-fin heat sinks is approximately three times greater
than for the microchannel heat sinks. The machining time is
directly related to the length of the tool path, and because
of the limited feed rates available at this time, the cost to
micromachine heat sinks is primarily a function of machining
time. Therefore, the cost to micro-end-mill a micro-pin-fin
heat sink is approximately three times greater than that for a
microchannel heat sink. As ongoing research enables increases
in spindle speeds and feedrates, the total cost to machine these
heat sinks will decrease and the difference in cost between the
two designs will decrease.
R EFERENCES
[1] T. M. Harms, M. J. Kazmierczak, and F. M. Gerner, Developing
convective heat transfer in deep rectangular microchannels, Int. J. Heat
Fluid Flow, vol. 20, no. 2, pp. 149157, 1999.
[2] K. Kawano, K. Minakami, H. Iwasaki, and M. Ishizuka, Microchannel
heat exchanger for cooling electrical equipment, presented at Proc.
ASME Int. Mech. Eng. Congr. Expo., Fairfield, NJ, 1998.
[3] P.-S. Lee, S. V. Garimella, and D. Liu, Investigation of heat transfer in
rectangular microchannels, Int. J. Heat Mass Transfer, vol. 48, no. 9,
pp. 16881704, 2005.
[4] W. Qu and I. Mudawar, Experimental and numerical study of pressure
drop and heat transfer in a single-phase microchannel heat sink, Int. J.
Heat Mass Transfer, vol. 45, no. 12, pp. 25492565, Jun. 2002.
[5] D. B. Tuckerman and R. F. W. Pease, High-performance heat sinking
for VLSI, IEEE Electron. Devices Lett., vol. 2, no. 5, pp. 126129,
May 1981.
[6] A. Kosar, C.-J. Kuo, and Y. Peles, Hydoroil-based micro-pin-fin heat
sink, in Proc. 2006 ASME Int. Mech. Eng. Congr. Expo., (IMECE06)
Microelectromech. Syst., Chicago, pp. 563570.
[7] A. Kosar, C. Mishra, and Y. Peles, Laminar flow across a bank of
low aspect ratio micro-pin-fins, J. Fluids Eng., Trans. ASME, vol. 127,
no. 3, pp. 419430, 2005.
[8] A. Kosar and Y. Peles, Thermal-hydraulic performance of MEMS-based
pin fin heat sink, J. Heat Transfer, vol. 128, no. 2, pp. 121131, 2006.
[9] Y. Peles and A. Kosar, Convective flow of refrigerant (R-123) across a
bank of micro-pin-fins, Int. J. Heat Mass Transfer, vol. 49, no. 1718,
pp. 31423155, 2006.
[10] Y. Peles, A. Kosar, C. Mishra, C.-J. Kuo, and B. Schneider, Forced
convective heat transfer across a pin fin micro heat sink, Int. J. Heat
Mass Transfer, vol. 48, no. 17, pp. 36153627, 2005.
[11] R. S. Prasher, Nusselt number and friction factor of staggered arrays
of low aspect ratio micro-pin-fins under cross flow for water as fluid,
J. Heat Transfer, vol. 129, no. 2, pp. 141153, 2007.
[12] A. Siu-Ho, W. Qu, and F. Pfefferkorn, Experimental study of pressure
drop and heat transfer in a single-phase micro-pin-fin heat sink, Trans.
ASME. J. Electron. Packaging, vol. 129, no. 4, pp. 479487, 2007.
[13] W. Qu and A. Siu Ho, Liquid single-phase flow in an array of micropin-fins. Part I. Heat transfer characteristics, J. Heat Transfer, vol. 130,
no. 12, pp. 122402-1122402-11, Dec. 2008.
[14] W. Qu and A. Siu Ho, Liquid single-phase flow in an array of micropin-fins. Part II: Pressure drop characteristics, J. Heat Transfer, vol. 130,
no. 12, pp. 124501-1124501-4, 2008.
[15] J. Eugene, F. Xi, B. Tan, and B. A. Jubran, An overview on micro-meso
manufacturing techniques for micro heat exchangers for turbine blade
cooling, Int. J. Manufacturing Res., vol. 3, no. 1, pp. 326, 2008.
[16] K. H. Ho and S. T. Newman, State of the art electrical discharge
machining (EDM), Int. J. Mach. Tools Manufacture, vol. 43, no. 13,
pp. 12871300, 2003.

159

[17] S. Kalpakjian and S. R. Schmid, Advanced manufacturing processes,


in Manufacturing Engineering and Technology, 5th ed. Upper Saddle
River, NJ: Pearson Education, Inc., 2006, ch. 27, pp. 846850.
[18] C. S. Lin, Y. S. Liao, and S. T. Chen, Development of a novel micro
wire-EDM mechanism for the fabricating of micro parts, presented at
Int. Conf. Advanced Manufacture, Taipei, Taiwan, 2005.
[19] X. Cheng, Development of ultraprecision machining system with
unique wire EDM tool fabrication system for micro/nano-machining,
CIRP Ann. Manufacturing Technol., vol. 57, no. 1, pp. 415420,
2008.
[20] D. T. Pham, S. S. Dimov, S. Bigot, A. Ivanov, and K. Popov, MicroEDM recent developments and research issues, presented at 14th Int.
Symp. Electromach., Edinburgh, Scotland, 2004.
[21] A. B. M. A. Asad, T. Masaki, M. Rahman, H. S. Lim, and Y. S. Wong,
Tool-based micro-machining, J. Materials Process. Tech., vol. 192
193, pp. 204211, 2007.
[22] M. M. Sundaram, S. Billa, and K. P. Rajurkar, Generation of high
aspect ratio micro holes by a hybrid micromachining process, presented
at ASME Int. Conf. Manufacturing Sci. Eng., Atlanta, GA, 2007.
[23] W. Ehrfeld, Micro electro discharge machining as a technology in
micromachining, in Proc. SPIE Int. Soc. Optical Eng., vol. 2879.
Austin, TX, 1996, pp. 332337.
[24] M. Sundaram and K. Rajurkar, Toward freeform machining by micro
electro discharge machining process, Trans. North Amer. Manufacturing
Res. Inst., vol. 36, pp. 381388, 2008.
[25] J. S. Coursey, J. Kim, and K. T. Kiger, Spray cooling of high aspect
ratio open microchannels, J. Heat Transfer, vol. 129, no. 8, pp. 1052
1059, 2007.
[26] M. J. Madou, Lithography, in Fundamentals of Microfabrication, 2nd
ed. Boca Raton, FL: CRC Press, 2002, ch. 1, pp. 110, 4950.
[27] R. Muwanga and I. Hassan, Flow boiling oscillations in microchannel
heat sinks, presented at 9th AIAA/ASME Joint Thermophysics Heat
Transfer Conf., San Francisco, CA, 2006.
[28] S. Stefanescu, M. Mehregany, J. Leland, and K. Yerkes, Micro jet array
heat sink for power electronics, presented at Proc. IEEE Micro Electro
Mechn. Syst., Orlando, FL, 1999.
[29] X. Wei, C. H. Lee, Z. Jiang, and K. Jiang, Thick photoresists for
electroforming metallic microcomponents, J. Mech. Eng. Sci., vol. 222,
no. 1, pp. 3742, 2008.
[30] C. H. Ho, K. P. Chin, C. R. Yang, H. M. Wu, and S. L. Chen, Ultrathick
SU-8 mold formation and removal, and its application to the fabrication
of LIGA-like micromotors with embedded roots, Sensors and Actuators
A (Physical), vol. 102, no. 12, pp. 130138, 2002.
[31] H. Guckel, High-aspect-ratio micromachining via deep X-ray lithography, Proc. IEEE, vol. 86, no. 8, pp. 15861593, Aug. 1998.
[32] J. Li, D. Chen, J. Zhang, J. Liu, and J. Zhu, Indirect removal of SU-8
photoresist using PDMS technique, Sensors and Actuators A (Physical),
vol. 125, no. 2, pp. 586589, 2006.
[33] L. T. Romankiw, Path: From electroplating through lithographic masks
in electronics to LIGA in MEMS, Electrochimica Acta, vol. 42, no. 20
22, pp. 29853005, 1997.
[34] R. K. Kupka, F. Bouamrane, C. Cremers, and S. Megtert, Microfabrication: LIGA-X and applications, Appl. Surface Sci., vol. 164, no. 14,
pp. 97110, 2000.
[35] R. Engelke, Complete 3-D UV microfabrication technology on strongly
sloping topography substrates using epoxy photoresist SU-8, Microelectron. Eng., vol. 7374, pp. 456462, 2004.
[36] A. J. Pang, Design, manufacture and testing of a low-cost microchannel
cooling device, presented at 6th Electron. Packaging Technol. Conf.,
Piscataway, NJ, 2004.
[37] L. S. Stephens, K. W. Kelly, D. Kountouris, and J. McLean, A pin
fin microheat sink for cooling macroscale conformal surfaces under the
influence of thrust and frictional forces, J. Microelectromech. Syst.,
vol. 10, no. 2, pp. 222231, 2001.
[38] G. Baumeister, R. Ruprecht, and J. Hausselt, Microcasting of parts
made of metal alloys, Microsyst. Technol., vol. 10, no. 3, pp. 261264,
2004.
[39] B. S. Li, M. X. Ren, C. Yang, and H. Z. Fu, Microstructure of Zn-Al4
alloy microcastings by micro precision casting based on metal mold,
Trans. Nonferrous Metals Soc. China, vol. 18, no. 2, pp. 327332, 2008.
[40] G. Baumeister, K. Mueller, R. Ruprecht, and J. Hausselt, Production of
metallic high aspect ratio microstructures by microcasting, Microsyst.
Technol., vol. 8, no. 23, pp. 105108, 2002.
[41] G. Baumeister, R. Ruprecht, and J. Hausselt, Replication of LIGA
structures using microcasting, Microsyst. Technol., vol. 10, no. 6,
pp. 484488, 2004.

160

IEEE TRANSACTIONS ON COMPONENTS AND PACKAGING TECHNOLOGY, VOL. 33, NO. 1, MARCH 2010

[42] S. Chung, S. Park, H. Jeong, I. Lee, and D. Cho, Replication techniques


for a metal microcomponent having real 3-D shape by microcasting
process, Microsyst. Technol., vol. 11, no. 6, pp. 424437, 2005.
[43] J. Cao and N. Krishnan, Recent advances in microforming: Science,
technology and applications, in Proc. Materials Sci. Technol. 2005,
Pittsburgh, PA, pp. 225234.
[44] N. Krishnan, J. Cao, and K. Dohda, Study of the size effects on friction
conditions in microextrusion Part I: Microextrusion experiments and
analysis, J. Manufacturing Sci. Eng., Trans. ASME, vol. 129, no. 4,
pp. 669676, 2007.
[45] M. Geiger, M. Kleiner, R. Eckstein, N. Tiesler, and U. Engel, Microforming, CIRP Ann. Manufacturing Technol., vol. 50, no. 2, pp. 445
462, 2001.
[46] N. Krishnan, J. Cao, B. Kinsey, S. A. Paraslz, and M. Li, Investigation
of deformation characteristics of micro-pins fabricated using microextrusion, presented at ASME Int. Mech. Eng. Congr. Expo., Orlando,
FL, 2005.
[47] S. G. Li, Dimensional variation in production of high-aspect-ratio
micro-pillars array by micro powder injection molding, Appl. Physics
A-Materials Sci. Process., vol. 89, no. 3, pp. 721728, 2007.
[48] G. Fu, Injection molding, debinding and sintering of 316L stainless
steel microstructures, Appl. Physics a-Materials Sci. Process., vol. 81,
no. 3, pp. 495500, 2005.
[49] J. Chen, J. Yang, and T. Zuo, Micro fabrication with selective laser
micro sintering, presented at 1st IEEE Int. Conf. Nano Micro Eng.
Molecular Syst., Zhuhai, China, 2006.
[50] X. Li, H. Choi, and Y. Yang, Micro rapid prototyping system for micro
components, Thin Solid Films, vol. 420421, pp. 515523, 2002.
[51] Y. Jeon, and F. Pfefferkorn, Effect of laser preheating the workpiece
on micro-end-milling of metals, presented at Proc. ASME Int. Mech.
Eng. Congr. Expo., Orlando, FL, 2005.
[52] E. Oberg, F. D. Jones, H. L. Horton, and H. H. Ryffel, Aluminum
alloys, in Machinerys Handbook, 28th ed. New York: Industrial Press,
2004, p. 1014.
[53] S. Jahanmir, Ultra-high-speed micro-milling spindle, Poster
MSECICMP2008-72573 presented at ASME Int. Conf. Manufacturing
Sci. Eng., Evanston, IL, Oct. 710, 2008.
[54] T. Ozel and T. Altan, Process simulation using finite element method
prediction of cutting forces, tool stresses and temperatures in highspeed flat end milling, Int. J. Mach. Tools Manufacture, vol. 40, no. 5,
pp. 713738, 2000.
[55] M. P. Vogler, R. E. DeVor, and S. G. Kapoor, On the modeling and
analysis of machining performance in micro-end-milling, Part I: Surface
generation, Trans. ASME. J. Manufacturing Sci. Eng., vol. 126, no. 4,
pp. 685694, 2004.
[56] R. D. Blevins, Applied Fluid Dynamics Handbook, 1st ed. New York:
Van Nostrand Reinhold Company, 1984, ch. 6, pp. 38123.
[57] R. K. Shah and A. L. London, Laminar Flow Forced Convection in
Ducts: A Source Book for Compact Heat Exchanger Analytical Data,
Supl. 1, New York: Academic Press, 1978, ch. 7, pp. 196222.

Benjamin A. Jasperson received the B.S. degree


in mechanical engineering from the University of
Wisconsin, Madison in May 2008. He is currently
pursuing masters degree in mechanical engineering
at the University of Wisconsin, under the supervision
of Professors Pfefferkorn and Turner.
His research areas include micro heat flux sensors,
micro end milling and micro fabrication.
Mr. Jasperson is a member of Tau Beta Pi
(Wisconsin Alpha).

Yongho Jeon received the B.S. degree in mechanical


engineering from Ajou University, South Korea,
and the Illinois Institute of Technology, in 2003.
He received the M.S.M.E. and Ph.D. degrees from
the University of Wisconsin, Madison, in 2005 and
2008, respectively.
He is currently the Manager of fundamental manufacturing engineering development team at Hyundai
Motors, Seoul, South Korea.

Kevin T. Turner received the B.S. degree in


mechanical engineering from Johns Hopkins University, Baltimore, MD, in 1999, and the S.M. and
Ph.D. degrees in mechanical engineering from the
Massachusetts Institute of Technology, Cambridge,
in 2001 and 2004, respectively.
Since 2005, he has been a Faculty Member in the
Department of Mechanical Engineering, University
of Wisconsin, Madison. His primary research interests are the mechanics and design of MEMS and
semiconductor manufacturing processes.
Dr. Turner is a member of American Society of Mechanical Engineers and
the Materials Research Society. In 2008, he received the ASEE Ferdinand
P. Beer and E. Russell Johnston, Jr. Outstanding New Mechanics Educator
Award.

Frank E. Pfefferkorn received the B.S.M.E. degree


from the University of Illinois, Urbana-Champaign,
in 1994, and the M.S.M.E. and Ph.D. degrees
in mechanical engineering from Purdue University,
West Lafayette, IN, in 1997 and 2002, respectively.
He has been a Faculty Member in the Department
of Mechanical Engineering, University of Wisconsin, Madison, since 2003. His primary research interest is in developing a science-based understanding
of manufacturing processes, including heat transfer
problems, micro end milling, friction stir welding,
thermally-assisted manufacturing, and laser micro-polishing.
Dr. Pfefferkorn is a member of the American Society of Mechanical
Engineers and the Society of Manufacturing Engineers. He is the recipient of
a Research Initiation Award and the 2007 Kuo K. Wang Outstanding Young
Manufacturing Engineer Award from the Society of Manufacturing Engineers.

Weilin Qu received the B.E. and M.S. degrees


in engineering thermophysics in 1994 and 1997,
respectively, both from Tsinghua University, Beijing,
China, and the Ph.D. degree in mechanical
engineering in 2004 from Purdue University,
West Lafayette, IN.
He joined the Department of Mechanical Engineering, University of Hawaii at Manoa, Honolulu as an
Assistant Professor in 2004, where he established the
MicroScale Thermal/Fluid Laboratory. His research
has been focused on microscale thermal/fluid transport processes, boiling and two-phase flow, high-heat-flux thermal management, and electronic cooling. His doctoral research involved experimental
study, theoretical modeling, and numerical analysis of the various transport
phenomena associated with single-phase liquid flow and forced convective
boiling in microchannels.
Dr. Qu is a member of the American Society of Mechanical Engineers.

Vous aimerez peut-être aussi