scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Experimental investigation into vortex structure and pressure drop across microcavities in 3D integrated electronics

12 Apr 2011-Experiments in Fluids (Springer-Verlag)-Vol. 51, Iss: 3, pp 731-741
TL;DR: In this paper, a microcavities with cylindrical micropin fin arrays simulating a single layer of a water-cooled electronic chip stack is investigated experimentally.
Abstract: Hydrodynamics in microcavities with cylindrical micropin fin arrays simulating a single layer of a water-cooled electronic chip stack is investigated experimentally. Both inline and staggered pin arrangements are investigated using pressure drop and microparticle image velocimetry (μPIV) measurements. The pressure drop across the cavity shows a flow transition at pin diameter–based Reynolds numbers (Re d ) ~200. Instantaneous μPIV, performed using a pH-controlled high seeding density of tracer microspheres, helps visualize vortex structure unreported till date in microscale geometries. The post-transition flow field shows vortex shedding and flow impingement onto the pins explaining the pressure drop increase. The flow fluctuations start at the chip outlet and shift upstream with increasing Re d . No fluctuations are observed for a cavity with pin height-to-diameter ratio h/d = 1 up to Re d ~330; however, its pressure drop was higher than for a cavity with h/d = 2 due to pronounced influence of cavity walls.

Summary (3 min read)

1 Introduction

  • High-performance next-generation multi-core processors require new architectures and advanced packaging.
  • Brunschwiler et al. (2009) investigated various pin fin arrangements with different pin diameter, pitch and cavity heights for 3D integrated chip cooling.
  • They found a vortex shedding activity moving upstream with increasing pitch between the cylinders at a fixed Reynolds number.
  • The velocity measurements are performed using microparticle image velocimetry (lPIV) for flow visualization at the microscale (Santiago et al.
  • Detailed velocity data were obtained at different locations in micropin fin arrays, and vortex structures were analyzed in order to explain the pressure drop trend along the flow cavity.

2 Experimental setup

  • The microfluidic chips were prepared using standard microfabrication techniques.
  • The pressure drop along the fluid path was probed using four pressure ports located at 0, 1, 5, and 10 mm along the flow direction (see Fig. 1a).
  • The authors found that adjusting the water pH to *9 using sodium hydroxide minimized the accumulation of anionic microspheres on micropins and cavity walls.
  • The time difference between the images of any pair was adjusted between 2 and 10 ls to yield a maximum particle displacement of approximately 1/4th of the initial interrogation window size of 256 9 256 pixels.
  • The intensity patterns from the fluorescent particles associated with the laser pulse was recorded on individual frames.

3 Results and discussion

  • The hydrodynamics of the flow through the chips was investigated through pressure drop and velocimetry measurements as described next.
  • The pressure drop measurements across the entire chip for two different micropin fin arrangements are shown in Fig. 3 as a function of the Reynolds number, Red defined in terms of pin diameter d as Red ¼ vmd=m; ð2Þ where m is the kinematic viscosity of water and vm the mean velocity of the fluid between the pins.
  • The pressure drops near the inlet and outlet are measured across five transversal pin rows, whereas nearly 20 transversal pin rows each are covered for two measurements in the middle.
  • The pressure drop measurement conducted at the outlet over five transversal pin rows revealed more details about the gradual upstream movement of the flow transition, which is marked by vortex shedding (see the following section).
  • Note that the onset of the transition at the outlet (Red & 180) is hardly visible in the overall pressure drop measurement across the entire chip (c.f. Fig. 3).

3.2.1 Steady flow

  • In order to understand the sharp transition in pressure drop with flow rate, the lPIV measurements were performed on the pin fin arrays before and after the transition.
  • The steady nature of the flow before the transition can be further appreciated by looking at the streamwise velocity profiles, for different locations along the flow direction, plotted in Fig. 6a.
  • The profiles show the streamwise velocity of the channellike flow at Red = 143 averaged over a distance of four pins in flow direction at different positions on the chip.
  • The velocity profiles evaluated by ensemble averaged crosscorrelation and double-shot cross-correlation are compared in Fig. 6b.
  • Therefore, the velocimetry analysis itself is expected to be more accurate than the error bars in Fig. 6b indicate.

3.2.2 Transition flow

  • After the transition, the flow pattern becomes very complex and the velocity field is strongly disturbed by transversal fluctuations as shown in Fig. 7 (for more details, see supplementary movie).
  • This means that the authors obtain a good picture of the instantaneous flow field; however, the information about the temporal evolution of the flow field is limited by a low acquisition frequency of image pairs (4 Hz in their experiments).
  • Given this limitation, the characteristic time scale of the fluctuations could not be measured.
  • The sample instantaneous images shown in Fig. 7b suggest that with time the bimodal flow structure, consisting of symmetric vortices and a microchannel-like flow in between the pins, gives way to a single vortex followed by a complete collapse of recirculation.
  • Due to the vortex structure described above, the streamwise velocity profile between pins also becomes significantly altered compared to the profiles for steady flow and becomes time dependent.

3.2.3 Flow impingement

  • The transient vortex structure and the associated higher fluctuation in flow result in increased viscous dissipation.
  • A second reason for enhanced post-transition pressure drop is flow impingement on to the pin walls as shown in Fig. 10.
  • Such flow oscillations start gradually and increase with increasing Red; Fig.
  • 7 a Streamlines showing the instantaneous vortex structure developed at the center of the PFI chip at Red = 250.
  • B Selected samples of the different observed vortex structures between the pins.
  • Beyond the critical Red, however, the flow oscillations become strong enough to initiate flow impingement on the pins.

3.2.4 Confinement effects

  • For integrated electronic cooling, a small cavity is desirable to facilitate higher packaging density and shorter TSV length.
  • Recently, Patil and Tiwari (2010) numerically investigated the flow past a square cylinder for different aspect ratios 0.5 B h/d B 3.
  • The oscillation induces an inflection point in the profile since the maximum velocity is shifted closer to the pin in (i) and (iii) compared to the intermediate microchannel-like flow in (ii) Fig. 10 a Flow impingement on pins contributing to the pressure drop transition, b zoomed in view of flow impingement on the pins.
  • The presence or absence of vortex shedding can be determined from the RMS value of the velocity fluctuations.
  • It is worth noting that for a cavity height h = 200 lm, the contribution of the additional wall friction arising from vortex shedding is still smaller compared to the viscous losses due to increased confinement for a cavity height h = 100 lm.

4 Conclusion

  • The geometry simulates a typical single cavity out of vertically stacked, TSV-connected layers of 3D chips with integrated cooling structure to address the needs for next-generation integration challenge.
  • Before the transition, the flow field was stationary with a characteristic channel-like flow between the pin rows in the flow direction and a stable vortex pair in between pins.
  • For this steady flow, the double-shot (image pair) and ensemble averaged crosscorrelations produced acceptably accurate velocity profiles (error \4.5%).
  • Moreover, for the chip with a 1:1 height-to-pin diameter ratio, no vortex shedding was observed for Red up to 330 indicating a strong confinement effect of cavity walls.
  • This work was partially supported by the Swiss Confederation through the SNSF evaluated RTD project nr. 618_67— CMOSAIC—funded by Nano-Tera.ch.

Did you find this useful? Give us your feedback

Figures (7)

Content maybe subject to copyright    Report

ETH Library
Experimental investigation into
vortex structure and pressure
drop across microcavities in 3D
integrated electronics
Journal Article
Author(s):
Renfer, Adrian; Tiwari, Manish K.; Brunschwiler, Thomas; Michel, Bruno; Poulikakos, Dimos
Publication date:
2011-09
Permanent link:
https://doi.org/10.3929/ethz-b-000039358
Rights / license:
In Copyright - Non-Commercial Use Permitted
Originally published in:
Experiments in Fluids 51(3), https://doi.org/10.1007/s00348-011-1091-5
This page was generated automatically upon download from the ETH Zurich Research Collection.
For more information, please consult the Terms of use.

RESEARCH ARTICLE
Experimental investigation into vortex structure and pressure
drop across microcavities in 3D integrated electronics
Adrian Renfer
Manish K. Tiwari
Thomas Brunschwiler
Bruno Michel
Dimos Poulikakos
Received: 24 November 2010 / Revised: 12 February 2011 / Accepted: 28 March 2011 / Published online: 12 April 2011
Ó Springer-Verlag 2011
Abstract Hydrodynamics in microcavities with cylindri-
cal micropin fin arrays simulating a single layer of a water-
cooled electronic chip stack is investigated experimentally.
Both inline and staggered pin arrangements are investigated
using pressure drop and microparticle image velocimetry
(lPIV) measurements. The pressure drop across the cavity
shows a flow transition at pin diameter–based Reynolds
numbers (Re
d
) *200. Instantaneous lPIV, performed using
a pH-controlled high seeding density of tracer microspheres,
helps visualize vortex structure unreported till date in
microscale geometries. The post-transition flow field shows
vortex shedding and flow impingement onto the pins
explaining the pressure drop increase. The flow fluctuations
start at the chip outlet and shift upstream with increasing Re
d
.
No fluctuations are observed for a cavity with pin height-to-
diameter ratio h/d = 1uptoRe
d
*330; however, its pres-
sure drop was higher than for a cavity with h/d = 2 due to
pronounced influence of cavity walls.
1 Introduction
High-performance next-generation multi-core processors
require new architectures and advanced packaging. Verti-
cal integration improves the bandwidth from core to cache
memory, by reducing wiring length. Furthermore, it helps
in obtaining a larger processor area, which effectively
becomes the number of dies in the stack times the maxi-
mum projected lithography size for two-dimensional chips
(Ulrich and Brown 2006). Therefore, the semiconductor
industry is investing heavily into the development of
through-silicon vias (TSV), which are the most important
building blocks toward three-dimensional (3D) integration.
However, in multi-layered, 3D packages, both heat flux and
thermal resistances accumulate, resulting in junction tem-
peratures above the reliability threshold. This constrains
the electrical design and is the reason why thermal man-
agement is one of the key challenges for future micropro-
cessor development. Interlayer water cooling is far superior
to traditional air cooling techniques since it can remove the
dissipated heat directly between the individual chip layers,
promising performance increases for several device gen-
erations. The superior heat transfer capability of water
notwithstanding the pumping power associated with the
pressure drop across the chip is a criterion contributing to
the overall energy costs for cooling 3D electronic chip
stacks. To seal the electrical interconnects from water, the
TSVs are embedded into silicon pins and the overall
arrangement consists of a micropin fin array inside a mi-
crocavity. Therefore, the hydrodynamic investigation into
Electronic supplementary material The online version of this
article (doi:10.1007/s00348-011-1091-5) contains supplementary
material, which is available to authorized users.
A. Renfer M. K. Tiwari D. Poulikakos (&)
Department of Mechanical and Process Engineering,
Laboratory of Thermodynamics in Emerging Technologies,
ETH Zurich, ML J 36, 8092 Zurich, Switzerland
e-mail: dimos.poulikakos@ethz.ch;
dimos.poulikakos@sl.ethz.ch
A. Renfer
e-mail: arenfer@ethz.ch
M. K. Tiwari
e-mail: mtiwari@ethz.ch
T. Brunschwiler B. Michel
Advanced Thermal Packaging, IBM Research Laboratory,
8803 Rueschlikon, Switzerland
e-mail: tbr@zurich.ibm.com
B. Michel
e-mail: bmi@zurich.ibm.com
123
Exp Fluids (2011) 51:731–741
DOI 10.1007/s00348-011-1091-5

microfluidic chips with micropin fin arrays inside micro-
channels is of paramount importance in developing the
next-generation integrated liquid cooling of 3D chip stacks.
Classically, for macroscale flows across a single cylin-
der, a critical Reynolds number marks the inception of
vortex shedding. For an array of cylinders on the other
hand, the flow field characteristics are more complicated,
mainly due to interacting wakes (Ziada and Oengo
¨
ren
1993). In low aspect ratio (h/d) pin fin array cavities, the
effect of confinement is expected to play an additional role,
resulting in different flow regimes than in flows across
cylinder arrays and cannot be described solely through a
Reynolds number based on the cylinder diameter.
Flow across large-scale tube bundles with various con-
figurations and shapes has been extensively investigated
experimentally by several researchers using hot-wire or
laser Doppler anemometry, PIV as well as pressure mea-
surements. An overview of selected studies can be found in
Paul et al. (2007). Iwaki et al. (2004) conducted PIV
experiments using inline and staggered tube bundles
(d = 15 mm) at Reynolds numbers of 5,400 and higher.
They could identify three vortex structures behind the
tubes; however, the vector fields were obtained by
ensemble cross-correlation over 200 images. These studies
focus on macroscale tubes, but the detailed hydrodynamics
in micropin fin arrays confined in microchannels remains
unexplored. Flows with low aspect ratio pin fins are likely
to be different due to pronounced effects of cavity wall
damping (Brunschwiler et al. 2009). Kos¸ar et al. (2005)
performed an experimental study on 100-lm-long micropin
fin bundles with various configurations and reported on an
increasing pressure slope at higher flow rate because of
cylinder–wake interaction. These authors neither provided
a detailed explanation for the pressure transition nor any
flow visualization to support this observation. Prasher et al.
(2007) observed a pressure gradient transition in the
hydraulic performance of low aspect ratio micropin fins but
did not provide a physical explanation. Brunschwiler et al.
(2009) investigated various pin fin arrangements with dif-
ferent pin diameter, pitch and cavity heights for 3D inte-
grated chip cooling. For inline and staggered distorted
designs with a pitch and cavity height of 200 lm, they
observed a clear flow regime transition through pressure
measurements. No flow measurements were performed to
investigate this transition. Alfieri et al. (2010) developed a
hydrodynamics and conjugate heat transfer model of a
single row of inline micropins in order to obtain correla-
tions for modeling a 3D chip stacks simulator as a
non-equilibrium porous media. No flow transition was
considered. Liang et al. (2009) presented a numerical
approach to simulate laminar flow across a row of six
cylinders for different longitudinal pitches. They found a
vortex shedding activity moving upstream with increasing
pitch between the cylinders at a fixed Reynolds number. In
the study of Nishimura et al. (1993), an upstream devel-
opment of the transition to vortex shedding with increasing
Reynolds number was observed in an array of inline
and staggered large diameter tubes (d = 15 mm, Re
d
=
60–3,000). No experimental visualization and measure-
ment of the transition flow in micropin fin arrays confined
in cavities has been reported to this date.
Herein, we report velocity and pressure drop measure-
ments in micropin fin chips in inline and staggered dis-
torted arrangement for flows with pin diameter–based
Reynolds number (Re
d
) up to 290. The velocity measure-
ments are performed using microparticle image velocime-
try (lPIV) for flow visualization at the microscale
(Santiago et al. 1998; Wereley and Meinhart 2010).
Instantaneous velocity field measurements required high
microparticle seeding, which are prone to undesirable
coagulation and sticking to solid boundaries of the flow
domain (Lindken et al. 2009). In spite of these limitations,
unsteady and/or turbulent flow dynamics in simple geom-
etries such as micron size channels (Li and Olsen 2006a, b;
Angele et al. 2006; Blonski et al. 2007; Natrajan and
Christensen 2010), capillaries (Natrajan and Christensen
2007), obstacle-type valveless micropumps (Sheen et al.
2008), and inkjet print heads (Meinhart and Zhang 2000)
have been previously reported by measuring instantaneous
velocity fields. However, a complex geometry such as the
microcavity with the micropin fin array considered here
requires special care and techniques to avoid unwanted
particle sticking. A novel pH-controlled, higher seeding of
microparticles is employed to enhance signal-to-noise ratio
in lPIV and to obtain instantaneous velocity measure-
ments. The instantaneous lPIV measurements enabled us
to perform unsteady flow characterization in such a com-
plex microscale geometry as opposed to more common
ensemble averaged lPIV measurements, which are only
suited for steady state flows (Raffel et al. 2007). Detailed
velocity data were obtained at different locations in micr-
opin fin arrays, and vortex structures were analyzed in
order to explain the pressure drop trend along the flow
cavity. Furthermore, we show a position dependence of the
pressure transition along the flow direction and the absence
of vortex shedding for a cavity height equals pin diameter,
in the Reynolds number domain investigated. The geo-
metric sizes and flow rates selected to be studied are spe-
cifically suitable for integrated chip cooling. The
corresponding pressure drop in such microcavities is
essentially a necessary penalty of the chip cooling process.
Our measurements show that the vortex shedding leads to a
sharp rise in the pressure drop across the chip. However,
vortices will also lead to enhanced mixing and heat trans-
fer, thereby requiring a trade-off in the design of such
chips. Through our work, we have identified these
732 Exp Fluids (2011) 51:731–741
123

competing aspects in designing microcavities with micr-
opin fin arrays for integrated cooling of electronic chips.
2 Experimental setup
2.1 Microfluidic chip fabrication
The microfluidic chips were prepared using standard
microfabrication techniques. Figure 1a, b shows pictures
and schematics of the chips used in the study, respectively.
The cylindrical micropin fins as well as the fluid ports were
etched into silicon with deep reactive ion etching (DRIE)
followed by anodic bonding of a 500-lm-thick glass plate
to the top of the chip in order to form the cavity and still
allowing optical access for lPIV. The resulting microflu-
idic chip cavity was square shaped in top-view
(10 9 10 mm
2
) with a height of 200 ± 10 lm. The pres-
sure drop along the fluid path was probed using four
pressure ports located at 0, 1, 5, and 10 mm along the flow
direction (see Fig. 1a). The pressure ports at 0 and 10 mm
are located within the cavity, 1 mm from the inlet/outlet as
shown in the scanning electron microscope (SEM) image
in the inset of Fig. 1a. The chip cavity was connected to a
water loop using two rectangular 1-mm-wide inlet and
outlet slots.
Two different cavity designs with 50 9 50 pins in the
flow area of the chip were used as shown in Fig. 1b. The
first is a pin fin inline (PFI) and the second a staggered
distorted cell (PFS-dc) arrangement of pins; both with a pin
pitch p of 200 ± 10 lm and a pin diameter d of
100 ± 5 lm(p/d = 2). Note that for the staggered dis-
torted cell, every second column of pins is shifted by half a
pitch in the flow direction.
2.2 Measurement method and experimental conditions
The flow was driven with an impeller pump (Conrad
Electronic SE, Germany) to minimize pulsation, which
could be the source of transient effects, at a maximum flow
rate of 190 ml/min. The variation of the flow rate above
50 ml/min was less than 2%. The flow was measured with
a laminar pressure gradient flow sensor (range: 0–500 ml/
min; Omega, USA) with a full scale accuracy of 2%. A
differential pressure gauge (range: 0–2.0 bar, accuracy:
0.2%; Omega, USA) was used to measure the pressure drop
between the ports. All the pressure drop measurements
were conducted at room temperature with de-ionized (DI)
water as the working fluid. However, the pH of the water
was adjusted using sodium hydroxide for lPIV experi-
ments for the reasons described below. The pressure
measurements were recorded with a delay of a few minutes
after setting the flow rate, to ensure stabilization of flow
through the flow loop. Figure 2 shows the flow loop and
connected lPIV system used for flow visualization. The
lPIV setup consisted of an epi-fluorescent microscope
system (FlowMaster Mitas, LaVision, Goettingen, Ger-
many) as shown in the figure. The test chips were placed on
Fig. 1 a Photograph and SEM
image of the microfluidic chip
with the pressure ports indicted
(top) and a SEM image of the
micropin fins in the flow cavity
(bottom). b Design of the inline
(PFI) and staggered distorted
cell (PFS-dc) chips used for the
experimental study
Exp Fluids (2011) 51:731–741 733
123

the 3D stage of the epi-fluorescent microscope in order to
change the location of the observed (focused) area. The
flow was seeded with fluorescent buoyant particles (1 lm
in diameter, Nile red FluoSpheres; Invitrogen, Carlsbad,
CA). The tracer particles were excited with a double-pulsed
532 nm Nd:YAG laser, and the emitted light (575 nm) was
recorded with a 2,048 9 2,048 pixel CCD camera. The
energy per pulse was adjusted to approximately 10 mJ
since the fluorescent signal saturated at higher pulse ener-
gies. Achieving instantaneous velocity fields from double-
shot images requires a high seed particle density so that
sufficient information is available for two-frame (double-
shot) cross-correlation. In our measurements, in the
absence of any further medication to DI water, particles
stuck heavily onto the walls of the pins and the chip cavity.
This is a common challenge faced in lPIV experiments
(Santiago et al. 1998) and result from Derjaguin and
Landau, Verwey and Overbeek (DLVO)-type interaction,
which accounts for both electric double-layer interaction
and van der Waals attraction, between the particles flow
boundaries (in our case pin surface and cavity walls) (Perry
and Kandlikar 2008). The microspheres were charge sta-
bilized by grafting proprietary polymers with pendant
carboxylic acid groups. It is known that a convenient
means to alter DLVO interaction is to alter the pH of the
solution, which alters the double-layer interaction, thus
stabilizing the charge-stabilized particles. We found that
adjusting the water pH to *9 using sodium hydroxide
minimized the accumulation of anionic microspheres on
micropins and cavity walls. The increased content of
hydroxide ions (due to high pH) deprotonate the carboxylic
groups on the microsphere surface such that the resulting
negative charge on the particle prevents agglomerations.
This step was crucial to obtaining instantaneous velocity
measurements by cross-correlating the image pairs at any
instant.
The timing of the laser illumination and frame acquisi-
tion were controlled using the LaVision software. The PIV
image pairs (i.e., the double-shot images) were acquired
with a frequency of 4 Hz. The time difference between the
images of any pair was adjusted between 2 and 10 lsto
yield a maximum particle displacement of approximately
1/4th of the initial interrogation window size of 256 9 256
pixels. This corresponds to a particle displacement of
18.8 lm. The intensity patterns from the fluorescent par-
ticles associated with the laser pulse was recorded on
individual frames. To compute the velocity vectors, an
adaptive multi-pass cross-correlation technique was used.
Starting from an interrogation window of 256 9 256 pix-
els, the vector fields computed at every iterative pass were
used to as to adjust the window shift for the following
passes. The final interrogation windows of size 64 9 64
(18.8 9 18.8 lm) or 32 9 32 pixels (9.4 9 9.4 lm) with a
50% overlap, in both cases, were used. The resulting vector
spacing was 9.4 lm (for 64 pixel windows) or 4.7 lm (for
32 pixel windows).The size of interrogation window is
specified in the caption of all figures in Results and dis-
cussion section showing velocimetry measurements. For
steady flows, ensemble averaged cross-correlation was
performed with 68–100 image pairs depending on the flow
situation and compared with instantaneous velocity fields
from double-shot images. To reduce image random noise,
the particle images were pre-processed with a 3 9 3 low-
pass filter, and a spatial sliding minimum subtraction was
used to remove background noise.
In this study, a 109 microscope objective with a
numerical aperture of 0.3 was used for imaging providing a
depth of correlation 2z
corr
= 27.6 lm (Olsen and Adrian
2000). The number of particles within an interrogation
window of 64 9 64 pixels was chosen to be n = 10.
Therefore, the volumetric particle concentration of the
seeding particle was computed as
c
vol
¼ 100cV
particle
; ð1Þ
where V
particle
denotes the particle volume and c the
number of particles in an interrogation depth, which was
computed as c ¼ n=ð2z
corr
A
int
Þ with A
int
designating the
interrogation area. Substituting the numerical values, we
obtain c
vol
= 0.054%.
Although the particle seeding was chosen to achieve a
final interrogation window size of 64 9 64 pixels, some
flow conditions allowed an additional refinement with a
Fig. 2 Schematic of the lPIV system with the microfluidic chip
connected to a water loop
734 Exp Fluids (2011) 51:731–741
123

Citations
More filters
Journal ArticleDOI
01 Feb 2021-Energy
TL;DR: In this article, the advantages and shortcomings of thermal enhancement technologies in different structural micro heat sinks are presented, and the barriers and challenges for the developments of thermal management of electronic devices by micro heat sink are discussed, and future directions of the research topic are provided.
Abstract: The electronic equipment developing towards miniaturization and high integration is facing the danger of high heat flux and non-uniform temperature distribution which leads to the reduction of life and reliability of electronic devices. The micro heat sinks have gained significant attention in heat dissipation of electronic devices with a high heat flux due to its large heat transfer surface to volume ratio, compact structure and outstanding thermal performance. In this review, we present the advantages and shortcomings of thermal enhancement technologies in different structural micro heat sinks. Moreover, the non-uniform temperature distribution which includes the temperature rising along the flow direction and hotspots, especially, the random hotspot with high heat flux, has been the serious issues in the thermal management of electronic devices. They are the main challenges for the efficient operation and service life of electronic components. Thus, it is urgent to develop an effective and economic process in automatic adaptive cooling of random hotspots. The purpose of this article is to introduce the existing thermal enhancement technologies in micro heat sinks and the reduction of non-uniform temperature distribution. Finally, the barriers and challenges for the developments of thermal management of electronic devices by micro heat sinks are discussed, and the future directions of the research topic are provided.

217 citations

Journal ArticleDOI
01 Jul 2012-Energy
TL;DR: In this paper, the authors report the energy and exergy efficiencies of Aquasar, the first hot water cooled supercomputer prototype, and establish hot water as a better coolant compared to air.
Abstract: We report the energy and exergy efficiencies of Aquasar, the first hot water cooled supercomputer prototype. The prototype also has an air cooled part to help compare the coolants's performances. For example, a chip/coolant temperature differential of only 15 °C was sufficient for chip cooling using water. The air cooled side, however, required air pre-cooling down to 23 °C and a chip/coolant temperature differential of 35 °C. Whereas extra exergy was expended for air pre-cooling, the higher thermal conductivity and specific heat capacity of water enabled coolant temperatures to be safely raised to 60 °C. Using such hot water not only eliminated the need for chillers, it also opened up the possibility of heat reuse. The latter was realized by using the hot water from Aquasar for building heating. A heat recovery efficiency of 80% and an exergetic efficiency of 34% were achieved with a water temperature of 60 °C. All these results establish hot water as a better coolant compared to air. A novel concept of economic value of heat was introduced to evaluate different reuse strategies such as space heating and refrigeration using adsorption chillers. It was shown that space heating offers the highest economic value for the heat recovered from data centers.

190 citations

Journal ArticleDOI
TL;DR: In this article, the effect of volume concentration (0.05, 0.1 and 0.15%) and temperature on viscosity and surface tension of graphene-water nanofluid has been experimentally measured.
Abstract: In the present study, the effect of volume concentration (0.05, 0.1 and 0.15 %) and temperature (10–90 °C) on viscosity and surface tension of graphene–water nanofluid has been experimentally measured. The sodium dodecyl benzene sulfonate is used as the surfactant for stable suspension of graphene. The results showed that the viscosity of graphene–water nanofluid increases with an increase in the volume concentration of nanoparticles and decreases with an increase in temperature. An average enhancement of 47.12 % in viscosity has been noted for 0.15 % volume concentration of graphene at 50 °C. The enhancement of the viscosity of the nanofluid at higher volume concentration is due to the higher shear rate. In contrast, the surface tension of the graphene–water nanofluid decreases with an increase in both volume concentration and temperature. A decrement of 18.7 % in surface tension has been noted for the same volume concentration and temperature. The surface tension reduction in nanofluid at higher volume concentrations is due to the adsorption of nanoparticles at the liquid–gas interface because of hydrophobic nature of graphene; and at higher temperatures, is due to the weakening of molecular attractions between fluid molecules and nanoparticles. The viscosity and surface tension showed stronger dependency on volume concentration than temperature. Based on the calculated effectiveness of graphene–water nanofluids, it is suggested that the graphene–water nanofluid is preferable as the better coolant for the real-time heat transfer applications.

146 citations

Journal ArticleDOI
TL;DR: In this paper, an experimental study on exergetically efficient electronics cooling using hot water as coolant was conducted and it was shown that water temperatures as high as 60°C are sufficient to cool microprocessors with over 90% first law (energy based) efficiency.
Abstract: We report an experimental study on exergetically efficient electronics cooling using hot water as coolant. It is shown that water temperatures as high as 60 °C are sufficient to cool microprocessors with over 90% first law (energy based) efficiency. The chip used in our experiment is kept at temperatures of 80 °C or below so as not to exceed any allowable industrial specifications for maximum microprocessor chip temperature. The use of hot water as coolant will eliminate the requirement for chillers typically used in air-cooled data centers and, therefore, significantly reduce the power consumption. An exergy analysis shows that a six fold rise in second law (exergy based) efficiency is achieved by switching the water inlet temperature from 30 °C to 60 °C. The resulting high exergy at the heat sink outlet is a measure of the potential usefulness of the waste heat of data centers, thereby helping to design data centers with minimal carbon footprint. A new metric for the economic value of the recovered heat, based on costs for electricity and fossil fuels, heat recovery efficiency and an application specific utility function, is introduced to underscore the benefits of hot water cooling. This new concept shows that the economic value of the heat recovered from data centers can be much higher than its thermodynamic value.

71 citations

Journal ArticleDOI
TL;DR: The present work paves the way toward improved performance in membraneless microfluidic flow cells for electrochemical energy conversion with the help of optimized herringbone flow promoting microstructures with an integrated separation zone.
Abstract: Enhancing mixing is of uttermost importance in many laminar microfluidic devices, aiming at overcoming the severe performance limitation of species transport by diffusion alone. Here we focus on the significant category of microscale co-laminar flows encountered in membraneless redox flow cells for power delivery. The grand challenge is to achieve simultaneously convective mixing within each individual reactant, to thin the reaction depletion boundary layers, while maintaining separation of the co-flowing reactants, despite the absence of a membrane. The concept presented here achieves this goal with the help of optimized herringbone flow promoting microstructures with an integrated separation zone. Our electrochemical experiments using a model redox couple show that symmetric flow promoter designs exhibit laminar to turbulent flow behavior, the latter at elevated flow rates. This change in flow regime is accompanied by a significant change in scaling of the Sherwood number with respect to the Reynolds number from Sh ~ Re0.29 to Sh ~ Re0.58. The stabilized continuous laminar flow zone along the centerline of the channel allows operation in a co-laminar flow regime up to Re ~325 as we demonstrate by micro laser-induced fluorescence (μLIF) measurements. Micro particle image velocimetry (μPIV) proves the maintenance of a stratified flow along the centerline, mitigating reactant cross-over effectively. The present work paves the way toward improved performance in membraneless microfluidic flow cells for electrochemical energy conversion.

62 citations

References
More filters
Book
11 Jun 2002
TL;DR: In this paper, the authors present a practical guide for the planning, performance and understanding of experiments employing the PIV technique, which is primarily intended for engineers, scientists and students, who already have some basic knowledge of fluid mechanics and nonintrusive optical measurement techniques.
Abstract: This practical guide intends to provide comprehensive information on the PIV technique that in the past decade has gained significant popularity throughout engineering and scientific fields involving fluid mechanics. Relevant theoretical background information directly support the practical aspects associated with the planning, performance and understanding of experiments employing the PIV technique. The second edition includes extensive revisions taking into account significant progress on the technique as well as the continuously broadening range of possible applications which are illustrated by a multitude of examples. Among the new topics covered are high-speed imaging, three-component methods, advanced evaluation and post-processing techniques as well as microscopic PIV, the latter made possible by extending the group of authors by an internationally recognized expert. This book is primarily intended for engineers, scientists and students, who already have some basic knowledge of fluid mechanics and non-intrusive optical measurement techniques. It shall guide researchers and engineers to design and perform their experiment successfully without requiring them to first become specialists in the field. Nonetheless many of the basic properties of PIV are provided as they must be well understood before a correct interpretation of the results is possible.

4,811 citations

Journal ArticleDOI
TL;DR: In this article, a micro-resolution particle image velocimetry (micro-PIV) system was developed to measure instantaneous and ensemble-averaged flow fields in micron-scale fluidic devices.
Abstract: A micron-resolution particle image velocimetry (micro-PIV) system has been developed to measure instantaneous and ensemble-averaged flow fields in micron-scale fluidic devices. The system utilizes an epifluorescent microscope, 100–300 nm diameter seed particles, and an intensified CCD camera to record high-resolution particle-image fields. Velocity vector fields can be measured with spatial resolutions down to 6.9×6.9×1.5 μm. The vector fields are analyzed using a double-frame cross-correlation algorithm. In this technique, the spatial resolution and the accuracy of the velocity measurements is limited by the diffraction limit of the recording optics, noise in the particle image field, and the interaction of the fluid with the finite-sized seed particles. The stochastic influence of Brownian motion plays a significant role in the accuracy of instantaneous velocity measurements. The micro-PIV technique is applied to measure velocities in a Hele–Shaw flow around a 30 μm (major diameter) elliptical cylinder, with a bulk velocity of approximately 50 μm s-1.

1,187 citations


"Experimental investigation into vor..." refers background or methods in this paper

  • ...The velocity measurements are performed using microparticle image velocimetry (lPIV) for flow visualization at the microscale (Santiago et al. 1998; Wereley and Meinhart 2010)....

    [...]

  • ...This is a common challenge faced in lPIV experiments (Santiago et al. 1998) and result from Derjaguin and Landau, Verwey and Overbeek (DLVO)-type interaction, which accounts for both electric double-layer interaction and van der Waals attraction, between the particles flow boundaries (in our case pin surface and cavity walls) (Perry...

    [...]

  • ...This is a common challenge faced in lPIV experiments (Santiago et al. 1998) and result from Derjaguin and Landau, Verwey and Overbeek (DLVO)-type interaction, which accounts for both electric double-layer interaction and van der Waals attraction, between the particles flow boundaries (in our case…...

    [...]

Journal ArticleDOI
TL;DR: In this article, the authors extended the theory of microscopic particle image velocimetry (PIV) to encompass this situation and derived an equation for a particle image intensity function that yields image diameter and intensity as a function of distance from the object plane.
Abstract: In microscopic particle image velocimetry (μPIV) the entire volume of a flowfield is illuminated, resulting in all of the particles in the field of view contributing to the image, either by forming discrete particle images or contributing to a background glow. The theory of PIV is expanded to encompass this situation. Equations are derived for a particle image intensity function that yields image diameter and intensity as a function of distance from the object plane, as well as an equation for a new quantity, termed particle visibility. The effect of changing experimental parameters is discussed. Next, the contribution of out-of-focus particles to the correlation function is addressed. A weighting function that can be used to calculate either velocity measurement bias or the distance from the object plane beyond which particles no longer significantly contribute to the correlation function is derived. A new experimental parameter, the depth of correlation, is then introduced, and its dependence on experimental parameters is discussed.

429 citations


"Experimental investigation into vor..." refers background or methods in this paper

  • ...In this study, a 109 microscope objective with a numerical aperture of 0.3 was used for imaging providing a depth of correlation 2zcorr = 27.6 lm (Olsen and Adrian 2000)....

    [...]

  • ...An overview of selected studies can be found in Paul et al. (2007). Iwaki et al....

    [...]

  • ...6 lm (Olsen and Adrian 2000)....

    [...]

Journal ArticleDOI
TL;DR: This review discusses the state of the art of the optical whole-field velocity measurement technique micro-scale Particle Image Velocimetry (microPIV), a useful tool for fundamental research of microfluidic applications in life science, lab-on-a-chip, biomedical research, micro chemical engineering, analytical chemistry and other related fields of research.
Abstract: In this review we discuss the state of the art of the optical whole-field velocity measurement technique micro-scale Particle Image Velocimetry (µPIV). µPIV is a useful tool for fundamental research of microfluidics as well as for the detailed characterization and optimization of microfluidic applications in life science, lab-on-a-chip, biomedical research, micro chemical engineering, analytical chemistry and other related fields of research. An in depth description of the µPIV method is presented and compared to other flow visualization and measurement methods. An overview of the most relevant applications is given on the topics of near-wall flow, electrokinetic flow, biological flow, mixing, two-phase flow, turbulence transition and complex fluid dynamic problems. Current trends and applications are critically reviewed. Guidelines for the implementation and application are also discussed.

354 citations


"Experimental investigation into vor..." refers background in this paper

  • ...Instantaneous velocity field measurements required high microparticle seeding, which are prone to undesirable coagulation and sticking to solid boundaries of the flow domain (Lindken et al. 2009)....

    [...]

Journal ArticleDOI
TL;DR: The fundamentals of the technique, its theoretical background, and several applications are discussed, including micro-particle image velocimetry, which can be used to characterize the performance of microfluidic systems with spatial resolutions better than one micron.
Abstract: Microfluidic devices are becoming increasingly common and are seen in applications ranging from biology to nanotechnology and manufacturing. Flow behavior in these small domains can often be counterintuitive because of the low Reynolds number or the relative importance of surface forces. Micro-particle image velocimetry (μPIV) is a quantitative method that can be used to characterize the performance of such microfluidic systems with spatial resolutions better than one micron. Illustrating the impact of this measurement technique, more than 100 journal articles are published per year that feature μPIV velocity measurements. This article discusses the fundamentals of the technique, its theoretical background, and several applications.

328 citations


"Experimental investigation into vor..." refers methods in this paper

  • ...The velocity measurements are performed using microparticle image velocimetry (lPIV) for flow visualization at the microscale (Santiago et al. 1998; Wereley and Meinhart 2010)....

    [...]

Frequently Asked Questions (21)
Q1. What are the contributions mentioned in the paper "Experimental investigation into vortex structure and pressure drop across microcavities in 3d integrated electronics" ?

Poulikakos et al. this paper investigated the hydrodynamics in microcavities with cylindrical micropin fin arrays simulating a single layer of a watercooled electronic chip stack. 

The superior heat transfer capability of water notwithstanding the pumping power associated with the pressure drop across the chip is a criterion contributing to the overall energy costs for cooling 3D electronic chip stacks. 

a possible mass flow rate contribution due to three-dimensional rotation about the main axial flow direction is not captured in the measurements. 

A novel pH-controlled, higher seeding of microparticles is employed to enhance signal-to-noise ratio in lPIV and to obtain instantaneous velocity measurements. 

Since the pressure drop transition was position dependent, the velocimetry was performed close to the inlet and outlet and at the center of the chip cavity in the streamwise direction at h/2. 

Due to the constraints imposed by the inlinegeometry, the vector angle is nearly zero at the pin locations (pin center at y/d = 0, 2, 4, 6) and the wavelength of the vector deviation angle is almost exactly equal to twice the pitch, i.e. k & 2p. 

Starting from an interrogation window of 256 9 256 pixels, the vector fields computed at every iterative pass were used to as to adjust the window shift for the following passes. 

for the chip with a 1:1 height-to-pin diameter ratio, no vortex shedding was observed for Red up to 330 indicating a strong confinement effect of cavity walls. 

A simple means to recognize this fluctuating flow pattern is to plot the variation in velocity vector angle from the streamwise direction at the point of strong shedding. 

The flow fluctuations were location dependent and occurred first (for smaller flow rates) near the outlet of the chip and moved upstream with increasing flow rate. 

The microchannel-like steady flow in between the pins before the transition appears to start oscillating post-transition and impinges alternatively on to the pins located on either side of it. 

the averaged velocity fields (computed from 100 instantaneous velocity fields across four pins) mass flow rate only deviated by 2% from the instantaneous microchannel-like mass flow. 

The authors found that adjusting the water pH to *9 using sodium hydroxide minimized the accumulation of anionic microspheres onmicropins and cavity walls. 

Koşar et al. (2005) performed an experimental study on 100-lm-long micropin fin bundles with various configurations and reported on an increasing pressure slope at higher flow rate because of cylinder–wake interaction. 

The strong slope increase after Re = 200 (Fig. 3) is problematic for microfluidic electronic cooling applications since it may result in an excessive overall pressure drop. 

As the boundary layer thickness is inversely related to the aspect ratio (Özdemir et al. 2009), for low h/d the boundary layer size can be of the order of the cavity height. 

For both cavities, the streamwise velocity profiles were averaged over 80 instantaneous velocity profiles between the pins (Fig. 11, dotted line in inset). 

the pressure drop associated with the pumping power to cool 3D electronic chipstacks is smaller for a 200-lm cavity for a wide range of flow rates, even though vortex shedding is present. 

The pressure drop measurements across the entire chip for two different micropin fin arrangements are shown in Fig. 3 as a function of the Reynolds number, Red defined in terms of pin diameter d asRed ¼ vmd=m; ð2Þwhere m is the kinematic viscosity of water and vm the mean velocity of the fluid between the pins. 

In low aspect ratio (h/d) pin fin array cavities, the effect of confinement is expected to play an additional role, resulting in different flow regimes than in flows across cylinder arrays and cannot be described solely through a Reynolds number based on the cylinder diameter. 

The deviation of the double-shot cross-correlations from ensemble averaged profile (indicated by error bars in Fig. 6b) is lower than 4.5%.