You are on page 1of 78

COMPARATIVE ANALYSIS OF MICROCHANNEL HEAT SINK CONFIGURATIONS SUBJECT TO A PRESSURE CONSTRAINT

BY DYLAN SEAN FARNAM B.S. Mechanical Engineering, Rensselaer Polytechnic Institute, 2005

THESIS Submitted in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering in the Graduate School of Binghamton University State University of New York 2007

UMI Number: 1449066

UMI Microform 1449066 Copyright 2007 by ProQuest Information and Learning Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code.

ProQuest Information and Learning Company 300 North Zeeb Road P.O. Box 1346 Ann Arbor, MI 48106-1346

Copyright by Dylan Sean Farnam 2007 All Rights Reserved

Accepted in partial fulfillment of the requirements for the degree of Master of Science in Mechanical Engineering in the Graduate School of Binghamton University State University of New York 2007

May 4th, 2007 Dr. Bahgat Sammakia ________________________________________ May 4th, 2007 Department of Mechanical Engineering, Binghamton University Dr. Harold Ackler ___________________________________________ May 4th, 2007 Department of Mechanical Engineering, Binghamton University Dr. Kanad Ghose ____________________________________________ May 4th, 2007 Department of Computer Science, Binghamton University

iii

Abstract

Leakage losses and ever-increasing power dissipation in the microprocessor are causing significant thermal, mechanical, and reliability problems. Aside from the issue of cooling chip hot spots in order to reduce stress-inducing thermal gradients, the traditional challenge of quelling overall operating temperatures remains. Conventional cooling methods are reaching their practical limits and new methods of lowering the operating temperature of microprocessors are being explored. Microfluidics-based cooling schemes are one approach being considered. Implementation of microchannels for forced convection at the chip level shows much promise, as the effective heat transfer surface area and attainable heat transfer coefficient are very favorable. A major design limitation to such an implementation is the pressure developed within such micro-flows, and the stresses that could result. In this study, multiple discrete microchannel heat sink configurations are analyzed computationally and compared in a cooling capability sense, while total pressure drop across the flows is carefully considered. A single cooling channel over an energy source is split into two smaller channels, and so on, while total pressure drop is maintained constant, and specified such that all flows remain in the laminar regime. It is shown that for the configurations analyzed, there exists a definitive height-dependent optimum cooling scheme. In addition, it is shown that a slimmer design may be implemented with very little effect on cooling capability. Furthermore, cooling capability dependence on total pressure drop of the flows is shown to be minimal for high-performing microchannel configurations of all heights.

iv

Dedicated to Dr. Gary L. Lehmann; tireless educator, proud father.

Acknowledgements

I would first like to thank Dr. Bahgat Sammakia for his enthusiastic and unwavering support and inspiration in all aspects of my studies. His advice and guidance in my education has been immeasurably appreciated. I would also like to thank Dr. Harold Ackler for the invaluable assistance and opportunities he provided me. Gratitude is also extended to Dr. Kanad Ghose for his unique perspective on the issues involved in this and further studies. Additionally, I would like to acknowledge the support and help of my colleagues and friends. Most importantly, I am grateful for the strength and guidance of my family, especially my sisters and parents. Their undying love and support is what keeps me going.

vi

Table of Contents Abstract Dedication Acknowledgements 1. Introduction 1.1. Overview 1.2. Motivation 1.3. Literature Survey 1.3.1. Thermal Challenges in Electronics Packaging 1.3.1.1. Implications 1.3.1.2. Solutions 1.3.2. Microchannel Heat Sinks 1.3.2.1. Microchannels 1.3.2.2. Merits 1.3.2.3. Apprehension 1.3.2.4. Previous Studies 1.3.3. Computational Fluid Dynamics & Heat Transfer 1.4. Current Study 1.4.1. Description 1.4.2. Differences 2. Numerical Solution 2.1. Description of Numerical Model 2.2. Numerical Validation 2.3. Grid Sensitivity Analysis 2.4. Baseline Case Results 3. Results & Conclusions 3.1. Discussion of Results 3.1.1. Quantity/Width Variation 3.1.2. Height Variation 3.2. Summary & Conclusions 4. Future Considerations Nomenclature References

Page iv v vi 1 1 1 2 2 2 6 8 8 9 10 10 20 22 22 27 29 29 31 34 36 45 45 45 52 61 63 65 66

vii

List of Tables 1.1 Classification of Temperature-Related Failures 1.2 Different cooling configurations studied 1.3 Constant material properties used in numerical model 2.1 Model resolutions employed in grid sensitivity analysis and Tavg,src 2.2 Results of case 1 for Pt of 1,2,3,4psi 3.1 Analytical determinations of ld for Pt = 1psi 3.2 Highest-performing configurations for multiple Pt

Page 5 25 27 35 43 46 52

viii

List of Figures 1.1 1.2 1.3 1.4 1.5 1.6 iNEMI 2007 Roadmap Illustration of thermomechanical deformation in solder joints Illustration of electronics packaging levels Partial view of power source and one possible heat sink configuration Unit Cell considered in this study First iteration of configuration comparison

Page 3 5 7 23 24 25 30 30 33 33 35 37 38 38 40 41 42 43 43 44 46 47 48 49 50 50 51 53 54 54 55 56 58 59 59 60 60

2.1 Symmetry conditions employed in all models 2.2 Adaptive meshing employed to enable high-resolution discretization 2.3 Numerical validation; pressure drop comparison 2.4 Numerical validation; outlet temperature comparison 2.5 Grid sensitivity analysis 2.6 Pt plotted axially at a channel wall location for case 1 at Pt = 1psi 2.7 u plotted axially at channel center for case 1 at Pt = 1psi 2.8 u half-profiles at vertical center of microchannel for case 1 at Pt = 1psi 2.9 Temperature contours of case 1 at Pt = 1psi 2.10 Temperature of case 1 at Pt = 1psi plotted axially 2.11 Surface Nu contours of microchannel walls for case 1 at Pt = 1psi 2.12 Vavg of microflow in case 1 for Pt of 1,2,3,4psi 2.13 Nuavg of microchannels of case 1 for Pt of 1,2,3,4psi 2.14 Tavg,src of case 1 for Pt of 1,2,3,4psi 3.1 Expected effects on cooling from case 1 to case 2 3.2 u plotted axially at channel center for cases 2-7 at Pt = 1psi 3.3 Temperature plotted axially at channel center for case 2-7 at Pt = 1psi 3.4 Vavg for all cases at multiple Pt 3.5 Nuavg for all cases at multiple Pt 3.6 Aeff for all cases 3.7 Tavg,src for all cases at multiple Pt 3.8 Vavg for different hch, for channels of w 87.5m (Case 4) at multiple Pt 3.9 Nuavg for different hch, for channels of w 87.5m (Case 4) at multiple Pt 3.10 Aeff for alteration of channel height, for channels of w 87.5m (Case 4) 3.11 Tavg,src for alteration of channel height of case 4 at multiple Pt 3.12 Temperature contours of case 4, altered to hch of 100m at Pt = 1psi 3.13 Tavg,src for alteration of channel height of cases 1-4 at various Pt 3.14 Surface plot of Vavg as a function of hch and Case for Pt of 4psi 3.15 Surface plot of Nuavg as a function of hch and Case for Pt of 4psi 3.16 Surface plot of Aeff as a function of hch and Case 3.17 Surface plot of Tavg,src as a function of hch and Case for Pt of 4psi

ix

Chapter 1 Introduction

1.1 Overview Thermo-mechanical issues due to increased power dissipation and other factors are endangering the function and reliability of modern-day electronic devices. In addition to the challenge of spot-cooling chip hot spots, overall operating temperature must be handled as well. The challenge will only increase, as the power density of devices is steadily increasing. Traditional methods of cooling, such as heat sink air cooling, are reaching their practical limitations, and new methods of lowering the operating temperatures of microprocessors are being explored. Microfluidics-based cooling options are one of the alternative methods being investigated. Implementation of microchannel heat sinks for convective-cooling at the device level shows much promise, as the attainable heat transfer can be very high. Tuckerman and Pease considered this application in 1981 and were able to experimentally dissipate 790 W cm 2 from a device. [1]

1.2. Motivation Many investigators have studied the cooling capability of microfluidic networks, constrained by velocity. However, a major design limitation to implementation

of these microfluidic-based cooling schemes is the pressure developed within such microflows and the stresses that could result in the surrounding architecture, compromising reliability. In this study, multiple microchannel cooling configurations are studied computationally and compared in a cooling capability sense. Quantity, width, and height of microchannels are varied accordingly, in an effort to distinguish and establish trends of effectiveness of cooling for the various configurations. Weighing the importance of the pressure developed in these schemes, applicable flows are subject to similar pressure constraints.

1.3. Literature Survey

1.3.1. Thermal Challenges in Electronics Packaging Increasing technological advances in the capability of electronic devices are resulting in more challenging thermal management issues that surpass the capacity of traditional cooling methods.

1.3.1.1. Implications Electrical devices require power to operate. The resistance to the flow of this electrical current through the various aspects of these devices generates significant internal heat within these components. The heat generated is usually detrimental to important operating aspects of these devices; most electrical components suffer from reliability and performance hindrance as operating temperature increases.

The evolution of the microprocessor is headed down an increasingly-dangerous thermal path. Traditional, 2-dimensional high performance microprocessors and other high-speed devices are operating at increased performance, while simultaneously their size is being reduced. That is to say, devices are operating at higher clock speeds with a higher density of process components, resulting in higher levels of heat generation and concentration. Furthermore, with the recent advent of 3-dimensional electronic circuit architectures, challenges for thermal management continue. 3-d architecture allows for the integration of logic with memory, and further features on a single chip. However, this integration results in further increased power density and heat concentration. Figure 1.1 shows how the International Electronics Manufacturing Initiative expects power levels of devices to grow even more in the coming years. [2] Additionally, as features of the microprocessor shrink in dimension, current leakage is becoming another critical thermal issue.
Automotive Portable & Consumer Office/Large Business/Communication Systems Defense/Aerospace Medical

350 300 250

watts(w)

200 150 100 50 0 2005 2007 2009 2011 2017

year

Figure 1.1 iNEMI 2007 Roadmap; Expected Thermal Design Power (Hottest Chip) (W) for different product sectors [2]

Left unchecked, the heat generated leads to circuit malfunction, decrease in performance, thermal runaways, and possible catastrophic device failure. As operating temperatures rise, failure mechanisms such as thermomechanical failure, metallization failure, corrosion, electromagnetic diffusion [3] , mechanical creep, parasitic chemical reactions, and dopant diffusion [4] accelerate. Thus, device reliability generally decreases as operating temperature increases. Thermomechanical failures arise from the stresses and strains generated within an electronic device due to thermal loading. As materials increase in temperature, they expand a certain amount, generally. The amount a material expands upon heating (and contracts upon cooling) is classified generally as its coefficient of thermal expansion (CTE). CTEs vary from material to material, and mismatches between different materials within a device and/or bonded materials create an issue. Geometric constraints, coupled with thermal gradients and CTE mismatch induce stresses and strains within the system. Material deformation, or worse, can occur. Figure 1.2 demonstrates thermomechanical deformation as a result of CTE mismatch. Catastrophic device failure such as fracture, delamination, melting, vaporization, or combustion of materials caused by thermal runaways may occur, usually as the result of a large uncontrolled temperature surge. Total device function may be lost, the consequences of which could be immeasurable. Sergent and Krums classification of thermally-induced failures by severity is shown in Table 1.1. [3]

Solder Joints Chip

Substrate Stress-Free

Solder Joints Chip

Substrate Stress Due to Elevated Temperature

Solder Joints Chip

Substrate Stress Due to Reduced Temperature

Figure 1.2 Illustration of thermomechanical deformation in solder joints

Table 1.1 Sergent and Krums Classification of Temperature-Related Failures [3] Failure Mode Characteristics Soft Failures Hard Failures (long term) Circuit continues to operate, but does not meet specification when the temperature is elevated beyond the maximum operating temperature. Circuit returns to normal operation when the temperature is lowered. Failure is due to changes in component parameters with temperature. Circuit does not operate. Circuit may or may not return to normal operation when temperature is lowered. Failure is likely due to component or interconnection breakdown, but may also be due to changes in component parameters with temperature. Circuit does not operate at any temperature. Failures are irreversible. Failures may be caused by corrosion, intermetallic formation, or similar phenomena. Failures may also be caused by mechanical stresses due to differences in coefficient of thermal expansion (CTE) between a component and the substrate.

Hard Failures (short term)

Thermal management has become a severe challenge to the growth in performance capability and compactness of the microprocessor. Electronic components

can perform reliably for many years, if proper thermal management is implemented. If insufficient cooling is provided, failure may occur. On the contrary, if designers overcompensate for cooling required, and overshoot acceptable operating temperatures significantly, a costly and unnecessarily large or heavy device may result.

1.3.1.2. Solutions In order to examine the traditional cooling methodologies implemented to protect microprocessors and other devices, it is necessary to first examine the packaging level distinction. Generally, packaging is divided into a number of different levels, as shown in Figure 1.3, which has been adapted from [4]. The chip/device package itself is thought of as the first level of packaging and the printed circuit board (PCB) makes up the second level of packaging. The third level of packaging is comprised of the motherboard and the fourth level of packaging often considered is the box or container that houses the entire assembly of a system; i.e. a computer case or telephone housing, etc. For the purposes of this study, first-level cooling is considered. First level cooling is traditionally thought of as aiding the conduction of heat from the device to the surface of the housing and on to the second level of packaging in order to lower device temperature. Thus, passive and active techniques are implemented at this level in an effort to lower the thermal resistance from the device to the second level. Some passive methods of lowering this thermal resistance include using a highconductivity adhesive, grease, or other material as a die-attach from the device housing to the second level. Phase change materials that wet surfaces more completely at operating temperatures may also be used as die-attach materials. Additionally, metal heat spreaders

or heat sinks may be attached to the device housing in an effort to quickly move heat from the device. Implementation of a heat sink increases the surface area for convective removal of heat. The convection may be a passive buoyancy-driven natural convection or an active technique.

(Level 1)

(Level 2)

(Level 3)

Figure 1.3 Illustration of electronics packaging levels, adapted from [4]

Active methods of heat sink convection most often include some sort of air jet impingement on the fins of the sink. Other active methods of first level cooling include direct attachment of heat pipes and dielectric liquid immersion. These methods are generally required in very high power applications, and dielectric liquid immersion is not always a viable option, considering environmental and application variables. These traditional methods of cooling are reaching their practical limitations. Some devices have measures implanted to slow down clock speeds when thermal issues become a threat to operation and/or reliability. This is a severe hindrance on attainable performance and is thus, an unattractive solution. Furthermore, following the trends of 7

electrical microdevices, the size of cooling solutions must decrease as the devices themselves do. [5] New methods of lowering the operating temperatures of microprocessors and other devices are being explored. Microchannel heat sinks are one of the alternative methods being explored.

1.3.2. Microchannel Heat Sinks Microchannel heat sinks usually involve networks of microchannels manufactured into the silicon of an electronic device, a substrate attached to a device, or sacrificial layer(s) of a device. These microchannels are used to circulate coolant, which transports heat away from the device via forced convection. Specifically, heat generated in a device moves via conduction to the solid regions encompassing the microchannels. The fluid then removes the heat from the walls via forced convection, and carries it away from the device. These networks are also sometimes referred to as micro-duct sinks, micro-heat sinks, and microchannel heat exchangers.

1.3.2.1. Microchannels Microchannels are enclosed channels with microscopic geometries that generally serve as a passage for internal fluid transport. Typically, the distinction of a microchannel from a traditional channel rests on the condition of having a microscopic hydraulic diameter (Dh). Hydraulic diameter serves as a characteristic length scale. Although there currently hasnt been any universal cutoff or distinction implemented between macro-, milli-, and microchannels, it is proven that as Dh becomes smaller and smaller, less than 1mm, certain scaling effects can affect the behavior of fluid flow (and

with it; heat transfer). Similarly, some traditionally neglected aspects of macro-fluidic analysis can no longer be justifiably neglected when dealing with microflows. Morini identifies the following scaling effects for single-phase flows in microchannels; axial heat conduction, conjugate heat transfer, temperature-dependence of material properties, channel wall roughness, and viscous dissipation. [6] The observable effects of these parameters on fluid behavior and heat transfer depend on not only channel geometry, but fluid properties, flow regime and intensity, and other variables.

1.3.2.2. Merits Implementation of microchannel heat sinks for forced convection at the chip level shows much promise, as the effective heat transfer surface area (Aeff) and attainable heat transfer coefficient are very favorable. Theory indicates that more than 1000 W cm 2 can be removed from 10mm by 10mm devices utilizing microchannel heat sinks. [1] Microchannel heat sinks have the valuable ability to be fabricated or installed millimeters or even micrometers from the heat source. Thus, the thermal resistance for heat removal via microchannel heat sinks can be incredibly small. Additionally, the large surface area to volume ratio, small heat sink mass and volume, small coolant inventory [7], high materials compatibility, and economic and process-compatible fabrication methodology [8] are all advantages that have drawn considerable interest to the technology.

1.3.2.3. Apprehension Microchannel heat sinks do face justifiable apprehension towards implementation. As mentioned previously, the pressure developed in microchannel flows can cause potentially large pressure gradients which can lead to warpage of device components or delamination of device layers. These stresses induced in the chip architecture can compromise performance, reliability, or directly and instantaneously lead to catastrophic failure. Additionally, the coolant must be carefully accounted for, both in installation, storage, and operation. Leakage could also have disastrous consequences regarding circuit reliability. The conductivity of the coolant may also induce undesired electrical disturbances, if careful design isnt performed. Technology drivers mentioned earlier, such as 2-D device speed and density trends, the advent of 3-D architectures, and current leakage from shrinking feature size continue to build the case for working effectively against these apprehensions, never-theless.

1.3.2.4. Previous Studies Over the past decade and more, multitudes of studies have focused on the analysis and behavior of fluid flow and heat transfer of various types in microchannels and microchannel heat sinks. Experimental, analytical, and numerical studies have been performed for both laminar and turbulent flows, for circular and non-circular channels, and for channels with characteristic length scales (Dh) on the order of 1mm and 1m. Velocity-bounded flow and pressure-bounded flows have been analyzed, and various assumptions suitable for standard macrofluidic theory have been made and not made.

10

Both unexplained behavior and explainable scaling effects with respect to standard macro-analysis [9] have been found. Many authors have encountered disparities between experimentally-observed microfluidic behavior and traditional macrofluidic derived predictions. Ho and Tai remarked that observation of microflows often results in the appearance of unexpected phenomena. [10] The ratio between surface and body forces and ratio between device and intrinsic length scales are two of the scaling effects that Ho and Tai observed. The authors claim differences between measured microfluidic behavior and those expected from macrofluidic correlations are the result of unknown physical mechanisms. Surface forces generally neglected in macrofluidic analyses are what Ho and Tai believe to be most responsible for this divergence. Peng and Piao performed a comprehensive review on the physical nature of microfluidic flow and heat transfer. [11] Citing the lack of an unequivocal agreement in comprehension of the phenomena occurring in microflows, they make a few points. Accurate experimental measurement techniques need to be universally developed and employed in order to enable effective research moving forward. Furthermore, the authors emphasize the importance of the careful selection of appropriate correlating parameters, in order to justify results. Peng and Piao also cite unfamiliar phenomena that occur in microflows at solid/fluid interfaces, most likely the result of the increase in the ratio of solid-fluid interface per unit volume, but say that sufficient knowledge regarding these is unavailable. Peng and Peterson [12] experimentally studied single-phase forced convection in rectangular microchannels and concluded that the flow exhibited strong differences from

11

that of the macrofluidic range. They observed that the critical Reynolds number (ReCr) was lower than the conventional value of ~2300. The authors observed transitional flow beginning at around 300, with fully turbulent flow occurring at around Re of 1000. They cite the importance of thermophysical properties and channel geometry as further influences on the development of flow regimes. Wang and Peng [13] and Harms, et al. [14] also found that fully turbulent flow was achieved at Reynolds numbers ranging from 1000-1500. Guo and Li reviewed and discussed the physical mechanisms of scaling effects of single-phase flow and heat transfer in microchannels. [15] The authors classified scaling effects of microflows into two categories. First, the gas rarefaction effect occurs when the continuum assumption breaks down as the scale is reduced. Secondly, variations in the dominance of factors change the relative importance of various phenomena regarding flow and heat transfer, as the scale of the channels reduce. Since the continuum assumption is usually indisputably valid for microchannels of characteristic length of 1m or higher, the second classification accounts for the majority of scaling effects. Among these phenomena of the second classification, the authors note that the high surface area to volume ratios of microchannels cause increased importance of terms related to the surface area. Surface friction induced compressibility and surface roughness are two such factors, and are theorized to cause flatter velocity profiles and early turbulence transition, respectively. Xu, Ooi, Wong, and Choi studied liquid flows in microchannels with two different experimental methods. [16] The first method involved microchannels machined (via micro-end-mills) into an aluminum plate, with a plexiglass cover plate and a thin

12

sealing gasket; method A. The second experimental method investigated involved microchannels etched onto silicon wafers. A Pyrex glass cover plate was then bonded to the wafer via anodic bonding; method B. Various dimensioned microchannels were fabricated for each of the two experimental techniques such that channels of similar scale were present in each of the two methods. Xu, et al.s results showed that the experimental data gathered from experimental method B matched conventional Navier-Stokes predictions well. Although method A involved channels of similar scale, the experimental results gathered from method A did not match as well with conventional macrofluidic predictions. The authors eventually concluded that unsmooth surfaces, coupled with the cover plate sealing technique used in method A were most likely responsible for leakage out of the channel. That is to say that a gap was developed between the cover plate and the supposed top of the microchannels and the flow cross-sectional area was inflated undesirably. Xu, et al. cite works by other authors who use similar experimental methods as method B and also find that their results match well with conventional macrofluidic predictions. G.L. Morini has studied the role that oft-neglected viscous heating (viscous dissipation) plays in microchannel flow. [6] Fluid flowing through a microchannel with a relatively small hydraulic diameter can experience significant heat generation due to viscous forces, the author states. A temperature rise can be experienced even if the channel is adiabatic. This temperature rise affects the thermophysical properties of the fluid, most importantly, viscosity. Morini demonstrates that some unexpected experimental results can be explained when viscous heating effects are considered; namely, the decrease in friction factor with increasing Reynolds number observed by

13

some works. The author also suggests a criterion to gauge the limit for which viscous heating effects can no longer be neglected and retains that both the continuum approach and Navier-Stokes equations hold for microchannels. Morini also conducted a thorough review of many experimental results obtained for single-phase convective heat transfer in microchannels. [17] The author notes that although many cases show that the experimental data for friction factor and Nusselt number in microchannels disagree with conventional theory, these cases show very little agreement with one-another as well. Reasons for these differences, including rarefraction and compressibility effects, viscous dissipation effects, electro-osmotic effects (EDL), property variation effects, channel surface conditions, and experimental uncertainties, have been proposed in the open literature to explain these differences. Morini shows how the validity of the dismissal of conventional theorys applicability to microfluidics is an open question at the current time. Some authors find good agreement with microfluidic experimental results and conventional theory, while others find strong disagreements for the same length scales. Perhaps the most intriguing point discussed by the author concerns the chronological decrease in disagreement of conventional theory and microfluidic experimental results. Deviations of observed microfluidic dissent from traditional macrofluidic behavior are decreasing with time. Explanation for this convergence can best be explained by the technological advances in microfabrication techniques, and the resulting decrease in behavior-influencing surface roughness and increase in control over the channel cross-sections, the author states. Furthermore, technological advances in experimental measurement have resulted in more accurate/reliable published results in recent years.

14

Herwig and Hausner identified five scaling effects that may be present in microfluidic devices; [9] axial heat conduction (small Peclet numbers), conjugate heat transfer (relatively thick walls), temperature dependent properties (large axial temperature gradients), pressure dependent properties (large axial pressure gradients), and wall roughness (specific wall roughness distribution). The authors cite recently published results in which experimental differences of microfluidic forced convection from conventional predictions were explained by secondary Brinkman effects. Herwig and Hausner proceeded to develop a numerical model for the previously published experiment in order to account for axial heat conduction and conjugate heat transfer. Taking these scaling effects into account, the results match conventional theory fairly accurately, and hypothesize that an erroneous temperature interpolation in the previous authors experimental method resulted in skewed results. The authors propose a more systematic approach to channel analyses in which studies that neglect all effects that are of minor importance in macro-dimensioned channels should be referred to as standard macro-analysis and the aforementioned effects should be referred to as scaling effects with respect to standard macro-analysis. Bavire, Favre-Marinet, and Le Person performed a recent experimental and numerical study of heat transfer measurements in microchannel flows with polished walls. [18] Their experimental results match well with standard macrofluidic theory and thus, their findings also suggest validation of the classical transport laws to the scale used. They also found that transitional regimes and turbulent regimes were initiated at similar Reynolds numbers as conventional macrochannels. The authors suggest that a bias effect on the solid/fluid interface temperature measurement may be a source of

15

discrepancy that is falsely reported as a scaling effect in previous works by different authors. Large inconsistencies in experimental microfluidic pressure drop data motivated Kohl et al. to develop a microchannel experimental platform that is capable of unobtrusively obtaining internal pressure measurements. [19] Most experimental microfluidics data obtained pressure measurements in plenums located upstream and downstream of the microchannels, with very little agreement. Kohl et al.s system utilizes pressure-sensing membranes and allows for post-fabrication calibration and adjustment, with calibration uncertainties as low as 2%. The system showed that standard models with appropriate assumptions are capable of accurately predicting the measured pressure data. Owhaib and Palm conducted an experimental investigation of single-phase convective heat transfer in circular microchannels. [20] The authors compared their measured experimental results to the classical correlation as well as numerous microfluidics-intended correlations published in recent literature, for both laminar and turbulent flows. The classical model and experimental results showed good agreement. On the contrary, none of the microchannel correlations showed agreement. It is important to note, however, that the diameters of the channels studied were 0.8mm, 1.2mm, and 1.7mm; substantially high for microchannel designation. Owhaib and Palm also observed that the heat transfer coefficients for the three channels observed were nearly identical in the laminar regime. Qu, Mudawar, Lee, and Wereley studied the flow development and pressure drop in a microchannel for various flow rates both experimentally and computationally, also in

16

an effort to examine the validity of the conventional Navier-Stokes equations for accurately predicting fluid behavior on the micro-scale. [7] Adiabatic, single-phase liquid water flows for Reynolds numbers ranging 196-2215 were studied in a microchannel of w,h, and l 222m, 694m, and 12cm respectively. Experimentally, the velocity fields and pressure drops were measured using a micro-particle image velocimetry (micro-PIV) system and pressure transducers, respectively. Their numerical model was discretized according to the finite difference (FD) method, and solved using the SIMPLE (SemiImplicit Method for Pressure-Linked Equations) algorithm. Agreement between their measured and predicted microchannel flow behaviors and pressure drops was very good and the authors concluded that the conventional Navier-Stokes equations predicted liquid microchannel flow fairly accurately. Qu and Mudawar performed a three-dimensional numerical analysis of heat transfer in a previously published heat sink of which experimental results were available. [21] The authors built a full three-dimensional finite difference model in order to account for conjugate heat transfer and axial heat conduction, and applied the conventional Navier-Stokes and energy equations. They cite previous microfluidic analyses in which the classical fin analysis has been applied and suggest numerous assumptions of this approach that compromise accuracy; fully developed (hydrodynamic and thermal) flow, complete thermal mixing, and constant heat transfer coefficient. Additionally, they observe several trends regarding heat transfer and temperature; temperature increases approximately linearly with flow direction and heat transfer coefficient and Nusselt Number (Nu) vary three-dimensionally, both axially and on the channel perimeter. They

17

also report some expected effects, such as the increase in length of the developing region and increase in heat transfer (and pressure drop) with increased Reynolds number. Li, Peterson, and Cheng performed a numerical study of heat transfer of the same microchannel heat sink Qu and Mudawar investigated in order to explore the impact of geometric and thermophysical parameters of the fluid on the flow behavior and heat transfer. [8] The authors acknowledge the scattered results obtained in past microfluidics studies from various authors and point out the strong differences of derived empirical correlations. Correlations founded in many previous works may only be applicable to their specific geometry. These differences are especially true when comparing works that studied single microchannels to those who considered entire microchannel heat sinks. Thus, the authors noted a need to develop numerical models that provide more insight into the fundamental physics of the transport processes involved. Li, Peterson, and Cheng utilized the classical Navier-Stokes and energy equations to their model. The authors validated model showed that thermophysical properties of the operating fluid significantly influence both the fluid flow and heat transfer of the heat sink. The aforementioned work demonstrates that microchannel heat sinks are no doubt an extremely attractive cooling solution. Their design is multi-facetted however, and presents an intimidating optimization problem. Channel geometric parameters such as channel length, width, and height, channel spacing, separating wall thickness, and separation distance from the heat source all affect cooling capability and difficulty of fabrication and/or design. Furthermore, limits on pumping power, channel velocity, channel pressure drop, temperature, and heat flux further complicate design optimization.

18

Fisher and Torrance numerically studied conjugate heat transfer in channels of non-rectangular curved boundaries utilizing the complex variable boundary element method (CVBEM). [5] The authors studied these channels as opposed to the oft-studied rectangular cross-sectional channels for a number of reasons; fabrication of rectangular channels often results in non-exact, rounded corners, true sharp corners inhibit local heat transfer by restricting fluid flow, and curved boundaries can increase surface efficiencies in high-aspect ratio fins. The authors are able to define thermal resistance as a function of curvature and other geometric parameters and obtain optimum values for various situations. Xie, He, and Tao numerically studied the effects of channel dimensions, channel fin (separating wall) thickness, bottom thickness, and flow parameters on the capability of a minichannel heat sink. [22] The authors chose to implement minichannels for forced convection, rather than microchannels, in order to lower the pressure drop required for operation of the heat sink. The heat flux, however, will be comparatively lower. In this parametric study, one parameter was varied at a time, while others kept constant. Once an optimum value for the varied parameter was found, it was assumed the best value for all other variable combinations, and the study proceeded to vary the next parameter. There are inherent flaws in this methodology in that not nearly all combinations are considered and the optimization can only be considered approximate. However, the authors were able to conclude that pressure drop is a strong function of channel geometry, and that narrow, deep channels outperform wide, shallow channels. Additionally, the authors were able to find nearly optimum values for the channel separating wall thickness and other parameters. They obtained an approximately optimized configuration for a 20mm by

19

20mm millichannel heat sink that is capable of a heat flux of 256 W cm 2 for an inlet velocity of 1.5 m s . The sink is also capable of a maximum heat flux of 137 W cm 2 for a substantially slower inlet velocity of 0.1 m s . Sasaki and Kishimoto comparatively studied the relationship between cooling capability and microchannel width for a microchannel heat exchanger both experimentally and numerically. [23] Substrates 24mm by 24mm, with heights of 2mm were constructed and attached to a 2x2 array of 8mm chips. The substrates had microchannel arrays of widths 70m, 140m, 340m, and 20mm, and channel heights were fixed at 900m. Using water as the coolant, the authors were able to fit their threedimensional finite difference model to their experimental results with good agreement. The model predicted optimal channel widths of 400m and 250m for pressure drops of 200 kg m 2 and 2000 kg m 2 , respectively. The authors then presented a numericallyderived correlation for optimum microchannel widths for cooling substrates of different sizes. While the results provided by Sasaki and Kishimoto are interesting and show very reasonable behaviors and trends, very little information on the experimental and numerical methodologies and accuracies is provided.

1.3.3. Computational Fluid Dynamics & Heat Transfer Fluid flow and related behavior are described by partial differential equations, which cannot be solved analytically, except for special cases. Computational fluid dynamics and mass and heat transfer (CFD) methods involve approximating fluidic and mass and/or heat transfer behavior computationally (numerically) when analytical solutions are unavailable. [24] The partial differential equations that govern fluid and 20

mass/heat transfer are approximated as a system of algebraic equations through various discretization methods. These algebraic approximations are applied to discrete domains in space and/or time and provide results at these discrete locations. There are multiple methods of discretization, solving, and linearization enacted, depending on the issue at hand, available computing time and resources, and accuracy desired. Popular methods of discretization include the finite difference (FD), finite volume (FV), and finite element (FE) methods. The FD method is generally held as the easiest implemented discretization for simple geometries. A mesh grid is implemented through the problem domain and the governing partial differential equations are approximated as a system of algebraic equations for each grid node. The variables of interest at the principal node and its surrounding nodes in the approximate algebraic equations are held as unknowns. Taylor series expansions or polynomial approximations are used to develop approximations to derivatives of the variables of interest in the algebraic equations implemented. The FD approach is generally limited to simple geometries, which is disadvantageous when complex flow and heat/mass transfer solutions are desired. The FV method utilizes a mesh grid to divide the problem domain into discrete control volumes. The variables of interest are calculated at the centroids of the respective control volumes from the conservation equations. From these centroid values, nodal values on the control volumes surfaces are interpolated. Through integral approximations, an algebraic equation results for each control volume, which contains neighbor nodal values. Unlike the FD approach, the FV method can handle complex geometries. The numerical (computational) models devised in this study are discretized and solved according to the finite volume approach.

21

Like the FV method, the FE method divides the problem domain into discrete volumes, or elements. The governing equations are scaled by a weight function prior to being integrated over the problem domain however. The behavior of the variables of interest is approximated between mesh points, in each finite element. Grids are easily refined for the FE method through subdivision and it is able to deal with arbitrary geometries.

1.4. Current Study

1.4.1. Description In this study, multiple microchannel heat sink configurations are studied computationally (using the FV method) and compared in a cooling capability sense. Weighing the importance of the pressure developed in these schemes, applicable flows are subject to similar pressure constraints. A single cooling channel transports water directly above a power source at a specified total pressure drop (Pt). In the next instance, two channels of the same height (h), but smaller width (w) run above the power source, with the same total pressure drop. Thus, the large single channel is replaced with two smaller channels, separated by a sidewall, in order to gauge the effect on cooling capability. This iteration is then repeated for more, smaller yet channels, and so on. Furthermore, for cases involving 1-4 cooling channels, the height is reduced, also maintaining synonymous pressure constraints, in order to evaluate the cooling costs for a slimmer design and implementation.

22

The power source on which all microchannel heat sinks are implemented is approximated as a silicon cuboid. Microchannel heat sinks consisting of microchannels, silicon sidewalls (behaving as fins), and a silicon cap, are implemented directly above this power source as seen in Figure 1.4.

Cap (Si)

Sidewalls (Si)

Microchannel Heat Sink

y x z

Power source (Si) Microchannels (H2O)

Figure 1.4 Partial view of power source and one possible microchannel heat sink configuration

Contrary to modeling an entire microprocessor of w and length (l) ~12mm by 12mm, and the entirety of the microchannels in the cooling scheme, a smaller portion (unit cell) is modeled for this study in order to keep modeling resolution and computation time realistic, as seen in Figure 1.5. This unit cell approach has been used by many authors in the past, including Qu and Mudawar [21] and Li, Peterson, and Cheng [8].

23

50m 500m 50m Si

Si 300m

H2O

Si

100m
y x

Power Source (Si)

600m
z

Figure 1.5 Portions of power source and microchannel heat sink considered in this study (unit cell). Inlet (-x) view of case 1 is shown in this figure.

Specifically, the power source is taken to be a cuboid with w, h, l = 600m, 100m, and 5mm respectively. The microchannel heat sink configurations are implemented in a volume above the power source, which has w, h, l = 600m, 350m, and 5mm. Distinction of the different heat sink configurations studied is based solely on the number of channels implemented and their respective widths. In essence, the baseline (case 1) microchannel heat sink configuration consists of a single water cooling microchannel of width 500m and height 300m above the source, operating at a specific total pressure drop. The next configuration consists of dual cooling microchannels of w 225m and h 300m above the source, meeting the same pressure drop constraint as the baseline cooling configuration, as seen in Figure 1.6. This iteration is then repeated for 3-7 sequentially smaller distinct channels, as shown in Table 1.2. Sidewall widths are maintained a constant 50m. Flows constrained by several different total pressure drops are studied in all cases in order to study further parametric effects.

24

Furthermore, channel h of cases 1-4 are lowered from 300m to 25m iteratively, also maintaining synonymous pressure constraints, in order to evaluate the cooling costs for implementation of a slimmer design. Cooling capability is quantitatively defined in this study as the average temperature of the power source (Tavg,src). The lower Tavg,src, the higher the cooling capability of a microchannel heat sink configuration.

50m 500m 50m Si


50m

50m 225m

50m 225m Si

50m

Si 300m

H2O

Si
300m

Si

H2O

Si

H2O

Si

100m
y x

Power Source (Si)


y

100m

Power Source (Si)

600m
z

600m
z

Figure 1.6 Inlet (-x) Views of Case 1 and 2, demonstrating first iteration of configuration comparison.

Table 1.2 Different cooling configurations studied

Case 1 (Baseline) 2 3 4 5 6 7

Number of Channels 1 2 3 4 5 6 7

Channel w (m) 500.0 225.0 133.3 87.5 60.0 41.7 28.6

The fluid flows in this study are assumed to be steady, incompressible, and laminar. The traditional value of 2300 is accepted as the critical Reynolds number (Recr). The coolant water enters at the channel inlets at a temperature of 300K with a uniform velocity profile. Thus, developing regions are considered. Suitably, radiation heat transfer

25

is neglected. The continuum assumption is accepted for the fluid studied here, as it is liquid water. [15,25] The Navier-Stokes equations are assumed accurate predictors for the microfluidic flows considered. [6,7,9,16,18,19,20,21]. In addition, the no-slip boundary condition on all fluid/solid interfaces is invoked. Gauge total pressure is specified at all channel inlets, where total pressure is defined as

Pt = Pstc + Pdyn

(1)

where Pstc is the static pressure and Pdyn is the local dynamic pressure. At all channel outlets, gauge Pstc is designated as 0 lb in 2 ( psi) . The power density generated in the power source is equivalent to that of a 12mm by 12mm by 100m device generating 100W; 6.94 109 W m3 . All external boundaries are assumed adiabatic. Because the numerical models constructed in this study are completely 3dimensional in both the solid and fluidic regions, axial heat conduction and conjugate heat transfer are accounted for in the current study. [9,21] All fluid and solid material properties are approximated as constant values, shown in Table 1.3. While implementation of temperature dependence of thermophysical properties would be more realistic, proper implementation into the numerical model devised herein is not plausible at this point. Channel walls are assumed smooth and viscous dissipation is not considered in this study.

26

Table 1.3 Constant material properties used in numerical model

Material Property Density ( kg m3 ) Specific Heat, CP ( J kg K ) Thermal Conductivity, k (W m K ) Viscosity (kg m s)

Silicon, Si 2330 700 149 N/A

Liquid Water, H2O 998.2 4182 0.6 0.001003

1.4.2. Differences Distinction of the current study from past analyses is attributed to a few important points. The majority of microchannel forced convection studies done in the past have constrained the flow with velocity conditions. [10,11,26] A more suitable bounding constraint is the pressure developed from the flow; total pressure drop. The pressures developed from fluid flow in microchannels can significantly affect the performance and/or reliability of microdevices. This was discussed more thoroughly in section 1.3.2.3. While no claims of optimization are asserted in this study, the iterative modification and comparative nature of this work draws significant similarities to optimization studies performed. The optimization study performed by Fisher and Torrance [5] focused on duct shapes with curved boundaries, as an approximation to account for imperfect fabrication techniques (among other reasons), whereas this study focuses on smooth rectangular channels, which is not out of the question as fabrication methods advance technologically. Fisher and Torrances numerical model did not account for axial heat conduction, citing a previous work which claimed it as negligible. On the contrary, many recent works demonstrate the importance of axial heat conduction in numerical analyses of microchannel heat sinks. [6,9] Xie, He, and Tao also performed 27

a numerical optimization study on channel heat exchanger design. [22] The authors studied a water-cooled millichannel heat sink, substantially larger than the channels studied herein. Sasaki and Kishimoto performed a notable experimental and numerical optimization study of the widths of microchannels in a heat sink. [23] While their results are very interesting, very little information is provided regarding their experimental and numerical methodologies and procedure. Further information regarding validation, resolution, etc. would substantially strengthen their results. The iterative reduction of heat sink height performed in this study is distinctive from other studies in its knowingly sacrifice of cooling capability in order to reduce total heat sink volume; an important design consideration because of the scarcity of available real estate at the chip level in todays applications. Furthermore, the resources available have resulted in very high discretization resolution of the numerical models in this study; this is discussed further in chapter 2.

28

Chapter 2 Numerical Solution

2.1. Description of Numerical Model Microchannel heat sink configurations are modeled in three dimensions in the Gambit 2.2.30 preprocessor and analyzed at steady state operation utilizing FLUENT 6.2.16 Computational Fluid Dynamics (CFD) software, which utilizes the Finite Volume Method approach to simultaneously solve the Navier-Stokes and energy equations. [27,28] The implicit segregated solver is employed in this study. Symmetry at the central axial vertical plane is applied to all cases (as seen in Figure 2.1) in order to reduce computation time and allow for higher resolution of the discretization of the volumes studied, mainly the fluidic volumes in the microchannels. Adaptive meshing is also employed in order to more accurately model the fluidic behavior in the microchannels, which can be seen in Figure 2.2. The average number of control volumes being used in the fixed-height models is 1,396,513. Computing time is very reasonable for this resolution. For further validation of the modeling resolution employed, a grid sensitivity analysis was performed, as discussed in section 2.3.

29

25m 225m Si

50m

50m

Plane of Symmetry

Si

H2O

Si

300m

Power Source (Si)


y x

100m

300m
z

Figure 2.1 Symmetry conditions are employed in all models. Inlet (-x) view of case 2 is shown in this figure.

Figure 2.2 Adaptive meshing is employed to enable high-resolution discretization of fluidic regions. Inlet (-x) view of case 4 is shown in this figure.

30

2.2. Numerical Validation It is necessary to validate the current studys numerical methodology by comparing its results with some available in the open literature. Qu, Mudawar, Lee, and Wereley experimentally and numerically studied flow development and pressure drop in a rectangular microchannel. [7] Their work was discussed in more detail in section 1.3.2.4. A numerical model was developed in this study to geometrically match that of which Qu et al. studied. Water flows of Re 196, 1021, 1895, and 2215 were studied in an adiabatic microchannel 222m wide, 694m high, and 120mm long. Figure 2.3 shows the fine agreement of the current studys numerical methodology with Qu et al.s experimental and numerical results, when comparing pressure drop of the channel flows. At the highest Reynolds number studied, the current study is a substantially more accurate predictor of the experimental results, than Qu et al.s numerical study. Comparing fully-developed centerline x-components of velocity derived from the current numerical model to Qu et al.s numerical and experimental results, good agreement is seen as well, with an average percent difference of ~5% from Qu et al.s experimental results. In order to further validate the numerical model used, results involving heat transfer and the energy equation were again compared to those available in the open literature. Qu and Mudawar devised a 3-dimensional numerical model in order to study the heat transfer in a previously experimentally-studied microchannel heat sink. [21] Again, this work was discussed in more detail in section 1.3.2.4. The authors broke the sink into a unit cell involving solid regions and a single microchannel of w,h, and l = 57m, 180m, and 10mm respectively. A constant heat flux was provided at the base of

31

the sink, and water at an inlet temperature of 20C was forced through the microchannel. The problem is formatted in such a way that full 3-D heat transfer is considered, both in the fluidic and solid regions. Among other effects, the authors studied the effects on heat transfer of variance of the Reynolds number. A numerical model complying with the current studys methodology was again constructed to match the scenario Qu and Mudawar studied. Figure 2.4 shows average fluid temperature at the channel outlet as a function of Re, as determined by Qus validated model and the model used herein. Good agreement is again seen for this comparison, with an average percent difference of 5.17% from the previously-validated numerical model. Validation for the numerical methodology utilized in the current study was successful, based on two different studies. Solutions involving solely microfluidic flow and microfluidic flow coupled with full 3-dimensional heat transfer in both solid and fluidic regions agree well with previously published results.

32

Qu et al., Experimental 2.5

Qu et al., Numerical

Farnam, Numerical

1.5

P(bar)
1 0.5 0 100

600

1100

1600

2100

Re

Figure 2.3 Numerical Validation - Comparison of Qu et al. experimental and numerical pressure drop [7] and the current studys numerically-determined pressure drop
Qu & Mudawar, Numerical
50 45 40 35 30 25 20 15 10 5 0 140

Farnam, Numerical

T,avg,outlet (deg-C)

340

540

740

940

1140

1340

Re

Figure 2.4 Numerical Validation - Comparison of Qu & Mudawar validated numerical average outlet temperature [21] and the current studys numerically-determined average outlet temperature

33

2.3. Grid Sensitivity Analysis In addition to the numerical validation performed, a grid sensitivity analysis was performed on one of the microchannel heat exchanger configuration models (case 1, specifically) in order to justify the resolution of discretization of the models. Clearly, the solution process of a low-resolution model will run faster than that of a highresolution model, but at the cost of accuracy. Models constructed must be of high enough resolution to instill confidence in the results, but not so high as to make computation effort and time unreasonable. The grid sensitivity analysis employed in this study compared the resulting variable of most interest, the average temperature of the power source (Tavg,src) from simulations for case 1 at a total pressure drop (Pt) of 1psi, involving different resolution models, as summarized in Table 2.1. Tavg,src and other variables of interest will be discussed further in section 2.4. dz, dy, and dx represent the lengths in each respective direction, of a single control volume in the model. dx is given as an average value because adaptive meshing is employed in the axial (x) direction, in order to provide higher resolution of the fluidic regions at the channel inlet and outlet. The model of highest resolution, with dz,dy,dx = 6.25,6.25,20.00m respectively, is the model actually used in this study. Other heat sink configurations modeled have variable dz and dy of similar scale (generally smaller), depending on the number of channels considered, and the same dx, with variable levels of bias regarding adaptive meshing. Comparing the resolution of the various microchannel heat exchanger configurations; as the number of channels increases, so does the number of control volumes, in order to accurately

34

describe the full profiles of motion in the fluidic regions. Figure 2.5 shows the resulting Tavg,src plotted as a function of the number of control volumes of the different models.

Table 2.1 Model resolutions employed in grid sensitivity analysis and resulting variable of interest

dz (m) dy (m) Avg. dx (m) No. of Volumes Tavg,src(K) 50.00 50.00 156.25 1728 316.2765 25.00 25.00 79.37 13608 314.3514 12.50 12.50 40.00 108000 315.1889 10.00 10.00 31.25 216000 315.3345 6.25 6.25 20.00 864000 315.1557

320

319

318

Model Used Model Used

317

T,avg,src (K)

316

315

314

313

312

311

310 0 100000 200000 300000 400000 500000 600000 700000 800000

Number of Control Volumes

Figure 2.5 Grid sensitivity analysis - Tavg,src plotted as a function of model resolution (number of control volumes)

From figure 5, it is clear that the model used operates in the region where the variable of interest is independent of model discretization resolution. This fact, coupled with the reasonable computation time encountered for solution of models in the particular range, 35

is enough to conclude that the discretization used is acceptable for our numerical models, assuming necessary adjustments for other cases are intelligently implemented.

2.4. Baseline Case Results As previously described in section 1.4.1. (and shown in Figure 1.5), the baseline microchannel heat exchanger configuration of this study consists of a single water cooling microchannel with w, h = 500m, 300m. The initial total pressure drop of the channel flow studied for the baseline case was specified as 1psi. The solutions resulting Pt plotted axially at a channel wall locale is shown in Figure 2.6. It is observable that the total pressure gradient at the channel inlet is higher, and then tapers to an approximately steady drop as the point of observation moves axially down the channel. The x-component of velocity (u) plotted axially at the centerline of the channel is shown in Figure 2.7. Observing that u never develops into a constant value, it is apparent that for the geometric and flow parameters of this configuration, the length of the hydrodynamic developing region (ld) exceeds the length of the microchannel itself. This trend is not uncommon of flows in microchannels. Figure 2.8 shows u half-profiles at the vertical center at several axial locations and also shows how the flow is not able to develop fully before the outlet of the channel. The length of the hydrodynamic developing region is verified analytically via the following relation approximated by Shah and London [29]

ld = (0.06 + 0.07 0.04 2 ) Re d h

(2)

36

where is channel aspect ratio. The analytical determination of ld can be seen in Table 2.2. It will be shown in the following chapter that subsequent cooling schemes have substantially shorter hydrodynamic developing regions.

Figure 2.6 Pt plotted axially in streamwise direction at a channel wall location for case 1 at Pt = 1psi

37

Figure 2.7 u plotted axially in streamwise direction at channel center for case 1 at Pt = 1psi

Figure 2.8 u (X-Velocity) half-profiles at vertical center of microchannel, for axial locations of 0,1,2,3,4mm for case 1 at Pt = 1psi

38

Figure 2.9 shows temperature contours of the baseline cooling scheme and power source from two views, (a) standard isometric and (b) +x (view of outlet). The isometric view shows the approximate linear temperature rise in the flow direction of the solid and fluid regions, a behavior noted by Qu and Mudawar. [21] Figure 2.10 more closely examines the axial temperature rise. The outlet view shows that the maximum temperature (Tmax) of the power source occurs at the lowest center point, at the outlet, as expected. For Pt = 1psi, the maximum temperature of the power source is reported as 319.9K. For this case, it is also observable that the heat flux at the channel walls varies peripherally. Heat is conducted from the power source, to the Si sidewalls and the Si cap, but the flux provided to the channel is a maximum at the bottom surfaces of the channel, the closest proximity to the power source. Furthermore, it is observable that a large central slug of fluid in the microchannel remains unheated. This is a sign that thermal mixing is very incomplete, expected because of the laminar flow assumptions of this model. This observation was also noted by Qu and Mudawar [21], and is a major contradiction to the assumptions of the classical fin analysis employed by many past works. Figure 2.10 shows the incomplete thermal mixing as well, as the center of the channel experiences no temperature rise.

39

(a)

(b)

Figure 2.9 Temperature (K) contours of case 1 at Pt = 1psi at (a) Isometric view and (b) Outlet (+x) view

40

Figure 2.10 Temperature (K) of case 1 at Pt = 1psi plotted axially at a channel sidewall location , channel center

Figure 2.11 shows the surface Nusselt number (Nu) on the walls of the channel. As expected, the Nu is higher at the bottom and lower portions of the sides of the channel, where in closer proximity to the power source. In addition to this behavior, the Nu approaches zero in the corners of the channel. These corner areas of no-convection expand slightly when observation moves in the flow direction and are the result of the noslip bounded flow approaching the corners from both the y and z directions. Furthermore, Nu is substantially higher at the areas near the channel inlet and gradually tapers lower towards the outlet. These observations are again consistent with those reported by Qu and Mudawar. [21] It is also possible to observe small areas of reduced-convection in proximities of the applied symmetry boundary condition. These are most likely the result of numerical noise and are shown to not significantly affect cooling results. Average

41

Nusselt numbers (Nuavg) for affected channels are on average only 1.43% lower than those of unaffected channels.

Figure 2.11 Surface Nu contours of microchannel walls for case 1 at Pt = 1psi. Channel inlet is in the foreground of this figure.

Results of case 1 for Pt of 1, 2 , 3, and 4psi are tabulated in table 2.2. As expected, larger Pt accommodates larger average velocity (Vavg) . Additionally, Re, ld, and Nuavg are also directly related to Pt. Average Temperature of the power source (Tavg,src) is also reported and serves as the method of measuring cooling capability for this study; The lower Tavg,src, the more effective the cooling configuration. As Pt increases, cooling capability increases, as expected. Figures 2.12-2.14 demonstrate these trends graphically. Similar results for cases 2-7 of this study are presented in chapter 3.

42

Table 2.2 Results of case 1 for 4 total pressure drops

Pt (psi) 1 2 3 4

Vavg (m/s) Re ld (mm) Nuavg Tavg,src (K) 2.207 927.3 30.5 6.94 315.16 3.301 1386.7 45.6 7.62 312.60 4.137 1738.2 57.1 8.00 311.35 4.900 2058.5 67.6 8.10 311.10

V,avg(m/s)

0 1 1.5 2 2.5 3 3.5 4

P,t (psi)

Figure 2.12 Vavg of microflow in case 1 for Pt of 1,2,3,4psi


9 8 7 6

Nu,avg

5 4 3 2 1 0 1 1.5 2 2.5 3 3.5 4

P,t (psi)

Figure 2.13 Nuavg of microchannels of case 1 for Pt of 1,2,3,4psi

43

316 314 312

T,avg,src (K)

310 308 306 304 302 300 1 1.5 2 2.5 3 3.5 4

Pt (psi)

Figure 2.14 Tavg,src of case 1 for Pt of 1,2,3,4psi

44

Chapter 3 Results & Conclusions

3.1 Discussion of Results

3.1.1. Quantity/Width Variation The cases presented in Table 1.2 are analyzed for the same Pt of which the baseline case was observed. Consider moving from case 1 to case 2; since Pt will remain constant, two effects are expected when we move from a single cooling channel with w, h of 500m, 300m to dual cooling channels with w, h of 225m, 300m. First, the velocities in the channels of case 2 are lower than the velocity observed in the baseline case. Since velocity is directly related to Nu, this has a negative effect on cooling capability. However, effective heat transfer surface area (Aeff) is higher in case 2, which will have a positive effect on cooling capability. Figure 3.1 demonstrates these points. Figure 3.2 shows u plotted central-axially for cases 2-7 for a Pt of 1psi. It is apparent that the length of the hydrodynamic developing region is larger than the actual length of the channel for case 2, as it was for case 1. Cases 3-7 however, reach fullydeveloped flow (case 3 only for a brief stint). The numerical results indication of flow development shows good agreement when compared with the analytical correlation

45

provided earlier, as equation 2 in section 2.4. Table 3.1 summarizes the analytical determinations of ld.
50m 500m 50m Si 50m 50m 225m Si 50m 225m 50m

Si 300m

H2O

Si 300m

Si

H2O

Si

H2O

Si

100m
y x

Power Source (Si)


y

100m

Power Source (Si)

600m
z

600m
z

For Equal Pt: Effect of Alteration (Case 12) Expected Cooling Effect? Velocity Decreases, Cooling Decreases, Surface Area Increases, Cooling Increases,
Figure 3.1 Expected effects on cooling when microchannel heat sink configuration is altered from case 1 to case 2

Table 3.1 Analytical determinations of ld for Pt = 1psi from equation 2

Case 1 (Baseline) 2 3 4 5 6 7

ld (mm) 30.500 11.900 4.000 1.300 0.360 0.087 0.019

Figure 3.3 shows temperature plotted central-axially for cases 2-7 for a Pt of 1psi. It is observable that the centerline of case 2 experiences no temperature increase for the length of the microchannel, again, as it was with case 1. Case 3 shows a slight temperature increase in proximity to the channel outlet, and central-heating steadily rises

46

as the cases advance. The quasi-linear axial temperature rise noted by Qu and Mudawar [21] and Li et al. [8] is also apparent from Figure 3.3.

Fig 3.2 u (X Velocity) plotted axially in streamwise direction at channel center for case 2-7 at Pt = 1psi.; top-left: case 2, top-right: case 3, middle-left: case 4, middle-right: case 5, bottom-left: case 6, bottom-right: case 7. Note that scale of X Velocity axis is not uniform from graph to graph.

47

Fig 3.3 T(K) plotted axially in streamwise direction at channel center for case 2-7 at Pt = 1psi.; topleft: case 2, top-right: case 3, middle-left: case 4, middle-right: case 5, bottom-left: case 6, bottomright: case 7. Note that scale of T axis is uniform from graph to graph.

Figures 3.4 and 3.5 show the behavior of Vavg and Nuavg as the number of channels is increased from 1 to 7 at Pt of 1, 2, 3, 4psi. Figure 3.6 shows how total Aeff increases as the number of channels is increased from 1 to 7 at various Pt. We can see in Figures 3.4 and 3.5 that as the channels become progressively smaller at all Pt, Vavg and Nuavg are converging to a common limit, 0. As the cross-sectional area (Ac) for each microchannel decreases, the implemented Pressure constraint will drive Vavg (and thus

48

Nuavg) to 0. An interesting difference between the behaviors in Figures 3.4 and 3.5 is observable. While Vavg changes substantially when different Pt are considered for the same case, Nuavg does not vary on the same scale. For instance, Vavg increases 122% for Case 1, when Pt is increased from 1psi to 4psi. However, Nuavg increases only 17% for the same alteration. Thus the dependence of Nuavg on Vavg no doubt exists, but a smaller Vavg (slower flow) may be implemented if a low Pressure application is desired, with small effect on Nuavg. Figure 10 clearly shows the linear increase in total Aeff as the number of channels is increased, in spite of the fact that the channels become sequentially smaller. As stated earlier and shown in Figure 3.6, increasing Aeff will continue to have a positive impact on cooling capability.

4 psi 5 4.5 4 3.5 V,avg (m/s) 3 2.5 2 1.5 1 0.5 0 1 2 3

3 psi

2 psi

1 psi

Number of Channels (Case)


Figure 3.4 Vavg for all cases at multiple Pt

49

4 psi 9 8 7 6 Nu,avg 5 4 3 2 1 0 1 2 3

3 psi

2 psi

1 psi

Number of Channels (Case)


Figure 3.5 Nuavg for all cases at multiple Pt

2.500E-05

2.000E-05

A,eff (m^2)

1.500E-05

1.000E-05

5.000E-06

0.000E+00 1 2 3 4 5 6 7 Number of Channels (Case)


Figure 3.6 Aeff for all cases

50

Cooling capability, represented by Tavg,src, for all cases studied is presented in Figure 3.7. It is observable that cooling capability initially increases, tapers off, and then decreases dramatically as the different cases are observed, for all Pt studied. Thus, considering the variables altered from case to case, an optimum microchannel heat sink configuration does indeed exist for this application, for various pressures. Additionally, as the trend of Nuavg shown in Figure 3.5 insinuated, the effect of Pt on cooling capability for cases 1-5 is minimal. The same cannot be said for cases 6 and 7 though, as the effect is becoming more pronounced, especially when moving from a Pt of 2psi to 1psi.

4 psi 350 345 340 335 T,avg,src (K) 330 325 320 315 310 305 300 1 2 3

3 psi

2 psi

1 psi

Number of Channels (Case)


Figure 3.7 Tavg,src for all cases at multiple Pt

Table 3.2 presents details on the highest-performing cooling configurations for each Pt studied. While 5 cooling channels of w 60.0m performs the best for Pt of 3 and 4psi, 51

the gain over 4 cooling channels of w 87.5m is minimal. Therefore, if a cooling configuration encompassing all Pt studied is to be chosen, case 4 is the optimum configuration.

Table 3.2 Highest-performing microchannel heat sink configurations for multiple Pt

Pt # of Channels (psi) (Case) 1 4 2 4 3 5 4 5

Vavg wch (m) (m/s) 87.5 87.5 60.0 60.0 0.836 1.446 1.158 1.482

Re 126.9 219.5 129.7 166.0

Nuavg Tavg,src (K) 1.56 1.70 1.13 1.17 309.6 307.7 306.8 306.2

3.1.2. Height Variation Compactness being as important as it is for chip level cooling designs, the height of the channels in cases 1-4 are reduced iteratively from the original 300m to 25m. This alteration is not claimed as further design optimization (for such an iterative optimization assumption would not consider all variation possible) but rather to gauge the effects of a more compact design on cooling capability, for these cases. Pt of 1, 2, 4, and 8psi were considered for the alteration of case 4 and Pt of 1, 2, and 4psi were considered for cases 1-3 (in order to keep Re in the traditionally-bounded laminar regime). Figures 3.8 and 3.9 show how Vavg and Nuavg behave for these height-variable configurations for case 4. Vavg decreases as Pt is held constant and hch decreases, as expected. Nuavg actually increases slightly when channel height is decreased to 200m. This is not entirely surprising when we recall that Nu is shown to be substantially higher at the bottom and lower portions of the side surfaces of the channel. Trimming the top 100m

52

of the channel would decrease the thermal path to the top and upper side surfaces where Nu is generally lower, and may work to inflate Nuavg. This inflation is eventually overpowered by the continually decreasing fluid velocity and Nuavg begins to approach 0 as expected when hch is 25m. Figure 3.10 shows the linear decrease in Aeff of case 4 configurations as the channel height is decreased and the channels become sequentially smaller.

1 psi 4.5 4 3.5 V-avg (m/s) 3 2.5 2 1.5 1 0.5 0 25 75 125

2 psi

4 psi

8 psi

175

225

275

Channel Height (m)


Figure 3.8 Vavg for different hch, for channels of w 87.5m (Case 4) at multiple Pt

53

1 psi 2.5

2 psi

4 psi

8 psi

Nu,avg

1.5

0.5

0 25 75 125 175 225 275 Channel Height (m)


Figure 3.9 Nuavg for different hch, for channels of w 87.5m (Case 4) at multiple Pt

1.80E-05 1.60E-05 1.40E-05 1.20E-05 A,eff (m^2) 1.00E-05 8.00E-06 6.00E-06 4.00E-06 2.00E-06 0.00E+00 25 75 125 175 225 275 Channel Height (m)
Figure 3.10 Aeff for alteration of channel height, for channels of w 87.5m (Case 4)

54

Figure 3.11 reports cooling capability, represented by Tavg,src for the alteration of channel height for case 4 configurations. Tavg,src increases only slightly as hch is lowered from 300m to 100m, similar to the behavior of Nuavg in Figure 3.9. As hch is lowered beyond 100m, Tavg,src increases exponentially to dangerous temperatures at Pt of 1,2, and 4psi. It is important to note that as the boiling point of water is approached, model accuracy must be questioned. The Tavg,src remaining when case 4 is altered to channels of h 100m very well may be affordable cooling losses when compared with the design volume saved. Reducing the channel height from 300m to 100m reduces the height of the cooling scheme over 57%, while Tavg,src only increases an average of 10.9K. Additionally, it can be noted that the effect Pt has on cooling capability is minimal for hch of 100m and above. The effect becomes significant as channel height is lowered below 100m. This behavior is similar to that observable in Figure 3.7.
1 psi 500 480 460 440 T,avg,src (K) 420 400 380 360 340 320 300 25 75 125 175 225 275 Channel Height (m)
Figure 3.11 Tavg,src for alteration of channel height of case 4 at multiple Pt

2 psi

4 psi

8 psi

55

Figure 3.12 shows Temperature contours for case 4 altered to a hch of 100m, with Pt of 1psi. The sidewalls and cap volumes are at similar temperature as the power source, thus we are seeing more symmetric heating of the cooling fluid with regard to the channel perimeter. We can also see more complete heating of the cooling fluid, when compared to that of Figure 2.9.

Figure 3.12 Temperature (K) contours of case 4, altered to hch of 100m at Pt = 1psi at Outlet (+x) view

Figure 3.13 compares the cooling capabilities of the height-variable microchannel heat sink configurations (cases 1-4) for Pt of 1, 2, and 4psi. It is very important to observe that as channel h of the configurations decreases, the optimum cooling configuration changes; for each Pt studied, there is a height-dependent point at which case 4 ceases to be the highest performing cooling scheme. The baseline cooling scheme (case 1) begins 56

to outperform other cases, and configurations of case 2 and 3 follow suit, completely reversing the trend seen for higher heights. This transition first occurs in the results obtained for heights of 100m for all Pt. The transition can be interpolated to heights of ~165m, ~135m, and ~120m for Pt of 1, 2, and 4psi respectively. Thus, it can be inferred that as total pressure drop increases, the transition of optimum cooling configurations will occur at lower height. Figures 3.14-3.17 display results for Vavg, Nuavg, Aeff, and Tavg,src as a function of both channel height and heat sink configuration for Pt of 4psi. Please note that Figure 3.17, pertaining toTavg.src, is rotated to provide best insight to the slope of the surface, and thus the orientation of axes for hch and case are not uniform to the other figures. Vavg behaves as expected, achieving a maximum value at the baseline case, with the highest height, 300m and decreases as the number of channels increases and/or hch decreases. Nuavg follows suit. Aeff achieves a maximum value at case 4, with h of 300m and decreases as hch and/or the number of channels decreases. The surface plot of Tavg.src clearly shows the transition of optimum cooling configurations, as 2 local surface minima are observable at 4 channels of h 300m and 1 channel of h 50m.

57

1 Channel
440

2 Channels

3 Channels

4 Channels

420

400

T,avg (K)

380

360

340

320

300 25 75 125 175 225 275

Channel Height (m) 1 Channel


440

2 Channels

3 Channels

4 Channels

420

400

T,avg (K)

380

360

340

320

300 25 75 125 175 225 275

Channel Height (m) 1 Channel


440

2 Channels

3 Channels

4 Channels

420

400

T,avg (K)

380

360

340

320

300 25 75 125 175 225 275

Channel Height (m)

Figure 3.13 Tavg,src for alteration of channel height of cases 1-4 at Pt of 1, 2, and 4psi; Top: 1psi, Middle: 2psi, Bottom: 4psi

58

Figure 3.14 Surface plot of Vavg(m/s) as a function of hch(m) and No. of Channels (Case) for Pt of 4psi.

Figure 3.15 Surface plot of Nuavg as a function of hch(m) and No. of Channels (Case) for Pt of 4psi.

59

Figure 3.16 Surface plot of Aeff as a function of hch(m) and No. of Channels (Case).

Figure 3.17 Surface plot of Tavg,src(K) as a function of hch(m) and No. of Channels (Case) for Pt of 4psi.

60

3.2. Summary & Conclusions In this study, multiple microchannel heat sink configurations subject to identical pressure constraint were studied numerically and compared in terms of total cooling capability. The validated numerical model shows that the microchannel heat sink considered in this application is capable of solitarily cooling a 100W device of 12mm x 12mm x 100m, with flows of Pt = 1-4psi, resulting in operating temperatures of only 306.2K-309.6K respectively. Thus, microchannel heat sinks will continue to be effective methods of first-level cooling for the foreseeable future as processor power dissipation and compactness continues to increase. The numerical model also shows that for the application studied, a heightdependent optimum cooling scheme exists. Four microchannels of w 87.5m (case 4) at various Pt generally outperform the other cases for the same Pt when a hch greater than ~200m is feasible. When hch is lowered beyond this Pt-dependent limit, a shift in optimum heat sink configuration is experienced; a single channel of w 500m outperforms all other cases considered. This shift in optimum configuration suggests that the optimum cooling configuration will also depend on other variables maintained constant in this study, such as channel length, power dissipated, etc. Furthermore, results show that for the application studied, hch can be lowered to at least 100m, with small losses in cooling capability. A reduction of channel height from 300m to 100m reduces the height of the microchannel heat sink over 57%, at an average temperature rise of only 10.9K. When the height of the microchannel heat sink is lowered, results show more symmetric heating and increased heat saturation of the coolant.

61

Additionally, cooling capability dependence on Pt is shown to be minimal for high-performing microchannel configurations of all heights. While a higher Pt does indeed provide for higher cooling capability, the gains are minimal for high-performance cooling configurations. For example, increasing Pt from 1psi to 4psi (300% increase) for case 4 at the initial hch of 300m lowers Tavg,src only 3.2K. The same cannot be said for schemes that result in less effective cooling; Tavg,src dependence on Pt is much more substantial for these configurations.

62

Chapter 4 Future Considerations

Future work in this area should continue to consider pressure constraints as a primary concern, as eventual microfluidics-cooling implementation will be governed by such constraints. Pt and the resulting stresses are a large source of apprehension toward microchannel heat sink feasibility, and thus should be carefully and thoroughly considered. Studies should continue to focus on feasible applications and operating conditions in order to hasten large-scale acceptance and installation of this promising technology. It would also prove advantageous to perform comparative analyses of microchannel heat sinks with other emerging first-level cooling methods, such as microjet impingement. Furthermore, isolated spot cooling and overall chip cooling should be classified to proper distinction and cooling technology must be adapted in order to provide the simplest, most effective solutions to particular thermal management issues. Additionally, appropriate scaling effects and other assumptions implemented in this numerical study, should be reevaluated, and if needed, appropriate adjustments made. Perhaps most important, temperature-dependence of thermophysical properties involved in analyses should be implemented in order to strengthen accuracy and confidence of the models. Furthermore, as channels in particular configurations become increasingly

63

smaller, viscous dissipation effects may need to be considered in the conventional theory in order to maintain accuracy. If a true, universal design optimization of microchannel heat sinks is to be developed based on the many parameters involved, an analytical correlation verified by experimental and/or numerical analyses would be the most efficient and justifiable approach. An experimental or numerical approach to such a complex optimization study would complicate exponentially as new factors are considered.

64

Nomenclature

Symbol A dh h l Nu P P Re T u V w Greek Symbol Subscript avg c ch cr d dyn eff max src stc t

Description Area Hydraulic Diameter Height Length Nusselt Number Pressure Pressure Drop Reynolds Number Temperature Velocity Component in x-direction Velocity Width Description Aspect Ratio Description Average Cross-Sectional Channel Critical Hydrodynamic Developing Region Dynamic Effective (Heat Transfer) Maximum Power Source Static Total

65

References [1] [2] [3] [4] [5] [6] [7] Tuckerman, D.B. and Pease, R.F.W., High-performance heat sinking for VLSI, IEEE Electron.. Dev. Lett. EDL-2 (1981) 126-129. iNEMI Technology Roadmaps 2007, Appendix A, January 2007. Sergent, J.E. and Krum, A., Thermal Management Handbook for Electronic Assemblies, McGraw-Hill 1998. Tummala, R.R., Fundamentals of Microsystems Packaging, McGraw-Hill, 2001. Fisher, T.S. and Torrance, K.E., Constrained optimal duct shapes for conjugate laminar forced convection, Int. J. Heat and Mass Transfer (2000) 43 pp.113-126. Morini, G.L., Viscous heating in liquid flows in micro-channels, Int. J. Heat and Mass Transfer (2005) 48 pp. 3637-3647. Qu, W., Mudawar, I., Lee, S., and Wereley, S.T., Experimental and Computational Investigation of Flow Development and Pressure Drop in a Rectangular Micro-channel, ASME J. Electronic Packaging (2006) 128 pp. 1-9. Li, J., Peterson, G.P., Cheng, P., Three-dimensional analysis of heat transfer in a micro-heat sink with single phase flow, Int. J. Heat and Mass Transfer (2004) 47 pp. 4215-4231. Herwig, H. and Hausner, O., Critical view on new results in micro-fluid mechanics: an example, Int. J. Heat and Mass Transfer (2003) 46 pp. 935-937. Ho, C. and Tai,Y., Micro-electro-mechanical-systems (MEMS) and fluid flows, Annu. Rev. Fluid Mech. (1998) 30 pp. 579-612. Peng, X.F. and Piao, Y., Characteristics of fluid flow and convection heat transfer in microchannels, Adv. in Fluid Mech. (2002) 32 pp. 85-94. Peng, X.F. and Peterson, G.P., The effect of thermofluid and geometrical parameters on convection of liquids through rectangular microchannels, Int. J. Heat and Mass Transfer (1995) 38 pp. 755-758. Wang, B.X. and Peng, X.F., Experimental investigation on liquid forcedconvection heat transfer through microchannels, Int. J. Heat and Mass Transfer (1994) 37 pp. 73-82. Harms, T.M., Kazmierczak, M., Gerner, F.M., Holke, A., Henderson, H.T., Pilchowski, J., Baker, K., Experimental Investigation of Heat Transfer and 66

[8]

[9] [10] [11] [12]

[13]

[14]

Pressure Drop through Deep Microchannels in a (110) Silicon Substrate, ASME HTD (1997) 351 pp. 347-357. [15] [16] Guo, Z. and Li, Z., Size effect on single-phase channel flow and heat transfer at microscale, Int. J. Heat and Fluid Flow (2003) 24 pp. 284-298. Xu, B., Ooi, K.T., Wong, N.T., and Choi, W.K., Experimental investigation of flow friction for liquid flow in microchannels, Int. Comm. Heat Mass Transfer (2000) 27 pp. 1165-1176. Morini, G.L., Single-phase convective heat transfer in microchannels: a review of experimental results, Int. J. Thermal Sci. (2004) 43 pp. 631-651. Bavire, R., Favre-Marinet, M., and Le Person, S., Bias effects on heat transfer measurements in microchannel flows, Int. J. Heat and Mass Transfer (2006) 49 pp. 3325-3337. Kohl, M.J., Abdel-Khalik, S.I., Jeter, S.M., and Sadowski, D.L., A microfluidic experimental platform with internal pressure measurements, Sensors and Actuators (2005) A 118 pp. 212-221. Owhaib, W. and Palm, B., Experimental investigation of single-phase convective heat transfer in circular microchannels, Exp. Thermal and Fluid Sci. (2004) 28 pp. 105-110. Qu, W. and Mudawar, I., Analysis of three-dimensional heat transfer in microchannel heat sinks, Int. J. Heat and Mass Transfer (2002) 45 pp. 3973-3985. Xie, X.L., He, Y.L., and Tao, W.Q., Numerical study of laminar heat transfer and pressure drop characteristics in a water-cooled minichannel heat sink, THERMES 2007: Thermal Challenges in Next Generation Electronic Systems, pp.179-186. Sasaki, S. and Kishimoto, T., Optimal Structure For Microgrooved Cooling Fin for High-Power LSI Devices, Electron. Lett. (1986) 22 pp. 1332-1334. Ferziger, J.H., Peri, M., Computational Methods for Fluid Dynamics, 3rd Edition, Springer 2002. Zohar, Y., Heat Convection in Micro Ducts, Kluwer Academic Publishers 2003, pp. 13. Sobhan, C.B. and Garimella, S.V., A Comparative Analysis of Studies on Heat Transfer and Fluid Flow in Microchannels, Microscale Thermophysical Engineering (2001) 5 pp. 293-311. FLUENT Version 6.2.16, ANSYS Inc., Canonsburg, PA, U.S.A. 67

[17] [18]

[19]

[20]

[21] [22]

[23] [24] [25] [26]

[27]

[28] [29]

FLUENT 6.2 Users Guide, ANSYS Inc., Canonsburg, PA, U.S.A. Shah, R.K. and London, A.L., Laminar Flow Forced Convection in Ducts: A Source Book for Compact Heat Exchanger Analytical Data, Academic, New York 1978. Farnam, D.S., Sammakia, B., Ackler, H., and Ghose, K., Comparative Analysis of Microchannel Cooling Schemes Subject to a Pressure Constraint, Proceedings of 5th International Conference on Nanochannels, Microchannels, and Minichannels (2007) ICNMM2007-30187. Cengel, Y.A., Heat Transfer, A Practical Approach, McGraw-Hill 1998, pp.370. Adams, T.M., Abdel-Khalik, S.I., Jeter, S.M., and Qureshi, Z.H., An experimental investigation of single-phase forced convection in microchannels, Int. J. Heat and Mass Transfer (1998) 41 pp. 851-857.

[30]

[31] [32]

68

You might also like