0% found this document useful (0 votes)
73 views

Mahajan Dissertation

Uploaded by

hithere
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
73 views

Mahajan Dissertation

Uploaded by

hithere
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 123

Template C with Schemes v3.0 (beta): Created by J.

Nail 06/2015

Development of oscillating heat pipe for waste heat recovery

By
TITLE PAGE
Govinda Mahajan

A Dissertation
Submitted to the Faculty of
Mississippi State University
in Partial Fulfillment of the Requirements
for the Degree of Doctor of Philosophy
in Mechanical Engineering
in the Department of Mechanical Engineering

Mississippi State, Mississippi

December 2016
Copyright by
COPYRIGHT PAGE
Govinda Mahajan

2016
Development of oscillating heat pipe for waste heat recovery

By
APPROVAL PAGE
Govinda Mahajan

Approved:

____________________________________
Heejin Cho
(Major Professor)

____________________________________
Scott M. Thompson
(Minor Professor)

____________________________________
Pedro J. Mago
(Committee Member)

____________________________________
Rogelio Luck
(Committee Member)

____________________________________
Keith Walters
(Graduate Coordinator)

____________________________________
Jason Keith
Dean
Bagley College of Engineering
Name: Govinda Mahajan
ABSTRACT
Date of Degree: December 9, 2016

Institution: Mississippi State University

Major Field: Mechanical Engineering

Major Professor: Dr. Heejin Cho

Title of Study: Development of oscillating heat pipe for waste heat recovery

Pages in Study: 104

Candidate for Degree of Doctor of Philosophy

The development and implementation of technologies that improves Heating

Ventilation & Air Conditioning (HVAC) system efficiency, including unique waste heat

recovery methods, are sought while considering financial constraints and benefits. Recent

studies have found that through the use of advanced waste heat recovery systems, it is

possible to reduce building’s energy consumption by 30%.

Oscillating heat pipes (OHP) exists as a serpentine-arranged capillary tube,

possesses a desirable aerodynamic form factor, and provides for relatively high heat

transfer rates via cyclic evaporation and condensation of an encapsulated working fluid

with no internal wicking structure required. In last two decade, it has been extensively

investigated for its potential application in thermal management of electronic devices.

This dissertation focuses on the application of OHP in waste heat recovery systems. To

achieve the goal, first a feasibility study is conducted by experimentally assessing a nine

turn copper-made bare tube OHP in a typical HVAC ducting system with adjacent air

streams at different temperatures.

Second, for a prescribed temperature difference and volumetric flow rate of air, a

multi-row finned OHP based Heat Recovery Ventilator (OHP-HRV) is designed and
analyzed for the task of pre-conditioning the intake air. Additionally, the energy and cost

savings analysis is performed specifically for the designed OHP-HRV system and

potential cost benefits are demonstrated for various geographical regions within the

United States.

Finally, an atypically long finned OHP is experimentally investigated (F-OHP)

under above prescribed operating condition. Helical fins are added to capillary size OHP

tubes at a rate of 12 fins per inch (12 FPI), thereby increasing the heat transfer area by

433%. The coupled effect of fins and oscillation on the thermal performance of F-OHP is

examined. Also, F-OHP’s thermal performance is compared with that of bare tube OHP

of similar dimension and operating under similar condition. It was determined that OHP

can be an effective waste heat recovery device in terms of operational cost,

manufacturability, thermal and aerodynamic performance. Moreover, it was also

determined that OHP-HRV can significantly reduce energy consumption of a commercial

building, especially in the winter operation.


ACKNOWLEDGEMENTS

I would like to express my deepest gratitude to my advisor, Dr. Heejin Cho, who

has provided the support, encouragement, and guidance to complete this work. I would

like to genuinely thank my committee, Dr. Scott M. Thompson, Dr. Rogelio Luck, and

Dr. Pedro Mago, for all of their support and guidance throughout my career as a student.

Additionally, I would like to thanks Mechanical Engineering Department at Mississippi

State University for providing me with the funding to complete this degree. Finally, I

would like to thank my family who have been a great source of motivation in this

educational endeavor.

ii
TABLE OF CONTENTS

ACKNOWLEDGEMENTS ................................................................................................ ii

LIST OF TABLES .............................................................................................................. v

LIST OF FIGURES ........................................................................................................... vi

NOMENCLATURE .......................................................................................................... ix

CHAPTER

I. INTRODUCTION .................................................................................................1

1.1 Waste Heat Recovery ................................................................................1


1.2 Conventional Heat Pipe .............................................................................3
1.3 Enthalpy Wheel .........................................................................................6
1.4 Fixed-Plate Heat Exchanger ......................................................................7
1.5 Oscillating Heat Pipe .................................................................................8
1.5.1 Operating Mechanism of OHP ..........................................................10
1.5.2 Influence of tube internal diameter....................................................11
1.5.3 Influence of Fill ratio .........................................................................12
1.5.4 Influence of Input Heat Flux .............................................................13

II. LITERATURE REVIEW ....................................................................................14

2.1 Experimental studies on OHP .................................................................16


2.1.1 Influence of OHP internal diameter ..................................................17
2.1.2 Influence of number of turns .............................................................19
2.1.3 Influence of working fluid and its filling ratio ..................................20
2.1.4 Influence of Channel shape and its configuration .............................21
2.1.5 Influence of Heat flux input ..............................................................23
2.2 Theoretical studies on OHP .....................................................................27
2.2.1 Start-up condition of OHP .................................................................28
2.2.2 Studies predicting hydro-thermodynamic behavior of OHP .............28
2.3 Application of OHP .................................................................................32
2.3.1 Application as an electronic cooling device ......................................32
2.3.2 Application in solar water heater .......................................................33
2.3.3 Application in waste heat recovery device ........................................34

III. EXPERIMENTAL SETUP AND PROCEDURE ...............................................37


iii
3.1 Experimental setup ..................................................................................37
3.2 Common Peripheral Devices ...................................................................39
3.3 Uncertainty Analysis ...............................................................................40

IV. FEASIBILITY STUDY OF OHP AS AN WASTE HEAT RECOVERY


DEVICE ..............................................................................................................42

4.1 OHP Fabrication ......................................................................................42


4.2 Working fluid selection ...........................................................................43
4.3 Mathematical Model ................................................................................44
4.4 Results & Discussion ...............................................................................47

V. ENERGY AND COST SAVING POTENTIAL OF OSCILLATING


HEAT PIPE BASED HEAT RECOVERY VENTILATOR ..............................54

5.1 OHP-HRV design concept and operating environment ..........................54


5.2 Heat transfer and pressure-drop analysis .................................................57
5.3 Pressure drop through OHP-HRV ...........................................................60
5.4 Energy and Cost Analysis .......................................................................61
5.5 Result and Discussion..............................................................................63
5.5.1 Heat transfer performance .................................................................63
5.5.2 Pressure drop .....................................................................................64
5.5.3 Design ................................................................................................65
5.5.4 Start-up behavior of OHP-HRV ........................................................66
5.5.5 Energy and Cost Saving Analysis Result ..........................................70

VI. EXPERIMENTAL INVESTIGATION OF ATYPICALLY LONG


FINNED OHP AS A WASTE HEAT RECOVERY DEVICE ..........................75

6.1 Finned OHP (F-OHP) design and construction .......................................75


6.2 Fin characteristics ....................................................................................77
6.3 Result and Discussion..............................................................................78
6.3.1 Thermal performance of Finned OHP (F-OHP) ................................79
6.3.2 Thermal performance of bare tube OHP (B-OHP)............................84
6.3.3 Comparative study of F-OHP and B-OHP ........................................86

VII. SUMMARY, CONCLUSION, AND FUTURE WORK ....................................90

7.1 Summery and Conclusion........................................................................90


7.2 Future work .............................................................................................92

REFERENCES ................................................................................................................. 94

iv
LIST OF TABLES

2.1 Characteristics of OHPs investigated in experimental studies


mentioned above .............................................................................................25

3.1 Additional information about the experimental apparatus and the


OHPs investigated in the present work ..........................................................40

3.2 Estimated uncertainty for input variables .......................................................41

4.1 Thermo-physical properties of air at film temperature [108] .........................45

4.2 Time averaged, steady state, thermal performance metric of OHP with
varied fill ratio of n-pentane ...........................................................................47

4.3 Experimental condition at the start-up of OHP with FR = 70 % n-


Pentane ...........................................................................................................50

5.1 OHP-HRV dimension and operating conditions ............................................57

5.2 Price of natural gas and electricity for different locations [122] and
climate [123]...................................................................................................71

6.1 Construction features of F-OHP .....................................................................77

6.2 Construction feature of Bare tube OHP (B-OHP). .........................................78

v
LIST OF FIGURES

1.1 Consumption of energy in U.S. infrastructure [1] ............................................1

1.2 Schematic of conventional heat pipe ................................................................4

1.3 Illustration of typical limitations of conventional heat pipe in different


operational regime [21]. ...................................................................................5

1.4 Schematics of energy wheel .............................................................................6

1.5 Schematics of fixed plate heat exchanger ........................................................7

1.6 Schematic of a closed loop oscillating heat pipe showing liquid and
vapor plug .......................................................................................................10

1.7 Thermo-mechanical boundary conditions of OHP operation [29] .................11

2.1 Different configurations of Loop Type Heat Pipe as proposed by


Akachi [33] .....................................................................................................15

2.2 Oscillating heat pipe design as described by Akashi [32] ..............................16

2.3 OHP configurations investigated by Lui et al. [49] ........................................18

2.4 OHP of double sided triangular (top) and rectangular (bottom) cross-
sectional channel shape [58] ...........................................................................22

2.5 Three dimensional OHP with 20 turn (a and b) and 10 turn (c) [60] .............23

2.6 Three different heating configurations studied by Thompson and Ma


[67] .................................................................................................................24

2.7 Schematic of one-dimensional OHP (Top) and mass spring


mechanical system mimicking liquid plug and vapor plug (Bottom). ...........29

2.8 OHP based compact cooler for electronic devices [81] .................................33

2.9 OHP based solar water heater [88] .................................................................34

vi
2.10 Schematics (left) and illustration (right) of closed loop OHP used by
Supirattanakul et al. [96] ................................................................................35

2.11 Schematic of closed loop OHP (CEOHP) used as a waste heat


recovery device in a batch-type drying process [97] ......................................36

3.1 Test apparatus showing two air ducts with other major components. ............38

3.2 Approximate location of thermocouples (Ti) and pressure sensors (Pi)


along the air ducts (side view) ........................................................................39

4.1 (Left) Schematic of OHP, its dimension and position relative to cold
air stream and hot air stream. (Right) Photograph of OHP under-
investigation. ..................................................................................................44

4.2 Temperature response of OHP-HE with FR = 70% n-Pentane with


operating regime AB, BC, and CD indicated via shaded zones. ....................51

4.3 OHP-HE with FR = 70 % n-Pentane heat recovery rate vs time with


operating regime AB, BC, and CD indicated via shaded zones .....................52

4.4 Comparison of measured, steady state heat transfer through OHP


filled with FR 70 % n –pentane, heat transfer through empty OHP,
and the analytically-estimated maximum heat transfer possible ....................53

5.1 Front view (left) and isometric view (right) of unit row OHP-HRV in
adjacent ducts .................................................................................................55

5.2 OHP-HRV in adjacent ducts side view (left) and top view (right) ................56

5.3 OHP-HRV heat transfer rate vs number of OHP-HRV rows and fin
spacing. Isoplanes for pre-cooling temperature are also provided .................64

5.4 OHP-HRV pressure drop vs fin spacing and number of rows .......................65

5.5 Startup heat transfer of acetone OHP-HRV vs vapor temperature for


δl = 2 μm ..........................................................................................................67

5.6 Start-up heat transfer of acetone OHP-HRV vs vapor temperature for


δl = 10μm ........................................................................................................68

5.7 Startup heat transfer for acetone and water, δl = 6 μm, rn = 2 μm..................69

5.8 OHP-HRV heat transfer and effectiveness vs maximum temperature


difference ........................................................................................................70

5.9 Seasonal waste heat recovery through proposed OHP-HRV across


different cities in United State of America .....................................................73
vii
5.10 Percentage annual energy and cost saving from proposed OHP-HRV
for different cities in United State of America ...............................................74

6.1 Photograph of 9 turn finned OHP showing cut-out view of fins with
FPI equal to 12................................................................................................76

6.2 Pressure drop in cold air stream across F-OHP and B-OHP ..........................79

6.3 Temperature time history of four tests of F-OHP conducted


subsequently with an interval of 24 hour .......................................................81

6.4 Temperature response of F-OHP filled with n-pentane with fill ratio
70 % ................................................................................................................82

6.5 Time-wise average heat recovery rate and thermal resistance of F-


OHP filled with n-pentane with fill ratio 70 % ..............................................84

6.6 Temperature response of bare tube OHP filled with n-pentane with
fill ratio 70 % ..................................................................................................85

6.7 Heat addition rate to cold air stream ( heat recovery rate ) via B-OHP
filled with n-pentane with fill ratio 70 % .......................................................86

6.8 Comparison of time-averaged heat addition rate to cold air stream


through four configurations of OHP...............................................................87

6.9 Comparison of time-wise average thermal resistance of four


configurations of OHPs ..................................................................................88

viii
NOMENCLATURE

A Area (m2)

Adiff Area of diffuser (m2)

Aohp,h Total external surface area of OHP intersected by hot stream

Atotal Total heat transfer area of OHP-HRV (m2)

Amin Minimum free flow area of OHP-HRV (m2)

Boc Critical bond number

C, Cr Ratio of minimum and maximum heat capacity

Cc Total heat capacity of cold air stream (W/K)

Ch Total heat capacity of hot air stream (W/K)

Cmin Heat capacity of hot air stream or cold air stream, whichever is

minimum (W/K)

cp Specific heat capacity (J/kg·K)

cpc Specific heat of cold air stream (kJ/kg·K)

chc Specific heat of hot air stream (kJ/kg·K)

D Diameter (m)

Dc Collar diameter (m)

ix
Efan Fan energy consumption (W)

f Fanning friction factor

Fp Fin pitch (m)

FPI Fin per inch

g Acceleration due to gravity (m/s2)

G Mass flux rate (kg/s·m2)

hfg Latent heat of vaporization (J/kg)

hh Average heat transfer coefficient for hot air stream (W/m2∙K)

j Colburn j-factor

k Thermal conductivity (W/m·K)

L1 Height of OHP-HRV (m)

L2 Depth of OHP-HRV (m)

L3 Width of OHP-HRV (m)

m Fin heat transfer parameter (m-1)

𝑚̇ Mass flow rate (kg/s)

N Number of fins

Nf Number of fins per inch

Nr Number of rows in OHP-HRV

Nt Number of turns in OHP-HRV

x
NTU Number of transfer units

Nu Number of turns in OHP-HRV

Pr Prandtl number

Q Heat transfer rate (W)

Qdelivered Hourly heating/cooling load (W)

Qrcv Hourly waste heat recovery rate (W)

qth Theoretical heat transfer rate between hot and cold air stream

through OHP (kW)

qexp Experimental heat transfer rate between hot and cold air stream

through OHP (kW)

qgain Amount of heat gained by the OHP at evaporator section (kW)

qmax Maximum heat transfer rate possible between the cold and hot air

stream (kW)

R, R” Thermal resistance (K/W)

Re Reynolds number of OHP tube

ri Bubble radius (m)

rn Inner radius (m)

S Tube pitch (m)

T Temperature (°C)

Tair,i Temperature of air stream at condenser or evaporator inlet (oC)


xi
Tair,o Temperature of air stream at condenser or evaporator outlet (oC)

Tv Vapor bubble temperature (°C)

Tci,avg Average temperature measured by the thermocouples located on

OHP condenser surface (oC)

Thi,avg Average temperature measured by the thermocouples located on

OHP evaporator surface (oC)

Tf Film temperature near OHP surface (oC)

TOAT Hourly averaged outdoor air temperature (°C)

TSAT Supply air temperature (°C)

Ts,c Time-averaged, surface temperature of OHP centermost tube (oC)

U Overall heat transfer coefficient (W/m2·K)

Ucp Estimated uncertainty in specific heat capacity of cold air stream

Umc Estimated uncertainty in mass flow rate of cold air stream

Uq Estimated uncertainty in experimental heat transfer rate between

OHP evaporator and condenser

UΔT Estimated uncertainty in temperature difference across the OHP in

cold air stream

V Air velocity (m/s)

v Specific volume, m3/kg

𝑉̇ Volumetric air flow rate (m3/s)


xii
ΔEcomp Hourly energy reduction while cooling (kW)

ΔEcooling Hourly energy savings while cooling (kW)

ΔEheating Hourly energy reduction while heating (kW)

ΔEfurnace Hourly energy reduction while heating (kW)

ΔEfan Increase in hourly fan consumption (kW)

ΔEheating Hourly energy savings during winter operation (kW)

ΔEtotal Total hourly energy savings (kW)

∆P Pressure difference, kPa

∆T Air stream temperature difference across the OHP-HRV (°C)

ΔCostcooling Hourly cost savings while cooling ($)

ΔCostheating Hourly cost savings while cooling ($)

ΔCostel Electricity cost for each location (cent/kW∙hr)

ΔCostng Natural gas cost for each location ($/28316.8 L) or ($/1000 cu. ft.)

Tc Temperature difference across the OHP in cold air stream (oC)

TOHP Difference between the average temperatures measured by the

thermocouples located on OHP evaporator and condenser surface

(oC)

xiii
Greek symbols

δl Film thickness (m)

ε OHP-HRV effectiveness

η Efficiency

𝜂𝑓𝑢𝑟𝑛𝑎𝑐𝑒 Furnace efficiency

μ Dynamic viscosity (N·s/m2)

ν Kinematic viscosity (m2/s)

ρ Density (kg/m3)

σ Surface tension (N/m)

μ Dynamic viscosity (N∙s/m2)

Subscripts

1 Before heat exchanger

2 After heat exchanger

c Condenser

cri Critical

diff Diffuser

D Diagonal

drive Fan belt drive

xiv
diff Diffuser

gain Heat transfer gain by the OHP through hot stream

h Hydraulic diameter

exp Experimental

l Liquid

th Theoretical

v Vapor

xv
CHAPTER I

INTRODUCTION

1.1 Waste Heat Recovery

The power invested for operating heating, ventilation and air conditioning

(HVAC) systems accounts for approximately 40 % of the U.S. total building energy

consumption as depicted in Figure 1.1 [1]. In addition, the recent realization of fresh

outdoor air in maintaining healthy indoor environment have further increased the

building HVAC load. Fresh air load accounts about 20 - 40 % of overall energy

consumption of HVAC systems and can be higher in case 100 % fresh air ventilation is

required (places like sport complex, hospitals, kitchen, etc) [2], [3]. Thus, increasing the

efficiency of the current energy infrastructure and heightening the appeal of

renewable/alternate energy harvesting systems has become a global priority.

Figure 1.1 Consumption of energy in U.S. infrastructure [1]

1
Recent studies have found that through the use of currently available energy

recovery/management technologies, it is possible to reduce the building’s energy

consumption by up to 50% [4]. Additional reductions of up to 30% can be realized

through the use of advanced waste heat recovery/ventilation technologies integrated

holistically with the building design - for utilizing the often neglected temperature

potential between cold/hot air streams in HVAC duct systems. This temperature potential

can drive heat transfer through, for example, air-to-air heat exchangers; thus allowing for

air pre-cooling and/or pre-heating for reducing air conditioning (AC) unit and/or heater

power demand.

Waste heat recovery technologies for comfort to comfort type application, such as

heat recovery ventilators (HRV), can play a critical balancing role in maintaining indoor

air quality and promoting energy conservation by reducing the HVAC load [5]. Heat

recovery ventilators (HRV), often referred as energy recovery ventilators, are the air-to-

air heat exchangers that per-conditions the fresh supply air so as to reduce the HVAC

load. Energy recovery ventilator and heat recovery ventilators differs in the way that the

former transfer not only sensible heat, but latent heat also. Studies suggest that

HRVs/ERVs have the capability to reduce building energy consumption of an air tight

building by 30 % [6], [7], [8]. In fact, HRVs/ERVs could be cost-effective, even if it is

used with air conditioners of low air flow rate configuration [9]. Examples of waste heat

recovery/ventilation technology include: fixed-plate heat exchangers (FP-HEs), enthalpy

(or energy) wheels, heat pipe heat exchangers (HPHEs) and oscillating heat pipe heat

exchangers (OHP-HEs). The next sections discusses the above mentioned type of waste

heat recovery systems in brief.

2
1.2 Conventional Heat Pipe

Conventional heat pipe (CHP), also referred as a heat pipe, deserve a special

mention in the field of waste heat recovery due to their wide acceptability and operational

characteristics. The basic understanding of the conventional heat pipe and its operating

mechanism is vital in light of present work. The difference between oscillating heat pipe

and conventional heat pipe will be shown in subsequent chapters along with how

oscillating heat pipes can overcome some of fundamental operational and physical

limitations of CHP.

The development of heat pipe was instigated by Jacobs Perkins in 1839 [10] with

the invention of hermetic tube boiler. Later, his son Angier March Perkins modified the

technology and patented with the name “Perkin’s Tube”, which led to the development of

modern heat pipe. Perkin’s Tube is a closed loop two phase thermosiphon that used water

as a working liquid. It underwent several modifications before it was successfully used in

a baker oven as a waste heat recovery device. George Grover, a physicist working at Los

Alamos National Laboratory inspired by Perkin’s Tube, developed the first working

model of heat pipe in 1963 [11], although Richard Gaugler from General Motors got the

patent for the idea of heat pipe in 1942 [12].

Conventional heat pipe, as shown in Figure 1.2, consists of a thermally

conductive, metallic shell (i.e. pipe) with an internal wicking structure, such as sintered

mesh screens or particles and/or co-axial grooves [13]. The CHP is evacuated of air and

then filled with a refrigerant or common fluid such as water. It can be observed in Figure

1.2 that one end of the CHP is located near a heat source, i.e. the evaporator, while the

other end is located near a heat sink, i.e. the condenser. When introduced to a sufficient

3
temperature difference, the encapsulated fluid undergoes phase-change and vapor

expansion/contraction, with condensate being returned to the evaporator via gravity and

structure-induced capillary forces. This cyclic evaporation/condensation cycle allows the

CHP to have an ultra-high heat transport capability, and this has allowed its broad

application and appeal in the electronics industry for the past 50+ years [14].

Figure 1.2 Schematic of conventional heat pipe

However, the CHP’s performance has known operating hindrances, such as the

capillary, sonic, and boiling limitations [13]. As shown in Figure 1.3, these limitations

affects the performance of CHP in different operational regime of input heat flux and

evaporator temperature. When the heat added to the evaporator is too high, the bubble

formation restricts liquid to travel back to evaporator through the wick structure. This

results in formation of hot spot and stoppage of CHP operation. This limitation of CHP is

called boiling limitation. CHP under low-temperature operation often ceases to circulate

the given working liquid. This happens because at low-temperature the pumping potential

4
is insufficient to drive the liquid to evaporator section. This phenomenon is called

capillary limitation. When the heat is added to the CHP, its evaporator and condenser

section undergoes continuous mass exchange of liquid plug and vapor bubble. If the

velocity of vapor travelling from evaporator to condenser is too high, it causes choking in

the CHP. This phenomenon typically occurs during the CHP start-up and is called sonic

limitation. Moreover, the cost of construction of CHP is relatively high and that’s one of

the reason why it is losing its market share [4].

A HPHE typically consists of an array of CHPs [15] and are widely investigated

for their application in HVAC systems [16]–[20]. HP-HEs used in HVAC systems are

generally made of copper or aluminum [15]. It comprises of multiple CHPs meandering

through evaporator and condenser section repeatedly. In general, the HP-HE operating at

an effectiveness of 50-80% results in a pressure drop of 100-500 Pa for a face velocity of

2 to 4 m/s [4].

Figure 1.3 Illustration of typical limitations of conventional heat pipe in different


operational regime [21].

5
1.3 Enthalpy Wheel

Enthalpy wheels as shown in Figure 1.4 are externally powered waste heat

recovery device that has slow rotating cylinder filled with air permeable material. The

rotating cylinder absorbs heat from the exhaust in the half of rotation and deliver it to the

supply air in next half. The enthalpy wheels are coated with desiccant that enables it to

transfer both, sensible heat and latent heat. For an equal mass flow rates of 1000 cfm in

counter-flow direction, enthalpy wheel can achieve a sensible effectiveness of ~80 %

with a pressure drop of 100-150 Pa [22]. Air moving through enthalpy wheel suffers from

cross contamination while recovering waste heat, hence it cannot be used in places like

hospitals, sport complex, etc.

Figure 1.4 Schematics of energy wheel

6
1.4 Fixed-Plate Heat Exchanger

Fixed-Plate Heat Exchangers (FP-HEs) are generally made up of aluminum, steel

or copper. As shown in Figure 1.5, FP-HEs consist of series of plates placed equidistant

to each other and are sealed either by welding, gluing, or folding. The supply air and

exhaust is made to flow in alternate channels between the plates and thus the heat transfer

take place between the two air streams. FP-HE used in HVAC system generally have

plates made up of aluminum and can achieve effectiveness as high as 80-90 %, while

having a pressure drop of 24-374 Pa [15] for a flow rate of 1000 cfm. They require less

maintenance than enthalpy wheels as they possess no moving parts, but are expensive

[15].

Figure 1.5 Schematics of fixed plate heat exchanger

7
1.5 Oscillating Heat Pipe

The OHP is a heat transfer device that functions via thermally excited oscillating

motion induced by the cyclic phase change of an encapsulated working fluid” [23]. It

consists of serpentine-arranged tube meanders to-and-through a hot and cold section

repeatedly. As shown in Figure 1.6, this allows for “U-sections” (or turns) to exist in the

heat reception (evaporator) and heat sink (condenser) regions. The OHP is first evacuated

and then partially filled with working liquid. The internal diameter of the tubes is at the

capillary scale; hence, upon filling, a natural liquid-vapor/slug-plug system is formed.

The encapsulated working fluid gains heat from evaporator region, which results in

increase in vapor volume and pressure in the region. Simultaneously, the heat in been

rejected in condenser region, which causes vapor volume and pressure to decrease. This

continual decrease and increase in vapor volume and pressure difference creates an

oscillatory motion in the working fluid. Since the heat transfer is due to this oscillatory

motion, the device is called Oscillating heat pipe (OHP). It is generally agreed in the

research community that most of heat transfer across the OHP is sensible [24]–[26]

From thermodynamic perspective, OHP is also been described as a heat engine

[23]. It converts energy gained in the evaporator section in work and heat. The work is

being utilized by vapor plugs to pump the liquid plug to condenser region. The liquid

plug performs sensible heat transfer in the condenser region. The net work output or input

from the OHP is zero and that why OHP is passive heat transfer device [23].

As mentioned earlier, the working fluid in the OHP exists in form of train of

liquid plug and vapor bubble. The vapor bubble has thin liquid film around it and is in

contact with the tube surface. When heat is added to the wall, it gets distributed partly to

8
the liquid plug and partly to the vapor. Since, the liquid film encompassing the vapor

plug is very thin, the heat transfer resistance is very small, thereby yielding exceptionally

high evaporating heat transfer coefficient, heat transport capability, and thermal

conductivity.

Contrary to CHP, OHP does not require wicking structure for the movement of

working fluid in the tube. Typically, the inner surface of the OHP is an unaltered, smooth

surface. Because of this, OHP has a high manufacturability and can be made with a wide

variety of metals and non-metals ranging from copper to titanium, plastic, glass etc. OHP

can be used with a wide variety of working liquids. This allows the OHP to be functional

in a wide range of temperatures. The OHP thermal performance is gravity dependent

[27], but recent studies have suggested that by increasing the number of turns and

selecting an optimum channel diameter, an OHP’s thermal performance can be rendered

gravity independent [28].

9
Figure 1.6 Schematic of a closed loop oscillating heat pipe showing liquid and vapor
plug

1.5.1 Operating Mechanism of OHP

OHP, although simple in construction, yet have very convoluted operational

mechanism. The oscillatory motion of the working fluid and the resultant heat transfer

depends upon a multitude of parameters. As shown in Figure 1.7, successful operation of

an OHP requires at least three thermo-mechanical boundary condition to be satisfied,

namely, input heat flux, filling ratio, and internal tube diameter [29]. The following

sections discusses the effect of each thermo-mechanical boundary condition in brief.

10
Figure 1.7 Thermo-mechanical boundary conditions of OHP operation [29]

1.5.2 Influence of tube internal diameter

As discussed in previous section, the OHP is first evacuated and then partially

filled with a suitable working liquid. The internal diameter of the tubes is at the capillary

scale; hence, upon filling, a natural liquid-vapor/slug-plug system is formed. Sustainable

operation of the OHP requires coexistence of series of distinct liquid plug and vapor plug.

The interface that develops between liquid plug and vapor plug plays a vital role in

maintaining the liquid-vapor segregation, and its existence is strongly linked to the

surface tension of working fluid and the critical radius of channel. The channel radius

smaller than the critical radius insures the domination of surface tension in liquid-vapor

interface and thus its existence. Mathematically, this condition is shown by Equation 1.1.

Bond number, as shown in Equation (1.1) is a dimensionless group that represents the

11
ratio of gravitational force and body forces. A Bond number less than 1 implies that, for

the specified internal diameter, the surface tension force dominates the body forces in the

interface region and vice versa.

𝜎
𝑟i = Boc √𝑔(𝜌 −𝜌 (1.1)
l v)

where 𝑟i is the inner radius of the OHP, σ is the surface tension of the working liquid, 𝜌l

is the density of working fluid in liquid phase, 𝜌v is the density of working fluid in vapor

phase, and 𝑔 is the acceleration due to gravity

1.5.3 Influence of Fill ratio

Apart from input heat flux and tube internal diameter, filling ratio of the working

fluid is another important parameter in successful operation of the OHP. It is defined as

the ratio of volume of working liquid filled and the OHP’s total internal volume. During

operation, OHP with low fill ratio will have too many vapor slugs and less liquid plugs.

This will cause quick dry out condition and untenable oscillations. On the other hand,

very high fill ratio will result in very few vapor slug, i.e. less pumping agent for

perturbation, which will eventually lower the performance. OHP with 0 % fill ratio will

have no working liquid and the heat transfer through it will only be via conduction. Its

thermal conductivity would be similar to that of material of the tube. OHP with 100 % fill

ratio will have total internal volume filled with working liquid. It will operate like single-

phase thermosyphon [26]. Heat transfer through it will be via convection mode governed

by buoyancy induced liquid circulation. In either case (0 % or 100 % fill ratio) there will

be no bubbles in the OHP and hence no oscillations.

12
1.5.4 Influence of Input Heat Flux

The input heat flux is a vital parameter in the functioning of OHP. It strongly

impacts the start-up condition of the OHP and sustainability of the oscillation. When heat

is added to the OHP, a pressure difference is created between the evaporator and

condenser section, which forces the working liquid to oscillate. The oscillating fluid

transport the heat from one section to another. Hence, uninterrupted heat flux is the

prerequisite for existence of oscillation in the OHP. Input heat flux below a critical level

will not be able to provide enough driving potential for the working fluid to start

oscillating in the OHP. On the other hand, exceedingly heat input heat flux would result

in an evaporator dry-out condition. The OHP operating between these two input heat flux

boundary condition would have very high thermal resistance. An optimum input heat flux

is a unique characteristics of every OHP and, till now it is only determined

experimentally [30]. It varies with the OHP’s orientation, working fluid, operating

temperature etc.

13
CHAPTER II

LITERATURE REVIEW

Oscillating heat pipe (OHP) was invented by Hisateru Akachi in the year 1991

[31]. The patent referred it as “Loop type Heat Pipe (LHP)” and claimed to be suitable in

areas where CHP cannot be used. Akashi defined it as an airtight closed-ended

continuous capillary passage comprising of a heat reception area, a heat rejection area,

and a flow control device. The LHP was fabricated with copper tube of internal diameter

2 mm and charged with working fluid Freon-11and Freon-114 with 60 % fill ratio.

Akashi demonstrated that LHP could achieve better heat transport capability than CHP.

The patent depicted 24 configurations of LHP out of which three are shown in Figure 2.1.

These configurations were equipped with a flow control device called the channel valve.

Later, the channel valve was removed in the refined version of LHP as it failed to assist

in achieving the desired thermal performance [32]. This version of LHP is now referred

as Pulsating Heat Pipe or Oscillating Heat Pipe (OHP) and is shown in Figure 2.2 [31],

[32].

14
Figure 2.1 Different configurations of Loop Type Heat Pipe as proposed by Akachi
[33]

After the introductory patent from Akachi, several investigations have been

conducted focusing on various aspect of OHP such as- understanding its thermodynamic-

hydrodynamic behavior; developing analytical models for predicting its heat transfer

performance, studying the role of various thermo-mechanical parameters (channel

diameter, working fluid, input heat influx, evaporator temperature etc.), and exploring its

application areas etc. The next section discusses relevant experimental studies on OHPs

that are conducted in last two decades.

15
Figure 2.2 Oscillating heat pipe design as described by Akashi [32]

2.1 Experimental studies on OHP

OHP with a right combination of working liquid, channel diameter, and input heat

flux can attain a thermal conductivity up to 10,000 W/m-K [34]. The most intriguing fact

about the OHP is the right combination varies with the application and configuration.

Many experimental studies have been conducted to characterize the thermal performance

of the OHP, and it is now well accepted that its performance is sensitive to its design,

working fluid selection [35]–[38], frequency of oscillation [39], volume fill ratio, internal

flow pattern [40], [41], operating temperature [42], OHP aspect ratio [43], operating

orientation, and more [44]. The following sections discusses some of these experimental

studies in brief. Table 2.1 shown at end of the chapter compiles the experimental studies

discussed in this section and gives insightful information about OHPs used in respective

studies.

16
2.1.1 Influence of OHP internal diameter

Internal diameter of OHP is closely related to the successful operation of OHP. It

not only affects the nature of oscillation, but it also affects its thermal performance. Yang

et al. [45], [46] studied the role of internal diameter using OHP of 1 mm and 2 mm

internal diameter. The study found that the thermal resistance of the OHP decreases with

the increase in internal diameter. The thermal resistant of 2 mm internal diameter OHP

was found to be 10 % less than that of OHP with 1 mm internal diameter. Similar

findings were reported by. Charoensawan et al. [42], and the study also proved that Yang

et al. [45] observation is also valid for OHPs operating in horizontal orientation. The

study also investigated the coupled effect of evaporator temperature and OHP internal

diameter on its thermal performance. It manifested that the decrease in the internal

diameter significantly increases the thermal resistance of an OHP for all evaporator

temperature. However, Saha et al. [47] upon comparing the thermal performance of 0.9

mm and 1.5 mm internal diameter open ended OHP found that the former has better

thermal performance then the latter. In addition, the choice of working liquid also

influence the effect of internal diameter on OHP’s thermal performance. For example,

Charoensawan et al. [42] highlighted that the OHP of 1 mm internal diameter gives a

better thermal performance with ethanol as a working liquid then water. On the other

hand, OHP of 2 mm internal diameter yields better thermal conductance with water.

Holley & Faghri [48] studied a unique kind of OHPs in which, the channel

diameter is varied along the flow path. The study revealed that varying the channel

diameter makes the OHP less sensitive to gravity and improve its heat load capability by

30 %. Liu et al. [49] studied the effects of internal diameter of the OHP on its thermal

17
performance by probing three novel configurations as shown in Figure 2.3. They are -a)

OHP with uniform internal diameter of 1.6 mm, b) Internal diameter was varied in

adjacent tubes between 2 mm and 1.6 mm. c) Internal dimeter of left –side tubes was kept

at 1.6 mm and right side tubes was kept at 2.0 mm. They found that while configuration

(b) has least thermal resistance and configuration (c) completely prevent reversal of the

flow, both the configurations increase the likelihood of circulatory flow and improved

overall thermal performance.

Figure 2.3 OHP configurations investigated by Lui et al. [49]

(a) OHP with uniform diameter equal to 1.6 mm; (b) OHP with diameter 1.6 mm and 2
mm alternately (c) OHP with 1.6 mm diameter on left side and 2 mm on right side

From above discussion, it can be established that internal diameter of the OHP

significantly affects its thermal performance, but this influence is strongly interconnected

with other parameters such as operating condition, choice of working fluid and its fill

ratio etc. Hence, the impact of internal diameter on OHP’s thermal performance cannot

18
be generalized for all configurations and operating conditions. More research is needed to

reveal the coupling effect of other parameter with OHP’s internal diameter.

2.1.2 Influence of number of turns

Number of turns of the OHP directly impact its thermal performance and is

critical for instigating and prolonging the oscillations. Quan et al. [50] pointed out that

increasing the number of turns increases the perturbations inside the OHP. The rise in

perturbation improves the internal pressure distribution, which leads to a better thermal

performance. Several studies underpins the existence of critical number of turn of OHP

[50], [51], [40]. Khandekar and Groll [40] illustrated that increasing number of turns of

the OHP beyond a critical limit would result in reduction of pressure differential in the

OHP . Consequently, this would discount the driving potential of the OHP and causes

low thermal performance. Akashi [32] determined the critical number of turns to be 81

for the investigated OHP. Charoensawan el al. [52], [42] found that the critical number of

turn is subject to several local parameters such as fill ratio of working liquid, evaporator

temperature, and internal diameter of the OHP.

Number of turns of the OHP also affects the suitability of a fill ratio of a working

fluid. For example, Khandekar and Groll [40] demonstrated a single turn OHP filled with

R-123 with fill ratio less than 50 % frequently experiences a complete decay of

oscillations, whereas this phenomenon is rare with a multi-turn OHP used with same

working liquid and fill ratio. Besides, higher number of turns have proven to lessen the

effect of gravity on OHP’s thermal performance of OHP [52],[53]. Thompson el al. [28]

studied the thermal performance of a flat-plate OHP under high-gravity loading. The

study indicated that although, the OHP’s operation is highly influenced by gravity but
19
increasing the number of turns convert the OHP into a gravity independent device. The

study experimentally showed that the performance of OHP in an anti-gravity mode is

equivalent to that of gravity-assisted mode when the number of turns are very high.

2.1.3 Influence of working fluid and its filling ratio

Working fluid and its filling ratio is strongly linked to start up condition,

continuous operation, and the thermal performance of OHP. OHP is shown to be

operational with wide range of working fluid and its filling ratio. The filling ratio of a

typical operational OHP ranges from 20 % to 80 % [30]. The optimum filling ratio of

OHP that yields the best thermal performance varies with the working fluid along with

many other parameters such as OHP configuration, evaporator temperature, heat flux etc.

Charoensawan et al. [54] studied the influence of filling ratio on a horizontally orientated

OHP using water and ethanol as working liquid. The study found that the thermal

resistance of 150 mm long OHP is more susceptible to choice of working fluid than the

50 mm long OHP. The study also found the optimum fill ratio for 150 mm long OHP to

be as low as 30 %, whereas the 50 mm long OHP performed similarly with 30% and 50

% fill ratio.

Li & Yuan [55] underlined that the thermal resistance of the OHP increases with

increase in fill ratio and found an optimum fill ratio to be 50% with ethanol as working

fluid. Mahajan et al. [56] investigated a 64 cm long OHP for low-grade waste heat

recovery application. The study used n-pentene as working liquid and found the optimum

fill ratio to be 70 %. Im et al. [36] studied low temperature heat transfer characteristics (<

290K) of OHP filled with three refrigerants (R-22, R290, R-142b) and varied its fill ratio

20
between 40 % and 50 %. The study reported that among all the refrigerants tested R-22

with 40 % fill ratio has least thermal resistance.

2.1.4 Influence of Channel shape and its configuration

OHPs have been fabricated and tested with channels of many cross-sectional

shapes other than circular cross-section. Zhou et al. [57] examined OHPs of rectangular

and triangular cross sections under same operating condition and found out that the OHP

with rectangular cross-sectional channel is more thermally resistant than that of triangular

cross-section. Results also showed that the thermal performance of both type of OHPs is

highly influenced by gravity. Similar results where obtain by Xiahou et al. [58], who

probed flat-plate OHP with double sided rectangular and triangular channel cross-section

as shown in Figure 2.4. Aboutalebi et al. [59] explored the thermal performance of

rotating OHP with varying rotational speed. The OHP was rotated up to a maximum

speed of 800 rpm and the coupling effect of working liquid fill ratio with rotation was

observed. The study revealed that the centrifugal force generated by the rotation

facilitated in increasing the thermal efficiency of the system for all filling ratio. Also, the

optimum fill ratio for distill water charged OHP for all the rotational speed was found to

be 50 %.

21
Figure 2.4 OHP of double sided triangular (top) and rectangular (bottom) cross-
sectional channel shape [58]

Three dimensional OHPs have multi-layer or multiple rows interconnected to

each other. Borgmeyer et al. [60] conducted experimental and visualization study on the

two layer OHP using neutron radiography technique. Figure 2.5 shows 10 turn and 20

turn OHPs that are three dimensionally wrapped around two copper spreader that acts as

an evaporator and condenser section. The study found that thermal resistance of

investigated multi-layer OHP’s sharply decreases with the increase in heat load.

Hathaway et al. [61] experimentally investigated the gravity effect on a novel three

dimensional OHP with uneven turn. The OHP has 20 turn in evaporator section and 14

turn in condenser The study showed that under low power inputs, the OHP in upside

down configuration did not perform as well as it did in gravity assisted mode. But, with

higher power inputs, the performance in upside down configuration (anti-gravity) was

very similar to that of in gravity assisted mode.

22
Figure 2.5 Three dimensional OHP with 20 turn (a and b) and 10 turn (c) [60]

2.1.5 Influence of Heat flux input

Input heat flux is one of most critical operational parameter for the OHP. As

discussed in previous chapter, input heat flux directly affects the OHP thermal

performance and start-up condition. Several studies [27], [62]–[64] have resonated that

input heat flux directly affects the thermal resistance of the OHP. In addition, the input

heat flux also manipulates the impact of other operational parameter on OHP’s thermal

performance. For example, Thompson et al. [27] highlighted that higher input heat flux

reduces the OHP’s susceptibility to inclination angle. Researchers have also made the use

of several visualization techniques to delve deeper into the role of input heat flux [65],

[66]. Wilson et al. [66] visualize the oscillating motion of working fluid inside the OHP

using neutron radiography technique. They concluded that increase in the heat flux
23
increases the flow velocity of working liquid whereas at low heat flux the direction of

oscillation is random. Thompson and Ma [67] conducted an investigation on the effect of

localized heating on FP-OHP’s thermal performance. As shown in Figure 2.6, the FP-

OHP was tested under three different heating configurations. Heating configuration (a) is

called the cover-plate mating heating (31.75 mm x 63.50 mm) where the width of heater

matches that of the FP-OHP. Heating configuration (b) is called the OHP-array mating

heating (25.40 mm x 44.45 mm) where width and the length of heater is slightly smaller

than that of the FP-OHP, while heating configuration (c) is called the spot heating

(diameter = 11.28 mm). Results indicate that decreasing the heating area causes an

increase in thermal resistance and the amplitude of oscillation.

Figure 2.6 Three different heating configurations studied by Thompson and Ma [67]

(a) Cover-plate-mating heating, (b) Array-mating-heating, (c) Spot heating

24
Table 2.1 Characteristics of OHPs investigated in experimental studies mentioned above

Author Type Inclination Material Inner Number of Working fluid Charge Total length
angle diameter turns ratio
Yang et al. [45], [46] Closed, 90o Pyrex glass 1 mm & 10 R-123 50 % 120 mm
Circular 2mm
Charoensawan et al. Closed, 0o Copper 1 mm, 1.5 5,11,16,26 Water 30-80 % 150 mm,
[56] Circular mm, 2 mm Ethanol 450 mm
Saha et al. [47] Open, 0o, 90o Copper 0.9 mm and 6 Water, 50 % 105 mm
circular 1.5 mm Ethanol,
Acetone
Holley & Faghri Closed, 0o-90o- 180o Copper 1 mm - 2 mm 1 Water 40 % 250 mm,
[48] circular 750 mm
Liu et al. [49] Closed, 90o Pyrex glass 1.6 mm - 2 4 R-12 40 % 125 mm

25
circular mm
Quan et al. [50] Closed, 90o Copper 2 mm 4 Acetone 30 %, 48 250 mm
Triangle, %
Rectangle
Khandekar and Groll Closed 90o Copper/ 2 mm 1 Ethanol 20-80 % 190 mm
[42] glass
Thompson el al. [28] Multi-layer - Copper 1.175 mm 6 HPLC-grade 73 % 130 mm
flat plate, Water/
Closed Acetone
Jia et al. [57] Open, 90o Copper 2 mm 9 Water, Ethanol 50 % 120 mm
circular
Mahajan et al. [56] Closed, 90o Copper 1.58 mm 9 n-Pentane, iso- 60 %, 70 630 mm
circular Pentane %
Im et al. [36] Closed, 90o Copper 1.4 mm 140 R-22, R-142b, 40 % 770 mm
circular R-290
Table 2.1 (continued)

Aboutalebi et al. [59] Closed, - Copper 2 mm 4 Distilled water 25-75 % 500 mm


Rotating
Borgmeyer et al. [60] Multi-layer, 90o Copper 1.65 mm 10, 20 HPLC-grade 50 % 254 mm,
circular water 202 mm
Hathaway et al. [61] Multi-layer, 90o, 180o Copper 1.65 mm Uneven Water, Acetone 62 % 62 mm
circular
Wilson et al. [66] Open, 90o Copper 1.65 mm 6 Water, Acetone 45 % - 51 80 mm
Closed, %
circular
Thompson and Ma Multi-layer - Aluminum 0.76 mm 15 HPLC-grade 73 % 101 mm
[67] flat plate acetone
An inclination angle of 90o refers to bottom heating and gravity assisted mode, whereas 0o inclined OHP refers to horizontally
orientated OHP.

26
2.2 Theoretical studies on OHP

Since the OHP does not require internal wicking structure, its construction is

fairly uncomplicated, but the working principle is equally complex. The heat transfer and

fluid motion in the OHP is due to non-linear hydro-thermodynamic coupling. Several

studies have attempted to model it analytically by applying various simplification

schemes. Some of the most widely adopted simplification schemes can be summarized as

following.

(a) Considering distinct adjacent liquid plugs and vapor bubbles as a spring-mass

damper system and performing kinematic and dynamical analysis (Zuo et al.

[68], Peng et al. [69], Peng et al. [70]).

(b) Considering a control volume encompassing adjacent liquid plug and vapor

bubble. Then applying governing equations of thermo-fluid dynamics and

solving it numerically (Shao and Zhang [71], Dilawar et al. [72], Sandia

National Laboratory [24], Yuan et al. [37], Arabrejad et al. [73], Zhang et al.

[74]).

(c) Simulating OHP as a conventional heat pipe and applying effectiveness-NTU

method to determine its effectiveness (Khandekar et al. [29], Mahajan et al.

[56])

Theoretical studies on OHP are primarily focused areas such as predicting the

necessary condition required for the start of oscillatory motion [75], understanding the

characteristics of oscillations, factors influencing the behavior of the oscillatory motion

[76], [77], performing non-linear analysis on temperature oscillations of the OHP [78],

[79]. Following sections discuss some of these studies in brief.


27
2.2.1 Start-up condition of OHP

Qu and Ma [75] performed a theoretical analysis to predict the start-up condition

of OHP. Based upon experimental visualization, the model classified vapor bubbles into

two types: long column (Taylor) bubble, spherical (Globe) bubble. The model proposed

Equation (2.1) to estimate the start-up heat flux required to start oscillation in an OHP.

The study found that start-up heat flux requirement can be reduced by increasing the

surface roughness of the inner surface. Furthermore, the type of bubble prevailing in the

OHP also affects the start-up performance of OHP. The globe-type vapor bubble showed

better start-up performance than Taylor-type bubble.

𝑘𝑙 𝑇𝑣 𝐴c,h 2 𝑅𝑇 2𝜎
𝑄̇ ≈ (1⁄{1 − ℎ 𝑣 ln [1 + 𝜌 ]} − 1) (2.1)
𝑟𝑖 ln[𝑟𝑖 ⁄(𝑟𝑖 −𝛿𝑙 )] 𝑓𝑔 𝑣 𝑅𝐺 𝑇𝑣 𝑟𝑛

where 𝑘𝑙 is the thermal conductivity of working liquid, 𝐴c,h is the area of condenser and

evaporator surface, 𝑟𝑖 is the radius of inner wall, 𝛿𝑙 is the liquid film thickness between

capillary wall and Taylor bubble, R is the universal gas constant, ℎ𝑓𝑔 is the specific

enthalpy of evaporation of working liquid, 𝑟𝑛 is radius of perfect spherical vapor bubble

trapped in the cavity and 𝑇𝑣 is its corresponding temperature, 𝑘𝑙 is the thermal

conductivity of working liquid, 𝜌𝑣 is the density of working fluid in vapor form, 𝑅𝐺 is the

radius of globe bubble, 𝜎 is the surface tension of working liquid

2.2.2 Studies predicting hydro-thermodynamic behavior of OHP

Cheng and Ma [69] considered the OHP as a mass - spring mechanical system and

developed Equation (2.2) to simulate the occurrence of oscillation in the OHP. The

equation factors several parameters such as filling ratio, operation temperature, gravity

etc. The model assumes the OHP’s internal fluid motion as one-dimensional and

28
simulates the train of liquid plug and vapor bubble as a linear spring-mass system as

shown in Figure 2.7. The model also assumes equal number of liquid plug and vapor

bubbles in the OHP and treated vapor plug as an ideal gas. The model suggested that

there exist a peak value of the sensible heat ratio for an OHP operating under given input

power. This peak value depicts OHP’s balanced state and estimation of thermal

performance based upon the peak value yields better predictions.

Figure 2.7 Schematic of one-dimensional OHP (Top) and mass spring mechanical
system mimicking liquid plug and vapor plug (Bottom).

(Chang & Ma 2002)


In the schematic of one dimensional OHP, xi represents the displacement of ith liquid plug
and Li and ρi as its length and density.

𝑑2 𝑥𝑛 𝜇𝑙 𝐿𝑙,𝑛 𝑑𝑥𝑛 1 1
𝜌𝑙 𝐿𝑙,𝑛 + (𝑓𝑅𝑒𝑙 ) + 𝑅𝜌𝑣0 𝑇0 ( − )𝑥 −
𝑑𝑡 2 2𝐷ℎ2 𝑑𝑡 𝐿𝑛−1,0 𝐿𝑛,0 𝑛

𝑥 𝑥 𝑅 𝑞 𝑞 𝛿𝑊𝑛−1→𝑛 𝛿𝑊𝑛→𝑛
𝑅𝜌𝑣0 𝑇0 (𝐿 𝑛−1 − 𝐿 𝑛 ) = 𝐶 {(𝐿 𝑛−1 − 𝐿 𝑛 ) 𝛿𝑡 + ( − )} (2.2)
𝑛−1,0 𝑛,0 𝑝𝐴 𝑛−1,0 𝑛,0 𝐿𝑛−1,0 𝐿𝑛,0

29
Later, Pang et al. [70] introduced a more comprehensive, fully non-linear thermo-

mechanical finite-element model to study the effect of multitude parameters such as fill

ratio, property of working liquid, difference between evaporator and condenser, internal

diameter, orientation etc. In contrast to Cheng and Ma [69] model, the spring-mass

system mimicking the liquid-vapor slug plug system was considered to be non-linear.

Assuming that there exists a linear relationship between fluid slug’s displacement and

vapor pressure, the study conducted a linear modal analysis to study the modal shapes

and natural frequency of the system. The study concluded that the oscillating frequency is

mainly a function of liquid plug mass and ratio of vapor-plug length and channel cross-

sectional area.

Zuo et al. [68] presented a simplified hydrodynamic model by mimicking the

OHP’s liquid-vapor slug plug system as a spring-mass damper system. The governing

principle of liquid-vapor movement was stated by Equation (2.3), a 2nd order

homogenous differential equation similar to that of mechanical vibration with viscous

damping. The spring constant term was considered a space and time dependent quantity,

which led the model to suggest ever-increasing spring-stiffness coefficient without any

decrease in amplitude. This refutes the experimental results presented in the study and

hence, the model has limited applicability.

𝑑2 𝑥 8𝜇𝑙 𝑃𝜙𝑜 𝑑𝑥 2𝐴2 𝑅𝑇𝑠𝑎𝑡 𝐿𝐴𝜌𝑙 (1−𝜙𝑜 ) 𝑄


2 +( ) +
𝜌𝑙 𝐷𝐴 𝑑𝑡 𝐿𝐴𝜌 (1−𝜙 ) 2
(
2
+ ℎ 𝑒 𝑡) 𝑥 = 0 (2.3)
𝑑𝑡 𝑓𝑔
𝐿𝐴𝜌𝑙 𝜙𝑜 { 𝑙2𝜌 𝑜 }
𝑣

Shao and Zhang [71] proposed a mathematical model to predict the heat transfer

characteristics of OHP. The model applied conservation of momentum and energy

principles to a control volume that has liquid plug sandwiched between 2 vapor plugs.

30
The model takes into account the axial variation in surface temperature, initial

temperature, and pressure lose in the U-turns. Numerical schemes were used to solve the

hydrodynamic governing equations, and the model was successful in proving that the

significant portion of heat transfer through the OHP is sensible heat transfer. However,

the model assumes a constant temperature of vapor plugs, which causes unrealistic

predictions in thermal performance. This assumption was taken care by Dilawar et al.

[72]. The study considered the temperature variation in vapor plug, and the temperature

of interface (surface between vapor plug and liquid plug) was assumed to be equal to

saturation temperature. The model proved that the pressure lose in the U-turns negatively

affects the amplitude and frequency of oscillation in the OHP, thereby reduces the heat

transfer rate of the OHP. Givler et al. [24] unlike above mentioned complex models, used

a homogenous bubble model to simplify the computation. Apart from solving the

simplified model similar to that of Shao and Zhang, [24] also modeled OHPs of multiple

configurations ranging from one turn OHP to four turns OHP. It also introduced a scaling

coefficient that controls the rate of mass exchange between liquid and vapor phase.

Khandekar et al. [80] proposed an analytical model to predict the heat transfer

effectiveness of the OHP. The heat transfer through the OHP was approximated using the

effectiveness-NTU method. Since the thermal resistance of the OHP

evaporator/condenser is much lower than that of the external convection, the OHP was

idealized as a device exchanging heat with surroundings via condensation and

evaporation of working fluid exclusively. The study found that the model prediction is in

closed agreement with the experimental results and concluded that the NTU-effectiveness

31
method can be applied for estimating thermal performance of the OHP with a reasonable

accuracy.

2.3 Application of OHP

OHP’s ability to operate passively in varied orientation[28], and its ease of

construction [43] has made it an attractive heat transfer device in the last two decades for

a variety of applications [44]. It has been widely investigated for diverse heat transfer

applications such as cooling electronic devices [81]–[83], fuel cell stack[84],

superconducting magnets [85], [86] etc. Furthermore, OHP’s have been employed in

application such as solar water heater collector [87]–[92], pumping water [93], extended

surface like fins [94], thermal management for hybrid vehicle [95], waste heat recovery

[96]–[98], and even power generation [99]. The next section discusses some of these

studies in brief.

2.3.1 Application as an electronic cooling device

Over the past decade, many studies have attempted to utilize OHP’s high heat

transport capability in cooling electronic devices. Maydanik et al. [81] developed a

compact heat exchanger for cooling electronic devices using a three dimensional OHP.

As shown in Figure 2.8, the OHP is wrapped around a computer’s central processing unit

(CPU) fan with impeller diameter of 92 mm. The evaporator section of OHP, located in

the center, is soldered to a copper plate measuring 40 x 35 x 0.5 mm. Plate type fins were

added to the condenser, which is located on the opposite side of fan. Results shows that

with a constant airflow rate of 32 cfm, the device was able to dissipate up to 250 W of

heat.

32
Figure 2.8 OHP based compact cooler for electronic devices [81]

2.3.2 Application in solar water heater

Solar water heaters are often faced with the issue of large space requirement for

installation. Arab et al. [87] successfully compacted a solar water heater by integrating it

with two OHPs. The evaporator section and the condenser section of each OHP was

embedded into the collector and the water tank respectively. The study reported that the

efficiency of the OHP based solar water heater to be as high as 54%. Rittidech et al. [88]

engineered a solar water heater in which the collector tube was made by OHP as shown

in Figure 2.9. The configuration of evaporator and condenser sections was similar to that

of [87]. The study found that that OHP based solar water heater efficiency is comparable

to the CHP based solar water heater, i.e. around 62 %, but it also offers corrosion free

33
operation and eliminate winter icing problems. In subsequent study, Rittidech et al. [91]

inserted check valves in the OHP and was able to increase its efficiency by 14%.

Figure 2.9 OHP based solar water heater [88]

(a) Closed loop OHP flat plate solar collector (evaporator) ; (b) water tank (condenser);
(c) wooden assembly

2.3.3 Application in waste heat recovery device

As discussed above, the majority of OHP applications has employed constant heat

flux or constant-temperature evaporator conditions, to obtain a high level of heat transfer

control and to simulate electronics-cooling applications. Very few studies have focused

on using heating/cooling configurations reminiscent of heat recovery in a HVAC

environment. Supirattanakul et al. [96], [100] reported a 14.9% increase in the coefficient

of performance (COP) of a split type AC system by embedding an OHP (with check

34
valves) as shown in Figure 2.11. The OHP has 56 turn, 220 mm long condenser and

evaporate or section, and an adiabatic section of length 190 mm.

Figure 2.10 Schematics (left) and illustration (right) of closed loop OHP used by
Supirattanakul et al. [96]

(1) Outdoor air, (2) return air, (3) pre-cooled air, (4) cooling coil, (5) cooled and
dehumidified air, (6) reheated air, (7) evaporating section of heat pipe, (8) condensing
section of heat pipe, (9) insulator, and (10) check valve.

Khandekar and Gupta [100] integrated a 150 mm long OHP in a space radiator

system and concluded that it is advantageous only in conditions where the base plate is

made of material with low thermal conductivity. Rittidech et al. [97] constructed an

OHP-based air preheater for waste heat recovery in a batch-type drying process. The

OHP preheater consisted of 32 rows of 8 turn copper-made OHPs. The OHP preheater

was shown to be capable of achieving an effectiveness of 0.52 when R123 was used as

the working fluid at a 50% fill ratio. Meena et al. [98] demonstrated the performance

enhancement in the cooling coil of a drying system with the application of OHPs with

35
check valves. The OHP used in the study has a footprint of 0.4 m2 and was made up of

copper tube of 2 mm inner diameter. The study reported an increase in effectiveness from

54% to 72 % upon increasing the hot-air temperature from 50 oC to 70 oC.

Figure 2.11 Schematic of closed loop OHP (CEOHP) used as a waste heat recovery
device in a batch-type drying process [97]

36
CHAPTER III

EXPERIMENTAL SETUP AND PROCEDURE

3.1 Experimental setup

Figure 3.1 shows the test section used in all the experimental investigations

presented in this dissertation. The ‘waste’ exhaust hot stream and cold/fresh incoming air

stream were simulated using two cross-flow air ducts equipped with air heaters (Brasch

Electroduct AG2002-001075 and TUTCO Heat Pack/81-0578-00) with combined heating

capacity of 5 (± 0.05) kW. Both the air ducts were equipped with two centrifugal fan

units (Continental/AXC150B-ES) located upstream and downstream from the OHP. Each

air duct consisted of 15.24 cm (6 in.) diameter circular aluminum ducts attached with

custom-made diffusers/nozzles to provide for HVAC-typical exit plane dimensions of

45.71 cm x 60.1 cm (18 in x 25 in). A 61 cm (24 in.) deep test section enclosing the OHP

is sandwiched between two diffusers having walls made up of 2.54 cm (1 in.) thick

polystyrene foam board (R-4.35). Fiber blanket insulation (R-19) was added around the

ducts and diffuser/nozzle to reduce inter-stream heat exchange. Total heat loss from the

hot stream during experimentation was estimated to be approximately 5% that of OHP

heat transfer. An operating environment typical of a US South-East winter was simulated

for present investigations. The cold air stream was taken at an ambient temperature of 10

(± 1) oC in the topmost air duct and the hot air stream was taken at a room temperature of

37
21 (± 1) oC and heated to 45 (± 1) oC in bottom air duct. Both air streams had a

volumetric flow rate of 0.19 (±0.01) m3/s.

As a part of the dissertation, a 127 cm (50 in) long finned OHP was

experimentally investigated under above mentioned operating condition. Its results are

discussed in Chapter VI. To accommodate the finned OHP, minor modifications were

made in the experimental setup. The diffusers were rotated by one quarter of turn, thereby

making it of dimension equal to 60.1 cm x 45.7 cm (25 in x 18 in.). The modification was

accompanied by lowering the position of the bottom air duct, which in turn resulted in an

additional head loses in the hot air stream. Volumetric flow rate of the hot air stream was

reduced to 0.13 (±0.01) m3/s due to the modification.

Figure 3.1 Test apparatus showing two air ducts with other major components.

38
3.2 Common Peripheral Devices

To measure the temperature of oscillating fluid in a Eulerian frame of references ,

six type-T thermocouples (Omega/5SRTC-GG-T-20-72) were used, three of which were

taped on the three center-most tubes along the OHP evaporator and other three similarly

positioned on the OHP condenser. Additional type-T thermocouples were placed in the

air duct to measure the temperature in the air streams. Pressure sensors (Omega/PX653-

0.255V) were used to measure the pressure changes in the two air streams and thereby its

volumetric flow rate was estimated. These sensors were placed in the middle of the air

streams and were connected to a data acquisition (DAQ) system (National Instruments

TC-2095/SCXI-1100). Figure 3.2 shows the approximate locations of the thermocouples

and pressure sensors

Figure 3.2 Approximate location of thermocouples (Ti) and pressure sensors (Pi) along
the air ducts (side view)

39
All of the thermocouple and pressure sensor connections were inspected prior to

testing. A portable calibrator (Omega/CL3515R) was used to calibrate all thermocouples

and the DAQ system; an offset, if found, was added for data correction. Air prevailing

inside air ducts was forced to exit the system by starting the air circulation 15 minutes

prior to start of experiment. Signal Express 2013 (National Instruments) DAQ software

was used to record measurements during experimentation. Temperature measurements

were recorded at a frequency of 10 Hz and sampling rate of 100 per second. During OHP

testing, sufficient time was allowed to pass until the OHP temperature oscillations

reached a steady mean temperature. Figure 3.1 provides additional information about

experimental apparatus.

Table 3.1 Additional information about the experimental apparatus and the OHPs
investigated in the present work

Material of OHP tube Copper Alloy 122


Outer diameter of OHP tube 3.18 mm
Inner diameter of OHP tube 1.65 mm
Area of evaporator section 0.29 m2 (0.45 x 0.63 m2)
Area of condenser section 0.29 m2 (0.45 x 0.63 m2)
Area between evaporator and condenser 0.016 m2 (0.025 x 0.63 m2)
Air flow rate 0.2 kg/s

3.3 Uncertainty Analysis

The uncertainty related to the measured OHP heat recovery rate was found using

linear uncertainty propagation (i.e., the first order Taylor series expansion approximation)

[101]–[103]. The data reduction equation, relating the heat recovery rate (qexp) with

measured variables during experimentation, Xi, is shown in Equation (3.1) and the
40
uncertainty in qexp was estimated using Equation (3.2). Note that mass flow rate of cold

air stream, 𝑚̇c , and temperature difference in cold air stream across the OHP, Δ𝑇c , are the

measured variables.

𝑞exp = 𝑚̇c 𝑐pc 𝛥𝑇c (3.1)

2 2 2
𝜕𝑞 𝜕𝑞 2 𝜕𝑞
𝑈q2 = (𝜕𝑚 ) 𝑈mc
2 2
+ (𝜕𝑐 ) 𝑈cp + (𝜕ΔT) 𝑈ΔTc (3.2)
c pc

In Equation (3.2), Uq is the total systematic uncertainty associated with qexp and

𝜕𝑞
are the absolute sensitivity coefficients of Xi. In this analysis, only the systematic type
𝜕𝑋i

of uncertainty was considered. The systematic uncertainty in estimating q can arise from,

for example: (1) air pressure and temperature measurements from sensors, (2) assuming

average air specific heat capacities in Equation (3.1) and, (3.2) assuming constant air

density in estimating mass flow rate.

The estimation of uncertainties arising from air pressure and temperature

measurements involves determination of data acquisition error1, spatial temperature non-

uniformity error, and calibration error. Uncertainty due to the assumption of average

specific heat capacity is negligible in comparison to uncertainty arising from other

components. Table 3.2 shows the estimated uncertainty value for each input variable

Table 3.2 Estimated uncertainty for input variables

Input variables Estimated uncertainty


Mass flow rate, 𝑚̇c 0.007 kg/s
Specific heat, 𝑐pc 0 kJ/kg·K
Temperature difference across the 0.283 oC
OHP in cold air strem, 𝛥𝑇c

1
The data acquisition error is calculated using the guidelines and accuracy of sub-components stated by
National Instrument- http://digital.ni.com/public.nsf/allkb/8BA2242D4BCC41B286256D1D00815B90

41
CHAPTER IV

FEASIBILITY STUDY OF OHP AS AN WASTE HEAT RECOVERY DEVICE

The first phase of the development of OHP as a waste heat recovery device is

presented in this chapter. An experimental feasibility study was conducted to assess a

nine turn copper-made tubular OHP for passive waste heat recovery via air-to-air heat

exchange. The configuration is similar to a typical HVAC ducting system with adjacent

air streams at different temperatures. The experiment was designed to demonstrate the

OHP’s ability to utilize otherwise wasted thermal energy to pre-heat or pre-cool air in

order to reduce building energy consumption. Additionally, the OHP’s thermal resistance

and other heat transfer characteristics were compared to an empty/evacuated OHP with

same overall dimensions. The OHP aerodynamic performance, in terms of pressure drop,

was evaluated and juxtaposed with the heat transfer characteristics. The OHP, existing as

a serpentine-arranged capillary tube, possesses a desirable aerodynamic form factor and

provides relatively high heat transfer rates via cyclic evaporation and condensation of an

encapsulated working fluid with no internal wicking structure required.

4.1 OHP Fabrication

A unique, large form factor OHP was fabricated for the present feasibility study.

A copper capillary tube (i.e., inner diameter 𝐷i = 0.165 cm (0.062 in) and outer diameter

𝐷o = 0.317 cm (0.125 in.)) was used for constructing a 9-turn OHP with tube-sections

approximately separated by a distance of 22 mm (1in.) (nominal). Since the current OHP


42
is designed for implementation in HVAC systems, its geometric footprint of 60 x 45 cm2

was chosen for its direct integration within typical-sized air ducts (each of size 63 x 51

cm2). Figure 4.1 provides a schematic of the OHP-to-duct integration and a photograph of

the copper OHP investigated in this study. The evaporator and condenser region in the

test section was separated by two layer of 2.54 cm (1 in.) thick polystyrene foam board

with R-value = 4.35. The OHP passes through the holes in foam board and fiber blanket

insulation with R value = 19 was added in the space between the inner surface of holes

and walls of tube to minimize the direct heat transfer between two air streams.

4.2 Working fluid selection

A novel working fluid with properties unique for low-grade heat flux conditions,

i.e. n-pentane, has been investigated for characterizing the OHP’s adaptability to waste

heat recovery applications. This specific working fluid was selected for its relatively low

boiling point, low dynamic viscosity and latent heat of evaporation. In addition, n-

pentane is non-toxic [104] and non-ozone-depleting [105] making it a feasible candidate

for HVAC application. Also, the selection of working liquid is strongly influenced by

inner diameter of OHP and operating temperature, as required by Equation (1.1). The

Bond number for the presented case was found to be ranging between 0.9 and 0.98. This

implies that the selected working fluid operating in the present OHP under the given

temperature difference will be able to persist in form of distinct vapor and liquid plug.

The OHP was first evacuated of air using a dual–seal vacuum pump (Welch

Scientific Co., Model 1400) to an approximate vacuum pressure of 6.67 (±0.6) Pa (50 ±5

mTorr). After evacuation, the OHP was filled with n-pentane (99%, reagent grade, ACS)

43
at a liquid-to-OHP volume ratio (fill ratio) of either 0 (+2) % (empty), 60 (±2) % or 70

(±2) %.

Figure 4.1 (Left) Schematic of OHP, its dimension and position relative to cold air
stream and hot air stream. (Right) Photograph of OHP under-investigation.

4.3 Mathematical Model

A mathematical model was developed to aid in quantifying the thermal

performance of the OHP-HE. The heat transfer through the OHP was approximated using

the effectiveness-NTU method [106], [107]. Since the thermal resistance of the OHP

evaporator/condenser is much lower than that of the external convection, the OHP was

idealized as a device exchanging heat with surroundings via condensation and

evaporation of working fluid exclusively (i.e. no sensible heating/cooling of working

fluid). All fluid properties were evaluated at a surface film temperature, 𝑇f , shown in

44
Equation (4.1). Major thermophysical properties of the hot air stream, for the investigated

conditions, are shown in Table 4.1

(𝑇air,i +𝑇air,o )⁄2+𝑇̅sc


𝑇f = 2
(4.1)

where for either the hot or cold stream, 𝑇air,i is the temperature of air at the inlet, 𝑇air,o is

the temperature of air at the outlet and 𝑇̅sc is the surface temperature of centermost tube

of the OHP averaged over time.

Table 4.1 Thermo-physical properties of air at film temperature [108]

Density, 𝜌ℎ 1.143 kg/m3


Prandtl number, Prh 0.711
Thermal conductivity, kh 26.8 mW/m·K
Dynamic viscosity, µℎ 185 x 10-7 N·s/m2
Isobaric specific heat capacity, cph 1.006 kJ/kg·K

The air flow over the OHP was modeled as single-phase, cross flow over

cylinders with the mass flow rate (𝑚̇ h ) and the Reynolds number (Reh ) of the hot air

stream calculated using Equations (4.2) and (4.3)

𝑚̇h = 𝐴diff 𝑣diff 𝜌h (4.2)

𝜌h 𝑣diff 𝐷h
Reh = (4.3)
𝜇ℎ

where 𝐴diff is the area of the diffuser exit plane, 𝑣diff is the velocity of air stream at the

diffuser exit plane, and 𝐷h is the hydraulic diameter of the diffuser exit plane. The

Nusselt number (Nuh ) and average heat transfer coefficient (ℎ̅h ) were determined using

Equations. (4.4) and (4.5) [109]. For a unit row of OHP under investigation, the heat

transfer coefficient is approximately the same as that of flow over a cylinder in a cross

flow [110], [111].


45
4/5
1/2
Prh 1/3 5/8
̅̅̅̅h = [0.3 + (0.62Reh
Nu h
[1 + 282000]
Re
)] (4.4)
2/3 1/4
0.4
[1+( ) ]
Prh

̅̅̅̅
Nuh kh
ℎ̅h = 𝐷o
(4.5)

where Prh is the Prandlt number of the hot air stream and k h is the thermal conductivity

of the hot air stream (mW/m∙K).

The number of transfer units (NTUh ), and evaporator-side effectiveness (𝜀h ) were

estimated using Equations (4.6) and (4.7), respectively.


̅h
Aohp,h ℎ
NTUh ≅ (4.6)
𝐶ph 𝑚̇h

𝜀h ≅ 1 − e−NTUh (4.7)

where 𝐴ohp,h is the heat transfer surface area of the OHP in the hot stream and 𝑐ph is the

specific heat of the hot air stream. The heat transfer from the OHP condenser was

calculated using a similar procedure outlined by Equations (4.1) – (4.7). Equations (4.8) -

(4.11) were used to estimate the overall theoretical effectiveness of the OHP, 𝜀th i.e.

𝐶c ≅ 𝑚̇ c 𝑐pc (4.8)

𝐶h ≅ 𝑚̇ h 𝑐ph (4.9)

min(𝑐ph ,𝑐pc )
𝐶≅ (4.10)
max(𝑐ph ,𝑐pc )

1 C −1
𝜀th = [min (𝜀 + max (𝜀 ] (4.11)
h ,𝜀c ) h ,𝜀c )

where 𝑐ph and 𝐶h are the specific heat and heat capacity rate of the hot air stream,

respectively, 𝑐pc and 𝐶c are the specific heat and heat capacity rate of the cold air stream,

respectively, 𝑚̇h is the mass flow rate of the hot air stream, 𝑚̇ c is the mass flow rate of

46
cold air stream, 𝜀h is the effectiveness of heat transfer between the OHP and hot air

stream and 𝜀c is the effectiveness of heat transfer between the OHP and cold air stream.

The maximum heat transfer attainable by the OHP-HE, 𝑞max was found using

Equations. (4.12). Equation (4.13) was used to estimate the theoretical actual heat

transfer, or waste heat recovery rate, 𝑞th to the cold air stream through the OHP.

𝑞max = 𝐶min (𝑇hi − 𝑇ci ) (4.12)

𝑞th = 𝜀th 𝑞max (4.13)

where, 𝑇hi is the average temperature of hot inlet air stream, 𝑇ci is the average

temperature of cold inlet air stream and 𝐶min is the minimum heat capacity rate between

the hot and cold air streams.

4.4 Results & Discussion

Theoretical and experimental results are now presented. The effectiveness,

effective thermal resistance, and heat recovery rate of the OHP for the various

investigated fill ratios (FRs) of n-pentane (i.e., 0%, 60% and 70%) are summarized in

Table 4.2. The pressure drop across the OHP and nozzle/diffuser, for the investigated

flow rate of 350 (± 20) CFM, was found to be approximately 62 (± 3) Pa. This pressure

drop value also includes the losses related to the flow in diffuser-duct configuration

Table 4.2 Time averaged, steady state, thermal performance metric of OHP with
varied fill ratio of n-pentane

Thermal performance FR = 70% FR = 60% FR = 0 %


OHP-HE effectiveness 0.05 0.04 0.02
Effective thermal resistance (ºC/W) 0.11 0.13 0.25
Heat recovery rate (W) 225 170 110

47
As shown in Table 4.2, the OHP performed best with a FR = 70% of n-pentane.

This can be attributed to the higher FR providing the OHP a steady-state fluid distribution

with more liquid present, on average, in the evaporator during operation. This allows for

more consistent OHP operation with fewer vapor pressure balances along the turns –

which can result in degradation of OHP heat transfer due to localized or global cessation

of internal fluid motion. With higher FRs, the probability for liquid volume to

prevent/alleviate pressure balancing is generally higher. With excessively-high FRs, i.e. >

0.85, in which the evaporator becomes overloaded with liquid, fluid oscillation within the

OHP is reduced [112] and the heat transfer is representative of that found in a capillary-

scale thermosyphon. Since it was observed that the OHP filled with n-pentane performed

better at a FR = 70%, many of the presented results are for this experimental condition

only.

As shown in Table 4.3Table 4.3, the effectiveness of the n-pentane OHP-HE was

measured to be twice that of the empty capillary tube (FR = 0%) – indicating the heat

transfer enhancement capability of the OHP. Relative to more proven HPHEs [20], [113],

[114], the OHP-HE effectiveness presented herein is low; however, the investigated

prototype was designed for proof-of-concept and was not necessarily optimized for final

application. The current OHP-HE consists of only a single, un-finned tube. A more

optimized design, with higher effectiveness, can be achieved by employing multiple

finned OHPs to form an array. In addition, the OHP wetted frontal area, or footprint, was

approximately half of the available diffuser cross-sectional area. Although this is

beneficial for reducing the OHP-HE pressure drop, a significant portion of waste heat

was ‘un-recovered’. The theoretical effectiveness of the OHP-HE at the evaporator side,

48
εh, was found to be 0.12; implying that ~12% of available heat is being gained by the

OHP tube structure within the given hot stream. Finned surfaces should increase this

value.

As shown in Table 4.3, the effective thermal resistance of the FR = 70% n-

pentane OHP-HE was found to 0.11 ºC/W during its operation and this is approximately

60% less than the empty OHP-HE – which is 0.25 ºC/W. This indicates that the sensible

and latent (i.e. two-phase) heat transfer provided by the OHP exceeds that provided by

more-traditional, single-phase heat exchangers.

Figure 4.2 provides the temperature response of n-pentane OHP with FR = 70%,

where Tin is the temperature of inlet hot air stream, 𝑇̅ci,avg is the average OHP condenser

temperature and 𝑇̅hi,avg is the average evaporator temperature. Various operating regimes

of the OHP were observed and are labeled. Note that all other thermocouples located

along either the condenser or evaporator demonstrated similar temperature responses. As

shown in Figure 4.2, the working fluid in OHP-HE oscillated with respect to time after a

critical heat flux was achieved and this indicates its ability to function via forced

convection with air and HVAC-type heating/cooling boundary conditions. The results

demonstrate that when the presented OHP-HE achieves a minimum start-up heat flux, the

internally filled n-pentane starts to pulsate at a relatively low frequency – indicating the

presence of cyclic evaporation and condensation. The consistent temperature oscillations

with respect to time indicate reliable OHP operation. Marked temperature drops in the

evaporator were observed to correspond to temperature increases within the condenser.

The OHP-HE heat recovery rate from the hot air duct to the cold air duct was found to

vary with time, but ranged between 200-250 W.

49
The benefit of two-phase heat transfer provided by the OHP, as evidenced by its

relatively low effective thermal resistance, is realized upon its excitation or start-up

which occurs when a critical heat flux is achieved through the OHP tube/fluid structure.

This critical heat flux is a function of the working fluid’s thermophysical properties (i.e.

latent heat of vaporization, specific heat capacity) and the surface roughness of the

OHP’s internal surface [75]. Hence, the OHP design and working fluid selection should

be carefully considered for ensuring that this critical heat flux is achievable for a given

HVAC application. Table 4.3 summarizes the start-up conditions measured for the

investigated OHP-HE while filled with n-pentene at FR = 70%.

Table 4.3 Experimental condition at the start-up of OHP with FR = 70 % n-Pentane

Time elapsed for the OHP-HE to start-up ~500 s

Temperature of hot air stream at start-up 38.2 ± 0.28 oC

Room temperature at start-up 21 ± 0.28 oC

Temperature difference across the OHP-HE at start-up 30 ± 0.28 oC

As shown in Figure 4.2, the OHP displays three different operating regimes (or

regions): AB, BC, and CD - with each time frame possessing distinct temperature

oscillations, and thus, internal fluidic behavior. Regime AB corresponds to heat transfer

via conduction only. No temperature oscillations were observed during this time. Point B

is the time elapsed for a critical heat flux to be established through the OHP-HE for

accomplishing ‘start-up’. It can be inferred that Region BC is dominated by bubble flow

and Region CD corresponds to a fully-developed slug flow condition [29]. The sudden
50
jump of heat transfer at point B provides evidence that the OHP-HE does facilitate the

heat transfer between the two air streams.

Figure 4.2 Temperature response of OHP-HE with FR = 70% n-Pentane with


operating regime AB, BC, and CD indicated via shaded zones.

Figure 4.3 provides the heat recovery rate (heat transfer from hot to cold air

stream) by the OHP-HE (n-pentane FR = 70%). The heat transfer is oscillatory due to the

cyclic phase-change heat transfer within the OHP. Heat exchange to the cold air stream

from the OHP-HE is less oscillatory due to convection and thermal capacitance effects.

The average heat recovery rate was approximately 225 W, with a peak heat rate of 250

W; the heat transfer was observed to increase with respect to time and then plateauing

after approximately 5000 s. The OHP heat transfer was found to vary with operating

regime – being lowest prior to its start-up and highest during region CD. Sensor noise
51
was found to contribute to the observed oscillatory heat recovery rate; hence, a Savitzky-

Golay filter [115] (i.e., a generalized moving average smoothing technique) was applied

for visualizing its trend. The filtered heat transfer rate, as shown in Figure 4.3, clearly

indicates a continual increase in heat addition to the cold stream, which implies that the

oscillation of working fluid in OHP-HE aids in decreasing the overall thermal resistance

of the copper tube.

Figure 4.3 OHP-HE with FR = 70 % n-Pentane heat recovery rate vs time with
operating regime AB, BC, and CD indicated via shaded zones

Figure 4.4 compares the theoretically estimated maximum heat transfer rate with

the actual heat transfer rate through the OHP-HE with and without the working fluid (FR

= 70% n-pentane) during steady-state operation. It may be observed that the OHP-HE

heat transfer, while filled with working fluid, is more than twice the heat transfer

achievable when the OHP-HE is empty. Moreover, Figure 4.4 also demonstrates that the
52
theoretically estimated maximum heat transfer rate lies within the range of uncertainty of

the experimentally measured heat transfer rate. Hence, it justifies the assumptions made

in mathematical modeling section of idealizing the OHP as a CHP for the ε-NTU

analysis. From Figure 4.4, it may be seen that the OHP, while filled with n-pentane can

achieve high heat transfer effectiveness.

Figure 4.4 Comparison of measured, steady state heat transfer through OHP filled
with FR 70 % n –pentane, heat transfer through empty OHP, and the
analytically-estimated maximum heat transfer possible

53
CHAPTER V

ENERGY AND COST SAVING POTENTIAL OF OSCILLATING HEAT PIPE

BASED HEAT RECOVERY VENTILATOR

This chapter discusses the energy and cost saving potential of oscillating heat pipe

based heat recovery ventilator (OHP-HRV). Heat recovery ventilators (HRVs) are air-to-

air heat exchangers that perform sensible waste heat recovery in residential, commercial,

and industrial applications [15]. They pre-condition building supply air by utilizing

otherwise wasted temperature gradients between air supply and exhaust. The current

investigation focuses on modeling and predicting the heat transfer and aerodynamic

performance of the OHP-HRV, while also considering potential energy and cost savings.

The evaporation and condensation heat transfer within the OHP-HRV is modeled and the

effect of working fluid on OHP-HRV thermal performance is demonstrated. The energy

and cost savings analysis is performed for a specifically-designed OHP-HRV system

operating in typical HVAC environments. The potential cost benefits are demonstrated

for various geographical regions within the United States, in which feasible operating

climates are elucidated.

5.1 OHP-HRV design concept and operating environment

Figure 5.1 shows the physical characteristics of a unit row OHP-HRV, which is

intended for waste heat recovery via air-to-air heat exchange. The OHP-HRV is

positioned within the cross-sectional space of two typically-sized HVAC ducts each with
54
dimensions of 60.96 cm x 45.72 cm. Counter-flowing air - at a volumetric flow rate of

1.18 m3/s (2500 CFM) is assumed and the outdoor air temperature varies between -8 °C

(17.6 ºF) and 48.9 ºC (120 ºF) [29]. The proposed OHP-HRV is assumed to be fabricated

from copper capillary tubing (outer diameter = 0.318 cm, internal diameter = 0.165 cm)

and 0.397 mm thick aluminum fins. It is confirmed that the internal diameter of the OHP

tube facilitates the capillary action of acetone within the tube during standard operating

conditions.

Figure 5.1 Front view (left) and isometric view (right) of unit row OHP-HRV in
adjacent ducts

55
The OHP-HRV can exists as a system of independent, finned OHPs with

successive OHPs arranged in a staggered fashion (for increased heat transfer) while

sharing plate fins aligned parallel to the air flow. A side view and a top view of the OHP-

HRV is provided in Figure 5.2, where it may be seen that each row (in the axial direction)

of the OHP-HRV is a single OHP that shares a common blade-type fin that is

perpendicular to the air flow direction. The adiabatic sections (regions with negligible

heat transfer), as shown in Figure 5.2, were assumed to be of negligible length for the

analytical study.

Figure 5.2 OHP-HRV in adjacent ducts side view (left) and top view (right)

As shown in Figure 5.2, two independent geometric characteristics of the OHP-

HRV are the transverse pitch, SS, and the longitudinal (axial) pitch, SL. The diagonal

pitch, SD, is defined as:

56
1⁄2
𝑆𝑆 2
𝑆𝐷 = (𝑆𝐿2 +(2) ) (5.1)

where SS possesses a minimum value to avoid OHP tube pinching during manufacture,

and herein is set to 1.3 cm. For the current analysis, the transverse pitch and axial pitches

were held constant at 2.18 cm and 1 cm, respectively. Upon considering the minimum

turn radii of the OHPs, this results in each individual OHP having 20 turns. A summary

of design parameters is provided in Table 5.1.

Table 5.1 OHP-HRV dimension and operating conditions

Dimension Value Units


OHP tube inner diameter 1.65 mm
OHP tube outer diameter 3.18 mm
OHP turn radius 1.09 cm
OHP number of turn 20 -
Aluminum fin thickness 0.397 mm
Duct width 45.72 cm
Duct height 60.96 cm
Transverse pitch 2.18 cm
Axial pitch 1.00 cm
Evaporator-side inlet T 37.78 °C
Condenser-side inlet T 21.11 °C
Volumetric flow rate 1.18 m3/s

5.2 Heat transfer and pressure-drop analysis

The OHP-HRV shown for the heat transfer and aerodynamic analysis was

approximated as two tube-and-plain-fin heat exchangers linking two airstreams. The heat

transfer was calculated using Equations (5.2) - (5.14). The OHP-HRV was designed to

achieve maximum effectiveness while maintaining a pressure drop less than 200 Pa. The

evaporator and condenser were assumed to be two linked air-to-liquid heat exchangers
57
[116], [117]. The overall heat transfer coefficient was found as the inverse of the sum of

major thermal resistances, i.e.:

𝑈 = (𝑅𝑜 + 𝑅𝑤 + 𝑅𝑖 )−1 (5.2)

The outer OHP-HRV thermal resistance is given by:


1
𝑅𝑜 = 𝜂 (5.3)
𝑠𝑢𝑟𝑓𝑎𝑐𝑒 ∙ℎ𝑎𝑖𝑟

where, if uniform heat transfer coefficients and fluid properties on each side of the OHP-

HRV are assumed, hair is given by


2
ℎ𝑎𝑖𝑟 = 𝑗 ∙ 𝑃𝑟 −3 ∙ 𝐺 ∙ 𝑐𝑝 (5.4)

The Colburn j-factor required for Equation (5.4) is given by the following relation,

applicable only for 3 < NF < 20 [118]:

𝑗 = 0.195𝑅𝑒𝐿 −0.35 (5.5)

with the Reynolds number, 𝑅𝑒𝐿 defined as:


𝐺∙𝑆𝐿
𝑅𝑒𝐿 = (5.6)
𝜇

The tube wall thermal resistance is found using:


𝑡 ∙𝐴
𝑅𝑤 = 𝑘𝑤 ∙𝐴𝑜 (5.7)
𝑤 𝑖

where the wall thickness, tw and the wall thermal conductivity, kw are 0.79 mm and 401

W/m·K, respectively. Finally, the convection thermal resistance inside the OHP-HRV

tubes was found via:


𝐴
𝑜
𝑅𝑖 = 𝐴 ∙ℎ (5.8)
𝑖 𝑖

where, heat transfer coefficient, hi for the current investigation was assumed to be 17.5

and 25.0 kW/m2∙K in the condenser and evaporator, respectively. These selected, order-

58
of-magnitude values are representative of the evaporation and condensation heat transfer

within an OHP [112]. For the heat exchanger analysis, the ε-NTU method was utilized,

with the number of transfer units (NTU) being defined as:


𝑈∙𝐴𝑜
𝑁𝑇𝑈 = 𝐶
(5.9)

where C is ~1400 W/K corresponding to 2500 CFM of air throughput in the system.

Assuming the total heat capacity of the fluid inside the OHPs is much greater than the

heat capacity of the air, the effectiveness for one OHP row is found via:

𝜀1 = 1 − exp(−𝑁𝑇𝑈) (5.10)

which is used to calculate the effectiveness for 𝑛 rows:

𝜀𝑛 = 1 − (1 − 𝜀1 )𝑛 (5.11)

The effectiveness for the entire OHP-HRV, if 𝐶c > 𝐶h , was found using:

1 𝐶ℎ ⁄𝐶𝑐 −1
𝜀 = (𝜀 + ) (5.12)
ℎ 𝜀𝑐

and if 𝐶h > 𝐶c

1 𝐶𝑐 ⁄𝐶ℎ −1
𝜀 = (𝜀 + ) (5.13)
𝑐 𝜀ℎ

where 𝜀c and 𝜀h are calculated for n rows using Equation (5.11). The heat transfer

through the entire OHP-HRV was then calculated by:

𝑄 = 𝜀 ∙ 𝐶𝑚𝑖𝑛 ∙ (𝑇ℎ,𝑖𝑛 − 𝑇𝑐,𝑖𝑛 ) (5.14)

Due to the approximate nature of the employed heat transfer model, these results

do not take into account the axial thermal resistance of the fluid (i.e. in duct-to-duct

direction), or the vapor/liquid fraction in the evaporator.

59
5.3 Pressure drop through OHP-HRV

The pressure drop across each side of the OHP-HRV (i.e. evaporator and

condenser) was found using Equation (5.15) [119]:

𝐺2 𝜈 𝐴𝑡𝑜𝑡𝑎𝑙 𝜈𝑚
∆𝑃 = ∙ 𝜈1 ∙ [(1 + 𝜎 2 ) (𝜈2 − 1) + 𝑓 ∙ ∙ ] (5.15)
2 1 𝐴𝑚𝑖𝑛 𝜈1

where σ is the ratio of minimum free-flow area to frontal area, ν is the specific volume of

air. Note that the mass velocity (𝐺) is based upon minimum free flow area of OHP-HRV,

which is estimated using Equation (5.16).

𝐿
𝐴𝑚𝑖𝑛 = {( 𝑆3 − 1) ∙ 𝑧 + [(𝑆𝑠 − 𝐷𝑐 ) − (𝑆𝑠 − 𝐷𝑜 ) ∙ 𝑡 ∙ 𝑁𝑓 ]} ∙ 𝐿1 (5.16)
𝑠

where

𝑧 = 2𝑥 𝑖𝑓 2𝑥 < 2𝑦 𝑜𝑟 (5.17)

𝑧 = 2𝑦 𝑖𝑓 2𝑦 < 2𝑥 (5.18)

in which 𝑥 and 𝑦 are given by

(𝑆𝑠 −𝐷𝑐 ) − (𝑆𝐿 −𝐷𝑐 )∙𝑡∙𝑁𝑓


𝑥= (5.19)
2

𝑦 = (𝑆𝐷 − 𝐷𝑐 ) − (𝑆𝐿 − 𝐷𝑐 ) ∙ 𝑡 ∙ 𝑁𝑓 (5.20)

The total area of the OHP-HRV, as shown by Equation (5.21), which is sum of primary

area (area of un-finned tubes) and total fin area.

𝐴𝑡𝑜𝑡𝑎𝑙 = 𝐴𝑝 + 𝐴𝑓𝑖𝑛 (5.21)

The primary area and total fin area are estimated using Equation (5.22) and Equation

(5.23).

𝐴𝑝 = {𝜋 ∙ 𝐷𝑜 [(𝐿1 ) −∙ 𝑡 ∙ 𝑁𝑓 ∙ 𝐿1 ]} ∙ 𝑁𝑡 ∙ 𝑁𝑟 (5.22)

𝐷𝑜2
𝐴𝑓𝑖𝑛 = 2 (𝐿2 𝐿3 − 𝜋 𝑁𝑡 ∙ 𝑁𝑟 ) ∙ 𝑁𝑓 ∙ 𝐿1 + 2𝐿3 ∙ 𝑡 ∙ 𝑁𝑓 ∙ 𝐿1 (5.23)
4

60
The friction factor, f is given by Equation (5.24), in which ReDc is the Reynolds number

based on collar diameter, and Fp is the fin pitch.

𝑡 −0.104 𝐹𝑝 −0.197
𝑓 = 1.039𝑅𝑒𝐷−0.418
𝑐
(𝐷 ) 𝑁𝑓 −0.0935 (𝐷 ) (5.24)
𝑐 𝑐

5.4 Energy and Cost Analysis

This section illustrates the methodology to estimate energy and cost saving

potential of the proposed OHP-HRV system. The proposed system is designed to be used

in an air-handling unit (AHU) for commercial buildings in the U.S. In the analysis, it is

assumed that the system under consideration has no recirculation of air and it is a

dedicated outdoor air system (DOAS). The hourly heating/cooling energy delivered to a

building area to meet a room set-point (Qdelivered) was estimated using Equation (5.25).

𝑄𝑑𝑒𝑙𝑖𝑣𝑒𝑟𝑒𝑑 = 𝑉̇ 𝜌𝐶𝑝 (|𝑇𝑂𝐴𝑇 − 𝑇𝑆𝐴𝑇 |) (5.25)

where, 𝑉̇ is the volumetric air flow rate in m3/s, Cp is the heat capacity of the intake air, 𝜌

is the density of air at mean temperature, TOAT is the hourly averaged outdoor air

temperature, TSAT is the supply air temperature of a AHU and it is assumed to be 12 oC

for summer and 40 oC for winter in this analysis. The hourly waste heat recovery rate,

Qrcv, through the OHP-HRV was estimated using Equation (5.26).

𝑄𝑟𝑐𝑣 = 𝑉̇ 𝜌𝐶𝑝 ∆𝑇 (5.26)

where ∆𝑇 is the difference in temperature of air stream across the OHP-HRV either in

condenser or evaporator region. Note that Equation (5.26) was equated to Equation (5.14)

to obtain the downstream temperature of air streams across the OHP. Equation (5.27) was

used to estimate the hourly energy reduction through OHP-HRV for cooling.
𝑄𝑟𝑐𝑣
∆𝐸𝑐𝑜𝑚𝑝 = (5.27)
𝐶𝑂𝑃
61
where COP is the coefficient of performance for a chiller assumed to be 3 in this study

and ṁ is the mass flow rate of the intake air. Equation (5.28) was used to estimate the

hourly energy reduction through OHP-HRV for heating. The furnace efficiency,

∆𝐸furnace , was assumed to be 0.9 for the present analysis.

𝑄𝑟𝑐𝑣 𝑚̇𝐶𝑝 ∆𝑇
∆𝐸𝑓𝑢𝑟𝑛𝑎𝑐𝑒 = 𝜂 =𝜂 (5.28)
𝑓𝑢𝑟𝑛𝑎𝑐𝑒 𝑓𝑢𝑟𝑛𝑎𝑐𝑒

For a single-speed fan, the pressure drop is directly proportional to the fan energy

consumption as shown in Equation (5.29) [120]:

𝑉̇ ∙∆𝑃
∆𝐸𝑓𝑎𝑛 = 𝜂 (5.29)
𝑓𝑎𝑛 ∙𝜂𝑚𝑜𝑡𝑜𝑟 ∙𝜂𝑑𝑟𝑖𝑣𝑒

where ΔP is the pressure increase in Pa, ∆Efan is the fan energy consumption increase due

to the pressure increase by the OHP-HRV, 𝜂𝑓𝑎𝑛 is the fan energy efficiency, 𝜂𝑚𝑜𝑡𝑜𝑟 is the

fan motor efficiency, and 𝜂𝑑𝑟𝑖𝑣𝑒 is the belt drive efficiency. The hourly energy saving

using the proposed system for summer operation can be estimated as shown in Equation

(5.30).

∆𝐸𝑐𝑜𝑜𝑙𝑖𝑛𝑔 = ∆𝐸𝑐𝑜𝑚𝑝 − ∆𝐸𝑓𝑎𝑛 (5.30)

The hourly cost saving for cooling can be determined by Equation (5.31).

∆𝐶𝑜𝑠𝑡𝑐𝑜𝑜𝑙𝑖𝑛𝑔 = ∆𝐸𝑐𝑜𝑚𝑝 ∙ 𝐶𝑜𝑠𝑡𝑒𝑙 − ∆𝐸𝑓𝑎𝑛 ∙ 𝐶𝑜𝑠𝑡𝑒𝑙 (5.31)

where 𝐶𝑜𝑠𝑡𝑒𝑙 is the electricity cost for each location. The hourly energy saving using the

proposed system for winter operation can be estimated as shown in Equation (5.32).

∆𝐸ℎ𝑒𝑎𝑡𝑖𝑛𝑔 = ∆𝐸𝑓𝑢𝑟𝑛𝑎𝑐𝑒 − ∆𝐸𝑓𝑎𝑛 (5.32)

The hourly cost saving for cooling can be determined by Equation (5.33).

∆𝐶𝑜𝑠𝑡ℎ𝑒𝑎𝑡𝑖𝑛𝑔 = ∆𝐸𝑓𝑢𝑟𝑛𝑎𝑐𝑒 ∙ 𝐶𝑜𝑠𝑡𝑛𝑔 − ∆𝐸𝑓𝑎𝑛 ∙ 𝐶𝑜𝑠𝑡𝑒𝑙 (5.33)

62
where 𝐶𝑜𝑠𝑡𝑛𝑔 is the natural gas cost for each location. The total hourly energy saving

using the proposed system is given by Equation (5.34).

∆𝐸𝑡𝑜𝑡𝑎𝑙 = ∆𝐸𝑐𝑜𝑜𝑙𝑖𝑛𝑔 + ∆𝐸ℎ𝑒𝑎𝑡𝑖𝑛𝑔 (5.34)

Finally, the total cost saving can be calculated by Equation (5.35)

∆𝐶𝑜𝑠𝑡𝑡𝑜𝑡𝑎𝑙 = ∆𝐶𝑜𝑠𝑡𝑐𝑜𝑜𝑙𝑖𝑛𝑔 + ∆𝐶𝑜𝑠𝑡ℎ𝑒𝑎𝑡𝑖𝑛𝑔 (5.35)

5.5 Result and Discussion

5.5.1 Heat transfer performance

To determine the heat transfer performance of OHP-HRV and its variation with

the fin spacing and number of rows, the evaporator-side inlet temperature and condenser-

side inlet temperature was held constant at of 37.8 °C and 21.1 °C respectively. The heat

transfer performance of OHP-HRV was plotted vs. fin spacing and number of rows, is

provided in Figure 5.3. Isotherms that represent 5.56 ºC, 2.78 ºC and 0.556 ºC of hot-side

air pre-cooling are also provided for reference. From Figure 5.3, it is evident that the

number of rows has a greater effect on heat transfer than fin spacing. The heat transfer

levels off at the heat transfer rate, which corresponds to ε = 0.5 due to equal air flow rates

in the evaporator and condenser [21]. This indicates that an optimal design will approach

an effectiveness of 0.5 while still providing a relatively low-pressure drop. Adding more

rows or fins to achieve only a small effectiveness increase will only waste energy, in the

form of the extra fan power required to overcome the increased pressure loss. From

Figure 5.3 it may also be seen that in order to pre-cool incoming air by 5.56 ºC, about 8

rows of OHPs are required while having fin-to-fin spacing greater than 5 mm. For this

case, the heat transfer corresponds to approximately 6 kW.

63
Figure 5.3 OHP-HRV heat transfer rate vs number of OHP-HRV rows and fin
spacing. Isoplanes for pre-cooling temperature are also provided

5.5.2 Pressure drop

The condenser-side pressure drop vs. fin spacing and number of rows is plotted in

Figure 5.4. It may be seen that both design parameters have a strong effect on pressure

loss. Given that the number of rows has a greater effect on thermal performance for a

given pressure loss, a feasible design would incorporate relatively fewer fins and

relatively more rows. Pursuant to the above discussion of the asymptotic effectiveness

behavior, it is optimal to stop adding fins and rows when the effectiveness approaches

0.5. Further material additions will only cause a higher pressure drop, and thus more head

loss, with little increase in heat transfer capability.

64
Figure 5.4 OHP-HRV pressure drop vs fin spacing and number of rows

5.5.3 Design

Based on Figure 5.3 and Figure 5.4, an optimal OHP-HRV for this application

consists of 15 rows of 20-turns OHPs with the fins set 8 mm apart and the transverse

pitch set at 2.18 cm while completely spanning the entire duct width. These geometric

characteristics provide for a minimal pressure drop while still allowing for high heat

transfer. For this particular OHP-HRV design, the heat transfer is 10.76 kW,

corresponding to an evaporator temperature drop (pre-cooling) of 8.0 °C and a heat

exchanger effectiveness of approximately 0.48. The condenser pressure drop is 39.8 Pa,

and the evaporator pressure drop is 36.4 Pa. This design obtains effectiveness within 4%

of the theoretical maximum, while still achieving a low-pressure drop. These factors

indicate that the proposed design may be feasible in a wide variety of applications.

65
5.5.4 Start-up behavior of OHP-HRV

The OHP requires a minimum heat input (and/or temperature difference) to

initiate its internal fluid motion and become an ultra-conductive heat transfer device. Qu

and Ma [75] provides a means for theoretically predicting the start-up heat input based on

the initiation of bubble growth in the evaporator due to a sufficient superheat being

achieved. The equation for the minimum heat transfer rate is provided as:

𝑘𝑙 𝑇𝑣 𝐴c,h 2 𝑅𝑇 2𝜎
𝑄≈𝑟 (1⁄{1 − ℎ 𝑣 ln [1 + 𝜌 ]} − 1) (5.36)
𝑖 ln[𝑟𝑖 ⁄(𝑟𝑖 −𝛿𝑙 )] 𝑓𝑔 𝑣 𝑅𝐺 𝑇𝑣 𝑟𝑛

where ri is the internal radius of the OHP capillary tube, and Ac,h is the cross-sectional

area of the tube section in the evaporator, found as:

𝜋
𝐴𝑐,ℎ = 𝑁𝐿 𝑁𝑇 (4 𝐷𝑖 2 ) (5.37)

where NL and NT are the number of tube sections in the longitudinal and transverse

direction, respectively.

Based on Equation (5.36), a number of working fluid and material properties are

of interest for minimizing the start-up heat input of the OHP-HRV. The liquid thermal

conductivity, kl, and latent heat of vaporization, hfg, should be minimized, and, most

importantly, the surface roughness of the internal tube, rn, should be maximized. Using

saturated properties for acetone at a temperature of approximately 20 ºC, the

thermophysical properties were obtained. The startup heat transfer is plotted as a function

of the vapor temperature, Tv, for various surface roughness values and film thicknesses in

Figure 5.5 to Figure 5.7. The vapor temperature was assumed to be equal to the incoming

hot air temperature (~310 K) due to the low amount of superheat required to initiate

oscillating motion with acetone [75]. Note that this allows startup in the case of δl = 10

66
μm for acetone. Acetone’s low toxicity and required startup heat transfer make it a

reasonable choice of working fluid. For instance, Figure 5.7 provides a comparison of the

required startup heat transfer between acetone and water. It can clearly be seen that water

lacks the thermophysical properties required for this specific waste heat recovery

application.

Figure 5.5 Startup heat transfer of acetone OHP-HRV vs vapor temperature for δl = 2
μm

67
Figure 5.6 Start-up heat transfer of acetone OHP-HRV vs vapor temperature for δl =
10μm

68
Figure 5.7 Startup heat transfer for acetone and water, δl = 6 μm, rn = 2 μm

Figure 5.7 (plotted for rn = 2 μm and δl = 6 μm), shows that when the start-up heat

transfer is not achieved, the OHP-HRV will behave similar to a bundle of copper tubes,

with an effectiveness of approximately 0.04 and capable of transferring heat levels up to

400 W. This is how the pre-startup heat transfer per OHP in Figure 5.5 and Figure 5.6 is

calculated – the OHP is treated as a series of pure copper tubes of diameter Do across a

temperature difference of (37.8-21.1) = 16.7 °C. However, once the start-up power level

is achieved, the OHP-HRV becomes an order-of-magnitude more effective due to

increased effective thermal conductivity as shown in Figure 5.8.

69
Figure 5.8 OHP-HRV heat transfer and effectiveness vs maximum temperature
difference

5.5.5 Energy and Cost Saving Analysis Result

An energy and cost saving analysis of the proposed OHP-HRV system described

in Section 5.1 was performed, and the results are presented and discussed in this section.

The proposed OHP-HRV system was designed to recover energy in an air-handling unit

(AHU) that is commonly found in the U.S. commercial building to demonstrate its

energy and cost saving benefits. The AHU was assumed to be equipped with a constant

speed fan of 1.18 m3/s (2500 CFM), that has 𝜂𝑓𝑎𝑛 = 0.65, 𝜂𝑚𝑜𝑡𝑜𝑟 = 0.85, and 𝜂𝑑𝑟𝑖𝑣𝑒 =

0.8. It was also assumed that the building was under 24 hour operation and the cooling

and heating was provided by a chiller and a gas furnace respectively. The performance of

the proposed OHP-HRV was evaluated in eight different U.S. climate locations (i.e.,

Atlanta, GA, Phoenix, AZ, Denver, CO, Los Angeles, CA, Baltimore, MD, Chicago, IL,

70
Miami FL and Houston, TX). The outdoor air temperature information for these cities

were obtained from the typical meteorological year data sets (TMY-3) [121] and used as

inputs to determine the total energy savings by the OHP-HRV. City-wise variations in the

cost of retail electricity price and natural gas for the commercial building sector were

obtained from [122] (see Table 5.2) and used to determine cost savings using Equations

(5.31), (5.33) and (5.35). Table 5.2 also shows the classification of the climate of cities as

par as Koppen-Geiger Climate Classification System [123].

Table 5.2 Price of natural gas and electricity for different locations [122] and climate
[123].

Location Natural Gas Electricity Climate Classification


Cost ($/1000 Cost
cu. ft.) (cents/kWh)
(cent/kWh)
Atlanta, GA 9.08 9.90 Humid subtropical
Baltimore, MD 10.98 10.68 Humid subtropical
Chicago, IL 11.48 8.16 Humid continental climate
Denver, CO 9.41 10.00 Semi-arid continental
Houston, TX 7.50 8.02 Humid subtropical
Los Angeles, CA 7.68 14.50 Mediterranean
Miami, FL 10.92 9.39 Tropical monsoon
Phoenix, AZ 10.67 9.85 Hot desert

The result of city-wise energy and cost saving analysis is shown in Figure 5.9 and

Figure 5.10. Figure 5.9 shows the season-wise potential of waste heat recovery through

proposed OHP-HRV across different US cities. Results for different locations of U.S.A

demonstrate that in general the waste heat recovery from the proposed OHP-HRV is

higher for winter operation than that of summer operation. For example, sub humid

71
tropical climatic region such as Atlanta and Baltimore shows that the waste heat recovery

potential for winter operation accounts for more than 80 % of the total annual waste heat

recovery potential. Continental climatic region such as Chicago and Denver shows the

maximum waste heat recovery potential whereas a tropical monsoon climatic region such

as Miami and a Mediterranean climatic region such as Los Angeles has the minimum

waste heat recovery potential. This can be attributed to the fact that Chicago and Denver

has approximately 8 month of winter with a monthly average temperature less than 13 oC

[121]. On the other hand, Los Angeles and Miami have approximately 8 months that has

the monthly average temperature between 16 oC and 26 oC [121]. In other words, the

difference between SAT and OAT in Chicago and Denver is higher than 8 oC for most

winter days. Following Figure 5.8, the OHP-HRV will operate with a higher

effectiveness in Chicago and Denver, where as in cities such as Los Angeles and Miami

the effectiveness of OHP-HRV operation will be low. Among the cities investigated,

Phoenix-a region classified as hot desert is the only city where waste heat recovery

potential for summer operation is greater than that of winter operation.

72
Figure 5.9 Seasonal waste heat recovery through proposed OHP-HRV across different
cities in United State of America

Figure 5.10 indicates the city-wise annual energy and cost savings potential from

the proposed OHP-HRV. It can be inferred from the Figure 5.10 that for an AHU of

capacity 2500 CFM, installed in commercial building in these eight cities, the average

percentage energy reduction was found to be 16.5 % with an average annual saving of

$714. Figure 5.10 also implies that the utilities rates in the respective city plays a

significant role in realizing the cost saving potential of the proposed system. For example

energy saving potential in Chicago and Denver is almost similar, but due to the difference

in utilities rates in these cities, Chicago has higher cost saving potential than Denver.

Likewise, Miami has more potential of energy saving than Los Angeles, but Miami’s cost

saving potential is lower than that of Los Angeles. Baltimore and Phoenix has almost

same percentage of energy saving potential, but cost saving potential of Baltimore is 70

73
% more than that of Phoenix. Atlanta and Houston has almost similar annual energy

saving potential but the annual cost saving potential of Atlanta is $614 more than that of

Houston. Among the cities investigated, Houston has the cheapest utility rates and that’s

why the annual cost saving potential is comparatively lower than other cities.

Figure 5.10 Percentage annual energy and cost saving from proposed OHP-HRV for
different cities in United State of America

74
CHAPTER VI

EXPERIMENTAL INVESTIGATION OF ATYPICALLY LONG FINNED OHP AS A

WASTE HEAT RECOVERY DEVICE

This chapter presents an experimental investigation of a novel atypically long

finned OHP (F-OHP) as a waste heat recovery device. The aim is to study the effect of

extended surface on OHP’s heat transfer performance and contrast it with the heat

transfer performance of bare tube OHP of similar overall dimension. Additionally, the

gain in heat transfer due to extended surface is juxtaposed with increase in pressure drop

and fin characteristics such as fin effectiveness was also calculated. The heat transfer area

of capillary size OHP tubes was enhanced by adding extended surface (helical fins) at a

rate of 12 FPI, thereby increasing the heat transfer area by approximately 433%. Helical

fins at above mentioned fin pitch is customary to HVAC systems [9]. Although, the

technique is conventional, but it is still unique as putting fins on capillary size tube is

seldom executed. To the best of author knowledge, investigation of the 1/8 inch (3.175

mm) external diameter F-OHP with having fins on both evaporator and condenser side

has never been done before.

6.1 Finned OHP (F-OHP) design and construction

The section discusses F-OHP’s design criteria and construction features. Equation

(1.1) was used to determine the critical radius for present working fluid and temperature

regime. Accordingly, F-OHP inner tube diameter was selected. F-OHP is fabricated by an
75
assembly of seven ‘circuits’ of finned copper capillary tube with inner diameter, 𝐷i =

0.165 cm (0.062 in) and outer diameter 𝐷o = 0.317 cm (0.125 in). All the circuits are

connected to form a closed loop using six vacuum union fittings (Swagelok/SS-2-UT-6).

The charging port is made by connecting two circuits of opposite end with vacuum union

tee fitting (Swagelok/SS-2-UT-3). Figure 6.1 shows the image of investigated F-OHP

highlighting attached fins and vacuum union fittings that are used to form a closed

circuit. Table 6.1 lists the additional constructional details of F-OHP.

Figure 6.1 Photograph of 9 turn finned OHP showing cut-out view of fins with FPI
equal to 12.

76
Table 6.1 Construction features of F-OHP

Material of F-OHP Copper Alloy 122


Outer diameter of F-OHP 3.18 mm (0.125 in)
Inner diameter of F-OHP 1.65 mm (0.062 in)
Number of turns 9
Height of F-OHP 127 cm (50 in)
Working fluid n-pentane
Tube spacing 2.54 cm (1 in)
Number of fins per inch (FPI) 12
Type of fin Helical
Outer diameter of fins 6.35 mm (0.25 in)
Thickness of fins 0.381 mm (0.015 in)
Fin construction type Edge Tension

6.2 Fin characteristics

Wrap on fins or L footed fins of helical type are applied to capillary size OHP

tubes. These types of fins are most commonly made with aluminum and copper strips.

Thin copper strip is first welded at one end of tube. Then, it is tightly wrapped around the

tube using tension force, thereby creating interference fit type contact with the base

metal. Finally, after wrapping at the other end it is again welded. The wraps on fins are

suitable for low temperature application as it offers maximum heat transfer in low

temperature range. Also it most cost effective compared to others. It eliminates the need

of helical grooves on bare tube. This is why, it is most suitable for bare tube that with

very thin wall thickness. Wrap on fins offers unique advantage as if the fins fails, the bare

tube can be re-finned with no effect on the walls. However, the wrap on fins are weak to

resist mechanical damage and required intensive handling and care. To evaluate the fin

effectiveness of F-OHP and study its role on oscillation, a bare tube OHP of dimension

similar of that of F-OHP was fabricated and tested under same operation condition.

Following table shows that description of B-OHP.


77
Table 6.2 Construction feature of Bare tube OHP (B-OHP).

Material of OHP C122 Copper


Outer diameter of F-OHP 3.18 mm (0.125 in)
Inner diameter of F-OHP 1.65 mm (0.062 in)
Number of turn 9
Height of F-OHP 123 cm (48.5 in)
Working fluid n-pentane
Tube spacing 2.54 cm (1 in)
Total length of tube 22.12 m

6.3 Result and Discussion

The results of experimental investigation of novel F-OHP as a waste heat

recovery device is presented here. F-OHP’s thermal performance and subsequent

pressure drop was compared to that of a bare/un-finned OHP (B-OHP) of same

dimension and operating condition (operating temperature and air flow rate). From the

investigation presented in Chapter IV, it was found that an OHP of current inner

diameter, encapsulating n-pentane and operating under present condition, will have least

thermal resistance with the fill ratio of 70 %. Hence, in the present study, F-OHP and B-

OHP was partially filled with n-pentane with a fill ratio 70 %. Figure 6.2 shows the

pressure drop in cold air stream across F-OHP and B-OHP including pressure drop due to

friction and duct design. The pressure drop in the cold air stream due friction was found

to be approximately 10 (±2) Pa. It can inferred from the Figure 6.2 that the addition of

fins on OHP causes an additional pressure drop of 7 (±2) Pa.

78
Figure 6.2 Pressure drop in cold air stream across F-OHP and B-OHP

6.3.1 Thermal performance of Finned OHP (F-OHP)

As mentioned in the earlier section, the F-OHP form a closed system with 7

copper fin tubes connected to each other via vacuum fittings. The condenser section

comprises of cold ambient air flowing across F-OHP. The vacuum retaining capability of

F-OHP and the effect of cold weather on start-up condition was studied. A total of 7

consecutive test was conducted on F-OHP with an interval of 24 hours. Each test lasted

for 6500 seconds and temperature time history for first four tests are shown in Figure 6.3

It shows the temperature time history of OHP condenser and evaporator as recorded by a

pair of thermocouple located on the center most tube. It can be seen from tests that there

exists a waiting period to start the oscillatory motion. For Test 1 shown in Figure 6.3a,

the waiting period was brief (< 500 seconds), whereas in subsequent tests the waiting

period was found to be more than 2000 seconds. This can be explain as follows.

79
The F-OHP starts to oscillate when vapor pressure in evaporator is high enough to

move the train of liquid plug and vapor bubble. Besides, time taken by OHP to attain

start-up condition also depends upon heat addition rate and heat rejection rate in

evaporator and condenser region respectively [75]. Test 1 was conducted just after the

filling operation, i.e., the F-OHP filled with working liquid was at room temperature

(23±2 oC) at the beginning of the test. Hence, with a small addition of heat flux, it swiftly

attained suitable start up condition. Test 2 was conducted exactly after 24 hours of Test 1

and so with the Test 3 and Test 4. To simulate the real world scenario, the F-OHP was

exposed to ambient cold weather environment with an average ambient temperature

prevailing around 10 (±2) oC. Since the initial temperature of OHP during Test 2, 3, and 4

was that of ambient temperature, it took a longer time for it attain required heat flux rate

necessary for start-up condition. Results from test 4 are considered for further analysis

since it among the test which mimic the real world scenario.

80
(a)

(b)

(c)

(d)

Figure 6.3 Temperature time history of four tests of F-OHP conducted subsequently
with an interval of 24 hour

81
Figure 6.4 shows the temperature oscillation of F-OHP as recorded by three most

centered thermocouple located in the evaporator and condenser region. As it can be

observed from the figure that it took 2200 seconds for the OHP to start-up. Upon start-up,

thermocouples located on evaporator surface records a sudden drop in temperature and

synchronously a sudden rise in temperature average was recorded by thermocouples

located on condenser surface. The average change in evaporator and condenser surface

temperature is approximately 3 (±1) oC. It can also be observed that there exists an

unbalanced oscillation between condenser and evaporator. This is attributed to an unequal

air flow rate prevailing in condenser and evaporator section. The hot air stream and cold

air stream both are equipped with fans of similar capacity, however the hot air stream

suffer higher head losses at the upstream side due to inherent duct design as compared to

cold air stream. This results in lower heat transfer coefficient at evaporator side than at

condenser side, which causes unbalance oscillation.

Figure 6.4 Temperature response of F-OHP filled with n-pentane with fill ratio 70 %

82
Figure 6.5 shows the time-wise heat transfer rate to the cold stream (heat recovery

rate) via F-OHP. The figure depicts three operating zone: AB, BC, and CD. Zone AB

represents pre-start up period of F-OHP. This period lasted for about 2200 seconds. Zone

BC represents the start-up zone of F-OHP and zone CD corresponds to the steady state of

oscillation in the F-OHP. It can be observed that heat recovery rate in zone AB seems to

be plateauing around 230 (± 40) W. Upon the start of the oscillation (zone BC), there is a

sudden surge in heat recovery rate and at steady state (zone CD) it averaged at around

400 (± 40) W.

The secondary axis of Figure 6.5 shows the average thermal resistance time

history of F-OHP. The average thermal resistance in pre-startup period (zone AB) was

found approximately to be equal to 0.273 (± 0.01) K/W, whereas it reduces to an

averaged value of 0.04 ((±0.01) K/W during the steady state oscillation (zone CD). In

other words, the oscillatory behavior of working fluid lowers the overall thermal

resistance of F-OHP by almost 86 %. The sharp reduction in overall average thermal

resistance and rise in average heat transfer rate upon the start of oscillation is mainly

attributed to two reason – 1) before the start-up, the heat transfer to the working fluid is

sensible heat and latent heat of vaporization at liquid vapor interface. During oscillation,

nucleate boiling heat transfer occur which significantly increases the heat transport

capability [75]. 2) There exists a thin liquid film between the wall and the liquid plug,

movement of which during oscillation causes high heat transport capability [23].

83
Figure 6.5 Time-wise average heat recovery rate and thermal resistance of F-OHP
filled with n-pentane with fill ratio 70 %

6.3.2 Thermal performance of bare tube OHP (B-OHP)

To evaluate the fin effectiveness of F-OHP and study its impact on oscillation, a

bare tube OHP of dimension similar of that of F-OHP was fabricated and tested under

same operation condition. Figure 6.6 shows the temperature time history of B-OHP. It

can be observed that the start-up time for presented OHP is less than 500 s. The testing

was conducted just after filling operation, that is, the initial temperature of B-OHP was at

room temperature (21 (± 2) oC). Hence, fairly less amount of excitation force was

required by the working fluid to start oscillation. Similar to the F-OHP, B-OHP also

experienced unbalanced oscillation.

84
Figure 6.6 Temperature response of bare tube OHP filled with n-pentane with fill ratio
70 %

Figure 6.7 shows the time-wise heat addition rate to cold air stream via B-OHP.

Figure 6.7 also shows time-wise average thermal resistance of B-OHP on its secondary

axis. The average thermal resistance before the start–up condition was found to be 0.11

K/W, whereas upon the start of oscillation it drops to an average value of 0.07 K/W. It

can be notice clearly that the there is significant drop in average thermal resistance as

soon as the oscillation starts.

85
Figure 6.7 Heat addition rate to cold air stream ( heat recovery rate ) via B-OHP filled
with n-pentane with fill ratio 70 %

6.3.3 Comparative study of F-OHP and B-OHP

A comparative study of the F-OHP and B-OHP was conducted to assess the

coupled effect of fins and oscillation on OHP’s thermal performance. OHPs with four

configurations were investigated for this comparative study. They are -1) F-OHP filled

with n-pentane with FR of 70 %, 2) an empty F-OHP, 3) B-OHP filled with n-pentane

with FR of 70 %; 4) an empty B-OHP.

86
Figure 6.8 Comparison of time-averaged heat addition rate to cold air stream through
four configurations of OHP

The four configuration are finned & filled, finned & empty, bare & filled, bare & empty

Figure 6.8 shows the time-averaged heat recovery rate through of four

configurations of OHPs. The average heat transfer rate via F-OHP filled with 70 % n-

pentane was found to be almost 80 % high up when compared to B-OHP filled with same

working fluid with same fill ratio. In fact, when compared to an empty B-OHP of similar

dimension, the coupled effect of fins and working fluid oscillation enhanced the heat

transfer rate by almost 400 %.

87
Figure 6.9 Comparison of time-wise average thermal resistance of four configurations
of OHPs

Fin effectiveness is vital characteristics of any fin surface. Fin effectiveness is

defined as the “ratio of fin heat transfer rate to the heat transfer rate that would exist

without the fins” [110]. Accordingly, the fin effectiveness, at steady state, was estimated

as the ratio of time-averaged heat transfer rate of the F-OHP and B-OHP. The fin

effectiveness of empty F-OHPs was estimated to be 1.13, whereas for filled F-OHP the

fin effectiveness was 2.11. That is, the oscillation of working liquid enhanced the fin

effectiveness by almost 2 times. This can be attributed to the fact that the oscillation of

working fluid increases the overall thermal conductance of OHP, which therefore results

in higher fin effectiveness of F-OHP. This is clearly evident in Figure 6.9. It compares

the time-wise variation of average thermal resistance of four different configurations of

88
OHPs. Note that the data shown in the Figure 6.9 is smooth and filtered data obtained by

applying Savitzky-Golay filter (a generalized moving average smoothing technique) to

the raw data [115].

89
CHAPTER VII

SUMMARY, CONCLUSION, AND FUTURE WORK

7.1 Summery and Conclusion

This dissertation presents the on-going development of OHP as a waste heat

recovery device. To the best of author’s knowledge, this present work is first systematic

attempt to investigate OHP as a waste heat recovery device. Unique working fluids with

limited research inside OHPs, but with properties desirable for low grade heat fluxes, i.e.

n-pentane, iso-pentane, and acetone were chosen as the working fluid for the OHPs

investigated in this dissertation.

Chapter I examined the potential role of waste heat recovery devices in addressing

the heightened appeal of improving the efficiency of current energy infrastructure.

Conventional waste heat recovery systems such as CHP, enthalpy wheel, FP-HE were

introduced and their limitations were stated. The operational mechanism of OHP was

reviewed and its potential role as a waste heat recovery device was discussed. In addition,

the influence of three most fundamental parameters (input heat flux, channel diameter,

working fluid fill ratio) on OHP’s functioning was explained.

Chapter II presented a comprehensive literature review on OHP. Its evolution and

a wide range of theoretical and experimental studies focusing on its thermal performance

was also reviewed. In addition, this chapter also highlights the versatility of OHP and its

90
ease of construction. Chapter III showed the experimental setup and procedure adopted in

present investigations.

Chapter IV presented an experimental feasibility of a nine turn copper-made

tubular OHP for passive waste heat recovery via air-to-air heat exchange. The OHP

thermal resistance and total heat transfer for hot-environment HVAC operation was

benchmarked with an empty/evacuated OHP with same overall dimensions. The working

fluid, n-pentane, was utilized due to its suitable thermophysical properties and the OHP

fill ratio was varied between 0%, 60% and 70%. The results show that using a simple

OHP design (un-finned and unit row), up to 240 W of heat can be recovered from the

waste exhaust air stream, while just having a pressure drop of 62 Pa in the cold air

stream. It was also found that, among all the fill-ratio tested, n-pentane with a fill ratio of

70% is least thermally resistive. Moreover, the heat transfer through the OHP with 70%

n-pentane is more than two times the heat transfer through empty/evacuated OHP. The

oscillations in the OHP-HE with 70 % n-pentane start within 500 seconds and at 38.2 oC

hot air inlet temperatures, hence, it is suitable for the HVAC heat transfer task.

Chapter V provided a first-order analysis to describe the heat transfer

performance of oscillating heat pipes (OHPs) for air-to-air heat exchange in a typical air

conditioning system and environment. The results demonstrate that the OHP-HRV

comprising of 15 rows of 20 turn copper-made OHPs has the potential to pre-cool

incoming air by 8.0 °C, with an effectiveness on the order of 0.48 and a pressure drop of

approximately 40 Pa. The results from the annual energy and cost savings analysis show

that the OHP-HRV system can provide energy efficient and cost effective operation -

reducing total average annual energy consumption by 16 % and total annual operational

91
cost by $714 for an AHU with an outdoor intake air flow rate of 1.18 m3/s (2500 CFM)

that provides cooling/heating for a commercial building located in eight different cities

across U.S.A.

Chapter VI showed an experimental investigation of a novel finned oscillating

heat pipe (F-OHP) as a waste heat recovery devices. The heat transfer area of capillary

size tubes was enhanced by adding extended surface (helical fins) to the OHP at a rate of

12 fins per inch (12 FPI), thereby, increasing the heat transfer area by 433%. Helical fins

at above mentioned fin pitch is customary to HVAC systems. The thermal performance

of F-OHP and resultant pressure drop was equated with that of bare tube OHP (B-OHP)

of similar dimension. It was found that a unit row of F-OHP filled with n-pentane with

FR of 70 % can recover up to 400 (± 40) W of heat from a typical waste exhaust air

stream. The additional pressure drop due to fins was estimated to be 3.7 (± 2) Pa. The

average heat recovery rate via F-OHP was found to be almost 80 % more than B-OHP

filled with same working fluid with same FR. In fact, when compared to an empty B-

OHP under similar operating condition, the coupled effect of fins and working fluid

oscillation enhanced the heat recovery rate by almost 400 %. The fin effectiveness for the

case of filled F-OHP was found to be greater than 2, and hence fins on the present OHP is

justified.

7.2 Future work

This dissertation is stepping stone in development of OHP as a waste heat

recovery device. While the progress shown in present work shows OHP has a potential to

be an efficient waste heat recovery device, it would be naïve to claim the

comprehensiveness of the present investigation. The future work should be to continue to


92
develop OHP as an effective waste heat recovery device and following suggestion may

be helpful to take the lead in this direction.

1. The present investigation has mainly focused on application of OHPs for

low-grade heat transfer process i.e., comfort to comfort type waste heat

recovery. It can be further extended by examining the application of OHP

in other types of waste heat recovery such as process to comfort (medium

grade heat transfer processes) and process to process type (high-grade heat

transfer processes).

2. The present investigation has primarily shown the potential of a unit row

of OHP as a waste heat recovery devices. The investigation needs to be

extended to interconnected and separate multiple rows of OHPs for further

acceptance of the technology

3. Recent studies have suggested that adding nanoparticles to the base

working fluid of OHP significantly enhances its heat transport capability.

The resultant working fluid is called nanofluid. Its use in the field of waste

heat recovery can be a game changer and therefore needs to be examined.

93
REFERENCES

[1] L. Pérez-Lombard, J. Ortiz, and C. Pout, “A review on buildings energy


consumption information,” Energy Build., vol. 40, no. 3, pp. 394–398, 2008.

[2] W. Wu, Z. Fang, W. Ji, and H. Wang, “Optimal operation condition division with
profit and losses analysis of energy recovery ventilator,” Energy Build., Nov.
2015.

[3] L.-Z. Zhang and F. Xiao, “Simultaneous heat and moisture transfer through a
composite supported liquid membrane,” Int. J. Heat Mass Transf., vol. 51, no. 9–
10, pp. 2179–2189, May 2008.

[4] K. W. Roth, D. Westphalan, J. Dieckmann, S. D. Hamilton, and W. Goetzler,


“Energy consumption characteristics of commercial building HVAC systems
volume III: Energy savings potential,” Cambridge, MA, 2002.

[5] J. Zhang and A. S. Fung, “Experimental study and analysis of an energy recovery
ventilator and the impacts of defrost cycle,” Energy Build., vol. 87, pp. 265–271,
Jan. 2015.

[6] J. Dieckmann, “Improving Humidity Control With Energy Recovery Ventilation,”


ASHRAE J., vol. 45, no. 8, pp. 38 – 45, 2008.

[7] R. W. Besant and C. J. Simonson, “Air-To-Air Energy Recovery,” ASHRAE J.,


pp. 31–42, 2000.

[8] S. Delfani, H. Pasdarshahri, and M. Karami, “Experimental investigation of heat


recovery system for building air conditioning in hot and humid areas,” Energy
Build., vol. 49, pp. 62–68, Jun. 2012.

[9] R. M. Lazzarin and A. Gasparella, “Technical and economical analysis of heat


recovery in building ventilation systems,” Appl. Therm. Eng., vol. 18, no. 1–2, pp.
47–67, Jan. 1998.

[10] D. A. Reay, “The Perkins Tube-a noteworthy contribution to heat exchanger


technology,” Journal of Heat Recovery Systems, vol. 2, no. 2. pp. 173–187, 1982.

[11] “Inspired Heat-Pipe Technology, Los Alamos National Laboratory,” National


Security Science, 2011. [Online]. Available:
http://www.lanl.gov/science/NSS/issue1_2011/story6full.shtml.
94
[12] G. Richard, “Heat transfer device,” US2350348, 06-Jun-1944.

[13] G. P. Peterson, An Introduction to Heat Pipes. Modeling, testing, and


applications, 1st ed. New York, Chichester,: Wiley, 1994.

[14] G. M. Grover, T. P. Cotter, and G. F. Erickson, “Structures of very high thermal


conductance,” J. Appl. Phys., vol. 35, no. 6, pp. 1990–1991, 1964.

[15] ASHRAE, ASHRAE Handbook- HVAC Systems and Equipment. Atlanta, GA,
2012.

[16] M. A. Abd El-Baky and M. M. Mohamed, “Heat Pipe heat exchanger for heat
recovery in air conditioning,” Appl. Therm. Eng., vol. 27, pp. 795–801, 2007.

[17] Y. H. Yau and A. S. Tucker, “The performance study of a wet six‐row heat‐pipe
heat exchanger operating in tropical buildings,” Int. J. Energy Res., vol. 27, no. 3,
pp. 187–202, 2003.

[18] N. J. Lamfon, Y. S. H. Najjar, and M. Akyurt, “Modeling and simulation of


combined gas turbine engine and heat pipe system for waste heat recovery and
utilization,” Energy Convers. Manag., vol. 39, no. 1, pp. 81–86, 1998.

[19] D. Liu, G.-F. Tang, F.-Y. Zhao, and H.-Q. Wang, “Modeling and experimental
investigation of looped separate heat pipe as waste heat recovery facility,” Appl.
Therm. Eng., vol. 26, no. 17, pp. 2433–2441, 2006.

[20] H. Jouhara and R. Meskimmon, “Experimental investigation of wraparound loop


heat pipe heat exchanger used in energy efficient air handling units,” Energy, vol.
35, no. 12, pp. 4592–4599, 2010.

[21] B. Suman, “Modeling, Experiment, and Fabrication of Micro-Grooved Heat


Pipes: An Update,” Appl. Mech. Rev., vol. 60, no. 3, p. 107, May 2007.

[22] W. Shang and R. W. Besant, “Theoretical and experimental methods for the
sensible effectiveness of air-to-air energy recovery wheels,” HVAC&R Res., vol.
14, no. 3, pp. 373–396, 2008.

[23] H. Ma, “Oscillating Heat Pipes,” in Oscillating Heat Pipes, 1st ed., Springer-
Verlag New York, 2015.

[24] R. C. Givler and M. J. Martinez, “Modeling of Pulsating Heat Pipes,” Sandia


National Laboratories, Livermore, CA, 2009.

[25] S. Khandekar, N. Dollinger, and M. Groll, “Understanding operational regimes of


closed loop pulsating heat pipes: An experimental study,” Appl. Therm. Eng., vol.
23, pp. 707–719, 2003.

95
[26] M. Groll and S. Khandekar, “Pulsating heat pipe: progress and prospects,” in
International Conference on Energy and the Environment, Shanghai, China,
2003, vol. 29, no. 1, pp. 20–44.

[27] S. M. Thompson, P. Cheng, and H. B. Ma, “An experimental investigation of a


three-dimensional flat-plate oscillating heat pipe with staggered microchannels,”
Int. J. Heat Mass Transf., vol. 54, no. 17–18, pp. 3951–3959, Aug. 2011.

[28] S. M. Thompson, A. A. Hathaway, C. D. Smoot, C. A. Wilson, H. B. Ma, R. M.


Young, L. Greenberg, B. R. Osick, S. Van Campen, B. C. Morgan, D. Sharar, and
N. Jankowski, “Robust thermal performance of a flat-plate oscillating heat pipe
during high-gravity loading,” J. Heat Transfer, vol. 133, no. 10, p. 104504, Aug.
2011.

[29] S. Khandekar and Groll Manfred, “On The Defination of Pulsating Heat Pipes:
An Overview,” in 5th Minsk International Seminars(Heat Pipes, Heat Pumps and
Refrigerators), Minsk, Belarus, 2003.

[30] X. Han, X. Wang, H. Zheng, X. Xu, and G. Chen, “Review of the development of
pulsating heat pipe for heat dissipation,” Renew. Sustain. Energy Rev., vol. 59, pp.
692–709, 2016.

[31] H. Akachi, “Structure of a heat pipe,” US 4921041 A, 01-May-1990.

[32] H. Akachi, “L-type heat sink,” US 5490558 A, 13-Feb-1996.

[33] S. Khandekar, “Thermo-hydrodynamics of Closed Loop Pulsating Heat Pipes,” p.


154, 2004.

[34] S. M. Thompson, B. S. Tessler, H. Ma, D. E. Smith, and A. Sobel, “Ultrahigh


thermal conductivity of three-dimensional flat-plate sscillating heat pipes for
electromagnetic launcher cooling,” Plasma Science, IEEE Transactions on, vol.
41, no. 5. pp. 1326–1331, 2013.

[35] Y.-H. Lin, S.-W. Kang, and H.-L. Chen, “Effect of silver nano-fluid on pulsating
heat pipe thermal performance,” Appl. Therm. Eng., vol. 28, no. 11–12, pp. 1312–
1317, Aug. 2008.

[36] Y.-B. Im and J.-S. Kim, “Influence of working fluid to heat transfer
characteristics of oscillating heat pipe for low temperature,” in 13th International
Heat Pipe Conference, Shanghai, China, 2004, pp. 21–25.

[37] D. Yuan, W. Qu, and T. Ma, “Flow and heat transfer of liquid plug and
neighboring vapor slugs in a pulsating heat pipe,” Int. J. Heat Mass Transf., vol.
53, no. 7–8, pp. 1260–1268, Mar. 2010.

96
[38] P. Meena, S. Rittidech, and P. Tammasaeng, “Effect of evaporator section lengths
and working fluids on operational limit of closed loop oscillating heat pipes with
check valves (CLOHP/CV),” Am. J. Appl. Sci., vol. 6, no. 1, pp. 133–136, 2009.

[39] W. Qu, Y. Li, and T. Ma, “Frequency analysis of pulsating heat pipe,” in
Proceedings of ASME InterPACK Conference, British Columbia, Vancouver,
Canada, 2007, pp. 689–693.

[40] S. Khandekar and M. Groll, “An insight into thermo-hydrodynamic coupling in


closed loop pulsating heat pipes,” Int. J. Therm. Sci., vol. 43, pp. 13–20, 2004.

[41] N. Bhuwakietkumjohn and S. Rittidech, “Internal flow patterns on heat transfer


characteristics of a closed-loop oscillating heat-pipe with check valves using
ethanol and a silver nano-ethanol mixture,” Exp. Therm. Fluid Sci., vol. 34, no. 8,
pp. 1000–1007, 2010.

[42] P. Charoensawan and P. Terdtoon, “Thermal performance correlation of


horizontal closed-loop oscillating heat pipes,” Appl. Therm. Eng., vol. 28, pp.
460–466, 2008.

[43] T. Katpradit, T. Wongratanaphisan, P. Terdtoon, S. Ritthidech, P. Chareonsawan,


and S. Waowaew, “Effect of aspect ratios and bond number on internal flow
patterns of closed end oscillating heat pipe at critical state,” in Proceedings of the
13th International Heat Pipe Conference, Shanghai, China, 2004.

[44] Y. Zhang and A. Faghri, “Advances and unsolved issues in pulsating heat pipes,”
Heat Transf. Eng., vol. 29, no. 1, pp. 20–44, 2008.

[45] H. Yang, S. Khandekar, and M. Groll, “Operational limit of closed loop pulsating
heat pipes,” Appl. Therm. Eng., vol. 28, no. 1, pp. 49–59, 2008.

[46] H. Yang, S. Khandekar, and M. Groll, “Performance characteristics of pulsating


heat pipes as integral thermal spreaders,” Int. J. Therm. Sci., vol. 48, no. 4, pp.
815–824, Apr. 2009.

[47] M. Saha, C. M. Feroz, F. Ahmed, and T. Mujib, “Thermal performance of an


open loop closed end pulsating heat pipe,” Heat Mass Transf. und
Stoffuebertragung, vol. 48, no. 2, pp. 259–265, 2012.

[48] B. Holley and A. Faghri, “Analysis of pulsating heat pipe with capillary wick and
varying channel diameter,” Int. J. Heat Mass Transf., vol. 48, no. 13, pp. 2635–
2651, Jun. 2005.

[49] S. Liu, J. Li, X. Dong, and H. Chen, “Experimental study of flow patterns and
improved configurations for pulsating heat pipes,” J. Therm. Sci., vol. 16, no. 1,
pp. 56–62, Mar. 2007.

97
[50] L. Quan and L. Jia, “Experimental study on heat transfer characteristic of plate
pulsating heat pipe,” in ASME 2009 Second International Conference on
Micro/Nanoscale Heat and Mass Transfer, Volume 3, Shanghai, China, 2009, pp.
361–366.

[51] Q. Cai, R.-L. Chen, and C.-L. Chen, “An investigation of evaporation, boiling,
and heat transport performance in pulstating heat pipe,” in ASME 2002
International Mechanical Engineering Congress & Exposition, New Orleans,
Louisiana, USA, 2002, vol. 7, pp. 99–104.

[52] P. Charoensawan, S. Khandekar, M. Groll, and P. Terdtoon, “Closed loop


pulsating heat pipes - Part A: Parametric experimental investigations,” Appl.
Therm. Eng., vol. 23, no. 16, pp. 2009–2020, 2003.

[53] Y. H. Lin, S. W. Kang, and T. Y. Wu, “Fabrication of polydimethylsiloxane


(PDMS) pulsating heat pipe,” Appl. Therm. Eng., vol. 29, no. 2–3, pp. 573–580,
2009.

[54] P. Charoensawan and P. Terdtoon, “Thermal performance of horizontal closed-


loop oscillating heat pipes,” Appl. Therm. Eng., vol. 28, no. 5–6, pp. 460–466,
2008.

[55] J. Li and L. Yan, “Experimental research on heat transfer of pulsating heat pipe,”
J. Therm. Sci., vol. 17, no. 2, pp. 181–185, 2008.

[56] G. Mahajan, H. Cho, S. M. Thompson, H. Rupp, and K. Muse, “Oscillating heat


pipes for waste heat recovery in HVAC systems,” in ASME 2015 International
Mechanical Engineering Congress & Exposition, San Antonio, Texas, USA, 2015.

[57] Y. Zhou and W. Qu, “Experimental study on capillary structure and size effects of
pulsating heat pipes,” K. Cheng Je Wu Li Hsueh Pao/Journal Eng. Thermophys.,
vol. 28, no. 4, pp. 646–648, 2007.

[58] G. W. Xiahou, C. Y. Yang, and L. L. Chen, “Heat transfer performance of flat


plate pulsating heat pipe with double sides rectangular or triangular channel,”
Zhongnan Daxue Xuebao (Ziran Kexue Ban)/Journal Cent. South Univ. (Science
Technol., vol. 43, no. 5, pp. 1984–1989, 2012.

[59] M. Aboutalebi, A. M. Nikravan Moghaddam, N. Mohammadi, and M. B. Shafii,


“Experimental investigation on performance of a rotating closed loop pulsating
heat pipe,” Int. Commun. Heat Mass Transf., vol. 45, pp. 137–145, Jul. 2013.

[60] B. Borgmeyer, C. Wilson, R. A. Winholtz, H. B. Ma, D. Jacobson, and D.


Hussey, “Heat transport capability and fluid flow neutron radiography of three-
dimensional oscillating heat pipes,” J. Heat Transfer, vol. 132, no. 6, Jun. 2010.

98
[61] A. A. Hathaway, C. A. Wilson, and H. B. Ma, “Experimental Investigation of
Uneven-Turn Water and Acetone Oscillating Heat Pipes,” J. Thermophys. Heat
Transf., vol. 26, no. 1, pp. 115–122, Jan. 2012.

[62] Y.-H. Lin, S.-W. Kang, and H.-L. Chen, “Effect of silver nano-fluid on pulsating
heat pipe thermal performance,” Appl. Therm. Eng., vol. 28, no. 11–12, pp. 1312–
1317, Aug. 2008.

[63] L. Jia and Y. Li, “Experimental research on heat transfer characteristics of


pulsating heat pipe,” in ASME 2008 Heat Transfer Summer Conference,
Jacksonville, Florida, USA, 2008, vol. 2, pp. 375–379.

[64] P. Charoensawan and P. Terdtoon, “Thermal performance of horizontal closed-


loop oscillating heat pipes,” Appl. Therm. Eng., vol. 28, no. 5–6, pp. 460–466,
Apr. 2008.

[65] S. Thongdaeng, S. Rittidech, and B. Bubphachot, “Flow patterns and heat-transfer


characteristics of a top heat mode closed-loop oscillating heat pipe with check
valves (THMCLOHP/CV),” J. Eng. Thermophys., vol. 21, no. 4, pp. 235–247,
Nov. 2012.

[66] C. Wilson, B. Borgmeyer, R. A. Winholtz, H. B. Ma, D. L. Jacobson, D. S.


Hussey, and M. Arif, “Visual observation of oscillating heat pipes using neutron
radiography,” J. Thermophys. Heat Transf., vol. 22, no. 3, pp. 366–372, Jul. 2008.

[67] S. M. Thompson and H. B. Ma, “Effect of localized heating on three-dimensional


flat-plate oscillating heat pipe,” Adv. Mech. Eng., vol. 2, Jan. 2015.

[68] Z. J. Zuo, M. T. North, and K. L. Wert, “High heat flux heat pipe mechanism for
cooling of electronics,” IEEE Trans. Components Packag. Technol., vol. 24, no.
2, pp. 220–225, 2001.

[69] P. Cheng and H. Ma, “A Mathematical Model of an Oscillating Heat Pipe,” Heat
Transf. Eng., vol. 32, no. 11–12, pp. 1037–1046, 2011.

[70] H. Peng, P. F. Pai, and H. Ma, “Nonlinear thermomechanical finite-element


modeling, analysis and characterization of multi-turn oscillating heat pipes,” Int.
J. Heat Mass Transf., vol. 69, pp. 424–437, 2014.

[71] Wei Shao and Y. Zhang, “Thermally-induced oscillatory flow and heat transfer in
an oscillating heat pipe,” J. Enhanc. Heat Transf., vol. 18, no. 3, pp. 177–190,
2011.

[72] M. Dilawar and A. Pattamatta, “A parametric study of oscillatory two-phase flows


in a single turn Pulsating Heat Pipe using a non-isothermal vapor model,” Appl.
Therm. Eng., vol. 51, no. 1–2, pp. 1328–1338, Mar. 2013.

99
[73] S. Arabnejad, R. Rasoulian, M. B. Shafii, and Y. Saboohi, “Numerical
Investigation of the Performance of a U-Shaped Pulsating Heat Pipe,” Heat
Transf. Eng., vol. 31, no. 14, pp. 1155–1164, 2010.

[74] Y. Zhang, A. Faghri, and M. B. Shafii, “Analysis of liquid–vapor pulsating flow


in a U-shaped miniature tube,” Int. J. Heat Mass Transf., vol. 45, no. 12, pp.
2501–2508, Jun. 2002.

[75] W. Qu and H. B. Ma, “Theoretical analysis of startup of a pulsating heat pipe,”


Int. J. Heat Mass Transf., vol. 50, pp. 2309–2316, 2007.

[76] J.-S. Kim, Y.-B. Im, and N.-H. Bui, “Numerical analysis of pulsating heat pipe
based on separated flow model,” J. Mech. Sci. Technol., vol. 19, no. 9, pp. 1790–
1800, Sep. 2005.

[77] C.-M. Chiang, K.-H. Chien, H.-M. Chen, and C.-C. Wang, “Theoretical study of
oscillatory phenomena in a horizontal closed-loop pulsating heat pipe with
asymmetrical arrayed minichannel,” Int. Commun. Heat Mass Transf., vol. 39, no.
7, pp. 923–930, Aug. 2012.

[78] J. Qu, H. Wu, P. Cheng, and X. Wang, “Non-linear analyses of temperature


oscillations in a closed-loop pulsating heat pipe,” Int. J. Heat Mass Transf., vol.
52, no. 15, pp. 3481–3489, 2009.

[79] S. M. Pouryoussefi and Y. Zhang, “Numerical investigation of chaotic flow in a


2D closed-loop pulsating heat pipe,” Appl. Therm. Eng., Jan. 2016.

[80] S. Khandekar, “Pulsating heat pipe based heat exchangers,” Proc. 21st Int. Symp.
Transp. Phenom., no. 3, pp. 2–5, 2010.

[81] Y. F. Maydanik, V. I. Dmitrin, and V. G. Pastukhov, “Compact cooler for


electronics on the basis of a pulsating heat pipe,” Appl. Therm. Eng., vol. 29, no.
17–18, pp. 3511–3517, Dec. 2009.

[82] Y. Miyazaki, “Cooling of notebook PCs by flexible oscillating heat pipes,” in


ASME 2005 Pacific Rim Technical Conference and Exhibition on Integration and
Packaging of MEMS, NEMS, and Electronic Systems, 2005, pp. 65–69.

[83] G. Karimi and J. R. Culham, “Review and assessment of Pulsating Heat Pipe
mechanism for high heat flux electronic cooling,” in The Ninth Intersociety
Conference on Thermal and Thermomechanical Phenomena In Electronic
Systems (IEEE Cat. No.04CH37543), 2004, vol. 2, pp. 52–59.

[84] J. Clement and X. Wang, “Experimental investigation of pulsating heat pipe


performance with regard to fuel cell cooling application,” Appl. Therm. Eng., vol.
50, no. 1, pp. 268–274, Jan. 2013.

100
[85] K. Natsume, T. Mito, N. Yanagi, H. Tamura, T. Tamada, K. Shikimachi, N.
Hirano, and S. Nagaya, “Heat transfer performance of cryogenic oscillating heat
pipes for effective cooling of superconducting magnets,” Cryogenics (Guildf).,
vol. 51, no. 6, pp. 309–314, Jun. 2011.

[86] K. Natsume, T. Mito, N. Yanagi, H. Tamura, T. Tamada, K. Shikimachi, N.


Hirano, and S. Nagaya, “Development of cryogenic oscillating heat pipe as a new
device for indirect/conduction cooled superconducting magnets,” Appl.
Supercond. IEEE Trans., vol. 22, no. 3, Jun. 2012.

[87] M. Arab, M. Soltanieh, and M. B. Shafii, “Experimental investigation of extra-


long pulsating heat pipe application in solar water heaters,” Exp. Therm. Fluid
Sci., vol. 42, pp. 6–15, Oct. 2012.

[88] S. Rittidech and S. Wannapakne, “Experimental study of the performance of a


solar collector by closed-end oscillating heat pipe (CEOHP),” Appl. Therm. Eng.,
vol. 27, no. 11–12, pp. 1978–1985, Aug. 2007.

[89] K.-B. Nguyen, S.-H. Yoon, and J. H. Choi, “Effect of working-fluid filling ratio
and cooling-water flow rate on the performance of solar collector with closed-
loop oscillating heat pipe,” J. Mech. Sci. Technol., vol. 26, no. 1, pp. 251–258,
Jan. 2012.

[90] S. H. Kargar, M. Ghiasei, M. Mirzaei, and R. S. Nejad, “Experimental


investigation of the effect of using pulsating heat pipes on the performance of a
solar air heater,” in ASME 3rd International Conference on Mechanical and
Electrical Technology, China, 2011.

[91] S. Rittidech, A. Donmaung, and K. Kumsombut, “Experimental study of the


performance of a circular tube solar collector with closed-loop oscillating heat-
pipe with check valve (CLOHP/CV),” Renew. Energy, vol. 34, no. 10, pp. 2234–
2238, Oct. 2009.

[92] H. Kargarsharifabad, S. J. Mamouri, M. B. Shafii, and M. T. Rahni,


“Experimental investigation of the effect of using closed-loop pulsating heat pipe
on the performance of a flat plate solar collector,” J. Renew. Sustain. Energy, vol.
5, no. 1, p. 013106, Jan. 2013.

[93] R. T. Dobson, “An open oscillatory heat pipe water pump,” Appl. Therm. Eng.,
vol. 25, no. 4, pp. 603–621, Mar. 2005.

[94] A. Nuntaphan, S. Vithayasai, N. Vorayos, N. Vorayos, and T. Kiatsiriroat, “Use


of oscillating heat pipe technique as extended surface in wire-on-tube heat
exchanger for heat transfer enhancement,” Int. Commun. Heat Mass Transf., vol.
37, no. 3, pp. 287–292, Mar. 2010.

101
[95] G. Burban, V. Ayel, A. Alexandre, P. Lagonotte, Y. Bertin, and C. Romestant,
“Experimental investigation of a pulsating heat pipe for hybrid vehicle
applications,” Appl. Therm. Eng., vol. 50, no. 1, pp. 94–103, Jan. 2013.

[96] P. Supirattanakul, S. Rittidech, and B. Bubphachot, “Application of a closed-loop


oscillating heat pipe with check valves (CLOHP/CV) on performance
enhancement in air conditioning system,” Energy Build., vol. 43, no. 7, pp. 1531–
1535, Jul. 2011.

[97] S. Rittidech, W. Dangeton, and S. Soponronnarit, “Closed-ended oscillating heat-


pipe (CEOHP) air-preheater for energy thrift in a dryer,” Appl. Energy, vol. 81,
pp. 198–208, 2005.

[98] P. Meena, S. Rittidech, and N. Poomsa-ad, “Closed-loop oscillating heat-pipe


with check valves (CLOHP/CVs) air-preheater for reducing relative humidity in
drying systems,” Appl. Energy, vol. 84, no. 4, pp. 363–373, Apr. 2007.

[99] J. G. Monroe, E. S. Vasquez, Z. S. Aspin, K. B. Walters, M. J. Berg, and S. M.


Thompson, “Electromagnetic induction by ferrofluid in an oscillating heat pipe,”
Appl. Phys. Lett., vol. 106, no. 26, pp. 0–4, 2015.

[100] S. Khandekar and A. Gupta, “Embedded pulsating heat pipe radiators,” in 14th
International Heat Pipe Conference, Florianopolis, Brazil, 2007, vol. 1, no. 1, pp.
4–9.

[101] H. W. Coleman and W. G. Steele, Experimentation and Uncertainty Analysis for


Engineers, 2nd ed. John Wiley & Sons, Inc., 1999.

[102] H. Cho and N. Fumo, “Uncertainty analysis for dimensioning solar photovoltaic
arrays,” in 6th International Conference on Energy Sustainability & 10th Fuel
Cell Science, Engineering and Technology Conference, San Diego, CA, 2012, pp.
1–5.

[103] International Organization for Standardization (ISO), “Guide to the Expression of


Uncertainty in Measurement (corrected and reprinted 1995),” Geneva,
Switzerland, 1993.

[104] R. McKee, E. Frank, J. Heath, D. Owen, R. Przygoda, G. Trimmer, and F.


Whitman, “Toxicology of n-pentane (CAS no. 109–66–0),” J. Appl. Toxicol., vol.
18, no. 6, pp. 431–442, 1998.

[105] Pollution Prevention and Abatement- Ozone-Depleting Substance: Alternative,


World Bank Group, 1998, no. July. .

[106] S. Khandekar, “Pulsating heat pipe based heat exchangers,” Proc. 21st Int. Symp.
Transp. Phenom., no. iii, pp. 2–5, 2010.

102
[107] W. M. Kays and A. L. London, Compact heat exchanger, 2nd ed. New York:
McGraw-Hill Book Co, 1964.

[108] T. L. Bergman, A. S. Lavine, and F. P. Incropera, Fundamentals of Heat and


Mass Transfer, 7th Edition. John Wiley & Sons, 2011.

[109] S. W. Churchill and M. Bernstein, “A correlating equation for forced convection


from gases and liquids to a circular cylinder in crossflow,” ASME, Trans. Ser. C-
Journal Heat Transf., vol. 99, pp. 300–306, 1977.

[110] F. P. Incropera, Fundamentals of Heat and Mass Transfer. John Wiley & Sons
Australia, Limited, 1992.

[111] S. Y. Yoo, H. K. Kwon, and J. H. Kim, “A study on heat transfer characteristics


for staggered tube banks in cross-flow,” J. Mech. Sci. Technol., vol. 21, no. 3, pp.
505–512, 2007.

[112] H. B. Ma, B. Borgmeyer, P. Cheng, and Y. Zhang, “Heat transport capability in


an oscillating heat pipe,” J. Heat Transfer, vol. 130, no. 8, pp. 081501–081501–7,
2008.

[113] D. Liu, G.-F. Tang, F.-Y. Zhao, and H.-Q. Wang, “Modeling and experimental
investigation of looped separate heat pipe as waste heat recovery facility,” Appl.
Therm. Eng., vol. 26, no. 17–18, pp. 2433–2441, Dec. 2006.

[114] Y. H. Yau and A. S. Tucker, “The performance study of a wet six-row heat-pipe
heat exchanger operatiing in tropical buildings,” Int. J. energy Res., vol. 27, pp.
187–202, 2003.

[115] R. W. Schafer, “What Is a Savitzky-Golay Filter? [Lecture Notes],” Signal


Processing Magazine, IEEE, vol. 28, no. 4. pp. 111–117, 2011.

[116] E. Azad and F. Geoola, “A design procedure for gravity-assisted heat pipe heat
exchanger,” J. Heat Recover. Syst., vol. 4, no. 2, pp. 101–111, Jan. 1984.

[117] S. H. Noie, “Investigation of thermal performance of an air-to-air thermosyphon


heat exchanger using ε-NTU method,” Appl. Therm. Eng., vol. 26, no. 5–6, pp.
559–567, Apr. 2006.

[118] D. G. Rich, “The effect of fin spacing on the heat transfer and friction
performance of multi-row, plate fin-and-tube heat exchangers,” AsHRAE Trans,
vol. 79, no. 2, pp. 137–145, 1973.

[119] W. M. Kays and A. L. London, Compact heat exchangers: a summary of basic


heat transfer and flow friction design data. National Press, 1955.

103
[120] ASHRAE, “2013 ASHRAE Handbook- Fundamentals,” vol. 30329, no. 404,
Atlanta, GA, p. 404.

[121] W. Marion and K. Urban, “User’s Manual for TMY2s, National Renewable
Energy Laboratory,” Golden, Colorado, USA, 1995.

[122] “U.S. Energy Information Administration, 2013, Electric Power Annual 2012,
U.S. Department of Energy, Washington, DC, USA,” Washington, DC, USA.

[123] M. Kottek, J. Grieser, C. Beck, B. Rudolf, and F. Rubel, “World Map of the
Köppen-Geiger climate classification updated,” Meteorol. Zeitschrift, vol. 15, no.
3, pp. 259–263, Jun. 2006.

104

You might also like