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

Performance Assessment of

Single and Multiple Jet


Impingement Configurations in a
Pablo A. de Oliveira
POLO—Research Laboratories for Emerging
Refrigeration-Based Compact
Technologies in Cooling and Thermophysics,
Department of Mechanical Engineering,
Federal University of Santa Catarina,
Heat Sink for Electronics Cooling
Florianopolis 88040-900, Santa Catarina, Brazil The performance of a novel impinging two-phase jet heat sink operating with single and
1 multiple jets is presented and the influence of the following parameters is quantified: (i)
Jader R. Barbosa, Jr. thermal load applied on the heat sink and (ii) geometrical arrangement of the orifices
POLO—Research Laboratories for Emerging (jets). The heat sink is part of a vapor compression cooling system equipped with an R-
Technologies in Cooling and Thermophysics, 134a small-scale oil-free linear motor compressor. The evaporator and the expansion
Department of Mechanical Engineering, device are integrated into a single cooling unit. The expansion device can be a single ori-
Federal University of Santa Catarina, fice or an array of orifices responsible for the generation of two-phase jet(s) impinging
Florianopolis 88040-900, Santa Catarina, Brazil on a surface where a concentrated heat load is applied. The analysis is based on the ther-
e-mail: jrb@polo.ufsc.br modynamic performance and steady-state heat transfer parameters associated with the
impinging jet(s) for single and multiple orifice tests. The two-phase jet heat sink was
capable of dissipating cooling loads of up to 160 W and 200 W from a 6.36 cm2 surface
for single and multiple orifice configurations, respectively. For these cases, the tempera-
ture of the impingement surface was kept below 40  C and the average heat transfer coef-
ficient reached values between 14,000 and 16,000 W/(m2 K). [DOI: 10.1115/1.4036817]

Keywords: jet impingement, heat sink, boiling, phase change, refrigeration

1 Introduction them. Additional design benefits associated with this approach are
the reduction of weight and volume of the cooling system [8].
Throughout the years, ground breaking advances in design,
The viability of a direct liquid cooling configuration is highly
microfabrication, and performance of electronic systems (e.g., com-
dependent on its ability to produce very high convective heat
puter technology and power modules) have been accompanied by
transfer coefficients in order to compensate for the inferior
higher rates of heat dissipation (and temperature rise), which are a
thermophysical properties of dielectric coolants. Two-phase cool-
challenging obstacle for technological innovation and development
ing schemes, such as sprays, impinging jets, and boiling in micro-
[1]. According to the 2015 International Electronics Manufacturing
channels, have been attracting the attention of the research
Initiative (iNEMI) roadmaps, contemporary microprocessors for
community for a number of years due to their enhanced heat trans-
desktop machines are designed to operate above 150 W, server and
fer characteristics [1,7,9–15].
computer cluster chips are likely to exceed power dissipation of
Mechanical vapor compression refrigeration can lower the
500 W per chip, and the power dissipation of chips in automotive
component (e.g., junction) temperature to a value inferior to that
applications, whose dimensions are typically smaller than desktop
of its surroundings. The integration of vapor compression systems
computer and servers, is reaching 300 W with heat fluxes of
with two-phase enhanced heat transfer schemes (jets, sprays,
200–300 W/cm2 [2]. Besides, the demand for reducing the size of
microchannels) has proven to be an innovative and effective way
insulated gate bipolar transistor power modules for hybrid and elec-
of removing highly concentrated heat loads [16–20].
tric vehicles was expected to increase the volumetric power rating
Jet impingement cooling has been extensively studied over the
in 10–50 times compared to the larger modules commonly used in
past two decades. The effects of several parameters such as nozzle
railway applications [3].
shape and diameter, impingement height, impact velocity, nozzle
Several cooling technologies have been studied with a focus on
pitch distance, jet array, subcooling degree, surface structure and
advanced military avionics and high-performance computers and
aging, cross flow and drainage conditions on the fluid dynamics,
servers [4–6] since novel efficient alternatives for high heat flux
and heat transfer of single- and two-phase impinging jets have
removal are required in these applications. Among the existing
been widely reported [21–28]. The use of multiple impinging jets
thermal management solutions, direct liquid cooling [7] enables
has been proposed in many studies. However, this cooling princi-
the contact between a dielectric liquid coolant and the electronic
ple has not yet been integrated into mechanical vapor compression
components taking advantage of the complete elimination of the
refrigeration loops [29–40]. Impinging jets have also been com-
thermal resistances of the different layers of materials separating
bined with microchannels aiming to create high-performance
hybrid cooling schemes that benefit from the attributes of both
enhanced heat transfer techniques [41–46]. A few works have pre-
1
Corresponding author. sented mechanical vapor compression systems that use thermal
Contributed by the Electronic and Photonic Packaging Division of ASME for
publication in the JOURNAL OF ELECTRONIC PACKAGING. Manuscript received January
devices based on spray cooling, but were not purposely designed
25, 2017; final manuscript received April 24, 2017; published online June 28, 2017. to be compact or miniaturized [47–53]. Oil lubricated compres-
Assoc. Editor: Dong Liu. sors [48,51–53], ancillary expansion valves [47–49,51], and tube

Journal of Electronic Packaging Copyright V


C 2017 by ASME SEPTEMBER 2017, Vol. 139 / 031005-1

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


evaporators [52,53] were often needed to operate the active cool- 60.05 kg/h). The volumetric flow rate of the condenser secondary
ing system, as revised in Refs. [54,55]. fluid—a 90%/10% vol. mixture of distilled water and ethylene
A typical single-stage vapor compression-based processor cool- glycol (WEG)—was measured with a Seametrics SPX-038 paddle
ing system is composed of a compressor, a condenser, an expan- wheel flow meter (uncertainty of 60.07 L/min). Absolute pressure
sion device, and a heat sink or cold plate evaporator. As a heat transducers (uncertainties of 60.05 bar and 60.03 bar for the
transfer strategy, two-phase impinging jets offer the advantage of high- and low-pressure sides, respectively) and resistance temper-
combining micro-orifices as the expansion device and the evapo- ature detectors (RTDs) (uncertainty of 60.20  C) were used to
rator/heat sink into a single cooling unit, which goes toward mini- measure the local pressures and temperatures (and monitor in real
aturization [54]. Besides, the full potential size reduction of the time the thermodynamic states of the refrigerant) at the points
system can be achieved by the usage of a small-scale compressor. shown in Fig. 1. The reading, logging, and storage of data are per-
This paper advances the performance evaluation of the compact formed with a National Instruments data acquisition system
active cooling system introduced in Refs. [54–56] by experi- (SCXI-1000) connected to a computer equipped with the Lab-
mentally evaluating its operation with multiple jets and focusing VIEW software. The data acquisition frequency was set at 10 Hz
on the influence of distinct parameters, namely, the applied ther- for all variables.
mal load and the geometrical configuration of multiple orifices. An electrical trace heater wrapped around the compressor suc-
The analysis presented here is based on thermodynamic per- tion line was installed to guarantee a safe operation of the com-
formance and steady-state heat transfer parameters associated pressor, as the refrigerant may exit the test section (two-phase jet
with the impinging jet(s) for single and multiple jet cooling heat sink) with a significant liquid fraction. In a real application
tests. (as opposed to a lab test device such as the present apparatus), the
thermal energy required to superheat the refrigerant vapor enter-
2 Experimental Work ing the compressor may come from an internal heat exchanger
(using the condenser as a heat source), not necessarily adding to
2.1 Experimental Apparatus. A schematic diagram of the the electrical supply to the system. A fixed refrigerant superheat-
experimental apparatus is shown in Fig. 1. The compressors are ing degree was maintained at the compressor inlet. The tempera-
small-scale (0.27 cm3 maximum volumetric displacement, 340 Hz tures of point 1 (compressor inlet in Fig. 1) and of the calorimeter
operating frequency, 1.3 kg weight per unit) R-134a oil-free linear were accurately controlled by dedicated hardware and software,
motor units equipped with frequency inverters to control their i.e., a built-in LabVIEW proportional–integral–derivative control-
volumetric displacement (piston stroke). Although the two units ler code, which operates with a specially designed voltage-to-
were installed in parallel to generate higher flow rates, only one current converter electronic board. The room temperature was
compressor was operated in this study. The compressor electrical controlled by a split-type air conditioner equipped with a
power consumption was measured with a digital power meter variable-speed compressor. The complete specification of setup
(Yokogawa WT230). Inside the compressor calorimeter, direct components, measurement instrumentation, and variable control
current (DC)-powered centrifugal fans are utilized to homogenize strategies can be found in Ref. [54]. As seen in Fig. 1, the experi-
the air temperature. Their electrical power consumption was mental facility has a secondary evaporation circuit (assembled in
measured using digital multimeters (Minipa—ET2517A). A parallel with the primary circuit), which is composed of a meter-
brazed plate counter-flow condenser and two cascade thermal ing (expansion) valve and an evaporator. The reason for incorpo-
baths controlled the temperature of the hot reservoir (ambient). rating a second evaporation circuit in the experimental loop is
The R-134a mass flow rate was measured with a Siemens allowed for thermal performance tests using other high heat flux
MASS 2100 DI 1.5 Coriolis mass flow meter (uncertainty of removal devices, such as microchannel evaporators. By using the

Fig. 1 Schematic diagram of the experimental apparatus: (1) compressor inlet (suction),
(2) compressor outlet (discharge), (3) condenser inlet (refrigerant side), (4) condenser
outlet (refrigerant side), (5) jet cooler inlet, (6) jet cooler outlet, (7) condenser inlet (WEG
side), (8) condenser outlet (WEG side), and (9) secondary evaporator inlet (not used in
this study)

031005-2 / Vol. 139, SEPTEMBER 2017 Transactions of the ASME

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Fig. 4 Dimensions of the threaded screw (nozzle): (a) size and
(b) orifice diameters (in lm)

thermal insulation and fluid drainage, a reservoir to collect the


spent two-phase mixture, a film heater, and a copper block
that emulates the electronic component. The external dimensions
of the jet cooler unit are 80 mm (width)  80 mm (depth)
 112.5 mm (height).
Fig. 2 Two-phase jet cooler: assembly and main components
High-pressure subcooled liquid from the condenser flows
[55,56] through the metallic cap directly into the orifice plenum. The ple-
num, shown in detail in Fig. 3, consists of an AISI 316 stainless
steel outer orifice plate, a polyacetal resin polyoxymethylene
same cooling system at the same operating conditions, a fairer and (POM) inner orifice plate, and the expansion device itself, which
more robust basis for comparing different technologies can be is an array of POM 10 mm long threaded screws with the actual
established. However, this part of the loop was not used in this orifices drilled along their centerlines, as shown in Fig. 4. The ori-
study. fice plates are held together by the screws. Up to 13 orifices in a
staggered array can be accommodated in the plenum, covering a
2.2 Two-Phase Jet Heat Sink. The core idea of the novel 20 mm side square area. The low thermal conductivity of POM
heat sink is to combine the expansion device and the evaporator reduces the thermal losses from the refrigerant to the surroundings
into a single cooling module using two-phase impinging jet(s) to (inner orifice plate and screws) before the expansion.
remove the heat load from a small surface. The two-phase jet The use of hollow threaded screws as expansion devices is one
cooler is presented in Fig. 2 together with some ancillary compo- of the main features of the novel jet heat sink. Considering the
nents. It is composed of a metallic cap, a jet impingement cham- range of orifice diameters that can be drilled and the many alterna-
ber, an internal orifice plenum, a polymeric unit that serves for tives in terms of number of screws and their positions, several jet

Fig. 3 Orifice plenum: (a) assembled components and (b) outer and inner orifice plates

Journal of Electronic Packaging SEPTEMBER 2017, Vol. 139 / 031005-3

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Fig. 5 Assembly of the skin heater (with the Teflon base plate) and copper block in the insula-
tion and drainage unit

array configurations can be tested with the same orifice plenum. by a Teflon base plate. This arrangement guarantees a good ther-
As the subcooled liquid passes through the nozzle, it expands gen- mal contact between the heater and the block. The thermal load
erating a downward-oriented jet inside a polycarbonate jet cham- was finely controlled by a DC digital power supply (Agilent
ber. The jet length, i.e., the orifice-to-heater distance, is defined as N5770A). During the experimental runs, the thermal load was
the vertical distance from the outlet of the orifice to the base of augmented in steps by increasing the voltage or the current pro-
the chamber, which is at the same level as the impingement vided by the power supply.
surface. The copper block has six RTD wells, as shown in Fig. 6. Five
The impingement surface is the top surface of a heated cylindri- RTDs are used to measure the temperature and allow for an estimate
cal copper block mounted inside a polycarbonate bottom piece of the surface temperature to be made using Fourier’s law. The
(the so-called insulation and drainage unit). This unit was spe- remaining RTD is connected to a commercial proportional–integra-
cially designed to thermally insulate the lateral surface and the l–derivative controller that functions as a safety system cutting the
bottom of the copper block and facilitate drainage of the two- heating power supply in the event of a temperature runaway associ-
phase mixture from the jet chamber into the reservoir just below ated with the critical heat flux. In the present experiments, the cutoff
it. The assembly procedure of the copper block in the unit is temperature was set at 68  C. The area of the target surface is 6.36
shown in Fig. 5. The small gap between the copper block and the cm2 (D ¼ 28.54 mm) and the distance from the plane of the RTDs to
insulation and drainage unit is sealed by two O-rings (the grooves the impingement surface is 10.03 mm [54,55].
are shown in Fig. 6). This prevented the flow of refrigerant into The copper block was designed so that its diameter is circum-
the skin heater cavity. The heat load is provided by a 200 W thin scribed the skin heater surface area (20 mm  20 mm), as shown
film heater pressed firmly against the bottom of the copper block in Fig. 5(a). Numerical simulations of the heat transfer in the

Fig. 6 (a) Copper block and (b) cross-sectional view of the copper block (RTD’s plane, dimen-
sions in millimeters)

031005-4 / Vol. 139, SEPTEMBER 2017 Transactions of the ASME

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Fig. 7 Internal orifice plenum showing the (a) single and (b)–(d) multiple orifice configurations:
(a) single jet, (b) multiple jet array #1, (c) multiple jet array #2, and (d) multiple jet array #3

copper block/insulation and drainage unit (the skin heater was subcooling degree at the outlet of the condenser, vapor mass qual-
modeled and implemented as a source term) revealed that the dif- ity at the outlet of the two-phase jet cooler, heat transfer rates, and
ference between the power dissipated at the skin heater and the compressor power. Based on the output variables, the jet impinge-
heat transfer rate effectively dissipated at the jet impingement sur- ment heat transfer coefficient can be calculated.
face (top surface of the copper block) was lower than 5%. Further All output data were recorded during a 15 min time interval,
details regarding this analysis can be found in Ref. [54]. Ds, after guaranteeing that the following steady-state criteria were
simultaneously satisfied:
2.3 Experimental Conditions and Procedure. The experi- sðXÞjDs  jUðXÞj (1)
mental tests were run at the maximum compressor displacement.
The diameter of the orifices was 300 lm and the orifice-to-heater  
distance (jet length) was 28.84 mm. The secondary fluid tempera- @X
max Ds  jsð XÞ (2)
ture was set at 25  C in all tests. The refrigerant charge in the sys- @s Ds
tem was adjusted so that the tests for each multiple jet array had
the same refrigerant mass flow rate at the lowest thermal load where sðXÞ is the standard deviation of a generic output variable
(75 W). The mass flow rate of the WEG mixture was kept fixed X and U expresses its overall experimental uncertainty. j is an
at 180 kg/h. This value was sufficient to produce a high overall integer corresponding to each one of the three divisions of the
thermal conductance in the condenser, thus guaranteeing large Gaussian probability distribution, i.e., j ¼ 1, 2 or 3. The left-hand
condenser heat transfer rates (up to 301 W). The room temperature side of Eq. (2) was estimated numerically by a second-order cen-
and the calorimeter temperature were set at 25  C. The compressor tered finite difference scheme.
inlet superheat, defined as DTsh ¼ T1  Tevap , was kept fixed at Before recording the data, Eqs. (1) and (2) were verified during
10  C. three consecutive time intervals, Ds, to ascertain that the steady-
The influence of the following variables was investigated: (i) state was actually reached. The output data were recorded only
applied thermal load and (ii) orifice array configuration (for the when the left-hand side of Eq. (2) was lower than 2sðXÞ. In the
tests with multiple jets). Figure 7 presents the single and the mul- following, three distinct cases are presented as numerical exam-
tiple jet arrays explored, which are different combinations of a ples of the computation procedure. Each case considers a different
five-orifice configuration. For the single-jet test, the orifice screw output variable, i.e., mass flow rate, temperature, and pressure, for
was positioned at the center of the orifice plate assembly and the single-jet, 75-W test.
dummy screws occupied the remaining positions. Case 1: Mass flow rate: Considering that sðm_ r ÞjDs ¼ 0:006 kg/h,
All experimental runs were carried out in the heating-up mode, Uðm_ r Þ ¼ 60:05 kg/h, ðmaxð@ m_ r =@sÞÞjDs Ds  maxðDm_ r =2Þ ¼
i.e., increasing the heat load until the critical heat flux was 0:003 kg/h.
reached, which explains the different number of tests for the Case 2: Temperature: Considering that sðT2 ÞjDs ¼ 0:16  C,
single- and multiple-jet configurations. The critical heat flux was UðT2 Þ ¼ 60:16  C, ðmaxð@T2 =@sÞÞjDs DsmaxðDT2 =2Þ¼0:15  C.
characterized experimentally by a sudden increase of the heater Case 3: Pressure: Considering that sðP2 ÞjDs ¼ 0:016 bar, UðP2 Þ
surface temperature and was confirmed by visual observations to ¼ 60:05 bar, ðmaxð@P2 =@sÞÞjDs Ds  maxðDP2 =2Þ ¼ 0:013 bar.
coincide with a partial or total dryout of the liquid film on the
heater surface.
The output variables of the experimental apparatus are the pres- 2.4 Dependent Variables. The dependent variables were cal-
sures, temperatures, refrigerant mass flow rate, refrigerant culated based on the experimental measurements, as described in

Journal of Electronic Packaging SEPTEMBER 2017, Vol. 139 / 031005-5

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Refs. [54–56]. Energy balances between the inlet and outlet of the Table 1 Expanded uncertainties (maximum values) of the cal-
condenser and superheating line (see Fig. 1) can be written as culated parameters associated with the present experiments

Q_ cond ¼ m_ r ðh3  h4 Þ (3) Calculated parameter Expanded uncertainty

Q_ c (W) 2.01
Q_ sh ¼ m_ r ðh1  h6 Þ (4) W_ comp (W) 4.53
COP 0.49
where m_ r is the refrigerant mass flow rate and the refrigerant x6 0.013
enthalpy at the compressor inlet, h1 , condenser inlet, h3 , and con- hs (W/m2 K)
 433.0
denser outlet, h4 , were computed via REFPROP 8.0 [57] using the Ts ( C) 0.11
local experimental values of pressure and temperature. Tcond ( C) 0.09
Tevap ( C) 0.26
The refrigerant enthalpy at the outlet of the two-phase jet DTsh ( C) 0.31
cooler, h6 , was calculated from an energy balance on the cooler, DTsc ( C) 0.17
neglecting the kinetic and potential energy contributions. Thus DP (bar) 0.05

Q_ c
h6 ¼ h5 þ (5)
m_ r

where Q_ c is the cooling capacity and the refrigerant The temperature of the surface of the copper block, Ts , was
specific enthalpy at the inlet of the jet cooler, h5 , was also determined through a linear extrapolation of Fourier’s law consid-
determined from local measurements of pressure and temperature. ering one-diimensional heat conduction in the axial direction
At steady-state, the cooling capacity is equal to the thermal load

provided by the skin heater, considering a perfectly insulated test LQ_ c
section. Ts ¼ TRTD  (9)
ks As
The vapor mass quality at the exit of the jet cooler, x6 , was
determined from h6 and the measured outlet pressure, P6 . The sat- where L is the distance between the copper block surface and the
uration pressures were calculated directly from the experimental plane of the RTDs and ks is the thermal conductivity of copper
data, i.e., Pevap ¼ P6 and Pcond ¼ P4 , both taken at the outlets of evaluated at TRTD , which is the arithmetic mean of the five RTDs
the jet cooler and condenser, respectively. used to measure the copper block temperature.
The indicated power, i.e., the useful work performed on the
refrigerant by the piston per unit time, can be determined by the 2.5 Energy Balances and Experimental Uncertainties. As
following equation: discussed in Ref. [55], energy balances were executed on the
experimental apparatus for each test to verify the consistency of
W_ ind ¼ m_ r ðh2  h1 Þ (6) the experimental measurements. The energy balance residue, deb ,
is defined as the difference between the heat transfer rates out of
where h2 is the enthalpy at the compressor outlet computed using and into the refrigeration system
the local experimental values of pressure and temperature.
The coefficient of performance (COP) has been defined consid- 
jQ_ cond  Q_ c þ Q_ sh þ W_ ind j
ering the energy consumption strictly necessary to remove the deb ¼  100% (10)
imposed heat load upon the jet cooler Q_ c þ Q_ sh þ W_ ind

Q_ c As shown in Fig. 1, a compressor calorimeter has been specially


COP ¼ (7) designed and constructed to provide an indirect measure of the
W_ rate of heat dissipation by the compressor through its shell. A full
where W_ ¼ W_ comp þ W_ fan is the sum of the power consumptions description of the calorimeter is given in Refs. [54,55]. The calo-
of the compressor and fan located inside the calorimeter (recom- rimeter energy (heat) balance residue, dcal , is defined as the rela-
mended by the compressor manufacturer to cool the compressor tive error of the measurement of the heat transfer rate through the
shell). compressor shell
The jet impingement heat transfer coefficient, hs , is a
key parameter in the design of the two-phase jet cooler. It is jQ_ shell;cal  Q_ shell;eb j
dcal ¼  100% (11)
defined by Q_ shell;eb

Q_ c where Q_ shell;eb is calculated from the energy balance on the com-
hs ¼ (8)
As ðTs  Tevap Þ pressor, i.e., W_ comp  W_ ind , and Q_ shell;cal is computed according to
the procedure presented in Refs. [54,55].
where As is the impingement surface area, Tevap is the evaporat- For the tests reported in this paper, the average value of deb was
 1.4%, which confirms that the energy balance was fully satisfied
ing temperature, and Q_ c is the heat transfer rate at the copper
and that the experimental facility was properly insulated and the
block surface, i.e., the fraction of the input heat load that
energy transfer rates were measured accurately. The compressor
actually reaches the impingement surface. Considering steady-
calorimeter energy balance was also satisfied (despite the some-
state conditions and negligible thermal losses through the layers
what larger average residue of 6.4%), indicating that the energy
of elastomeric thermal insulation placed around the jet cooler
transfer rates to and from the oil-free compressor were also meas-
unit, it was assumed that the heat load provided by the skin
ured accurately.
heater is totally absorbed by the refrigerant flowing inside the
Table 1 presents the expanded uncertainties for the calculated
jet cooler, including the drainage channels. Therefore, its value
parameters, according to the procedure of Ref. [58]. As these
equals the cooling capacity of the system. The procedure to
 uncertainties vary according to each experimental run, only the
determine Q_ c was outlined in detail in Ref. [55], with the rela- maximum values are reported [54]. Table 2 presents some results

tive difference between Q_ c and the thermal load being of the of the energy balance calculations considering different power
order of 5% for all tests. loads and experimental tests.

031005-6 / Vol. 139, SEPTEMBER 2017 Transactions of the ASME

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Table 2 Energy balance calculations for different experimental conditions

Variable Single-jet test, 125 W Jet array test #1, 100 W Jet array test #2, 75 W Jet array test #3, 200 W

Q_ cond (W) 278.83 251.19 235.82 299.85


Q_ c (W) 125.17 99.99 74.98 199.88
Q_ sh (W) 148.20 138.19 149.36 90.16
W_ ind (W) 10.05 16.83 14.77 13.51
deb (%) 1.6 1.5 1.4 1.2
W_ comp (W) 74.48 57.60 48.88 54.53
Q_ shell;eb (W) 64.43 40.77 34.11 41.02
Q_ shell;cal (W) 58.00 38.71 36.15 39.93
dcal (%) 10.0 5.1 6.0 2.7

3 Results
The coefficient of performance is shown in Fig. 8. A quasi-
linear increase of COP with Q_ c is observed, which results mainly
from the increase of Q_ c since W_ comp increases only slightly with
the thermal load, as depicted in Fig. 9. The multiple jet strategy
not only allows for larger thermal loads to be removed from the
heated surface (up to 200 W), but it also does so with a higher
COP compared to the single-jet case.
The physical explanation for the lower values of compressor
power consumption in the multiple jet cases lies in the corre-
sponding smaller value of pressure lift in comparison to the single
jet case. The behavior of the pressure lift (the difference between
the suction and discharge pressures) is shown in Table 3 for all
cases. For all the multiple jet arrays, the pressure lift is approxi-
mately equal to 4 bar (considering a combined experimental
uncertainty of 60.05 bar), which is almost half the value of pres-
sure lift observed in the single-jet case. As more orifices are pres-
ent in the multiple jets, the restriction imposed on the expanding
fluid is lower than that provoked by the single orifice (300 lm
diameter). As a result, a smaller refrigerant mass is accumulated
in the components on the high-pressure side of the system,
which reduces the pressure upstream of the orifice. The lower DP
Fig. 8 Coefficient of performance as a function of the applied through the orifice results in a significantly higher pressure down-
heat load (cooling capacity) for single and multiple jet impinge- stream (evaporating pressure). The behavior of the saturation tem-
ment cooling peratures (evaporating and condensing) is shown in Fig. 10, which
obviously follows that of the suction and discharge pressures. As
the condensing temperature is significantly reduced for tests with
multiple jet arrays, the subcooling degree at the condenser outlet,
i.e., DTsc ¼ Tcond  T4 , drastically diminished from above 21  C,
for the single-jet tests, to values ranging from 0.2 to 3.5  C, for the
multiple-jet tests.
In Figs. 8 and 9, no noticeable differences were perceived
between the values of COP and W_ comp for the multiple-jet arrays
#2 and #3. However, a higher compressor power consumption
was observed for array #1. Although the differences between val-
ues of W_ comp for arrays #1 and #2 (from 4.1 to 7.5 W) and for
arrays #1 and #3 (from 3.7 to 7.7 W) were within the expanded
experimental uncertainty reported in Table 1 (64.53 W), the exis-
tence of a deviation is clear. As can be seen in Fig. 11, the refrig-
erant mass flow rate was very similar for arrays #1, #2, and #3, so
the higher power consumption behavior for jet array #1 can be
attributed to larger values of heat dissipation rate through the
compressor shell compared to the other multiple jet cases. The
exact reason for this higher heat transfer rate is unknown—a
slightly higher air temperature (1  C) inside the calorimeter
chamber is a possible cause—but the fact is that consistently
higher compressor gas discharge temperatures have been meas-
ured in the tests with array #1, in comparison with the other multi-
ple jet tests (see Table 4). These higher discharge temperatures
resulted in higher values of W_ ind , Q_ shell;eb , and Q_ shell;cal .
Fig. 9 Compressor power consumption as a function of the The behavior of the average heater surface temperature is
applied heat load (cooling capacity) for single and multiple jet shown in Fig. 12. As can be seen, a jet configuration with more
impingement cooling spaced orifices (arrays #2 and #3) is advantageous for reaching

Journal of Electronic Packaging SEPTEMBER 2017, Vol. 139 / 031005-7

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Table 3 Pressure lift of the refrigeration system (bar) per cooling capacity (W)

Q_ c (W) Single jet (bar) Multiple jet array #1 (bar) Multiple jet array #2 (bar) Multiple jet array #3 (bar)

75 7.63 4.16 4.00 4.04


100 7.54 4.15 3.98 4.07
125 7.75 4.08 3.97 4.05
150 7.80 4.07 3.96 4.09
175 7.86 4.05 3.95 4.07
200 — 4.08 3.94 4.08

Table 4 Calorimeter data associated with the multiple orifice


tests

Q_ c (W) W_ ind (W) Q_ shell;eb (W) Q_ shell;cal (W) Tsuc ( C) Tdis ( C)

Jet array test #1


75 16.95 39.26 37.57 18.81 42.01
100 16.83 40.77 38.71 20.10 42.32
125 15.84 41.19 39.75 21.87 42.51
150 15.41 42.46 39.94 23.05 42.79
175 15.51 43.80 43.04 23.82 43.17
200 15.35 — 42.86 25.22 43.87
Jet array test #2
75 14.77 34.12 36.15 19.64 40.75
100 14.66 35.39 34.17 20.88 41.11
125 14.56 36.54 34.60 22.16 41.67
150 14.16 38.18 35.34 23.45 42.04
175 13.88 39.38 35.82 24.52 42.45
200 13.76 40.33 37.64 25.42 42.84
Jet array test #3
75 14.67 33.93 36.78 19.84 40.99
100 14.29 35.65 36.39 21.34 41.48
125 14.13 36.79 35.22 22.34 41.76
Fig. 10 Saturation temperatures (evaporating and condensing) 150 13.85 38.43 36.43 23.61 42.31
as a function of the applied heat load (cooling capacity) for sin- 175 13.68 39.67 39.60 24.75 42.80
gle and multiple jet impingement cooling (error bars not shown 200 13.51 41.02 39.93 25.79 43.30
as the temperature expanded uncertainties are smaller than the
symbol size, as seen in Table 1)

Fig. 12 Surface temperature as a function of the applied heat


Fig. 11 Refrigerant mass flow rate as a function of the applied load (cooling capacity) for single and multiple jet impingement
heat load (cooling capacity) for single and multiple jet impinge- cooling (error bars not shown as the temperature expanded
ment cooling uncertainties are smaller than the symbol size, as seen in
Table 1)

lower surface temperatures, particularly at high cooling capacities. 150 W, it is possible to see that the three multiple jet arrays give
Since the multiple jet arrays #2 and #3 differ by a 45 deg rotation very similar results. However, a gradual temperature increase is
of the orifice configuration, the similar temperature behavior of clearly perceived for array #1 at higher cooling loads (175 and
these two cases was somewhat expected. For cooling loads up to 200 W). At the more extreme conditions, partial or total dryout of

031005-8 / Vol. 139, SEPTEMBER 2017 Transactions of the ASME

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Table 5 Typical values of parameters related to the heat trans-
fer coefficient and surface temperature typical test conditions

Single-jet Jet array Jet array Jet array


Variable test test #1 test #2 test #3

Applied thermal load: Q_ c ¼ 100 W


P5 (bar) 12.11 6.96 6.91 7.08
P6 (bar) 5.43 4.16 4.27 4.33
Tevap ( C) 18.3 10.1 10.9 11.3
Ts ( C) 30.0 22.6 24.0 23.9
hs (W/(m2 K))
 12,758 11,948 11,377 11,904
Applied thermal load: Q_ c ¼ 125 W
P5 (bar) 12.46 7.04 6.98 7.12
P6 (bar) 5.56 4.41 4.45 4.48
Tevap ( C) 19.1 11.9 12.1 12.4

Ts ( C) 32.0 26.4 26.3 26.3
hs (W/(m2 K))
 14,543 12,906 13,185 13,412
Applied thermal load: Q_ c ¼ 150 W
P5 (bar) 12.75 7.10 7.04 7.23
P6 (bar) 5.79 4.58 4.65 4.67

Tevap ( C) 20.4 13.0 13.5 13.6
Ts ( C) 34.7 29.5 28.6 28.9
Fig. 13 Heat transfer coefficient as a function of the applied hs (W/(m2 K))
 15,667 13,579 14,831 14,667
heat load (cooling capacity) for single and multiple jet impinge-
ment cooling

Table 5 for some typical conditions. Therefore, one cannot draw


the liquid film on the surface was observed to coincide with the any meaningful conclusion about the performance of different
increase in the heater surface temperature. Such a temperature configurations by comparing their surface temperatures alone.
excursion is immediately perceived by the safety system, which It is also clear that the higher heat transfer coefficients seen in
cuts the heating power supply. Therefore, for each curve in Fig. 13 for the single jet tests are a direct consequence of the
Fig. 12, the highest temperature corresponds to the temperature higher values of mass flow rate (Fig. 11), which is known to be a
just before the total dryout. In other words, the heat load could not dominating parameter in determining the magnitude of the heat
be raised further without a significant increase of the heater tem- transfer coefficient in sprays and impinging jets [59]. In the pres-
perature above the controller cutoff temperature. ent experiments, the higher mass flow rates in the single jet tests
A single (concentrated) jet is more prone to the occurrence of are due to the larger pressure difference across the orifice, as seen
liquid splattering and partial dryout at high heat fluxes because of in Fig. 10 and Table 5.
the liquid maldistribution [54]. This becomes evident when It should be mentioned that the present experimental tests are
comparing the maximum thermal load achieved with each orifice different from typical jet impingement boiling heat transfer tests,
configuration. In Fig. 12, the single jet had the lowest maximum where the mass flow rate and the thermal load can be varied inde-
thermal load/cooling capacity, while the multiple jet configura- pendently. Here, like the heater surface temperature, the evaporat-
tions could sustain a maximum heat load of 200 W. Nevertheless, ing and the condensing pressures, the mass flow rate is also a
multiple jet array #1 exhibited a gradual increase in surface tem- dependent variable.
perature starting at 175 W, while the remaining multiple jet arrays In the present tests, the mass flow rate increases with the ther-
(#2 and #3) maintained the same trend until the critical heat flux mal load, as described in Fig. 11. The thermal conductance of the
was reached. This particular behavior of jet array #1 (gradual tem- impinging jet flow increases with the mass flow rate and with the
perature increase leading to the critical heat flux) can be justified heat load itself (nucleate boiling contribution). However, from
based on the liquid flow distribution on the surface. Since the liq- one point onward, the increase in mass flow rate (which cannot
uid jet is concentrated in the center of the heater, the formation of be controlled independently) is not sufficient to compensate for
dry spots is somehow facilitated by the liquid flow maldistribu- the formation of dry patches and hot spots, which deteriorate the
tion. In a way, the gradual temperature increase before the critical heat transfer coefficient leading to the critical heat flux. In this
heat flux for jet array #1 reflects the different temperatures meas- sense, the better liquid distribution on the heated surface pro-
ured by the RTDs embedded in the copper block. vided by the multiple jet arrangements resulted in a more effec-
For jet arrays #2 and #3, partial dryout also occurred prior to tive performance with respect to the critical heat flux than the
the critical heat flux (a slight increasing temperature slope is single jet.
observed starting at 150 W), but a more effective liquid distribu- Regarding the multiple jet arrays specifically, an explanation
tion delays the appearance of hot spots, so the average surface for the heat transfer coefficient behavior for jet array #1 in com-
temperature is lower than for jet array #1 at 175 and 200 W. For parison with cases #2 and #3 can be entirely based on the tempera-
thermal loads larger than 200 W, no multiple jet configurations ture behavior described above in terms of the geometric
could sustain a heater surface temperature higher than the cutoff configuration of the jets and formation of dry spots on the heated
value of 68  C. surface. For multiple jet array #1, the jets are positioned closer to
The jet impingement heat transfer coefficients shown in Fig. 13 each other and the impingement is concentrated at the center of
were computed from Eq. (8) using the experimentally determined the heated surface—a geometry that resembles the single-orifice
average surface temperatures and the evaporating temperatures impingement from the point of view of the heat transfer interac-
associated with each experimental condition. Prima facie, the tion. On the other hand, multiple jets are expected to distribute the
higher values of the heat transfer coefficient for the single jet tests liquid more uniformly on the surface. This increases the critical
appear to be at odds with the higher surface temperatures seen in heat flux and gives rise to a less abrupt transition from a fully wet-
Fig. 12 for this configuration. This apparent inconsistency can be ted to a partially wetted surface regime (dryout). This became
explained by the fact that the evaporating temperatures were sig- apparent through the gradual reduction of the heat transfer coeffi-
nificantly higher for the single orifice tests in comparison with the cient for arrays #1, #2, and #3 and was also confirmed by visual
multiple orifice cases, as seen in Fig. 10 and shown in detail in observations during the tests.

Journal of Electronic Packaging SEPTEMBER 2017, Vol. 139 / 031005-9

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


Acknowledgment
This work was made possible through the financial investment
from the EMBRAPII Program (POLO/UFSC EMBRAPII Unit—
Emerging Technologies in Cooling and Thermophysics). In addi-
tion, financial supports from Embraco and CNPq (Grant No.
573581/2008-8—National Institute of Science and Technology in
Cooling and Thermophysics) are duly acknowledged. An abridged
version of this paper was presented at the 16th International
Refrigeration and Air Conditioning Conference at Purdue, West
Lafayette, IN, 2016. Parts of the paper were presented at the 15th
IEEE ITHERM Conference at Las Vegas, NV, 2016.

Nomenclature
As ¼ impingement surface area, cm2
D¼ copper block diameter, cm
h¼ specific enthalpy, J kg1
hs ¼ jet impingement heat transfer coefficient, W m2 K1
ks ¼ thermal conductivity of the heated surface, W m1 K1
L¼ distance between the impingement surface and the
RTDs’ plane, mm
m_ r ¼ refrigerant mass flow rate, kg h1
Fig. 14 Vapor mass quality at the outlet of the jet cooler (x6 ) as
a function of the applied heat load (cooling capacity) for single
P¼ pressure, bar
and multiple jet impingement cooling Q_ c ¼ cooling capacity, W
Q_ cond ¼ condenser heat transfer rate, W
Q_ sh ¼ superheating heat transfer rate, W
Q_ shell ¼ heat transfer rate through the compressor shell, W

The outlet vapor mass quality, x6 , is shown in Fig. 14. As Q_ c ¼ heat transfer rate at the copper block surface, W
expected, the vapor quality increases steadily, exhibiting very s¼ standard deviation
close values for all multiple jet array configurations as well as T¼ temperature,  C
reaching very high values (up to 75%). According to Ref. [50], Tamb ¼ temperature of the hot reservoir (ambient),  C
the outlet vapor quality can be interpreted as an evaporation effi- TRTD ¼ mean temperature of the RTDs in the copper block,  C
ciency, since the liquid that leaves the jet cooler does not produce Ts ¼ impingement surface temperature,  C
a cooling effect. Therefore, high evaporation efficiencies are U¼ expanded uncertainty
observed (up to 75%), showing that the designed heat sink was W_ ¼ total input work rate, W
capable of converting a large amount of the impinging liquid jet W_ comp ¼ electrical power consumption of the compressor, W
into vapor, which is the main physical mechanism responsible for W_ fan ¼ electrical power consumption of the fan, W
the heat removal from the test surface. W_ ind ¼ indicated power, W
x¼ vapor mass quality
4 Conclusions X¼ generic output variable
d¼ heat balance residue, %
A novel two-phase jet heat sink that integrates the evapora-
DP ¼ pressure drop through the jet cooler, bar
tor and the expansion device into a single cooling module was
DTsc ¼ refrigerant subcooling degree,  C
combined with a refrigeration system which operates with a
DTsh ¼ refrigerant superheating degree,  C
compact R-134a oil-free linear-motor compressor. The applic-
Ds ¼ sampling interval, s
ability of the system in the removal of highly concentrated
heat loads has been demonstrated. Experiments have been con-
ducted for distinct jet configurations, i.e., single and multiple Subscripts
jet arrays. The influence of the applied thermal load and orifice
cal ¼ calorimeter energy (heat) balance
geometric configuration on the system performance was quanti-
cond ¼ condensing
fied using thermodynamic performance and heat transfer
eb ¼ first-law energy (heat) balance
parameters. For a fixed orifice diameter and a fixed jet length,
evap ¼ evaporating
operating the active cooling system with multiple orifice con-
figurations resulted in a better thermodynamic performance
than the single orifice configuration. Although more research is References
needed to determine the optimal multiple jet configurations, [1] Smakulski, P., and Pietrowicz, S., 2016, “A Review of the Capabilities of High
these seem to be the way forward to achieve higher critical Heat Flux Removal by Porous Materials, Microchannels and Spray Cooling
heat fluxes and, therefore, higher evaporation efficiencies (out- Techniques,” Appl. Therm. Eng., 104, pp. 636–646.
let vapor qualities). [2] Bar-Cohen, A., and Holloway, C. A., 2014, “Thermal Science and Engineering—
From Macro to Nano in 200 Years,” 15th International Heat Transfer Conference
The two-phase jet heat sink was capable of dissipating cool- (IHTC-15), Kyoto, Japan, Aug. 10–15, Paper No. IHTC15-FL01.
ing capacities of up to 160 W and 200 W from a 6.36 cm2 sur- [3] Nakayama, W., Suzuki, O., and Hara, Y., 2009, “Thermal Management of
face for single and multiple orifice configurations (#2 and #3), Electronic and Electrical Devices in Automobile Environment,” Fifth Interna-
respectively. For these cases, the temperature of the impinge- tional IEEE Vehicle Power and Propulsion Conference (VPPC), Dearborn, MI,
Sept. 7–10, pp. 601–608.
ment surface was kept below 40  C and the heat transfer coeffi- [4] Mudawar, I., 2001, “Assessment of High-Heat-Flux Thermal Management
cient reached values between 14,000 and 16,000 W/(m2 K). Schemes,” IEEE Trans. Compon. Packag. Technol., 24(2), pp. 122–141.
The compact vapor compression cooling solution introduced [5] Chu, R. C., Simons, R. E., Ellsworth, M. J., Schmidt, R. R., and Cozzolino, V.,
here can be further developed for specific applications in ther- 2004, “Review of Cooling Technologies for Computer Products,” IEEE Trans.
Device Mater. Reliab., 4(4), pp. 568–585.
mal management of power electronics for a variety of station- [6] Kheirabadi, A. C., and Groulx, D., 2016, “Cooling of Server Electronics:
ary and mobile systems (for instance, hybrid and electric A Design Review of Existing Technology,” Appl. Therm. Eng., 105, pp.
vehicles). 622–638.

031005-10 / Vol. 139, SEPTEMBER 2017 Transactions of the ASME

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or


[7] Bar-Cohen, A., Arik, M., and Ohadi, M., 2006, “Direct Liquid Cooling of High [35] Buchanan, R. A., and Shedd, T. A., 2013, “Extensive Parametric Study of Heat
Flux Micro and Nano Electronic Components,” Proc. IEEE, 94(8), pp. 1549–1570. Transfer to Arrays of Oblique Impinging Jets With Phase Change,” ASME J.
[8] Mudawar, I., Bharathan, D., Kelly, K., and Narumanchi, S., 2008, “Two-Phase Heat Transfer, 135(11), p. 111017.
Spray Cooling of Hybrid Vehicle Electronics,” 11th Intersociety Conference on [36] Joshi, S. N., Rau, M. J., Dede, E. M., and Garimella, S. V., 2013, “An Experi-
Thermal and Thermomechanical Phenomena in Electronic Systems (ITHERM), mental Study of a Multi-Device Jet Impingement Cooler With Phase Change
Orlando, FL, May 28–31, pp. 1210–1221. Using HFE-7100,” ASME Paper No. HT2013-17059.
[9] Agostini, B., Fabbri, M., Park, J. E., Wojtan, L., Thome, J. R., and Michel, B., [37] Joshi, S. N., Rau, M. J., and Dede, E. M., 2013, “An Experimental Study of a
2007, “State of the Art of Heat Flux Cooling Technologies,” Heat Transfer Single-Device Jet Impingement Cooler With Phase Change Using HFE-7100
Eng., 28(4), pp. 258–281. and a Vapor Extraction Manifold,” ASME Paper No. IMECE2013-63249.
[10] Kandlikar, S. G., and Bapat, A. V., 2007, “Evaluation of Jet Impingement, [38] Joshi, S. N., and Dede, E. M., 2015, “Effect of Sub-Cooling on Performance of
Spray and Microchannel Chip Cooling Options for High Heat Flux Removal,” a Multi-Jet Two Phase Cooler With Multi-Scale Porous Surfaces,” Int. J.
Heat Transfer Eng., 28(11), pp. 911–923. Therm. Sci., 87, pp. 110–120.
[11] Ebadian, M. A., and Lin, C. X., 2011, “A Review of High-Heat-Flux Heat [39] Maddox, J. F., Knight, R. W., and Bhavnani, S. H., 2015, “Local Thermal Meas-
Removal Technologies,” ASME J. Heat Transfer, 133(11), p. 110801. urements of a Confined Array of Impinging Liquid Jets for Power Electronics
[12] Marcinichen, J. B., Olivier, J. A., Lamaison, N., and Thome, J. R., 2013, Cooling,” 31st Semiconductor Thermal Measurement, Modeling and Manage-
“Advances in Electronics Cooling,” Heat Transfer Eng., 34(5–6), pp. 434–446. ment Symposium, (SEMI-THERM), San Jose, CA, Mar. 15–19, pp. 228–234.
[13] Kadam, S. T., and Kumar, R., 2014, “Twenty First Century Cooling Solution: [40] Gould, K., Cai, S. Q., Neft, C., and Bhunia, A., 2015, “Liquid Jet Impingement
Microchannel Heat Sinks,” Int. J. Therm. Sci., 85, pp. 73–92. Cooling of a Silicon Carbide Power Conversion Module for Vehicle
[14] Cheng, W.-L., Zhang, W.-W., Chen, H., and Hu, L., 2016, “Spray Cooling and Applications,” IEEE Trans. Power Electron., 30(6), pp. 2975–2984.
Flash Evaporation Cooling: The Current Development and Application,” [41] Sung, M. K., and Mudawar, I., 2009, “CHF Determination for High-Heat Flux
Renewable Sustainable Energy Rev., 55, pp. 614–628. Phase Change Cooling System Incorporating Both Micro-Channel Flow and Jet
[15] Riofrıo, M. C., Caney, N., and Gruss, J.-A., 2016, “State of the Art of Efficient Impingement,” Int. J. Heat Mass Transfer, 52(3–4), pp. 610–619.
Pumped Two-Phase Flow Cooling Technologies,” Appl. Therm. Eng., 104, pp. [42] Sung, M. K., and Mudawar, I., 2008, “Single-Phase and Two-Phase Cooling
333–343. Using Hybrid Micro-Channel/Slot-Jet Module,” Int. J. Heat Mass Transfer,
[16] Barbosa, J. R., Jr., Ribeiro, G. B., and Oliveira, P. A., 2012, “A State-of-the-Art 51(15–16), pp. 3825–3839.
Review of Compact Vapor Compression Refrigeration Systems and Their [43] Sung, M. K., and Mudawar, I., 2008, “Single-Phase and Two-Phase Heat Transfer
Applications,” Heat Transfer Eng., 33(4–5), pp. 356–374. Characteristics of Low Temperature Hybrid Micro-Channel/Micro-Jet Impinge-
[17] Trutassanawin, S., Groll, E., Garimella, S. V., and Cremaschi, L., 2006, ment Cooling Module,” Int. J. Heat Mass Transfer, 51(15–16), pp. 3882–3895.
“Experimental Investigation of a Miniature-Scale Refrigeration System for Elec- [44] Barrau, J., Chemisana, D., Rosell, J., Tadrist, L., and Iba~ nez, M., 2010, “An
tronics Cooling,” IEEE Trans. Compon. Packag. Technol., 29(3), pp. 678–687. Experimental Study of a New Hybrid Jet Impingement/Micro-Channel Cooling
[18] Marcinichen, J. B., Thome, J. R., and Michel, B., 2010, “Cooling of Microproc- Scheme,” Appl. Therm. Eng., 30(14–15), pp. 2058–2066.
essors With Micro-Evaporation: A Novel Two-Phase Cooling Cycle,” Int. J. [45] Chien, L.-H., Liao, W.-R., and Liu, H.-Y., 2014, “An Experimental Study of
Refrig., 33(7), pp. 1264–1276. Two-Phase Convection in Micro-Channels With Impinging FC-72 Jets,” Appl.
[19] Marcinichen, J. B., Olivier, J. A., de Oliveira, V., and Thome, J. R., 2012, “A Therm. Eng., 67(1–2), pp. 159–167.
Review of On-Chip Micro-Evaporation: Experimental Evaluation of Liquid [46] Muszynski, T., and Andrzejczyk, R., 2016, “Heat Transfer Characteristics of
Pumping and Vapor Compression Driven Cooling Systems and Control,” Appl. Hybrid Microjet-Microchannel Cooling Module,” Appl. Therm. Eng., 93, pp.
Energy, 92, pp. 147–161. 1360–1366.
[20] Mancin, S., Zilio, C., Righetti, G., and Rossetto, L., 2013, “Mini Vapor Cycle [47] Yan, Z. B., Toh, K. C., Duan, F., Wong, T. N., Choo, K. F., Chan, P. K., and
System for High Density Electronic Cooling Applications,” Int. J. Refrig., Chua, Y. S., 2010, “Experimental Study of Impingement Spray Cooling for
36(4), pp. 1191–1202. High Power Devices,” Appl. Therm. Eng., 30(10), pp. 1225–1230.
[21] Womac, D. J., Ramadhyani, S., and Incropera, F. P., 1993, “Correlating Equa- [48] Chunqiang, S., Shuangquan, S., Changqing, T., and Hongbo, X., 2012,
tions for Impingement Cooling of Small Heat Sources With Single Circular “Development and Experimental Investigation of a Novel Spray Cooling System
Liquid Jets,” ASME J. Heat Transfer, 115(1), pp. 106–115. Integrated in Refrigeration Circuit,” Appl. Therm. Eng., 33–34, pp. 246–252.
[22] Garimella, S. V., and Rice, R. A., 1995, “Confined and Submerged Liquid Jet [49] Tan, Y. B., Xie, J. L., Duan, F., Wong, T. N., Toh, K. C., Choo, K. F., Chan, P.
Impingement Heat Transfer,” ASME J. Heat Transfer, 117(4), pp. 871–877. K., and Chua, Y. S., 2013, “Multi-Nozzle Spray Cooling for High Heat Flux
[23] Lienhard, V. J. H., 1995, “Liquid Jet Impingement,” Annual Review of Heat Applications in a Closed Loop System,” Appl. Therm. Eng., 54(2), pp. 372–379.
Transfer, Vol. 6, Begell House, New York, pp. 199–270. [50] Xie, J. L., Tan, Y. B., Wong, T. N., Duan, F., Toh, K. C., Choo, K. F., Chan, P.
[24] Pan, Y., and Webb, B. W., 1995, “Heat Transfer Characteristics of Arrays of K., and Chua, Y. S., 2014, “Multi-Nozzle Array Spray Cooling for Large Area
Free-Surface Liquid Jets,” ASME J. Heat Transfer, 117(4), pp. 878–883. High Power Devices in a Closed Loop System,” Int. J. Heat Mass Transfer, 78,
[25] Whelan, B. P., and Robinson, A. J., 2009, “Nozzle Geometry Effects in Liquid pp. 1177–1186.
Jet Array Impingement,” Appl. Therm. Eng., 29(11–12), pp. 2211–2221. [51] Xu, H., Si, C., Shao, S., and Tian, C., 2014, “Experimental Investigation on
[26] Lindeman, B. A., and Shedd, T. A., 2013, “Comparison of Empirical Correla- Heat Transfer of Spray Cooling With Isobutane (R-600a),” Int. J. Therm. Sci.,
tions and a Two-Equation Predictive Model for Heat Transfer to Arbitrary 86, pp. 21–27.
Arrays of Single-Phase Impinging Jets,” Int. J. Heat Mass Transfer, 66, pp. [52] Hou, Y., Liu, J., Su, X., Qian, Y., Liu, L., and Liu, X., 2015, “Experimental
772–780. Study on the Characteristics of a Closed Loop R-134a Spray Cooling,” Exp.
[27] San, J.-Y., and Chen, J.-J., 2014, “Effects of Jet-to-Jet Spacing and Jet Height Therm. Fluid Sci., 61, pp. 194–200.
on Heat Transfer Characteristics of an Impinging Jet Array,” Int. J. Heat Mass [53] Chen, S., Liu, J., Liu, X., and Hou, Y., 2015, “An Experimental Comparison of
Transfer, 71, pp. 8–17. Heat Transfer Characteristic Between R-134a and R-22 in Spray Cooling,”
[28] Qiu, L., Dubey, S., Choo, F. H., and Duan, F., 2015, “Recent Developments of Jet Exp. Therm. Fluid Sci., 66, pp. 206–212.
Impingement Nucleate Boiling,” Int. J. Heat Mass Transfer, 89, pp. 42–58. [54] Oliveira, P. A., 2016, “Development of a Two-Phase Jet Heat Sink Integrated
[29] Fabbri, M., Jiang, S., and Dhir, V. K., 2005, “A Comparative Study of Cooling With a Compact Refrigeration System for Electronics Cooling,” Ph.D. thesis,
of High Power Density Electronics Using Sprays and Microjets,” ASME J. Federal University of Santa Catarina, Florian opolis, Brazil.
Heat Transfer, 127(1), pp. 38–48. [55] Oliveira, P. A., and Barbosa, J. R., Jr., 2017, “Novel Two-Phase Jet Impinge-
[30] Meyer, M. T., Mudawar, I., Boyack, C. E., and Hale, C. A., 2006, “Single- ment Heat Sink for Active Cooling of Electronic Devices,” Appl. Therm. Eng.,
Phase and Two-Phase Cooling With an Array of Rectangular Jets,” Int. J. Heat 112, pp. 952–964.
Mass Transfer, 49(1–2), pp. 17–29. [56] Oliveira, P. A., and Barbosa, J. R., Jr., 2016, “Two-Phase Jet Impingement Heat
[31] Browne, E. A., Michna, G. J., Jensen, M. K., and Peles, Y., 2010, “Microjet Sink Integrated With a Compact Vapor Compression System for Electronics
Array Single-Phase and Flow Boiling Heat Transfer With R-134a,” Int. J. Heat Cooling,” IEEE Intersociety Conference on Thermal and Thermomechanical
Mass Transfer, 53(23–24), pp. 5027–5034. Phenomena in Electronic Systems (ITHERM), Las Vegas, NV, May 31–June 3,
[32] Wang, Z. Y., Wong, T. N., and Duan, F., 2011, “Submerged Liquid Jet pp. 976–986.
Impingement Cooling,” 13th Electronics Packaging Technology Conference [57] Lemmon, E. W., Huber, M. L., and McLinden, M., 2007, “NIST Reference
(EPTC), Singapore, Dec. 7–9, pp. 660–666. Fluid Thermodynamic and Transport Properties (REFPROP): Version 8.0,”
[33] Whelan, B. P., Kempers, R., and Robinson, A. J., 2012, “A Liquid-Based Sys- National Institute of Standards and Technology (NIST), Boulder, CO.
tem for CPU Cooling Implementing a Jet Array Impingement Waterblock and a [58] Coleman, H. W., and Steele, W. G., 2009, Experimentation, Validation, and
Tube Array Remote Heat Exchanger,” Appl. Therm. Eng., 39, pp. 86–94. Uncertainty Analysis for Engineers, 3rd ed., Wiley, Hoboken, NJ.
[34] Parida, P. R., Ekkad, S. V., and Ngo, K., 2012, “Impingement-Based High Per- [59] Hsieh, S.-S., Fan, T.-C., and Tsai, H.-H., 2004, “Spray Cooling Characteristics
formance Cooling Configurations for Automotive Power Converters,” Int. J. of Water and R-134a—Part I: Nucleate Boiling,” Int. J. Heat Mass Transfer,
Heat Mass Transfer, 55(4), pp. 834–847. 47(26), pp. 5703–5712.

Journal of Electronic Packaging SEPTEMBER 2017, Vol. 139 / 031005-11

Downloaded From: http://electronicpackaging.asmedigitalcollection.asme.org/pdfaccess.ashx?url=/data/journals/jepae4/936321/ on 06/28/2017 Terms of Use: http://www.asme.or

You might also like