integrated absorption refrigeration and thermoelectric

130
- The Pennsylvania State University The Graduate School Integrated Absorption Refrigeration and Thermoelectric Based Cascaded Waste Heat Recovery A Dissertation in Mechanical Engineering by Shahzaib B. Abbasi © 2020 Shahzaib B. Abbasi Submitted in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy December 2020

Upload: others

Post on 18-Dec-2021

13 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Integrated Absorption Refrigeration and Thermoelectric

-

The Pennsylvania State University

The Graduate School

Integrated Absorption Refrigeration and

Thermoelectric Based Cascaded Waste

Heat Recovery

A Dissertation in

Mechanical Engineering

by

Shahzaib B. Abbasi

© 2020 Shahzaib B. Abbasi

Submitted in Partial Fulfillment

of the Requirements

for the Degree of

Doctor of Philosophy

December 2020

Page 2: Integrated Absorption Refrigeration and Thermoelectric

ii

The dissertation of Shahzaib B. Abbasi was reviewed and approved by the following:

Alexander S. Rattner Assistant Professor of Mechanical Engineering

Dissertation Adviser

Chair of Committee

James D. Freihaut

Professor of Architectural Engineering

Bed Poudel

Associate Research Professor of Material Science and Engineering Special Committee Member

Stephen Lynch

Associate Professor of Mechanical Engineering

Matthew Rau

Assistant Professor of Mechanical Engineering

Daniel Haworth

Associate Head for Graduate Programs

Page 3: Integrated Absorption Refrigeration and Thermoelectric

iii

Abstract

Waste Heat Recovery (WHR) methods that use only a single process to recover heat may

be termed as single-pathway WHR methods. Common examples are technologies based on

Thermoelectric Generators (TEGs), Organic Ranking Cycles (ORC), or absorption-based

Thermally Activated Refrigeration (TAR). Most single-pathway WHR methods can only

effectively harness heat sources in certain temperature ranges. A system in which heat is recovered

as it cascades from a higher temperature to a lower temperature can be termed a cascaded WHR

system. Cascaded WHR methods are of a particular interest for applications like refrigerated

transport vehicles and industrial carburizing furnace operations where multiple outputs are needed,

such as electricity, refrigeration, and process heating.

This dissertation proposes a novel, integrated cascaded WHR system that uses a

temperature matched approach that can increase process efficiency by providing electrical power

and refrigeration through WHR. To investigate the potential of this approach, cycle models,

thermoeconomic studies, and an experimental investigation of a cascaded approach of WHR, is

performed.

Thermoelectric generators (TEGs) are solid state devices that perform at their optimum

when the temperature difference across their junctions is high. The efficiency of a TEG scales with

the hot junction temperature. However, the efficiency of TEGs is low (~5%), hence the TEG

pathway of WHR only harvests a small portion of a high-availability waste heat. Absorption

Refrigeration Systems (ARS), however, can operate efficiently with low temperature heat sources.

In this study, TEGs and Absorption Refrigeration (AR) subsystems are integrated using a coupling

fluid, such that high-grade waste heat cascades through the TEGs, and the low-grade waste heat

rejected by the TEG subsystem is used to operate an absorption refrigeration subsystem. Both

Page 4: Integrated Absorption Refrigeration and Thermoelectric

iv

systems have previously been investigated as single-pathway waste heat recovery methods but

owing their respective temperature ranges in which they operate efficiently (high for TEGs and

mid-to-low for Absorption Refrigeration), they are well-suited for integration.

A thermoeconomic study of the proposed cascaded WHR system is performed, and the

results from that study are compared to ORC based WHR systems. The study concluded that the

payback periods for TEGs and AR based cascaded WHR systems are comparable to those of ORC

based WHR systems.

In order to investigate the challenges of integrated, cascaded WHR and to provide a basis

for the thermoeconomic feasibility analysis, a 1/10th scale experimental facility was built based on

the results from an engineering model created for the refrigerated transport application. This facility

was experimentally tested at different inlet temperatures and flow rates for simulated exhaust and

coupling fluid, to simulate the operating conditions of the vehicle application.

An important part of the facility construction was the design and development of the Heat

Acquisition Unit which extracts waste heat from the exhaust stream and transfers it through the

TEGs to the coupling fluid, and the coupling fluid delivers heat to the absorption subsystem. Using

this experimental approach, different design tradeoffs like electrical power output vs. cooling

delivery and cost vs. WHR efficiency were explored. The experimental results were used for the

validation of subsystem models.

Page 5: Integrated Absorption Refrigeration and Thermoelectric

v

Table of Contents

List of Figures........................................................................................................................... vii

List of Tables ............................................................................................................................. ix

Acknowledgements ..................................................................................................................... x

Introduction ........................................................................................................... 1

1.1 Waste Heat Recovery (WHR) applications ................................................................... 2

1.1.1 Single-pathway WHR .......................................................................................... 3

1.1.2 Integration of TEG and ARS based pathways of WHR ......................................... 8

1.1.3 Integrated Cascaded Waste Heat Recovery Systems ........................................... 10

1.2 Heat Acquisition Unit (HAU) .................................................................................... 11

1.3 Summary of prior research into Waste Heat Recovery methods .................................. 13

1.3.1 Research Needs in Cascaded WHR .................................................................... 15

1.4 Goals of the Present research and Dissertation ........................................................... 15

Literature Review ................................................................................................ 20

2.1 Single-pathway, single-grade WHR studies ................................................................ 22

2.2 Cascaded WHR studies .............................................................................................. 25

2.3 Discussion ................................................................................................................. 26

Thermoeconomic Studies ..................................................................................... 27

3.1 Thermoeconomic Analysis of Vehicle and Furnace Application ................................. 28

3.1.1 System Modeling Results ................................................................................... 28

3.1.2 Thermoeconomic Analysis ................................................................................. 30

3.1.3 Conclusions ....................................................................................................... 37

3.2 ARS-TEG vs. ORC-VCC; Thermoeconomic Comparison .......................................... 38

3.2.1 ORC-VCC System Description .......................................................................... 39

3.2.2 Exhaust Temperature ......................................................................................... 40

3.2.3 Closure Parameters for Heat Exchangers ............................................................ 43

3.2.4 Isentropic Efficiency Assumptions and Working Fluids ...................................... 44

3.2.5 Cycle Model for ORC-VCC ............................................................................... 45

3.2.6 Results ............................................................................................................... 49

3.2.7 Economic Analysis ............................................................................................ 52

3.2.8 Discussion ......................................................................................................... 55

Heat Acquisition Unit Design, Development and Experimentation ....................... 57

Page 6: Integrated Absorption Refrigeration and Thermoelectric

vi

4.1 HAU Design Considerations ...................................................................................... 58

4.2 HAU Simple Engineering Model ............................................................................... 61

4.2.1 TEG Model ........................................................................................................ 61

4.2.2 Fin Array Resistance Model ............................................................................... 62

4.2.3 Spreading Thermal Resistance ........................................................................... 63

4.2.4 Model for Heat Lost to the Ambient ................................................................... 64

4.2.5 Stage-wise Total Resistance Calculations ........................................................... 65

4.2.6 Heat Transfer Calculations ................................................................................. 67

4.3 HAU Iterative Development and Experimental Procedure .......................................... 69

4.3.1 HAU Experimental Subsystem Description ........................................................ 71

4.3.2 Design improvements to HAU-2 based on HAU-1 performance ......................... 72

System Description and Cycle Modeling .............................................................. 80

5.1 System Description .................................................................................................... 81

5.2 Cycle Model .............................................................................................................. 84

5.2.1 Conservation Laws............................................................................................. 84

5.2.2 Closure Parameters ............................................................................................ 86

5.2.3 Results ............................................................................................................... 88

Experimental Results ........................................................................................... 90

6.1 Experimental Facility ................................................................................................. 91

6.2 Instrumentation .......................................................................................................... 92

6.3 Exergetic Efficiency .................................................................................................. 93

6.4 Thermoelectric Power ................................................................................................ 95

6.5 Comparison of HAU Model Predictions and Experimental Results............................. 96

6.6 Discussion ............................................................................................................... 101

Conclusions and Recommendations for Future Research .................................... 104

7.1 ARS-TEG thermoeconomic studies ......................................................................... 105

7.2 HAU design ............................................................................................................. 106

7.3 HAU Model prediction and experimental results ...................................................... 107

7.4 Recommendations for future research ...................................................................... 107

7.4.1 Advanced HAU design .................................................................................... 108

7.4.2 Experimental Investigations ............................................................................. 108

Appendix: Velocity Measurement Calibration ......................................................................... 109

Bibliography ........................................................................................................................... 111

Page 7: Integrated Absorption Refrigeration and Thermoelectric

vii

List of Figures

Figure 1.1 Waste heat recovery categorization according to the mode of WHR ............................ 3

Figure 1.2 Single-pathway WHR using Thermoelectric generators (TEGs) .................................. 4

Figure 1.3 Single-pathway WHR using Absorption Refrigeration System (ARS) ......................... 6

Figure 1.4 ARS and TEG performance at different WHR Source Temperature ............................ 8

Figure 1.5 (a) Individual, single-pathway WHR systems, (b) More compact, integrated cascaded

WHR system that eliminates the need for additional heat acquisition unit and air-coupled HX ... 11

Figure 1.6 An illustration of a typical Heat Acquisition Unit (HAU) used for integrated WHR .. 12

Figure 2.1 Venn-diagram showing different types of WHR Systems .......................................... 21

Figure 3.1 Cascaded WHR cycle diagram ................................................................................. 28

Figure 3.2 NPV versus time for the vehicle application ............................................................. 35

Figure 3.3 NPV versus time for the carburizing furnace application ........................................... 36

Figure 3.4 NPV versus time for the carburizing furnace application ........................................... 37

Figure 3.5 Schematic for ORC-VCC based WHR system .......................................................... 40

Figure 3.6 (a) Cumulative frequency of exhaust temperatures (b) Cumulative frequency of

available exhaust heat to operate a 5kW absorption refrigeration system for different driving

cycles [15] ................................................................................................................................ 41

Figure 3.7 Real-world velocity and SCR temperature data for a heavy-duty diesel vehicle [51] . 42

Figure 3.8 Frequenc distribution of exhasut temperatures for a heavy-duty refrigeated transport

truck [51] .................................................................................................................................. 42

Figure 4.1 Thermal Circuit for Heat Lost model ........................................................................ 65

Figure 4.2 Stage-1 Thermal Resistance Model ........................................................................... 66

Figure 4.3 Stage-2 Thermal Resistance Model ........................................................................... 67

Figure 4.4 (a) Cover-plate sub-assembly (b) Base-plate subassembly ........................................ 69

Figure 4.5 (a) TEGs mounted on the outside the HAU (b) Coupling plate mounted on top of the

TEGs (c) Complete assembly of the HAU-1 with oil-block cover plate ..................................... 70

Figure 4.6 Assembled HAU-1 ................................................................................................... 71

Figure 4.7 Experimental setup for HAU subsystem evaluation .................................................. 72

Figure 4.8 Equivalent Thermal Circuits:(a) No heat-spreader: Largest temperature drop occurs

between the air-side fins and TEG, (b) Copper heat-spreader: Spreading resistance drops and the

largest temperature drop occurs between the two sides of the TEG ............................................ 74

Page 8: Integrated Absorption Refrigeration and Thermoelectric

viii

Figure 4.9 (a) Thermal Oil side view of HAU-2 (b) Air side view of HAU-2 (c) Illustration of

HAU-2 thermal oil flow path .................................................................................................... 76

Figure 4.10 TEG Power Output Comparison between HAU-1 and HAU-2 ................................ 78

Figure 4.11 Oil Outlet Temperature Comparison between HAU-1 and HAU-2 .......................... 79

Figure 5.1 Cycle diagram for the proposed TEG and ARS based Integrated cascaded WHR

system ...................................................................................................................................... 81

Figure 6.1 Experimental Facility: (a) Experimental Facility highlighting experiment controls

(including DC load) and thermal system with the location of HAU highlighted, (b) Picture shows

the HAU-1 model and simulated exhaust (hot air blower) .......................................................... 91

Figure 6.2 Exergetic Efficiency of HAU-2 at different inlet conditions ...................................... 94

Figure 6.3 TEG Power produced by HAU-2 at different inlet conditions (uncertainty of ±0.2%) 95

Figure 6.4 Qoil Model vs Experiment Comparison ..................................................................... 98

Figure 6.5 Qair Model vs Experiment Comparison ..................................................................... 98

Figure 6.6 Toil Model vs Experiment Comparison ...................................................................... 99

Figure 6.7 Tair Model vs Experiment Comparison ...................................................................... 99

Figure 6.8 ΔPexp vs ΔPmodel comparison .................................................................................... 100

Figure A.1 Stage-2 Airflow, immediately at the exit of the HAU, is contracted due to the shape of

the HAU. This contraction leads to an increase in the velocity at the location of temperature and

velocity measurement ............................................................................................................. 109

Figure B.2 Stage-2 Vcorr (Velocity measurement downstream of original measurement) vs Vmsrd

(original measurement) ........................................................................................................... 110

Page 9: Integrated Absorption Refrigeration and Thermoelectric

ix

List of Tables

Table 1.1 Key parameters of applications considered for integrated WHR study ......................... 9

Table 1.2 Classification of WHR methods with respect to source type and grade ...................... 14

Table 3.1 The UA values for refrigerated vehicle and industrial carburizing furnace ................. 30

Table 3.2 Comparison of payback periods with previous studies ............................................... 31

Table 3.3 Comparison of payback periods with previous studies ............................................... 38

Table 3.4 Heat Exchanger Closure Parameters for ORC-VCC .................................................. 43

Table 3.5 Working Fluid pairs and isentropic efficiencies for different studies .......................... 44

Table 3.6 Isentropic Efficiencies of ORC-VCC components ..................................................... 45

Table 3.7 Properties at various state points for ORC-VCC ........................................................ 49

Table 3.8 Important properties of different components of ORC-VCC system........................... 50

Table 3.9 Properties at various state points for ARS-TEG ......................................................... 50

Table 3.10 Important properties of different components of ARS-TEG system.......................... 51

Table 3.11 U values for HXs....................................................................................................... 51

Table 3.12 Areas of the ORC-VCC HXs ................................................................................... 52

Table 3.13 Areas of the ARS-TEG HXs.................................................................................... 52

Table 3.14 Component cost relation matrix .............................................................................. 53

Table 3.15 Capital costs of ORC-VCC system .......................................................................... 54

Table 3.16 Capital cost of the ARS-TEG system ....................................................................... 54

Table 4.1 HAU-1 component details ......................................................................................... 70

Table 4.2 Results from HAU-1 experimental runs ..................................................................... 73

Table 4.3 Test Matrix for HAU-2 ............................................................................................. 77

Table 5.1 Heat Exchanger Closure Parameters for ARS-TEG ................................................... 86

Table 5.2 Summary of model results ......................................................................................... 88

Table 5.3 Point-wise results for ARS-TEG cycle ...................................................................... 89

Table 6.1 Control Matrix for Experimental Facility................................................................... 92

Table 6.2 Measurement types and uncertainties for ARS-TEG experimental facility ................. 93

Page 10: Integrated Absorption Refrigeration and Thermoelectric

x

Acknowledgements

Firstly, I would like to thank my advisor, Dr. Alexander S. Rattner, who has guided me

throughout my PhD research. He has been patient and extremely kind. I would also like to thank

all my past and present lab mates, especially Yue Cao and Chris Greer, for helping me with my

research.

I would also like to thank the U.S. Department of State, the Fulbright Program, the

International Institute of Education, the United States Educational Foundation in Pakistan and the

Higher Education Commission of Pakistan for their financial support through a Fulbright Grant.

I am really thankful to my committee members: Dr. James Freihaut, Dr. Bed Poudel, Dr.

Stephen Lynch and Dr. Matthew Rau.

Finally, I would like to thank my wonderful wife Britt, my parents and my sisters for their

ever-present, unwavering love and support.

Page 11: Integrated Absorption Refrigeration and Thermoelectric

xi

Disclaimer

Any opinions, findings, and conclusions or recommendations expressed in this publication

are those of the author and do not necessarily reflect the views of the U.S. Department of State, the

Fulbright Program, the International Institute of Education, the United States Educational

Foundation in Pakistan and the Higher Education Commission of Pakistan.

Page 12: Integrated Absorption Refrigeration and Thermoelectric

1

Introduction

Page 13: Integrated Absorption Refrigeration and Thermoelectric

2

1.1 Waste Heat Recovery (WHR) applications

Dependence on fossil fuels for energy and sustained population growth have resulted in

large amounts of greenhouse gas emissions [1]. This has led to a rise in global temperatures and

environmental degradation. Electricity production and transportation account for the major portion

of greenhouse gas emissions in the US, each contributing about 30% of the national total [2].

Industries like iron and steel, textiles, petroleum, and fertilizer production account for almost 40%

of global energy uses, and contribute to about 37% of global greenhouse emissions [3]. Two-thirds

of energy input for electricity generation in the US is lost as waste heat, and 12.5% of this generated

electricity is used for low availability tasks like water, space heating and low-temperature process

heating [4].

Waste Heat Recovery (WHR) technologies can help reduce emissions by increasing energy

efficiency while also reducing operating costs. Waste heat sources are often graded according to

their temperatures. For example, Jouhara et al. [5] surveyed industrial waste heat sources and

classified them as high-grade (>400°C), mid-grade (100°C – 400°C) or low-grade (<100°C) waste

heat .WHR technologies can be classified according to the mode of energy conversion, Figure 1.1

illustrates this classification and the typical technologies associated with these modes of energy

conversion.

According to a report from the US Department of Energy [6], the most frequently employed

use for WHR in US industries is process heating – specifically, the use of heat exchangers like

regenerators, recuperators and preheaters.

At lower temperatures in industrial exhaust waste heat streams, water vapor condensation

can lead to corrosion and, to achieve any significant heat transfer, large heat transfer surfaces are

Page 14: Integrated Absorption Refrigeration and Thermoelectric

3

needed. Combined with a source to end use temperature mismatch, low-to-mid grade waste heat is,

typically, not utilized.

WASTE

HEAT

HEATING COOLINGPOWER

GENERATION

• Space Heating

• Recuperators

• Regenerators

• Preheaters

• Heat pumps and

Transformers

• Thermally Activated

Refrigeration (TAR):

Absoption/Adsorption

• Desiccant

dehumidification

• Steam Rankine Cycles

(SRC)

• Organic Rankine

Cycles (ORC)

• Thermoelectric

Generators (TEGs)

Figure 1.1 Waste heat recovery categorization according to the mode of WHR

1.1.1 Single-pathway WHR

WHR systems that recover heat using one specific technology are called single-pathway

WHR systems. An example of such a system is a typical single-stage Thermoelectric Generator

(TEG)-based WHR system. TEGs are solid state devices that operate on the Seebeck effect in which

heat flow along dissimilar conductors (N-type and P-type) effects an electric current. Figure 1.2

shows a TEG based single-pathway waste heat recovery system.

Page 15: Integrated Absorption Refrigeration and Thermoelectric

4

TH

TC

Waste Heat in (Qin)

RL

Heat Rejected (Qout)

I

N P N P N P

Figure 1.2 Single-pathway WHR using Thermoelectric generators (TEGs)

The electrical power produced by a TEG based WHR system can be approximated using

Equation 1.1 [7]:

𝑷𝑻𝑬𝑮 = (𝑵𝑻𝑬𝑮𝑺∆𝑻𝑻𝑬𝑮)𝟐/(4𝑵𝑻𝑬𝑮𝑹𝑻𝑬𝑮) 1.1

Here, PTEG is the maximum power delivered by the TEGs to the load, in W, with resistance

RL = NTEG •RTEG in Ohms. NTEG is the number of TEGs modules in series, S is the Seebeck

Coefficient of a TEG module in V/K, ΔTTEG is the temperature difference across the TEGs in K,

Page 16: Integrated Absorption Refrigeration and Thermoelectric

5

and RTEG is the electrical resistances of one TEG module in Ohms. The model presented in Equation

1 assumes that S, RTEG and RL are constants and independent of temperature.

As can be seen from Equation 1.1, for a TEG based WHR system, the amount of waste

heat converted to electric power strongly depends on the temperature difference between the hot

and cold side of the TEGs. Therefore, TEG based, single-pathway, WHR systems can typically

only effectively harness high grade waste heat. Such systems are typically designed to reject heat

at the lowest feasible temperature (e.g., to the ambient) to maximize conversion efficiency.

Absorption Refrigeration Systems (ARS) are a type of Thermally Activated Refrigeration

(TAR) system that uses a refrigerant and solvent working fluid pair for operation (e.g., NH3-H2O,

H2O-LiBr, NH3+LiNO3). Figure 1.3 shows a schematic representation of a single effect ARS cycle.

The operation of a typical single-effect absorption refrigeration system is described here

assuming the NH3-LiNO3 working fluids under investigation in the present study. Waste heat is

supplied to the desorber (points 1→2), heating the dilute NH3-LiNO3 salt solution and desorbing

refrigerant vapor (NH3) from the solution along the desorber. The exiting concentrated salt solution

and refrigerant are separated in a separator tank. The concentrated solution enters the recuperative

solution heat exchanger (SHX) and cools as it preheats the dilute solution (Points 3→4).

Downstream of the SHX, the concentrated solution expands to a lower pressure through a valve

(Points 4→5) before entering the absorber at Point 6.

Page 17: Integrated Absorption Refrigeration and Thermoelectric

6

Figure 1.3 Single-pathway WHR using Absorption Refrigeration System (ARS)

The refrigerant vapor exiting the separator tank liquefies as it rejects its heat to the ambient

in the condenser (Points 9→10). The liquefied refrigerant is recuperatively subcooled as it flows

through the refrigerant pre-cooler (RPC) between Points 11→12 and it is then expanded through a

valve (Points 12→13). The refrigerant evaporates in the evaporator delivering cooling to the

conditioned space (Points 13→14). The refrigerant continues through the RPC, precooling the

refrigerant liquid between Point 14→15.

The refrigerant mixes with the low-pressure concentrated solution at the inlet of the

absorber (Point 5). The vapor absorbs into the solution in the absorber while rejecting heat to the

ambient (Points 6→7). The dilute solution exits the absorber, is pumped to a higher pressure (Points

Page 18: Integrated Absorption Refrigeration and Thermoelectric

7

7→8), and continues through the SHX where it is recuperatively preheated (Points 8→1),

completing the cycle.

The performance of a refrigeration system is typically rated using the Coefficient of

Performance (COP):

𝑪𝑶𝑷 = 𝑸𝒆𝒗𝒂𝒑/𝑸𝒊𝒏 1.2

Where Qevap is the cooling load in W and Qin is the heat input to the system in W.

The COP of ARS is typically lower than vapor compression refrigeration systems (0.7 for

ARS, compared to 3.5 for VCC) [8]. However, compared to conventional vapor compression

refrigeration systems, ARSs do not require a large power consuming compressor. Instead, they use

a pump which consumes much less electrical power (~20W for a pump and ~1kW for a

compressor). This is due to the fact that the refrigerant in an ARS exists as a solution in a salt-

solution pair, whereas the refrigerant in a VCC system is a gas that requires compression.

Typically, >95% of the input energy to ARSs is in the form of desorber heating to thermally

separate the refrigerant form the salt solution. Therefore, ARS systems are frequently used where

a heat source is readily available and low cost and refrigeration is required, but electricity may be

limited or expensive (e.g. off-grid solar absorption refrigeration [9], in refrigerated trucks [10], for

industrial waste heat recovery [11]).

Absorption refrigeration-based single pathway WHR methods have received renewed

interest in recent years as promising technologies for harnessing low-grade waste heat (60°C –

140°C) [12].

Page 19: Integrated Absorption Refrigeration and Thermoelectric

8

1.1.2 Integration of TEG and ARS based pathways of WHR

Figure 1.4 shows a comparison of COP of a representative water/lithium bromide based

ARS [13] (~6°C cooling delivery temperature and 25°C heat rejection temperature), and the

efficiency of a representative TEG module [14] (~30°C cold side temperature) vs. waste heat source

temperature.

Figure 1.4 ARS and TEG performance at different WHR Source Temperature

From the figure it can be seen that the representative ARS only operates effectively with

low-to-medium grade waste heat (maximum COP = 0.74 at 75°C). System performance actually

degrades with higher temperature heat input. According to Herold et al. [13] with an increase in

desorber/source temperature for single-effect ARS, the capacity of the ARS increases with

increased source temperature. This increase in capacity results in an increase in heat exchanger

duty in all heat exchangers of the ARS, this increase in duty is accompanied by greater heat transfer

irreversibilities in all of the heat exchangers. Therefore, an increase in desorber/source temperature

for single effect ARS can result in degradation of the COP.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

0

1

2

3

4

5

6

7

8

0 50 100 150 200 250 300

TA

R C

OP

TE

G E

ffic

ien

cy (

%)

Source Temperature (°C)

Efficiency COP

Page 20: Integrated Absorption Refrigeration and Thermoelectric

9

In contrast, the representative TEG system achieves increased efficiency with source

temperature. In the past, ARS and TEGs have been primarily studied for single pathway WHR

applications. In integrating the two systems, if the >90% of input heat rejected from the TEG system

is delivered to drive an absorption system, the TEG rejection temperature may only be increased

slightly. Both processes can operate efficiently through cascading the single heat source. The

present investigation seeks to explore this potential integration of these WHR pathways, using a

coupling fluid that transfer heat between the two sub-systems, which could result in a more

compact, economical design that may offer greater exergetic efficiencies than can be achieved with

either technology individually.

The two applications of interest for the current study are refrigerated transport vehicles and

carburizing furnaces. Table 1.1 lists the key representative of these application from previous WHR

studies:

Carburizing Furnace [14] Refrigerated Transport Vehicle [15]

Source Temperature 280°C 440-490°C

Rejection Temperature 30°C 20-30°C

Waste Heat Available 20kW 33kW

Single Pathway Type TEGs NH3-H2O ARS

Waste Heat Recovered 240W 5kW

Table 1.1 Key parameters of applications considered for integrated WHR study

For these applications an integrated, TEG and ARS based WHR system can deliver both,

cooling and electrical power. In the case of the carburizing furnace application the WHR system

can space cooling and electrical power for facility lighting. For the refrigerated transport truck

application, the WHR system can deliver refrigeration and supplement auxiliary power to the

truck’s battery. Therefore, the aim of current study is to explore the potential of this proposed

Page 21: Integrated Absorption Refrigeration and Thermoelectric

10

integrated waste heat recovery system to deliver the required cooling and electrical power for the

applications described above.

1.1.3 Integrated Cascaded Waste Heat Recovery Systems

An WHR system in which waste heat cascades through WHR pathways that harness, both,

high and low-grade waste heat is classified as an integrated cascaded WHR system. The benefits

of a cascaded approach to the integration of two single-pathway WHR methods can result in a more

compact, lightweight and economical overall system.

Figure 1.5 shows a comparison of individual, single-pathway ARS and TEG based WHR

systems and an integrated WHR system. Figure 1.5(A), shows the two WHR sub-systems

implemented as single-pathway recovery devices that could individually provide cooling or

electrical output. However, as can be seen in Figure 5(B), their integration could deliver both

cooling and electrical output. For the integrated system shown in Figure 5(B), heat cascades

through the TEG subsystem (high-grade WHR) to the ARS subsystem (low-mid-grade WHR), this

system can be classified as an integrated, cascaded WHR system. In an exhaust-gas-heated and air-

cooled WHR system, the heat acquisition (heat acquisition unit, HAU) and rejection heat

exchangers will be the largest components. As the proposed integrated configuration will have

comparable heat acquisition and rejection loads to either stand-alone unit, the overall system

volume need not increase significantly.

To integrate the two pathways, a subsystem is needed that extracts waste heat from the

exhaust stream and delivers heat to the TEG modules, and can deliver the heat rejected via the TEG

cold junctions to the ARS. Such a device can be called a Heat Acquisition Unit (HAU). The next

section describes this device.

Page 22: Integrated Absorption Refrigeration and Thermoelectric

11

Figure 1.5 (a) Individual, single-pathway WHR systems, (b) More compact, integrated

cascaded WHR system that eliminates the need for additional heat acquisition unit and air-

coupled HX

1.2 Heat Acquisition Unit (HAU)

Figure 1.6 shows an illustration of a HAU, for a TEG and ARS based WHR system. As

can be seen in the figure, an HAU for an integrated, cascaded WHR system, provides an enclosure

for exhaust gases to transfer heat using extended surfaces (finned surfaces), through a solid

interface (heat spreaders and TEGs) and into a coupling fluid using extended surfaces (water blocks

or finned surfaces). The heat spreader and TEG receive the highest grade of waste heat, whereas

Page 23: Integrated Absorption Refrigeration and Thermoelectric

12

the TEG cold side rejects heat to the coupling fluid at lower temperature as heat cascades through

the heat spreader and TEG.

Figure 1.6 An illustration of a typical Heat Acquisition Unit (HAU) used for integrated WHR

The design of a HAU greatly impacts the performance of the WHR system. Some of the

key challenges in designing a HAU for a WHR system are:

• Extended surface sizing: To transfer the required heat from an exhaust stream to

the WHR system. Therefore, an HUA must be designed with sufficient heat

exchange surface area on the exhaust stream and coupling fluid side.

• Pressure drop: In close relation to the extended surface sizing is the problem of

pressure drop in a HAU. The principal in this context is to not

Page 24: Integrated Absorption Refrigeration and Thermoelectric

13

• Spreading resistance: Thermal spreading resistance occurs at the interface of large

heat transfer surfaces and the TEGs. An effective HAU must minimize these

spreading resistances.

• Optimum number of TEGs: Adding more than an optimum number of TEGs can

result in lower HAU performance as an increase in TEGs results in an overall lower

ΔT.

• Single/Dual stage WHR: TEG are high thermal resistance devices that require large

heat exchange surfaces on either side to enable any significant heat transfer across

them. Dividing an HAU into one stage with TEGs and another without TEGs can

allow more waste heat recovery than with a single stage HAU. Therefore,

depending on the application requirements, an HAU design can either be single

stage TEG design or a dual stage TEG design.

These and other aspects of HAU design are discussed in detail in Chapter 5.

For the current study, the TEG subsystem is integrated into the HAU and as such, it can be

operated independently of the ARS subsystem by passing the coupling fluid through a heat

exchanger where it rejects heat to a recirculating cooler; this also allows for the control of the inlet

temperature of the coupling fluid to the HAU.

1.3 Summary of prior research into Waste Heat Recovery methods

Research into Waste Heat Recovery (WHR) has intensified dramatically in the last decade.

According to the publication data provided by Scopus[16], of the 7,100 journal articles published

on the topic from 1969 to 2019, over 50% of these articles were published from 2015-2019.

Page 25: Integrated Absorption Refrigeration and Thermoelectric

14

Waste heat recovery methods can be categorized with respect to the waste heat source

temperatures. US Department of Energy report [6] categorizes waste heat recovery sources

according to temperatures as under:

a. Low: Less than 230°C (e.g. Flue gas [17], Coolants[18])

b. Medium: 230° – 650°C (e.g. Process gases and vapors[19])

c. High: Greater than 650°C (e.g. Molten slag from steel industry[20], vehicle

exhaust[21])

Organic Rankine Cycles are well suited for low-to-medium range source temperature waste

heat recovery owing to the low boiling point of the organic working fluids [10,11]. Other low-to-

mid temperature waste heat recovery methods include thermally activated cooling (TAC) systems

like solid and liquid desiccant cooling, absorption and adsorption chillers [24]. For high

temperature WHR, methods like steam Rankine cycle (SRC), regenerative burners, waste heat

boilers (WHB), thermophotovoltaic (TPV) generators, thermoelectric generators (TEGs) and

thermionic generator can be used [5] (Table 1.2)

WASTE HEAT

GRADE

TEMPERATURE

RANGE SOURCES

SUITABLE WHR

METHODS

High 650°C and above Molten Steel Slag, Vehicles

Exhaust

SRC, TEGs, TPV,

Regenerators, RGB

Medium 230°C – 650°C Process gases and vapors ORC, TAC

Low Below 230°C Flue Gas, Coolants ORC, TAC

Table 1.2 Classification of WHR methods with respect to source type and grade

Of all the waste heat generated by energy consumption in industry, only 32% is recovered

[25]. Owing to its difficulty, low-grade waste heat accounts for nearly zero percent of this recovered

energy [26]. Therefore, low-to-mid grade waste heat recovery methods have received significant

attention in recent years [5], [19], [25], [27]–[29]. For a more detailed review of WHR state of the

art a review paper by Mahmoudi et al. [19] can be consulted.

Page 26: Integrated Absorption Refrigeration and Thermoelectric

15

1.3.1 Research Needs in Cascaded WHR

As discussed in the last section, single pathway WHR, has received significant research

attention. However, WHR systems that can harness both high-grade and low-to-medium grade

portions of waste heat sources have received limited attention. Prior studies into cascaded WHR

explored integrated ORC and SRC cycles [30], ORC and Vapor Compression Cycle (VCC) [31],

combined ARS and VCC systems [32], Cascaded Thermoelectric WHR [33] and other combined

power cycles [26,27,28]. These studies highlight benefits for a cascaded approach to low-to-mid

grade WHR, and motivate the current investigation into high-grade to low/medium grade waste

heat.

The current study proposes to experimentally investigate an integrated, cascaded WHR

approach. The waste heat sources, considered for this study, are the exhausts from refrigerated

transport vehicles and carburizing furnaces. In this approach, a Heat Acquisition Unit (HAU)

receives high-grade waste heat which cascades through TEG subsystem, which produces

electricity, and instead of rejecting the low-grade portion to the ambient, the ARS subsystem

recovers this cascaded low-grade portion of the waste heat and delivers cooling.

1.4 Goals of the Present research and Dissertation

In the proposed Ph.D. dissertation study, the following research questions will be

addressed:

1) What is the thermoeconomic feasibility of the proposed TEG and ARS

based integrated, cascaded WHR system?

Page 27: Integrated Absorption Refrigeration and Thermoelectric

16

Any proposed WHR system must be economically viable for significant adoption. This

study investigates the thermoeconomic feasibility of the proposed system by the following:

• Developing application-specific thermodynamic models and conducting a

payback period study

• Developing lab-scale model to provide practical insights into hardware cost,

sizing and thermoeconomic optimization

2) How does the proposed ARS-TEG WHR system compare with

conventional WHR systems?

This study also includes a comparative thermoeconomic study of the proposed ARS-TEG

WHR system with an integrated Organic Rankine Cycle (ORC) and Vapor Compression Cycle

(VCC). The comparative study explores this by means of:

• Developing an application specific thermodynamic model for each WHR

system based on real-world vehicle data and reasonable closure parameters

• Conducting a capital cost analysis

3) What are the design considerations for application specific ARS and TEG

based integrated cascaded WHR systems?

For a specific application of WHR, certain system tradeoffs must be considered during its

design phase:

• Technical tradeoffs in system performance: To recover sufficient heat from the

exhaust stream to operate the WHR system, finned surfaces are used.

Introduction of these finned surfaces results in an increase in system

backpressure. This increase in backpressure requires additional work from the

Page 28: Integrated Absorption Refrigeration and Thermoelectric

17

source to move the exhaust gases through the HAU. For each specific

application, the increase in exhaust backpressure must be such that it does not

decrease the waste heat source’s system efficiency. For example, Mikulic et

al. [37] state that fuel consumption of a Diesel engine increases by ~0.13% for

every kPa of backpressure. Using this ratio, the allowable backpressure for the

vehicle application can be calculated. Vehicle engines typically reject 33% of

fuel of energy as exhaust, if, for the cascaded system proposed in this study,

50% of this exhaust waste heat is recovered with TEG conversion efficiency

at 5% and COP of ARS is at 0.5, the overall increase in efficiency will be

around 8% i.e. for a diesel application the increase in backpressure for this

representative case should not exceed ~61.5kPa for the system to remain

efficient.

• Electrical Power vs. Cooling: In designing an integrated cascaded TEG and

ARS based WHR system, the competing nature of electrical power and cooling

requirements must be taken into account. As discussed, in in reference to

equation 1, thermoelectric power generation strongly depends on the

temperature difference across the two TEG junctions. For a specific average

exhaust temperature, the temperature difference across TEG junctions can be

increased by lowering the coupling fluid temperature or by increasing the

number of TEGs in the array. However, lowering the coupling fluid inlet

temperature to the HAU also lowers the desorber heat delivery temperature for

the ARS and if this temperature is lower than the temperature range of efficient

operation for the ARS subsystem, this will lead to an inefficient overall WHR.

The other option of increasing the number of TEG modules also has an upper

Page 29: Integrated Absorption Refrigeration and Thermoelectric

18

limit in terms of number of modules. This is discussed in detail in Chapter 5,

but, in short, increasing the number of TEGs beyond an upper limit brings the

temperatures of coupling fluid and exhaust streams closer together in HAU

and causes a “thermal short-circuit”, lowering TEG array performance.

4) Does the proposed WHR system have the capacity to meet cooling and axillary

power requirements for the proposed applications?

To ascertain the proposed WHR system’s viability for delivering useful and relevant

streams of electrical power and cooling, application specific thermodynamic models were created.

A lab scale facility has been built to assess these thermodynamic models and to investigate if the

proposed WHR system can recover from a heat source of approximately 2.0 kW at 400°C, electrical

energy of approximately 25W and deliver 0.8 kW of cooling at -15°C. These parameters represent

a 1/10th scale of a vehicle application’s waste heat requirements. This experimental portion of the

study will also provide insights into the practical challenges of system operation that are not be

captured in the modeling study portion. This will serve as a basis for hardware sizing and

thermoeconomic study.

5) What are the tradeoffs associated with the design of proposed system’s

Heat Acquisition Unit (HAU)?

Design of a HAU is of critical importance in assurance of exergetically efficient operation

of a cascaded WHR system. Two different HAU designs were developed for this study and

compared over a range of input waste heat stream flow-rate and temperatures. Their comparison

(Chapter 5) sheds light on the effects of spreading resistance, choice of flow direction, number of

TEGs and other design parameters on the overall exergetic efficiency of the WHR system.

Page 30: Integrated Absorption Refrigeration and Thermoelectric

19

6) How does the ARS sub-system’s cooling delivery performance vary at

different operating conditions?

This study also explores the cooling delivery performance of the absorption refrigeration

sub-system by varying different operating conditions like the heat input rate and temperature, and

the cooling delivery temperature.

The following chapters address these questions by first, presenting a review of recent

research in the area of WHR involving thermally activated refrigeration and thermoelectrics based

WHR methodologies, in Chapter 2. Thereafter, Chapter 3 presents details on the two

thermoeconomic studies conducted during the course of this research. Chapter 4 presents a detailed

discussion on design and development of the Heat Acquisition Units (HAUs) developed during this

research. This chapter also includes a performance comparison between the two HAUs developed

during this study (HAU-1 and HAU-2) in light of experimental results. Discussion regarding the

parameters of HAU design that have a controlling effect on its performance is also provided in

Chapter 4. Chapter 4 also includes a detailed engineering model of the HAU. In Chapter 5, cycle

model for the integrated ARS and TEG based WHR system is presented along with a discussion of

the closure parameters considered for the system and cycle model results for the two applications

considered for this study. Chapter 6 describes the experimental facility and presents the results from

the experiments performed on the lab-scale experimental facility. This chapter also presents a

comparison between the HAU model predictions and the experimental measurements. Chapter 7

provides a brief summary and key finding of this dissertation work alongside recommendations for

future research pathways.

Page 31: Integrated Absorption Refrigeration and Thermoelectric

20

Literature Review

Page 32: Integrated Absorption Refrigeration and Thermoelectric

21

Figure 2.1 shows a Venn diagram of different types of WHR systems. A WHR system can

be single pathway system (one WHR technology) that utilizes a single waste heat grade (Single-

pathway, single-source) or it can utilize multiple grades of waste heat (Single-pathway cascaded).

Examples of single-pathway, single-grade WHR systems are ORC and TAR based WHR systems.

Whereas, cascaded TEGs are an example of cascaded single-pathway WHR systems. Whereas, a

system that employs multiple WHR technologies and grades to harvest waste heat is called an

integrated WHR system. The present research is an ARS-TEG based integrated, cascaded WHR

study.

SINGLE PATHWAY

INTEGRATED

CASCADED

Figure 2.1 Venn-diagram showing different types of WHR Systems

In this chapter a review of prior relevant studies in WHR is presented. As described earlier,

prior WHR research has focused primarily single grade, single-pathway waste heat recovery.

Page 33: Integrated Absorption Refrigeration and Thermoelectric

22

However, some more recent theoretical studies have also considered cascaded WHR systems; both

single-pathway (one WHR technology) and integrated (multiple WHR technologies). This chapter

provides a summary of recent research into single-pathway and cascaded WHR studies.

2.1 Single-pathway, single-grade WHR studies

Li et al. [38] performed a theoretical investigation into using Absorption-Compression

Hybrid Refrigeration Cycle (ACHRC) for Coach Air Conditioning using exhaust gas WHR. Their

proposed system comprised of two subcycles; the absorption cycle would be thermally activated

using vehicle waste heat, whereas the vapor compression subcycle would use engine power.

According to the parameters considered in their investigation, at vehicle speeds exceeding 105

km·hr-1 the absorption subcycle could provide all of the required 30 kW of cooling load. At speeds

below 40 km·hr-1 all of the cooling could be provided by the compression subcycle. At speeds in

between 105 and 40 km·hr-1 the two subsystems would work in tandem. This study is an example

of a single-pathway ARS based WHR approach. The size of their system could be reduced by

eliminating the vapor compression cycle either by the use of PCMs or charging a refrigerant

reservoir at higher speeds which can then be utilized at lower speeds. Moreover, the high-grade

waste heat source (480°C) considered in their study may be utilized to produce electricity using

TEGs, which can then be cascaded to the ARS to produce the required cooling, thereby, increasing

exergetic efficiency compared to their proposed system.

Lu et al. [12] performed modeling and analysis of a large temperature span ARS based

WHR system aimed at recovering heat from flue gas during recovery process. In their system, they

employ temperature matching which they refer to as “continuous-temperature-changing generation

process”. This process takes place in the desorber, where the waste heat source and ammonia-water

Page 34: Integrated Absorption Refrigeration and Thermoelectric

23

mixture would be in counter-flow arrangement and the dilute mixture would be looped back into

the generator transferring heat to the incoming ammonia-water mixture and in turn, cooling down.

According to their study, this process could lead to a stable 5°C heat transfer temperature difference

between the waste heat source and solution, allowing for a waste heat temperature span of 67.5K.

For their proposed study, with a waste heat source of 10.16 kW at 150°C and 9 kW of cooling at -

15°C, a COP of 0.93 and an exergetic efficiency of 49.7% was predicted. This is an example of a

low-grade, single-pathway WHR method where the source of heat is flue gas with a large

temperature span. In their study, Lu et al. have used a relatively complex desorber, pre-mixer and

internal heat recovery system. For flue gases with higher temperatures (400°C), a TEG and ARS

based WHR approach will yield similar exergetic efficiencies and provide 1 kW of electrical energy

and a similar cooling capacity.

Sonthalia et al. [39] performed a theoretical study where waste heat from a single cylinder

Internal Combustion (IC) engine was recovered using vapor absorption refrigeration (VAR) and

thermoelectric converters. The electric power from the TEG modules was used to drive the VAR

pump. The maximum COP achieved was 0.28 with a 10% engine load at 2000 rpm. For their

proposed system, the TEG hot junction side would be in direct contact with the exhaust stream and

reject heat to a cooling chamber. The exhaust stream would continue to the desorber and supply

heat to the VAR subsystem. Their approach is an example of single-pathway WHR as the heat

rejected by the TEGs will would not be routed to the desorber but, instead, it will be rejected

through a cooling chamber. With the cascaded approach proposed in the present study, this rejected

heat can be harnessed to operate an ARS subsystem.

Meng et al. [40] performed a multiphysics modeling based investigation of TEG based

single-pathway, WHR from an automobile exhaust. Their key findings from their modeling results

are that the flow-arrangement between the hot stream and cold stream does not affect the overall

Page 35: Integrated Absorption Refrigeration and Thermoelectric

24

power output, however, using counterflow arrangements results in a more consistent temperature

difference across the number of TEG modules used. They also conclude that for a specific

geometrically constrained system, adding additional TEGs leads to a deterioration in power output.

Liu et al. [41] performed an experimental investigation where, using TEGs, 944 W of

electrical energy was recovered from an automotive engine’ exhaust waste heat of 51 kW at an

average temperature of 312°C and TEG average cold side temperature of 69°C. This cold side

temperature is sufficient to drive thermally activated refrigeration cycles. If such a system were

incorporated here in a cascaded approach with COP = 0.5, an additional ~12 kW of cooling could

be delivered.

Kaibe et al. [14] experimentally investigated WHR from a carburizing furnace where 240

W were recovered from an exhaust steam at of 20 kW at 240°C. The concluded that the TEG power

output could be increased with more effective heat collection as, of the 20 kW of waste heat

available, they were only able to recover 5 kW as TEG heat input. In their experimental study, the

TEGs were mounted on a heat sink board and flue gases from the carburizing unit directly impinged

on the board. On the cold side, a water-block was used to maintain an average temperature of 30°C.

As proposed above, their system could also be integrating with a bottoming thermally activated

refrigeration system that could meet other plant thermal needs. Such an application may result in

slightly reduced electrical conversion efficiency as the TEG cold side temperature would need to

be increased.

He et al. [42] performed a theoretical investigation on the influence of different cooling

methods for TEG based WHR systems. They considered co-flow and counter-flow arrangements

for air and water as coolants. The key findings from their study are that when using water as coolant,

the power output did not differ by much for counter and co-flow arrangements. For air flow,

Page 36: Integrated Absorption Refrigeration and Thermoelectric

25

counterflow arrangement resulted in greater power output for the same module areas. They also

observed that water cooling typically results in greater electrical power from the TEGs.

Rattner et al. [43] conducted a theoretical study to develop an analytical model for

optimally sizing TEG module arrays for WHR. Their study concludes that for a high-flux,

commercially available TEG, arranged in a spanwise array of co/counter-flow arrangement,

thermally connected to the hot and cold stream at isothermal temperatures via highly conductive

heat spreaders, there exits an optimum number of TEGs given by the “thermal-fluid figure of

merit”, HT.

2.2 Cascaded WHR studies

Many experimental studies in recent years have considered high to medium grade, single-

pathway WHR [25], [44]–[47]. Summaries of some of these studies are presented in this sub-

section.

Zhang et al. [25] performed an experimental study between a SRC and a cascaded WHR

system that employs SRC and ORC in tandem (S-ORC). They concluded that for exhaust

temperatures in the range of 150-210°C, ORC based WHR performs better than S-ORC, however,

within an exhaust temperature range of 210-350°C, S-ORC WHR systems would perform better

than ORC based WHR. This is due to better heat source temperature matching in the S-ORC based

WHR system. The cascaded WHR system built for their study had a peak thermal efficiency of

9.6%. The TEG and ARS based cascaded WHR system proposed the current study has the potential

to recover twice the amount of waste heat compared to their study.

More recently, theoretical studies of thermoelectric based WHR typically investigate the

effect of temperature matching of different TEG materials with their optimal temperatures of

Page 37: Integrated Absorption Refrigeration and Thermoelectric

26

operation. One such study was conducted by Shen et al. [7] where TEGs of three different material

composition were used to harness waste heat from methane burner. The high-temperature TEG

module receives heat directly from the methane burner’s exhaust, subsequently the high

temperature flue gases provide heat to the medium-temperature TEG module. For the low-

temperature TEGs heat is provided by the water coolant loop used to cool the high-temperature and

medium temperature TEG modules. Through this approach, they were able to achieve a WHR

system efficiency of 7% which is greater than single-stage TEG WHR (~5%).

Although cascaded TEG based single-pathway WHR system operate at a greater efficiency

than if only one type of TEG material were used, single-pathway thermoelectric based WHR reject

nearly 90% of the waste heat at a lower temperature, resulting in an overall low WHR efficiency.

2.3 Discussion

As can be seen form the discussion of past relevant studies provided above, single-pathway

ARS and TEG based WHR systems have been identified as promising pathways for WHR.

However, there is a need to demonstrated their potential as an integrated, cascaded WHR system.

Therefore, the current study proposes to explore the potential of an integrated cascaded, TEG and

ARS based WHR system.

Page 38: Integrated Absorption Refrigeration and Thermoelectric

27

Thermoeconomic Studies

Page 39: Integrated Absorption Refrigeration and Thermoelectric

28

3.1 Thermoeconomic Analysis of Vehicle and Furnace Application

A thermoeconomic study of cascaded WHR using TEGs and Absorption Refrigeration is

presented here for two power and cooling applications: (1) in a refrigerated transport truck and (2)

for an industrial carburizing furnace. For the vehicle application, TEGs augment the electrical

energy supply (battery) of the vehicle and the absorption refrigeration system provide the cooling

required at -15°C. For the carburizing furnace application, TEGs provide electrical power for the

plant and the absorption refrigeration system provides space cooling at 5°C.

Figure 3.1 Cascaded WHR cycle diagram

3.1.1 System Modeling Results

Figure 3.1 illustrates the integrated cascade refrigeration system. A steady-state cycle

model was developed that applies mass, species, and energy balances to each component assuming

representative closure parameters (e.g., heat exchanger approach temperatures, pump efficiency,

Commercial Off-the-Shelf (COTS) TEG performance parameters). The absorption system is

Page 40: Integrated Absorption Refrigeration and Thermoelectric

29

modelled assuming use of the NH3-LiNO3 working fluid pair, which avoids the need for a rectifier

and permits a low desorber temperature [10]. System operation is described below with

representative values for the refrigerated transport truck application.

Hot exhaust gas enters the heat acquisition heat exchanger (HX) (Point 1: 500°C and 0.007

kg s-1), where it rejects 25 kWth to the TEGs, reaching an exit temperature of T2 = 300°C. The HX

has TEGs mounted on the outside surface of the exhaust gas channel, which collect and convert a

portion of acquired heat into electricity. The array of TEG modules would produce 1.63 kW of

electric power, with an efficiency of approximately 6.5%. The hot side temperature of the TEGs is

425°C and the cold side temperature is 180°C. The remaining heat is rejected by the cold-side of

the TEGs directly to the absorption system desorber. Dilute ammonia-lithium nitrate solution

(NH3+LiNO3) solution enters the desorber (Pt. 8: 112°C, 1643 kPa, ψ = 62% LiNO3 mass fraction

in solution). NH3 is evaporated from the solution, bringing the concentrated outlet liquid stream to

T3 = 138°C and ψ3 = 69%. The concentrated LiNO3 solution enters the recuperative solution heat

exchanger (SHX) (Pt. 3) and cools to T4 = 55°C as it preheats the dilute solution. The concentrated

solution expands through a valve to reach (T5 = 55°C, P5 = 174 kPa).

The NH3 vapor is the refrigerant, and liquefies as it rejects heat to the ambient in the

condenser (Pt. 9: 140°C, 1643 kPa → Pt. 10: 40°C, 1643 kPa). The liquid refrigerant is subcooled

in the refrigerant pre-cooler (RPC, Pt. 11: 17°C, 1643 kPa), and is expanded through a valve to the

evaporator (Pt. 12: -22°C, 174 kPa). The NH3 evaporates, delivering cooling to the conditioned

space (Pt. 13: -20°C, 174 kPa). The refrigerant vapor continues through the RPC, precooling the

refrigerant liquid (exiting at Pt. 14: 30°C, 174 kPa).

The NH3 vapor (Pt. 14) mixes with the low pressure concentrated LiNO3 solution (Pt. 5) in

the absorber. The vapor absorbs into the solution rejecting heat to the ambient, and exits at Pt. 6

(40°C, 174 kPa, ψ6 = 62%). The liquid mixture is pumped to high pressure (Pt. 7: 40.5°C, 1643

Page 41: Integrated Absorption Refrigeration and Thermoelectric

30

kPa) and continues to the SHX where it is recuperatively preheated to Pt. 8 (112°C, 238 psi),

completing the cycle.

For the refrigerated transport truck application, 25 kWth of high temperature input heat is

acquired from the engine exhaust (T= 500 °C). The TEGs convert 1.63 kW to electricity (ηe =

6.5%), and deliver the remaining 23.4 kWth to the absorption system at 180°C. The absorption

system delivers 12.5 kWth of cooling to a refrigerated trailer at -15°C (COP = 0.53). This would

be sufficient to satisfy the cooling loads of typical large refrigerated transport trucks. This proposed

cascaded system can also provide a significant portion of the auxiliary electrical load of refrigerated

trucks (1.35 kWe net, accounting for pump motor power).

A similar analysis was performed to assess the performance of the proposed cascaded

WHR system using high temperature exhaust from an industrial carburizing furnace. In that

application, 50 kWth would be acquired at 500°C. The TEG stage would produce 3.3 kW of electric

power (ηe = 6.5%), which would result in a net power output of 2.9 kWe accounting for the

absorption pump requirement (377 W). The absorption system would deliver 26.6 kWth of cooling

at 5°C (COP = 0.6).

3.1.2 Thermoeconomic Analysis

3.1.2.1 Heat Exchanger Costs

For the two applications, the following table lists the calculated, required UA values:

Desorber Condenser RPC Evaporator Absorber SHX

UA for Vehicle (W K-1) 1248 1528 126 2585 2290 1615

UA for Furnace (W K-1) 9000 3053 172 5320 4145 1635

Table 3.1 The UA values for refrigerated vehicle and industrial carburizing furnace

Page 42: Integrated Absorption Refrigeration and Thermoelectric

31

Heat exchanger costs vary widely, depending on the working fluids, construction materials,

operating pressures, production volume, and heat exchanger design. Here, the heat exchanger

costing guides of Hewitt et al. [48] and Caputo et al. [49] were consulted to arrive at an estimated

air-coupled heat exchanger cost (absorber and condenser) of $3.50 W-1 K. The more complex heat

acquisition HX, which includes the TEGs and desorber channels, is estimated to be $4.50 W-1 K

(based on the air-side UA). TEGs are assumed to cost $0.27 W th-1 [50]. The liquid-liquid/working

fluid heat exchangers (SHX, RPC) are assumed to be compact and low cost compared with the

air/gas coupled heat exchangers, and are costed as a group with other system hardware.

3.1.2.2 Vehicle application

The thermoeconomic performance of the proposed cascaded WHR system is determined

by comparison with TEG-only and absorption refrigeration-only WHR systems. For the TEG-only

system, the heat acquisition UA is calculated to be 390 W K-1 ($1,365), accounting for both heat

acquisition and rejection stages. The TEG module cost is $6,750 and electric output is 1.88 kWe.

To calculate the payback period of TEG-only WHR system, the assumption used, along

with their references from published literature are presented in Table 3.2:

Parameter Qty Units Ref.

Duty hours 8 hr [51]

Duty days 330 hr -

Duty hours per year 2640 hr/yr -

Fuel cost 0.8 $/L [52]

Fuel consumption rate 0.4 L/kWh [53]

Table 3.2 Comparison of payback periods with previous studies

Based on the assumption listed in Table 3.2, yearly fuel consumption (1056 L/kW-1 yr-1)

and yearly cost of fuel ($844.8 $/kW-1 yr-1) can be calculated. Since the TEGs produce an average

total of 1.88 kWe, the savings can be calculated by multiplying the yearly cost of fuel to the power

generated by the TEGs ($1588.2 yr-1). The total capital investment (TEG+HX = $8115) can then

Page 43: Integrated Absorption Refrigeration and Thermoelectric

32

be divided by the yearly savings to calculate the payback period, which, for the TEG-only case is

5.1 years.

In the case of using only a WHR absorption refrigeration system (12.5 kWth cooling

capacity), the UA for heat acquisition system is 343 W K-1 and the condenser and absorber require

another 3818 W K-1. Assuming that additional $5,000 in costs will be required (compact HXs,

pump, framing), the total capital cost is $22,478. This can be compared with a conventional diesel-

powered refrigeration unit. Consumer costs for such units are ~$45,000, but assuming that the

manufacturing cost is 1/3rd of this, the conventional system capital cost would be $15,000[54]. The

COPs for the refrigeration systems of trucks typically range from 0.5–1.5 [55]. Assuming the COP

of a diesel-powered refrigeration system to be 1.5, the total power required to deliver a 12.5 kWth

cooling load can be calculated (8.3kW). By using the assumption made in Table 3.2, the total cost

of operation per year for the diesel engine powering the refrigeration system can be calculated

($7,011.8). This gives a simple payback period for the absorption WHR system of just 1.1 years.

For the proposed cascaded WHR system, the efficiency of the TEGs would drop from 7.5%

(standalone) to 6.5% because heat rejection to the desorber at 112-140°C would decrease the

temperature difference across the TEGs. The main HX costs would be $18,382 and additional

hardware costs would be $6,000. The TEG module cost is $6,750. The total capital cost is $31,132.

The TEGs produce 1.63 kW of power; therefore, they contribute $1,377.02 yr-1 in fuel savings. The

equivalent conventional system would have a capital cost of $15,000 and a yearly recurring cost of

$7,011.8. The payback period for this WHR system is calculated as 1.9 years.

The payback for the cascaded system is higher than absorption standalone system but it is

about 2.7 times lower than TEGs alone. Thus, for an application where electricity and refrigeration

are required, the cascaded system offers a significant advantage. Taking into account the factors of

Page 44: Integrated Absorption Refrigeration and Thermoelectric

33

“learning curve” [56] and “scale effects” [57] it can be surmised that future TEGs will cost

significantly less, thus, improving the attractiveness of this option.

3.1.2.3 Carburizing Furnace Application

The thermoeconomic performance of the proposed cascaded WHR is assessed by

comparison with TEG-only and absorption chiller-only WHR systems. For an input heat of 50 kW,

the TEG-only heat acquisition HX UA would be 1750 W K-1 ($6,125). The TEG module capital

cost would be $13,500. The total system capital cost would be $19,625. At ηe = 7.5%, the TEGs

would produce 3.75 kW. If this electricity were to be purchased from a utility at $0.11 kWh-1 [58],

the simple payback period would be 5.43 years. The reduced payback period compared with the

truck TEG WHR can be achieved because the industrial furnace is assumed to operate continuously

(vs. 2640 hrs yr-1 for the truck).

The absorption chiller-only WHR would require a UA of 300 W K-1 for the heat acquisition

HX and a UA of 8000 W K-1 for the condenser and absorber (combined). The total cost of these

HXs would be $33,741. Other system costs are estimated to be $15,000 (total cost: $47,940). The

chiller capacity would be 28.5 kW (233 MWh cooling per year). Producing the same cooling with

a conventional electric powered chiller (including capital costs), would cost $4,275 yr-1. The

payback period for this WHR system is 6.86 years. The projected payback period is sensitive to

assumed specific conventional chiller costs ($ kWth-1), which scale dramatically with overall

capacity.

For the integrated cascaded WHR system, the efficiency of the TEGs will drop to 6.5%, as

in the vehicle application case. Total HX costs (condenser, absorber and heat acquisition) would

be $38,516 and additional costs are estimated to be $16,000. The total system capital cost is $

68,016. The equivalent total conventional power and chiller system capital cost is $3,997. The

simple payback or payback for this integrated system is 6.7 years. The integrated system approach

Page 45: Integrated Absorption Refrigeration and Thermoelectric

34

has a lower payback period compared to the absorption system but it has a higher payback in

comparison to the TEGs alone.

3.1.2.4 Net Present Value Analysis

In addition to the simple payback period, a Net Present Value (NPV) analysis can lend

more insights into the economic competitiveness of the proposed WHR systems (integrated and

standalone) with each other. NPV is calculated using the formula presented in equation 3.1:

𝑁𝑃𝑉 = −𝑪0 + ∑𝐹𝑛

(1 + 𝑘)𝑁

𝑁

𝑖=1

3.1

In equation 3.1, C0 is the total capital investment of the WHR system and k is the interest

rate assumed to be 5% and Fn are the total savings. A non-negative NPV value indicates a profitable

state of operation, therefore, the earlier a WHR technology approaches a non-negative NPV value,

the more economical it is.

NPV analysis can be sensitive to the capital cost assumptions, therefore, a capital cost

uncertainty of 30% was added to the present analysis. It was also assumed that the WHR systems

have a lifespan of 15 years.

From Figure 3.2, it can be seen that the ‘TEG only’ approach takes the longest to achieve

a non-negative NPV for the vehicle application case. Whereas, the ‘ARS only’ approach is the

fastest to reach a non-negative NPV. The integrated approach, although, takes longer than the ‘ARS

only’ approach, however, in the long-term, within the margin of uncertainty in capital cost, it is

possible for the integrated approach to WHR to yield a greater NPV value.

In this analysis, the TEG power output corresponds to the efficiency of COTS TEG.

However, the analysis could yield different results if the efficiency of the TEG were to improve in

the future. To simulate the effects of such an improvement, for the ‘TEG only case’, a 50% increase

Page 46: Integrated Absorption Refrigeration and Thermoelectric

35

in TEG efficiency was assumed. Note that a 50% increase in TEG efficiency would result in a 25%

increase in total power output as half of the TEG power is consumed in overcoming TEG internal

resistance.

Figure 3.2 NPV versus time for the vehicle application

As can be seen in the graph shown in Figure 3.3, a 50% increase in the efficiency of the

TEG modules does not significantly impact the time at which the NPV value reaches a non-negative

number. Moreover, the overall NPV performance, over a period of 15 years, for the increased

efficiency ‘TEG only’ system is only slightly better than the present COTS TEG based system.

Therefore,

Page 47: Integrated Absorption Refrigeration and Thermoelectric

36

Figure 3.3 NPV versus time for the carburizing furnace application

Figure 3.4 shows the NPV values over time for the carburizing furnace application. Due

to the large capital cost of the ARS system, the ‘ARS only’ and integrated approaches of WHR take

longer to reach a non-negative NPV value. However, over the entire lifespan of the of the WHR

systems, the integrated approach yields a greater NPV value and is, therefore, the most economical

approach to WHR in a carburizing furnace.

Page 48: Integrated Absorption Refrigeration and Thermoelectric

37

Figure 3.4 NPV versus time for the carburizing furnace application

3.1.3 Conclusions

Integrated cascaded WHR systems allow use of both high- and low-grade portions of waste

heat. Such technologies could enable greater efficiencies than single pathway WHR. In this study,

cascaded TEG and absorption refrigeration WHR was modeled for two potential applications:

WHR from refrigerated transport truck exhaust and an industrial carburizing furnace exhaust.

Findings suggests that this cascaded WHR approach could offer a payback period of 1.9 years for

the refrigerated truck application, and 6.7 years for the carburizing furnace. These payback periods

are competitive with those of other waste heat recovery technologies reported in the literature

(Table 3.3)

Page 49: Integrated Absorption Refrigeration and Thermoelectric

38

Author System Type Net Input Energy (kW) Heat Source Payback

(Years)

Patel et al. [59] Trigeneration ORC 41.35 Waste Heat 6.2

Shu et al. [60],[61] Transcritical ORC 338 Waste Heat 7.8

Fontalvo et al. [62] Dual Pressure ORC 244 Waste Heat 8

Table 3.3 Comparison of payback periods with previous studies

3.2 ARS-TEG vs. ORC-VCC; Thermoeconomic Comparison

ORC based WHR systems have been a topic of great interest in recent years. According

the publication data from Scopus[16], of the nearly 1700 journal and conference articles discussing

WHR using ORC published in between the period of 1973 to 2020, over 90% were published during

the period between 2010 to 2020. A number of recent publications [63]–[66] have studied the

integration potential of ORCs with Vapor Compression Cycles (VCC) as an integrated WHR mode.

In this section, a comparative thermoeconomic analysis for the application case of

refrigerated transport truck [15] between ORC-VCC and the novel ARS-TEG system, will be

presented. The purpose of this section is to investigate if the proposed ARS-TEG system can

provide the required cooling and how its economic outcomes differ from the more conventional

ORC-VCC based WHR system.

First, a description of the ORC-VCC model is presented. Next, a discussion of the exhaust

temperature and flowrate is presented. Then the closure parameters, device isentropic assumptions,

and working fluid selection criterion is discussed, following which the cycle model for the ORC-

VCC system is explained. This discussion is followed by the presentation of the results from the

cycles of the respective WHR systems. Finally, an economic analysis of the cost of the respective

Page 50: Integrated Absorption Refrigeration and Thermoelectric

39

systems will be presented along with discussion of the costing guides in relevant published

literature.

3.2.1 ORC-VCC System Description

ORC-VCC based systems can have a number of different configurations. Several recent

studies have employed a varied number of configurations for the ORC-VCC systems, such as,

multiple regenerators [67] in the ORC subsystem or employed a single working fluid with

combined condenser [68]–[70], shaft connected expander and compressor [66], [71], have been

studied.

For the purposes of the present study, a basic ORC cycle is combined with a VCC by means

of an electricity generator. Figure 3.5 shows the schematic for the ORC-VCC based WHR system.

The pressure of the ORC working fluid in a subcooled state at 1 is increased by a pump

located between 1→2. Between 2→3, the fluid’s temperature rises at it receives heat form the

recuperator. The boiler receives heat from the vehicle exhaust which is transferred to the ORC fluid

between 3→4. The fluid leaves the boiler as a superheated vapor at point 4. Between 4→5 the ORC

fluid expands to a lower pressure as the Expander/Turbine produces mechanical power which is

converted to electricity by an electricity generator. Between 5→6 the ORC fluid is cooled down in

the Recuperator. A condenser between 6→1 cools down the ORC fluid to a subcooled state.

Page 51: Integrated Absorption Refrigeration and Thermoelectric

40

1

2

3 4

5 G

6

7

8

9

10

Tcool

Ec

ORCVCC

Boiler

Expander

Recuperator

Condenser

Pump

Evaporator

CompressorThrottling

Valve

Condenser

Figure 3.5 Schematic for ORC-VCC based WHR system

All of the portion of the electricity generated by the ORC subsystem is supplied to the

pump of the ORC subsystem and the remainder is delivered to the compressor of the VCC

subsystem.

In the VCC subsystem, a high pressure, subcooled, refrigerant expands through an

isenthalpic throttling valve between 7→8, to lower temperature and pressure. Between 8→9 the

refrigerant receives heat from a cooled space in the Evaporator as it gets superheated at 9. The

refrigerant is compressed form 9→10 to the subsystem high pressure. The refrigerant is subcooled

in a condenser from 10→7 as heat is rejected to the ambient, completing the cycle.

3.2.2 Exhaust Temperature

The vehicle exhaust temperature is a function of the fuel requirements during different

cycles of operation [15] (plain, city traffic or mountains). Within each cycle, the initial flowrate

and temperature are typically (in case of cold-start) not sufficient to operate the WHR system.

Page 52: Integrated Absorption Refrigeration and Thermoelectric

41

Moreover, in the case of city traffic operation cycle, the vehicle is mostly in an idling mode,

therefore, the exhaust temperature is not high enough for WHR operation.

(a) (b)

Figure 3.6 (a) Cumulative frequency of exhaust temperatures (b) Cumulative frequency of available exhaust heat to operate a 5kW absorption refrigeration system for different driving

cycles [15]

Figure 3.6 shows two graphs from the work of Koehler et al. [15] on WHR in refrigerated

transport vehicle using absorption refrigeration. Figure 3.6(a) shows a qualitative distribution of

cumulative frequency of exhaust temperature ranges. From this graph it can be seen that the higher

temperature frequency of plain cycle is the highest. The horizontal axis of Figure 3.6(b) shows

normalized generator/desorber heat inputs, where a value of unity represents a generator/desorber

heat input sufficient, to provide cooling of 5 kW. Negative values occur when the temperature is

lower than the desorption temperature of the refrigerant in the absorption system.

Figure 3.7 shows real-world velocity of a heavy-duty diesel vehicle and the selective

catalytic reduction (SCR) device inlet temperature data , with respect to time, from a cold-start

stage [51]. From the graph, clear fluctuations in exhaust temperature can be seen. To model the

ORC-VCC and ARS subsystems, an average temperature may be selected to conduct a comparison

Page 53: Integrated Absorption Refrigeration and Thermoelectric

42

Figure 3.7 Real-world velocity and SCR temperature data for a heavy-duty diesel vehicle [51]

Additionally, Figure 3.8 shows a frequency distribution graph for the exhaust temperature

of a heavy-duty refrigerated transport truck (~450hp) [51]. From the data presented in the graph,

an average exhaust temperature of ~250°C is chosen as the heat input temperature for ORC-VCC

and ARS-TEG systems for the purposes of thermoeconomic comparison.

Figure 3.8 Frequenc distribution of exhasut temperatures for a heavy-duty refrigeated transport

truck [51]

Page 54: Integrated Absorption Refrigeration and Thermoelectric

43

3.2.3 Closure Parameters for Heat Exchangers

Heat exchanger sizing is significant in establishing basis for thermoeconomic analysis and

comparison of any thermal cycles. For refrigeration systems, like the ARS subsystem discussed in

this study, AHRI [72], [73] guidelines and standards can be used to set key parameters like degree

of subcooling (condensers) and superheating (evaporators or boilers) and closest approach

temperatures (recuperators).

With respect to ORC, closure parameters were chosen by reviewing previous studies of

similar/proportionate source temperatures, sink temperatures, and input waste heat rate and power

generation. Majority of literature reviewed [66], [68], [74] refer to ‘Pinch Point Temperature’

(ΔTpp) for the ORC subsystem (typically 5-10 K). Other published works assumed no superheat

(evaporators/boilers) or subcooling (condensers) [69]. According to Park et al [75], ORC systems

with a power output of 3-50 kW are typically superheated types. To maintain consistency of

comparison between the two WHR systems, and based upon the preceding discussion, AHRI

guidelines were adopted for both ORC-VCC and ARS-TEG systems.

Table 3.4 lists the closure parameters for heat exchangers in the ORC-VCC system:

Subsystem Component Parameter

ORC

Boiler CAT 10 K

DoSH 2 K

Recuperator CAT 10 K

εHX 0.75

Condenser CAT 10 K

DoSC 2 K

VCC

Evaporator CAT 5 K

DoSH 2 K

Condenser CAT 10 K

DoSC 2 K

Table 3.4 Heat Exchanger Closure Parameters for ORC-VCC

Page 55: Integrated Absorption Refrigeration and Thermoelectric

44

In the above table CAT, DoSH and DoSC stand for ‘Closest Approach Temperature’,

‘Degree of Superheat’ and ‘Degree of Subcooling”, respectively. Moreover, the heat source

temperature, cooling delivery and the ambient temperature (250°C, -15°C and 30°C respectively)

are assumed to be the same for both the ORC-VCC and ARS-TEG systems.

3.2.4 Isentropic Efficiency Assumptions and Working Fluids

To estimate the performance of an ORC-VCC WHR, representative isentropic efficiencies

are assumed for the compressors, expanders, and pumps. Actual values of isentropic efficiencies of

these components vary with the size of the component and cycle pressure ratios, among other

parameters.

For example, according Park et al’s review [75] of experimental studies of ORC systems,

ORC based WHR systems that output 1-5kW of power, typically have an expander efficiency in

the range of 0.5-0.75. Table 3.5 lists a summary of isentropic efficiencies of different components

along with the waste heat input, net electric power out and cooling capacities for several ORC-

VCC theoretical studies.

Ref Qin

(kW)

Enet

(kW)

Qcool

(kW)

Working Fluid VCC ORC

ORC VCC Compressor Expander Pump

Type ηS Type ηS Type ηS

[76] 50 1.5 5 R245fa R245fa Piston 0.75 Scroll 0.6 - 0.8

[64] 40 0.6 5 R134a R134a - 0.8 - 0.8 - 0.8

[71] 1600 90 200 R245fa Butane - 0.75 - 0.8 - 0.7

[77] 21 0 4.2 R134a Isobutane - 0.8 - 0.8 - 0.8

[78] 10 0 5.3 R245fa R245fa - 0.8 Scroll 0.75 Piston -

Table 3.5 Working Fluid pairs and isentropic efficiencies for different studies

Page 56: Integrated Absorption Refrigeration and Thermoelectric

45

For selection of ORC working fluid, a number of published works were reviewed [70],

[77], [79]. Hærvig et al. [79], state that, to ensure sub-critical operation, the critical temperature of

the ORC working fluid should be, at most, less than 30°C of that of the heat source temperature.

For the refrigerated transport application (Tin = ~250°C), the ORC working fluid used for this

analysis is Cyclopentane which has a relatively high critical temperature (~240°C). For the vapor

compression cycle, according to Gopalnarayanan [80], for commercial, medium temperature

refrigeration applications (-10°C – -25°C), R134a is well suited.

Based on the literature reviewed, Table 3.6 lists the isentropic efficiencies and working

fluids used for the ORC-VCC model developed for the current study.

Subsystem Component Isentropic Efficiency Working Fluid

ORC Expander 0.6

Cyclopentane Pump 0.7

VCC Compressor 0.7 R134a

Table 3.6 Isentropic Efficiencies of ORC-VCC components

Additionally, the mechanical to electrical conversion efficiency of electricity generator

coupled with the expander is assumed to be 0.9.

3.2.5 Cycle Model for ORC-VCC

With the representative closure parameters selected in the preceding sections, a steady state

model was developed for the ORV-VCC based WHR system with the aim of modeling system

performance through parameters like thermal efficiency, Net Electric Power generated, COPVCC

and the overall COP. Additionally, the following assumptions were made:

• Energy and Mass are conserved and the system operate at a steady state.

Page 57: Integrated Absorption Refrigeration and Thermoelectric

46

• Negligible pressure losses occur in the heat exchangers.

• Friction, heat losses and kinetic energy changes are negligible.

• The turbine, pump, throttling valve and compressor operate adiabatically with respect to

the ambient.

• Boiler heat input from the exhaust is assumed to be 25 kW with exhaust inlet temperature

at 250°C and the cooling load is assumed to be 5kW.

Given the assumptions mentioned above, the mass flowrate of the ORC subsystem can be

calculated by the following equation:

𝑸𝒃 = ��𝑶𝑹𝑪(𝒉4 − 𝒉3) 3.2

Where Qb is the boiler heat input to the boiler form the exhaust and h4 and h3 are the

enthalpies of ORC working fluid at the boiler outlet and inlet, respectively and ṁORC is the mass

flowrate of the ORC subsystem.

Similarly, the VCC mass flow rate can be found by the following equation:

𝑸𝒄𝒐𝒐𝒍 = ��𝑽𝑪𝑪(𝒉9 − 𝒉8) 3.3

Where Qcool is the evaporator heat input from the refrigerated space form and h9 and h8 are

the enthalpies of the VCC working fluid at the boiler outlet and inlet, respectively and ṁVCC is the

mass flowrate of the VCC subsystem.

Turbine work is calculated using the following equations:

𝜼𝒔,𝒕 = (𝒉4 − 𝒉5)/(𝒉4 − 𝒉𝟓𝒔) 3.4

��𝒕 = ��𝑶𝑹𝑪(𝒉4 − 𝒉5) 3.5

Page 58: Integrated Absorption Refrigeration and Thermoelectric

47

Where ηs,t is the isentropic efficiency of the Turbine and Ẇt is the Turbine work rate, and

h4 and h5 are the enthalpies at the turbine inlet and outlet, respectively, and h5s is the isentropic

enthalpy of turbine outlet.

To find pump work, the following equations were used:

𝜼𝒔,𝒑 = (𝒉2 − 𝒉𝟏)/(𝒉𝟐 − 𝒉𝟏𝒔) 3.6

��𝒑 = ��𝑶𝑹𝑪(𝒉2 − 𝒉𝟏) 3.7

Where ηs,p is the isentropic efficiency of the pump and Ẇp is the pump work, and h1 and h2

are the enthalpies at the pump inlet and outlet, respectively, and h1s is the isentropic enthalpy of

pump inlet.

The following equation is used to calculate the heat transfer in the ORC subsystem

condenser.

𝑸𝒄𝒐𝒏,𝒑 = ��𝑶𝑹𝑪(𝒉𝟔 − 𝒉𝟏) 3.8

Where Qcon,p is the condenser heat rejection to the ambient and h6 and h1 are the enthalpies

of ORC working fluid at the condenser inlet and outlet, respectively, and ṁORC is the mass flowrate

of the ORC subsystem.

To find the rate of electric power generated by the electric generator, the following equation

is used:

��𝒈 = 𝜼𝒈 ∙ ��𝒕 3.9

Where ηg is the generator efficiency and Ėg is the electric power generated, and Ẇt is the

Turbine work.

In the VCC subsystem, compressor work rate can be calculated using Equations 3.10 –

3.111:

Page 59: Integrated Absorption Refrigeration and Thermoelectric

48

𝜼𝒔,𝒄 = (𝒉10 − 𝒉9)/(𝒉10𝒔 − 𝒉9) 3.10

��𝒄 = ��𝑽𝑪𝑪(𝒉10 − 𝒉9) 3.11

Where ηs,c is the isentropic efficiency of the compressor and Ẇc is the compressor work

rate, and h10 and h9 are the enthalpies at the compressor outlet and inlet, respectively, and h10s is

the isentropic enthalpy of compressor outlet.

To calculate the rate of heat rejected by the VCC condenser to the outlet Equation 3.12 is

used:

𝑸𝒄𝒐𝒏,𝒓 = ��𝑽𝑪𝑪(𝒉𝟏𝟎 − 𝒉𝟕) 3.12

Where Qcon,r is the condenser heat rejection to the ambient and h10 and h7 are the enthalpies

of the VCC working fluid at the condenser inlet and outlet, respectively, and ṁVCC is the mass

flowrate of the VCC subsystem.

The thermal efficiency of ORC subsystem can be calculated by using Equation 3.13:

𝜼𝑻 = ��𝒕/𝑸𝒃 3.13

The VCC COP is given by Equation 3.14:

𝑪𝑶𝑷 = 𝑸𝒄𝒐𝒐𝒍/��𝒄 3.14

To calculate the overall COP of the combined system, Equation 3.15 is used:

𝑪𝑶𝑷𝒐𝒗 = 𝑸𝒄𝒐𝒐𝒍/(��𝒄 + 𝑸𝒃) 3.15

Finally, with all the heat rates, and inlet and outlet temperatures calculated for the heat

exchangers of the combined VCC system, UA, for each heat exchanger can be calculated using

Equation 3.16:

Page 60: Integrated Absorption Refrigeration and Thermoelectric

49

𝑼𝑨 = 𝑸/𝑳𝑴𝑻𝑫 3.16

Where Q is the rate of heat transfer in a heat exchanger and LMTD is the log-mean

temperature difference of the heat exchanger in question.

3.2.6 Results

From the cycle model described above a code was developed in EES [81]. The results from

the code are presented in this section. The mass flowrate for the ORC and VCC systems were

calculated as 0.036 kg s-1 and 0.043 kg s-1, respectively.

Table 3.7 lists properties of ORC-VCC system at various state points corresponding to

Figure 3.2:

State Point T(°C) h(kJ/kg) P(kPa)

ORC

1 50 1.828 73.96

2 52.4 8.663 3531

3 89.57 84.61 3531

4 220 582.1 3531

5 114.1 486.8 73.96

6 64.21 410.8 73.96

VCC

7 50 123.5 1319

8 -22 123.5 121.7

9 -20 238.8 121.7

10 61.71 310.8 1319

Table 3.7 Properties at various state points for ORC-VCC

Table 3.8 lists the UA (product of overall heat transfer coefficient and Area of a Heat

Exchanger), the rate of work (Ẇ) and heat rate (Q) of different components of the ORC-VCC

system:

Component ṁ [kg s-1

] UA [kW/K] Ẇ [kW] Q [kW]

ORC Boiler

0.036 1.81 - 18.11

Condenser 1.49 - 14.89

Page 61: Integrated Absorption Refrigeration and Thermoelectric

50

Turbine - 3.47 -

Pump - 0.17 -

Recuperator 0.14 - 2.76

VCC

Condenser

0.043

0.81 - 8.123

Evaporator 1 - 5

Compressor - 3.12 -

Table 3.8 Important properties of different components of ORC-VCC system

The COP of the VCC subsystem is 1.6, whereas the overall COP of the ORC-VCC system

is 0.27.

Using the same inputs and closure parameters for ARS-TEG, the properties with respect to

the state point corresponding to Figure 3.1 are presented in Table 3.9:

State Point ṁ [kg s-1

] T [°C] P [kPa] X [-] H

[kJ/kg]

3 0.04123 137.7 1643 0.6879 23.8486

4 0.04123 55 1643 0.6879 -0.3678

5 0.04123 55.5 173.8 0.6879 -0.3678

6 0.04581 40 173.8 0.6191 -7.3976

7 0.04581 40.49 1643 0.6191 -7.1341

8 0.04581 112.4 1643 0.6191 14.6607

9 0.004581 139.7 1643 0 156.6

10 0.004581 40 1643 0 19.079

11 0.004581 16.39 1643 0 7.7246

12 0.004581 -22 173.8 0 7.7246

13 0.004581 -20 173.8 0 124

14 0.004581 30 173.8 0 135.3

Table 3.9 Properties at various state points for ARS-TEG

Table 3.10 lists the UA (product of overall heat transfer coefficient and Area of a Heat

Exchanger), the rate of work (Ẇ) and heat rate (Q) of different components of the ARS-TEG

system:

Component UA [kW/K] Ẇ [kW] Q [kW]

Desorber 0.51 - 10.29

Condenser 0.63 - 6.30

Page 62: Integrated Absorption Refrigeration and Thermoelectric

51

RPC 0.052 - 0.52

Evaporator 1.065 - 5.33

Absorber 0.94 - 9.44

SHX 0.67 - 9.98

Pump - 0.12 -

Table 3.10 Important properties of different components of ARS-TEG system

Meanwhile, 212 TEGs generate 0.81 kW of electrical power from 11.11 kW of heat input

from the exhaust, at a conversion efficiency of ~7%. Whereas the overall COP of the ARS-TEG

system is 0.51.

In order to calculate the area of each heat exchanger in both, ARS-TEG and ORC-VCC,

systems, an estimate of the overall heat transfer coefficient (U) is required. Based on the type of

heat exchanger (Gas-to-Liquid, Gas-to-Gas, plate, shell-tube etc.) exhaustive calculations can be

performed to obtain these estimates. However, for the purposes of current high-level comparison

between ARS-TEG and ORC-VCC systems, estimates from previous studies [82], [83] are listed

in Table 3.11.

Component U [Wm-2

K-1

]

Absorber 940

Boiler 150

Condenser 940

Desorber 150

Evaporator 940

Recuperator 580

SHX 580

RPC 580

Table 3.11 U values for HXs

Using the estimated values from Table 7.10, the areas of respective heat exchangers can

be calculated for the ARS-TEG and ORC-VCC systems. The area values for ORC-VCC, and ARS-

TEG heat exchangers are listed in Table 3.12. and Table 3.13, respectively.

Page 63: Integrated Absorption Refrigeration and Thermoelectric

52

Component UA [kW/K] A [m2]

ORC

Boiler 1.81 12.93

Condenser 1.49 1.58

Recuperator 0.14 0.25

VCC Condenser 0.81 0.86

Evaporator 1 1.06

Table 3.12 Areas of the ORC-VCC HXs

Component UA [kW/K] A [m2]

Desorber 0.51 3.43

Condenser 0.63 0.67

RPC 0.052 0.09

Evaporator 1.065 1.13

Absorber 0.94 1

SHX 0.67 1.14

Table 3.13 Areas of the ARS-TEG HXs

Using the values calculated in this section, costs for different components and devices of

the ARS-TEG and ORC-VCC systems can be calculated for comparison.

3.2.7 Economic Analysis

To object of the current economic analysis for the ARS-TEG and ORC-VCC systems is to

compare the capital investment required to achieve a cooling of 5kW for the refrigerated transport

truck application using waste heat.

The total cost of the systems is given by the Equation 3.17:

𝒁𝑻𝒐𝒕𝒂𝒍 = 𝒁𝑪 + 𝒁𝑳 3.17

Page 64: Integrated Absorption Refrigeration and Thermoelectric

53

Where ZTotal is the total cost in dollars and ZC and ZL are the component and labor costs in

dollars, respectively.

To estimate the costs, correlations from previously published studies were referenced [83]–

[85]. These studies rely upon data from CEPCI [86], with the latest data available from the year

2011. The relevant costing correlations are listed in Table 3.14.

Component Governing Variable Cost Year Cost Index

Compressor V [m3/s] (Volumetric

flowrate) 258.75 + 195.5�� 2011 585.7

Heat Exchangers AHX [m2] (Area of HX) 130(𝐴𝐻𝑋 0.093⁄ )0.78 2000 394.1

Pump Ẇ (Electric Power

input) 1035(�� 300⁄ )

0.25 2001 402.0

Turbine V [m3/s] (Volumetric

flowrate) 1.73(225 + 170��) 2001 394.3

Labor TCC (Total Capital

Cost) 0.3TCC - -

Table 3.14 Component cost relation matrix

To correct the cost for 2019 (latest available yearly average CEPCI cost index), Equation

3.18 is used:

𝒁𝟐𝟎𝟏𝟗 = 𝒁𝑩𝒂𝒔𝒆𝑪𝑰𝟐𝟎𝟏𝟗/𝑪𝑰𝑩𝒂𝒔𝒆 3.18

Where, Z2019 and ZBase are the costs for 2019 and year of reference, respectively, and CI2019

and CIBase are the cost indexes for the year 2019 and year of reference. The value of CI2019 is 607.5

[87].

For the present analysis, heat exchanger costs are calculated using the simplified relation

presented in Table 3.13. However, it should be noted that heat exchanger costs can vary from this

base cost based on the type of heat exchanger (plate, shell-tube, Finned-tube, etc.), material used

and the operational pressure.

Page 65: Integrated Absorption Refrigeration and Thermoelectric

54

Based on these data and correlations presented, total cost for each system can be calculated.

Table 3.15 lists the component costs for the ORC-VCC system:

Component V [m3/s] A [m

2] Ẇ [kW] Z2019 [$]

ORC

Boiler -

12.93 - 9,408.3

Condenser 1.58 - 1,825.66

Turbine 0.022 - 3.47 610

Pump - 0.17 241.3

Recuperator 0.25 - 433.4

VCC

Condenser -

0.86 - 1,136

Evaporator 1.06 - 1,337.2

Compressor 0.00071 - 3.12 398.8

Total: $ 15,390.7

Table 3.15 Capital costs of ORC-VCC system

The total labor costs can be calculated by using the relation provided in Table 3.13. The

total labor cost for the ORC-VCC system is calculated to be $4,617.2. The total capital cost of the

ORC-VCC is, therefore, 20,007.9.

Table 3.16 lists the component cost of the ARS-TEG based WHR system:

Component A [m2] Ẇ [kW] Z2019 [$]

Desorber 3.43 - 5012.85

Condenser 0.67 - 935

RPC 0.09 - 195.3

Evaporator 1.13 - 1,405.6

Absorber 1 - 1,277.8

SHX 1.14 - 1,415.3

Pump - 0.12 221.2

Total: $ 10,463.05

Table 3.16 Capital cost of the ARS-TEG system

Based on the component cost, the total labor cost can also be calculated as $3,138.9. The

TEG cost must also be included in the total calculated, according to the latest available TEG cost

Page 66: Integrated Absorption Refrigeration and Thermoelectric

55

rate [88], the cost of TEGs is calculated as $10,269.Therefore, the total capital cost of ARS-TEG

WHR system is $23,870.1.

3.2.8 Discussion

From the economic analysis conducted, it can be deduced that the cost of ARS-TEG WHR

system is only ~19% higher than the cost of ORC-VCC based WHR system. Moreover, the ARS-

TEG based WHR system delivers 0.81 kW of electrical power in addition to delivering 5 kW of

cooling.

The cost of ARS-TEG system can be significantly lowered by reducing the number of

TEGs. For example, by lowering the number of TEGs from 212 to 106, the total cost of the ARS-

TEG based WHR system can be reduced by $5,134.5 to be 18,736.5, with the TEG power being

reduced from 0.81 kW to 0.41 kW. This arrangement of ARS-TEG WHR will require a desorber

redesign, where half of the desorber heat from the exhaust stream will be delivered to the working

fluid through the TEG cold side, while the remaining heat would be recovered downstream of the

TEG stage.

In practice both WHR systems must be modified to meet the transient temperature

challenges that were alluded to in section 3.2.2 of this chapter. Specifically, the challenges of

delivering the required cooling when exhaust temperatures are either higher or lower than the

design temperature.

For the case of lower than required exhaust temperatures, electric heaters can be employed

to deliver heat to the desorber for a steady operation. In the case of higher than desired temperatures

of the exhaust gases, a bypass may be employed to divert the exhaust gases form the desorber.

Page 67: Integrated Absorption Refrigeration and Thermoelectric

56

Considering the requirement of an electric heater for steady operation of a refrigerated

transport vehicle’s WHR system, ARS-TEG based WHR system is preferable over ORC-VCC

based WHR system, as it can off-set the power required by the electric heater.

Page 68: Integrated Absorption Refrigeration and Thermoelectric

57

Heat Acquisition Unit Design, Development and

Experimentation

Page 69: Integrated Absorption Refrigeration and Thermoelectric

58

The HAU for the proposed integrated, cascaded TEG-ARS WHR system, provides an

enclosure for exhaust gases to transfer heat through the TEGs into the heat transfer fluid. Extended

heat transfer surfaces (fins) and heat spreaders are needed at the different interfaces for effective

operation. Design considerations for such an HAU involve flow arrangement of hot and cold stream

fluids (e.g., co- vs. counter-flow) and location and alignment of TEG modules with respect to the

exhaust stream flow direction. Extended surface sizing and the associated pressure drops, high

thermal resistance of TEGs, and spreading resistances are some of the technical challenges in

designing a HAU

This chapter will discuss these considerations and also provide details about the iterative

design and development of the HAU. That discussion will be accompanied by a comparative

analysis, based on experimental results, for the two HAU designs developed during this study.

4.1 HAU Design Considerations

Several different coupling-fluid flow arrangements have been proposed in previous studies

for optimal performance of fluid-heated TEG systems. Bejan et al. [89], using a first law analysis,

derived a power density comparison for co-flow and counterflow heat transfer to argue for a co-

flow arrangement as optimum for TEG systems. However, Meng et al. [40] and Bell et al. [90]

predicted better overall performance with counter-flow arrangements in their studies. He et al. [42]

found that flow direction did not significantly impact TEG system performance in the case of gas

heated and liquid cooled TEG systems. However, for gas cooled TEG systems, they found that

counterflow arrangements produced greater power for the same module areas.

The power generation of a thermoelectric module array can be estimated as[91]:

Page 70: Integrated Absorption Refrigeration and Thermoelectric

59

𝑷𝑻𝑬𝑮 = (𝑵𝑻𝑬𝑮𝑺∆𝑻𝑻𝑬𝑮)𝟐/(4𝑵𝑻𝑬𝑮𝑹𝑻𝑬𝑮) (4.1)

Equation 4.1 indicates that the power produced by the TEGs is proportional to the square

of the temperature difference between its hot and cold sides. Considering this, for maximum power

generation, the TEG array in an HAU should be positioned where the hot and cold stream have the

greatest temperature difference. The quadratic dependence favors co-flow arrangements which

have a greater temperature difference near the inlet over counter-flow arrangements with more

uniform, but lower peak temperature differences. The local temperature difference can be

maximized at the inlet of a co-flow arrangement of heating and cooling fluids.

Another important consideration for an efficient HAU design is in sizing of the extended

surfaces used in transferring heat from the hot gas stream into the cold liquid stream. Considering

the WHR application of these systems, increase in pressure drops associated with increased size of

extended surfaces, especially on the exhaust/gas side, must be taken into account. For a pin-fin

array type extended surfaces, as illustrated in Figure 1.6 in Chapter 1, the pressure drop can be

predicted with the following correlation [92]:

∆𝒑 = 𝑵𝑳 ∙ 𝝌 ∙ (𝝆 ∙ 𝑽𝒎𝒂𝒙𝟐 𝟐⁄ ) ∙ 𝒇 4.2

Where Δp is the pressure drop in Pa, NL is the number of lateral pin fins, χ is an empirical

correction factor, ρ is density in kg·m-1

, Vmax is the maximum velocity in m·s-1 and f is the friction

factor. For the air-side fins used in this study, f and χ have values of 0.5 and 1 respectively.

Quoilin et al. [84], using a modified form of Equation 4.2, conclude that increasing the

width of a heat exchanger drops 𝑉𝑚𝑎𝑥, consequently lowering the pressure drop, but it also increases

the required area for heat transfer. In light of this observation combined with the observation from

equation 1, a span-wise arrangement of TEGs, close to the inlet of the HAU would result in a more

efficient WHR.

Page 71: Integrated Absorption Refrigeration and Thermoelectric

60

On the cold stream/thermal oil side, the correct sizing of the extended surface/oil-blocks is

important for efficient TEG performance, since the cold side of the TEGs needs to be maintained

at lower temperature to ensure a large temperature difference across the TEGs.

Over-sizing the TEG module array will result in reduced average hot-to-cold side

temperature difference for a given heat transfer rate due to increased overall thermal conductance.

However, under-sizing the array will limit total heat transfer, and reduce available thermal power

for conversion. Rattner [43], using an analytical model, proposed a thermal-fluid figure of merit,

HT, which can be used to identify the optimal number of TEGs for a WHR system that maximizes

delivered electric power by balancing these factors:

𝑯𝑻 = [𝑲 ∙ (𝟏 𝜺𝑯 ∙ ��𝑯⁄ + 𝟏 𝜺𝑪 ∙ ��𝑪⁄ )]−𝟏

4.3

HT is the thermal-fluid figure of merit, K is the TEG module conductance in WK-1, εH and

εC are the thermal effectivenesses of hot and cold heat exchangers, respectively, and ĊH and ĊC are

hot and cold stream thermal capacities in WK-1, respectively.

According to Rattner [43], for a span-wise, single row TEG array, with hot and cold sides

at isothermal temperatures, and a specified overall UA, the optimal number of discrete TEG

modules is equal to HT.

The HAU design must also account for the thermal spreading resistance owing to the

relatively smaller contact surface area of the TEGs and thermal oil side fin blocks compared with

the large heat exchange surfaces of the gas side. Lee et al. [93] provide simple spreading resistance

correlations for uniform heat-flux source, according to which the spreading resistance, for relatively

thick plates, depends entirely on the contact size of the heat source.

To operate at their highest efficiency, TEGs require a high temperature across their hot and

cold sides. If the HAU were divided into two stages, in the first stage high availability waste heat

Page 72: Integrated Absorption Refrigeration and Thermoelectric

61

could be harnessed by the TEGs. While, in the second stage, the absence of high thermal resistance

of TEGs allows for low availability waste heat to be harnessed by the thermal oil stream. This

concept will be further discussed in the next section.

4.2 HAU Simple Engineering Model

This section will discuss the development of a simple engineering model for the HAU-2.

The model solves for the intermediate temperatures of the gas and liquid streams by stage-wise

modeling of the thermal resistances of different components of the HAU. The computation of the

engineering model was conducted using Engineering Equation Solver (EES)[81].

4.2.1 TEG Model

Thermoelectric power delivery can be estimated using the following equation:

𝑷𝑻𝑬𝑮 = (𝑵𝑻𝑬𝑮𝑺∆𝑻𝑻𝑬𝑮)𝟐/(4𝑵𝑻𝑬𝑮𝑹𝑻𝑬𝑮) 4.4

In the case of experiments conducted in this study, ΔTTEG is not known, however, a direct

measurement of the power provided to an external D.C load is recorded for each experimental run,

using which the total power generated by the TEG module array can be calculated. With PTEG

known, ΔTTEG can be inferred from Equation 4.4.

Once the temperature difference across the TEGs is known, the heat across the TEG

module array, QTEG, can be found using Equation 4.5:

𝑸𝑻𝑬𝑮 = ∆𝑻𝑻𝑬𝑮/(𝑵𝑻𝑬𝑮 ∙ 𝑹𝑻𝑬𝑮,𝒄𝒐𝒏𝒅) 4.5

Where, RTEG is the TEG conductive Resistance in K·W-1.

Page 73: Integrated Absorption Refrigeration and Thermoelectric

62

4.2.2 Fin Array Resistance Model

A model was developed to predict the thermal resistances of a fin block arrays, with respect

to geometry, on both air and thermal oil sides. Inlet flowrates and inlet temperatures can be set to

experimental values. The thermophysical properties of the fluid streams were calculated for their

average temperatures in stage 1, using built-in functions in EES.

The Nusselt number correlation used for this analysis is provided in Incropera et al. [92]:

𝑵𝒖 = 𝑪𝟏 ∙ 𝑹𝒆𝑫,𝒎𝒂𝒙𝒎 ∗ 𝑷𝒓𝟎.𝟑𝟔 ∙ (𝑷𝒓 𝑷𝒓𝒔)⁄ 𝟏/𝟒

4.6

Where Nu is the Nusselt number, ReD,max is the maximum Reynolds number occurring

within the fin blocks, Pr and Prs are the Prandtl numbers evaluated at mean temperature of flow

and surface temperature, respectively, C1 and m are a constants dependent on ReD,max. For the fin

array used in this investigation the C1 values for air and oil side were 0.8 and 0.9 respectively,

whereas the m value for both was 0.4.

Once the Nusselt number is calculated the heat transfer coefficient for the fin block array

can be calculated using the following equation:

�� = 𝑵𝒖 ∙ 𝒌𝒇 𝑫⁄ 4.7

Where h is the heat transfer coefficient in W·m-2·K-1, 𝑵𝒖 is the Nusselt number, kf thermal

conductivity of the fluid in W·m-1·K-1, and 𝐷 is the diameter of the pin fin in m.

The efficiency of the pins, 𝜼𝒑, fins was calculated by performing a 1-D heat transfer

analysis and assuming an adiabatic tip condition and the following equations:

𝜼𝒑 =𝐭𝐚𝐧𝐡 𝒎𝑳

𝒎𝑳 4.8

Page 74: Integrated Absorption Refrigeration and Thermoelectric

63

𝒎 =��𝑷

𝒌𝑨𝒄 4.9

Where P is the perimeter of the fin tip in m and Ac is the cross-sectional area of the fin in

the direction of heat flow in m2 and L is the length of the fin.

For the Overall efficiency of the Fin blocks, the following relation was used:

𝜼𝒐 = 𝟏 − [(𝑨𝒇 𝑨𝑻)⁄ ∙ (𝟏 − 𝜼𝒑)] 4.10

Where 𝜼𝒐 is the overall fin array efficiency, 𝑨𝒇, 𝑨𝑻 are the total areas of pin fins and total

area (Pin Fins + Base) in m2, respectively, and 𝜼𝒑 is the fin efficiency.

The overall thermal resistance of fin block array was found using this relation:

𝑹 = 𝟏 (𝑨𝑻 ∙⁄ �� ∙ 𝜼𝒐) 4.11

Where R is the overall fin array thermal resistance W·K-1

, AT is the total area (Pin Fins +

Base) in m2

, ℎ is the heat transfer coefficient in W·m-2

·K-1

, and 𝜼𝒐 is the overall fin array efficiency.

4.2.3 Spreading Thermal Resistance

Due to the difference between the contact surface areas of the heat exchange surfaces, the

engineering model must account for the thermal spreading resistances. The largest spreading

resistances occur at the interface of the copper spreader plate and the TEG module array in stage 1

for HAU-2.

Lee et al. [93] provide a simple spreading resistance model which was applied to the current

model:

Page 75: Integrated Absorption Refrigeration and Thermoelectric

64

𝑹𝒔𝒑 = 𝝓𝒎𝒂𝒙 𝒌𝒔𝒑 ∙ 𝑹𝟏 ∙ √𝝅⁄ 4.12

Where 𝑹𝒔𝒑 is the spreading resistance in W·K-1

, 𝑹𝟏 is the square root of the ratio of area

of TEGs to the constant 𝝅, 𝒌𝒔𝒑 thermal conductivity of the heat spreader in W·m-1

·K-1

and 𝝓𝒎𝒂𝒙

is a constant depending on the geometry of the heat spreader is given by the following equation:

𝝓𝒎𝒂𝒙 = 𝟏

√𝝅(𝟏 − 𝝐)𝚽𝒄 4.13

Where, ε is the dimensionless contact radius given by the ratio of ratio of the hydraulic

diameter of the TEG to the heat spreader and Φc is given by:

𝚽𝒄 = 𝒕𝒂𝒏𝒉 (𝝀𝒄𝝉) +

𝝀𝒄

𝑩𝒊

𝟏 + 𝝀𝒄𝑩𝒊 𝒕𝒂𝒏𝒉 (𝝀𝒄𝝉)

4.14

Where Bi is the Biot number of the heat spreader and τ is the dimensionless spreader

thickness calculated by dividing the thickness of the spreader by its hydraulic diameter. λc is an

empirical parameter given by the following equation:

𝛌𝒄 = 𝛑 + 𝟏

√𝝅𝝐 4.15

4.2.4 Model for Heat Lost to the Ambient

The mathematical model for the current study also accounts for the heat lost to the ambient.

The heat lost was modelled as being dependent on the difference between the temperatures of the

ambient and the average stage temperature on the air side. The following relation was used to

estimate the heat loss for each stage:

𝑸𝒍𝒐𝒔𝒕 = (𝟏 𝑹𝒍𝒐𝒔𝒕⁄ ) ∙ (𝑻𝒂𝒗𝒈 − 𝑻𝒔𝒖𝒓𝒓) 4.16

Page 76: Integrated Absorption Refrigeration and Thermoelectric

65

Where 𝑸𝒍𝒐𝒔𝒕 is the heat lost in W, 𝑅𝑙𝑜𝑠𝑡 is sum of all the thermal resistances of hot air, steel

wall, thermal insulation and the ambient air in W·K-1, 𝑻𝒂𝒗𝒈 is the average temperature for air for

the stage, and 𝑇𝑠𝑢𝑟𝑟 is the ambient air temperature.

As illustrated in Figure 4.9, the heat travels from the hot air stream and the thin steel plate,

in through the fiberglass insulation and to the ambient air. The total resistance of the resulting

thermal circuit is a sum of the convective resistance of the hot stream and ambient air, and the

conductive resistance of the steel wall and fiberglass insulation. The convective resistances were

calculated using the built-in EES functions.

Tst,i

Tst,o = Tfg,i

Tfg,o

Tair,av

Tsurr

Rair,av

Rst

Rfg

Rsurr

Tst,i

Tst,o = Tfg,i

Tfg,o

Tair,av

Tsurr

Qlost

Figure 4.1 Thermal Circuit for Heat Lost model

4.2.5 Stage-wise Total Resistance Calculations

With the total resistances for both fluid streams, the heat transfer and temperatures for each

stage can be calculated. For Stage 1 (Figure 4.10) the resistances are calculated as two parts of the

thermal circuit between the node TTEG,H:

𝑹𝒕𝒐𝒕𝒂𝒍,𝑯 = 𝑹𝒇𝒂,𝟏 + 𝑹𝒇𝒃𝒂,𝟏 + 𝑹𝑪𝒔𝒑 4.17

Page 77: Integrated Absorption Refrigeration and Thermoelectric

66

𝑹𝒕𝒐𝒕𝒂𝒍,𝑪 = 𝑹𝑻𝑬𝑮 + 𝑹𝑨𝒔𝒑,𝟏 + 𝑹𝒇𝒃𝒐,𝟏 + 𝑹𝒇𝒐,𝟏 4.18

Where 𝑹𝒕𝒐𝒕𝒂𝒍,𝑯 and 𝑹𝒕𝒐𝒕𝒂𝒍,𝑯 are the thermal resistances for the heat delivered to and from

the hot junction and cold of the TEG (QH & Qoil,1) respectively. 𝑹𝒇𝒂,𝟏, 𝑹𝒇𝒃𝒂,𝟏 are the Stage 1 thermal

resistances for air side fins and fin base, respectively. 𝑹𝑪𝒔𝒑, 𝑹𝑨𝒔𝒑,𝟏 are the Stage 1 spreading

resistances for aluminum and copper plates, respectively. 𝑹𝑻𝑬𝑮 is the TEG thermal resistance.

𝑹𝒇𝒃𝒐,𝟏, 𝑹𝒇𝒐,𝟏 are the Stage 1 thermal resistances for oil side fin base and fins, respectively. All

resistances have units of W·K-1.

Rfa,1 Rfba,1 RCsp RTEG RAsp,1 Rfbo,1 Rfo,1

PTEG

QH Qoil,1

Toil,in Toil,int

Tair,in Tair,intQH

Qoil,1

PTEG

TTEG,H TTEG,C

Qlost,1

STAGE 1

Figure 4.2 Stage-1 Thermal Resistance Model

For Stage 2 (Figure 4.11), total thermal resistance is calculated using the following

equation:

𝑹𝒕𝒐𝒕𝒂𝒍,𝟐 = 𝑹𝒇𝒂,𝟐 + 𝑹𝒇𝒃𝒂,𝟐 + 𝑹𝑨𝒔𝒑,𝟐 + 𝑹𝒇𝒃𝒐,𝟐 + 𝑹𝒇𝒐,𝟐 4.19

Here, 𝑹𝒕𝒐𝒕𝒂𝒍,𝟐 is stage 2 total thermal resistance, 𝑹𝒇𝒂,𝟐, 𝑹𝒇𝒃𝒂,𝟐 are the Stage 2 thermal

resistances for air side fins and fin base, respectively. 𝑹𝑨𝒔𝒑,𝟏 is the Stage 2 spreading resistances

for Aluminum plate, 𝑹𝒇𝒃𝒐,𝟐, 𝑹𝒇𝒐,𝟐 are the Stage 2 thermal resistances for oil side fin base and fins,

respectively.

Page 78: Integrated Absorption Refrigeration and Thermoelectric

67

Rfa,2 Rfba,2 RAsp,2 Rfbo,2 Rfo,2

Qoil,2

Toil,int Toil,out

Tair,int Tair,out

Qoil,2

Qlost,2

STAGE 2

Figure 4.3 Stage-2 Thermal Resistance Model

4.2.6 Heat Transfer Calculations

For Stage 1, the model solves the following equations simultaneously to find the stage exit

temperatures and heat transfers:

𝑸𝒂𝒊𝒓,𝟏 = 𝑸𝒐𝒊𝒍,𝟏 + 𝑸𝒍𝒐𝒔𝒕,𝟏 + 𝑷𝑻𝑬𝑮 4.20

𝑻𝒐𝒊𝒍,𝒂𝒗𝒈,𝟏 = (𝑻𝒐𝒊𝒍,𝒊𝒏 + 𝑻𝒐𝒊𝒍,𝒊𝒏𝒕)/𝟐 4.21

𝑻𝒂𝒊𝒓,𝒂𝒗𝒈,𝟏 = (𝑻𝒂𝒊𝒓,𝒊𝒏 + 𝑻𝒂𝒊𝒓,𝒊𝒏𝒕)/𝟐 4.22

𝑸𝒂𝒊𝒓,𝟏 = (𝑻𝒂𝒊𝒓,𝒂𝒗𝒈,𝟏 − 𝑻𝑻𝑬𝑮,𝑯)/𝑹𝒕𝒐𝒕𝒂𝒍,𝑯 4.23

𝑸𝒐𝒊𝒍,𝟏 = (𝑻𝑻𝑬𝑮,𝑪 − 𝑻𝒐𝒊𝒍,𝒂𝒗,𝟏)/𝑹𝒕𝒐𝒕𝒂𝒍,𝑪 4.24

𝑸𝒂𝒊𝒓,𝟏 = ��𝒂𝒊𝒓 ∙ [𝒉𝒂𝒊𝒓(𝑻𝒂𝒊𝒓,𝒊𝒏) − 𝒉𝒂𝒊𝒓(𝑻𝒂𝒊𝒓,𝒊𝒏𝒕)] 4.25

𝑸𝒐𝒊𝒍,𝟏 = ��𝒐𝒊𝒍 ∙ [𝒉𝒐𝒊𝒍(𝑻𝒐𝒊𝒍,𝒊𝒏𝒕) − 𝒉𝒐𝒊𝒍(𝑻𝒐𝒊𝒍,𝒊𝒏)] 4.26

𝑸𝒂𝒊𝒓,𝟏, 𝑸𝒐𝒊𝒍,𝟏, 𝑸𝒍𝒐𝒔𝒕,𝟏 are the Stage 1 heat Transfer rates for air, oil and heat loss,

respectively, in W. 𝒉𝒂𝒊𝒓(𝑻𝒂𝒊𝒓,𝒊𝒏𝒕), 𝒉𝒂𝒊𝒓(𝑻𝒂𝒊𝒓,𝒊𝒏) are the enthalpies for air exiting stage 1 and air

inlet, respectively, in J·kg-1. 𝒉𝒐𝒊𝒍(𝑻𝒐𝒊𝒍,𝒊𝒏𝒕), 𝒉𝒐𝒊𝒍(𝑻𝒐𝒊𝒍,𝒊𝒏) are the enthalpies for oil exiting stage 1 and

Page 79: Integrated Absorption Refrigeration and Thermoelectric

68

oil inlet, respectively, in J·kg-1, while ��𝒂𝒊𝒓 , ��𝒐𝒊𝒍 are the mass rates for air and oil, respectively, in

kg·s-1. While 𝑻𝒐𝒊𝒍,𝒂𝒗𝒈,𝟏 and 𝑻𝒂𝒊𝒓,𝒂𝒗𝒈,𝟏 are the average temperatures of oil and air side in K, within

the 1st stage.

A similar algorithm is used for stage 2 to find the temperatures of the liquid and gas streams

at the exit and the heat rates in stage 2.

Page 80: Integrated Absorption Refrigeration and Thermoelectric

69

4.3 HAU Iterative Development and Experimental Procedure

For the proposed study, two different HAU designs were developed, experimentally tested

and the results were compared to evaluate different HAU design tradeoffs. The first HAU (HAU-

1) was based on a simpler, single stage, compact design.

The HAU enclosure consists of a hexagonal base-plate sub-assembly with welded

sidewalls (Figure 4.4(b)) and a hexagonal cover-plate (Figure 4.4(a)) joined to it with standoffs.

Two staggered, pin-fin arrays, composed of three individual fin blocks, were screwed into each of

the two plates with a thermal contact compound applied at the mating surfaces.

10 cm

Figure 4.4 (a) Cover-plate sub-assembly (b) Base-plate subassembly

For this design, water-blocks were mounted using Aluminum coupling plates on either side

of the HAU, through which the cooling thermal oil flows in an oil-block. Manifolds were used for

thermal-oil flow distribution at both the inlet and outlet. The TEG arrays, comprising of three

individual commercially available TEG modules [94], on either side of this HAU were sandwiched

between the outer surface of the plates and the coupling plate (6 TEG modules in total).

Page 81: Integrated Absorption Refrigeration and Thermoelectric

70

Table 4.1 shows the details of the main components used in HAU-1:

Component Dimension (L x W x H) [mm] Material Qty

TEG [94] 56 x 56 x 3 Beryllium Telluride 6

Enclosure 400 x 250 x 100 Carbon Steel 1

Heatsink [95] 80 x 80 x 40 Aluminum 6

Coupling Plate 235 x 108 x 22 Aluminum 2

Oil-block 162 x 40 x 12 Aluminum 2

Table 4.1 HAU-1 component details

(a) (b) (c)

Figure 4.5 (a) TEGs mounted on the outside the HAU (b) Coupling plate mounted on top of the

TEGs (c) Complete assembly of the HAU-1 with oil-block cover plate

Figure 4.5 shows this assembly stack-up of HAU-1, with Figure 4.5-(a) showing the

location of three of the six TEGs and Figure 4.2-(b) shows the coupling plate that houses the water-

block, while, Figure 4.5-(c) shows the completed HAU-1 assembly.

The internal water-block structure and the complete HAU assembly is shown in Figure

4.6.

Page 82: Integrated Absorption Refrigeration and Thermoelectric

71

17 cm

4 cm

Coupling Plate

Water-block

internal

structure

Figure 4.6 Assembled HAU-1

4.3.1 HAU Experimental Subsystem Description

Figure 4.7 illustrates the schematic for the HAU evaluation experimental system.

Simulated hot exhaust gas (hot air) from a hot air blower enters the HAU at Point 1, where it

transfers heat to the TEG arrays and exits the system at a lower temperature at Point 2. The TEGs,

mounted on the outside surface of the HAU, convert a portion (~5%) of the acquired heat into

electricity. Much like any other D.C electric source, a TEG array’s maximum power output occurs

when it is connected to a matched DC load [96]. Therefore, the TEG array power is delivered to a

DC load that matches the internal electrical resistance of the TEG array. The cold junction sides of

the TEG modules are cooled by a coupling fluid (thermal oil) through aluminum heat sinks.

Between Points 3→4, the coupling fluid receives heat rejected by TEG array and its temperature

increases. By coupling the thermal oil with a laboratory chiller between 4→5 its temperature is

Page 83: Integrated Absorption Refrigeration and Thermoelectric

72

lowered to required inlet temperature to the HAU. A pump, located between 5→1, then recirculates

the coupling fluid through the system.

V

ΔP

P

T

T

T

Matched

DC Load

Pitot Tube Sensor

1

2

45

3

T

HAU Experimental Evaluation System

Lab Chiller

External Chiller

HAU & TEG

Hot Air

Blower

5.6975 V 39.998A

227.887 W 0.199Ω

ON

Figure 4.7 Experimental setup for HAU subsystem evaluation

This description of HAU refers to the first version of HAU (HAU-1) developed during this

study. In the second HAU version (HAU-2) the HAU is divided into two WHR stages. TEGs are

only present in the first stage and, in the second stage, heat is transferred directly between the

exhaust stream and coupling fluid through aluminum heat spreaders.

4.3.2 Design improvements to HAU-2 based on HAU-1 performance

HAU-1 was insulated using fiber glass insulation and its performance was evaluated by

performing experiments over a range of inlet simulated exhaust (air) temperatures and flow rates.

The simulated exhaust gas was provided by a hot air blower, which has controls built-in to regulate

Page 84: Integrated Absorption Refrigeration and Thermoelectric

73

the heat rate and flowrate of the air. The average ambient temperature was recorded as 20°C. The

oil mass flowrate, inlet temperature and air mass flowrate were held relatively constant (0.02 kg·s-

1, ~50°C and ~0.01 kg·s-1, respectively) for these experimental runs while the air inlet temperature

was varied. The experiment was performed under a ventilation hood and for each case, the HAU

was operated for 15-30 minutes at steady state and measurements were recorded every 2 seconds

and averaged at the end of the run. A summary of the results is presented in Table 4.1.

Max Blower Heat

& Flowrate

50% of Max Flowrate

& Max Blower Heat

50% of Max Blower

Heat & Max Flowrate

Tair,in (°C) 286±1 382±1 210±1

mair (kg s-1) 0.01±0.005 0.007±0.003 0.008±0.004

Toil,in (°C) 52.6±1 58.2±1 41.8±1

Toil,out (°C) 68.9±1 76.8±1 53.6±1

Qoil,in (kW) 0.67±0.06 0.81±0.06 1.034±0.05

moil (kg·s-1) 0.02±0.0002 0.02±0.0002 0.02±0.0002

ΔPHAU (Pa) 34.3±0.5 17±0.5 17.3±0.5

PTEG (W) 3.505±0.02 4.081±0.02 2.011±0.02

Table 4.2 Results from HAU-1 experimental runs

From the data presented in Table 4.2, it can be seen that the power generation by the TEGs

is lower than the design requirements (described in Section 1.4 → ~25W). Based on the results of

the experiments, the following observations were made:

• Spreading Resistance: As shown in Figure 4.8-(a), the mismatch between the

contact area of the TEG array relative to the larger area heatsink causes high

thermal spreading resistance (Rsp = 0.023 K·W-1). This results in lower hot

junction temperatures for the TEG, the efficiency of a TEG is higher at higher

Page 85: Integrated Absorption Refrigeration and Thermoelectric

74

hot junction temperatures for a given cold side temperature. In order to lower

this spreading resistance a thick copper heat spreader can be used (Figure 4.8-

(b)).

• Oil-block sizing: Another factor contributing towards a low TEG power

production is the high oil-block thermal resistance as shown in Figure 4.8 (Rob

= 0.38 K·W-1). This high thermal resistance results in higher than desired

temperatures on the TEG cold junction as the thermal oil cannot recover the

desired amount of heat. This results in a smaller temperature gradient across

the TEGs than desired, leading to a lower efficiency operation of the TEG

array.

Fin

blo

ck

TE

G

Wall

0.005

K/W

0.009

K/W0.0015

K/W

Oil

Block

ṁoil= 0.01 kg/s

ṁair= 0.001 kg/s

RfinRbRTEG

(b)0.38

K/W

Rob H

eat

Sp

read

erRsp

0.013

K/W

Fin

blo

ck

TE

G

Wall

0.005

K/W

0.023

K/W

0.0015

K/W

0.38

K/W

Oil

Block

ṁoil= 0.01 kg/s

ṁair= 0.001 kg/s

RfinRspRTEG Rob

(a)

Figure 4.8 Equivalent Thermal Circuits:(a) No heat-spreader: Largest temperature drop occurs

between the air-side fins and TEG, (b) Copper heat-spreader: Spreading resistance drops and the

largest temperature drop occurs between the two sides of the TEG

• TEG array size: Another contributing factor to the low temperature gradient

across the TEG array was identified as an oversized TEG array. For a higher

Page 86: Integrated Absorption Refrigeration and Thermoelectric

75

oil side thermal resistance and lower air side thermal resistance, increasing the

number of TEG results in more surface area for heat extraction, however, the

limiting heat transfer potential of the oil-side heat exchange surfaces results in

“thermal-short circuit” where the TEG’s hot and cold junction temperatures

come closer and little heat transfer can be achieved.

Based on the experimental results and observations from HAU-1, an improved design HAU

(HAU-2) was developed. The following major changes were made:

• The HAU was divided into two stages and the gas side fin-block arrays were

installed only on one side of the HAU.

• Cut-outs were made onto the baseplate sub-assembly allowing for gas side fin-

block bases to be directly mounted onto the coupling plates, thereby, reducing

thermal resistance.

• A high conductivity heat spreader plate, made of copper, was used in stage 1

to minimize spreading thermal resistance due to the difference between contact

surface area of the TEGs and heat exchange surfaces on the gas side.

• The sizing of the oil blocks, especially for stage 1, was increased to allow for

a lower temperature on the TEG cold side.

• The optimal TEG array size was determined based on the thermal-fluid figure

of merit parameter proposed by Rattner [43]. This led to a reduction in the

number of TEGs and a single array composed of two individual TEG modules

was used.

Page 87: Integrated Absorption Refrigeration and Thermoelectric

76

Figure 4.9 (a) Thermal Oil side view of HAU-2 (b) Air side view of HAU-2 (c) Illustration of

HAU-2 thermal oil flow path

The pictures in Figure 4.9(a) and Figure 4.9(b) show the dual stage arrangement on the

thermal oil side and simulated exhaust (air) side, respectively. Figure 4.9(c) shows an illustration

of the HAU, with the thermal oil flow path highlighted. The first stage, with the TEG array, has a

co-flow arrangement, whereas the second stage is in a counterflow arrangement.

All thermal contact surfaces were polished and cleaned to ensure good thermal contact.

Thermal paste and thermal adhesives were also applied where required.

60 cm

33 cm

Page 88: Integrated Absorption Refrigeration and Thermoelectric

77

For HAU-2, in the first stage, high temperature air passes through a set of fin blocks,

transferring heat to a copper coupling plate that servers as the heat spreader. The two TEG modules

yielded more electricity output than the six identical modules used in HAU-1. This electricity is

supplied to a variable DC load adjusted to match the internal resistance of the TEG array and

thereby operate at maximum power point. In the second stage (no TEGs) an Aluminum coupling

plate is used as spreading resistance is not as great a concern as in stage 1.

To performance of HAU-2 was evaluated over a range of simulated exhaust and thermal

oil input temperatures and flow rates. The tests were performed according to a primary test matrix

which is presented in Table 4.3.

TEST MATRIX

Quantities Varied

Tair,in mair Toil,in moil

250, 312, 400°C 0.005, 0.01 kg s-1 50, 75, 100°C 0.02, 0.01 kg s-1

Quantities Measure

Air and Oil Inlet Temperatures and Flow Rates, Air Side Pressure Drop, DC Load Power, Ambient Temperature

Quantities Calculated

Heat Input to TEG array, Heat Input to Thermal Oil, Heat Lost to Surroundings, Total TEG Power

Table 4.3 Test Matrix for HAU-2

Similar to the test runs for HAU-1, the hot air blower was used to moderate the input

flowrate and heat rate for air. The input temperature for the thermal oil was regulated using an

external recirculating chiller. The experiment was run under a ventilation hood and the average

ambient temperature for the thirty test runs was approximately 20°C. As in the HAU-1 tests, the

system was operated at a steady condition for 15-30 minutes and the measurements were then

recorded every 2 seconds and averaged at the end of each run.

Page 89: Integrated Absorption Refrigeration and Thermoelectric

78

A graph illustrating the electrical power delivered by HAU-1 and HAU-2 for different air

inlet temperatures (Toil,in = ~50°C ��𝒐𝒊𝒍 = 0.02 kg · s−1, ��𝒂𝒊𝒓 = 0.01 kg · s−1), is presented in

Figure 4.10:

200 220 240 260 280 300 320 340 360 380 400

2

4

6

8

10

12

14

16

18

20

22 PTEG,tot,HAU-1

PTEG,tot,HAU-2

PT

EG [

W]

Tair,in [°C]

TEG Power Output

Figure 4.10 TEG Power Output Comparison between HAU-1 and HAU-2

From the graph it can be seen that for similar air inlet temperatures and operating

conditions, HAU-2 design generates substantially greater power compared to HAU-1.

Figure 4.11 show the oil outlet temperatures for HAU-2, for the same inlet conditions and

flowrates as described above:

Page 90: Integrated Absorption Refrigeration and Thermoelectric

79

200 220 240 260 280 300 320 340 360 380 400

50

60

70

80

90

Toil,out,HAU-1

Toil,out,HAU-2

To

il,o

ut [

°C]

Tair,in [°C]

Outlet Oil Temperature

Figure 4.11 Oil Outlet Temperature Comparison between HAU-1 and HAU-2

The higher thermal oil outlet temperatures for HAU-2 indicate a greater amount of heat

recovered by HAU-2, compared to HAU-1. These comparisons show that the improvements made

to the HAU design have resulted in a better WHR performance efficiency for HAU-2.

Page 91: Integrated Absorption Refrigeration and Thermoelectric

80

System Description and Cycle Modeling

Page 92: Integrated Absorption Refrigeration and Thermoelectric

81

5.1 System Description

Figure 5.1 illustrates the proposed experimental WHR system cycle. A steady-state cycle

model was developed that applies mass, species, and energy balances to each component assuming

representative closure parameters (e.g., heat exchanger approach temperatures, pump efficiency,

COTS TEG performance parameters). Details on the modeling approach are discussed in Chapter

4.

Figure 5.1 Cycle diagram for the proposed TEG and ARS based Integrated cascaded WHR

system

The absorption system is modeled based on the NH3-LiNO3 working fluid pair, which

avoids the need for a rectifier and permits a low desorber temperature. According to a study by Sun

[97], at lower desorber temperatures (55°C-75°C), NH3-LiNO3 based ARS can achieve higher

COPs better than ammonia-water based ARS.

Page 93: Integrated Absorption Refrigeration and Thermoelectric

82

ARSs like ammonia-water based absorption refrigeration require an analyzer and a rectifier

to remove water vapor from the refrigerant mixture leaving the desorber. Since, with the NH3-

LiNO3 working fluid pair, the absorbent is non-volatile, the analyzer and rectifier are not required.

The TEG subsystem shown in Figure 5.1, can be operated separately from the ARS

subsystem by connecting the coupling fluid stream to a heat exchanger coupled to a laboratory

chiller. In this manner the TEG subsystem’s performance can be evaluated independently of the

ARS subsystem.

When the ARS and TEG subsystems are thermally connected by the coupling fluid,

simulated hot exhaust gas (hot air) from a hot air blower enters the HAU at Point 1, in which it

transfers heat to the TEG arrays and exits the system at a lower temperature at Point 2. The TEGs

are mounted on the outside surface of the HAU. The TEGs convert a portion of the acquired heat

into electricity. Much like any other D.C source, a TEG array’s maximum power output occurs

when it is connected to a matched DC load [96]. Therefore, the TEG array power is delivered to a

DC load that matches the internal electrical resistance of the TEG array. The cold junction sides of

the TEG modules are cooled by a coupling fluid (thermal oil) through aluminum heat sinks.

Between Points 3→4, the coupling fluid receives heat rejected by TEG array and its temperature

increases. This description of HAU refers to the first version of HAU (HAU-1) developed during

this study. In the second HAU version (HAU-2), heat transfer to the thermal oil is divided into two

stages. TEGs are only present in the first stage and in the second stage, heat is transferred directly

between the exhaust stream and coupling fluid through aluminum heat spreaders. A more detailed

discussion of HAU design is provided in Chapter 5.

The coupling fluid transfers heat to the pressurized salt solution (NH3+LiNO3) as it flows

through the desorber (Points 4→5). The cooled oil is pumped back into the HAU (Point 3).

Page 94: Integrated Absorption Refrigeration and Thermoelectric

83

Dilute, high-pressure, ammonia-lithium nitrate solution (NH3+LiNO3) enters the desorber

at Point 6 and heat transfer from the oil stream desorbs ammonia vapor (NH3). At the outlet of the

desorber, the ammonia vapor and concentrated solution are separated in a small separator tank. The

concentrated NH3+LiNO3 solution enters the recuperative solution heat exchanger (SHX, Point

7→8) and cools as it preheats the dilute solution (Points 6→12). Downstream of the SHX, the

concentrated solution expands through a valve to a lower pressure before entering the absorber at

Point 10.

The NH3 vapor exiting the separator tank liquefies as it rejects heat to the ambient in the

condenser (Points 13→14). The liquid refrigerant is recuperatively subcooled as it flows through

the refrigerant pre-cooler (RPC, Points 14→15). It is then expanded through a valve between Point

16→17. The NH3 refrigerant evaporates in the evaporator, delivering cooling to a thermal load

(Points 17→18). The refrigerant vapor flows through the RPC, precooling the refrigerant liquid

and it exits the RPC at Point 19.

In the experimental facility, the evaporator thermal load is simulated by a stream of 50/50

ethylene glycol-water mixture (between Pts. 20→21). The inlet temperature of the ethylene glycol

mixture is controlled with an electric heater driven by an automated process controller.

The NH3 vapor exiting the evaporator mixes with the low pressure concentrated

NH3+LiNO3 solution (Point 9) at the inlet to the absorber (Point 10). The vapor absorbs into the

solution, rejecting heat to the ambient, and exits at Point 11. The liquid mixture is pumped to high

pressure (Points 11→12), and continues to the SHX where it is recuperatively preheated to Point

6, completing the cycle.

Page 95: Integrated Absorption Refrigeration and Thermoelectric

84

The absorber and condenser reject heat to the ambient air (nominally at room temperature)

with the aid of fans mounted on these heat exchangers. The fan speeds can be continuously varied

in the experimental facility.

The next section of this chapter discusses the instrumentation, measurement uncertainties

and data collection method.

5.2 Cycle Model

For the proposed study, steady state thermodynamic models for refrigerated transport

vehicle, carburizing furnace application and the scaled experimental facility were developed using

the Engineering Equation Solver (EES) [81] software. In case of vehicle application model and

experimental model, the cooling delivery temperature is -15°C and the ambient temperature is

30°C, and for the carburizing furnace model the cooling delivery temperature is 5°C. A constant

exhaust inlet temperature of 500°C was also assumed. The cooling delivery temperature of -15°C

for the refrigerated transport truck is the temperature of the refrigerated enclosure This chapter

discusses the approach used in developing the three models.

5.2.1 Conservation Laws

The conservation laws applied in these models are described below:

• Mass conservation:

All three models apply mass conservation law to the overall system and at the component

level (heat exchanger, pump, valve, etc.). When applied to a component of the ARS, the total mass

Page 96: Integrated Absorption Refrigeration and Thermoelectric

85

entering each component equals the mass flow out. This is represented mathematically in Equation

5.1:

∑ ��𝒊𝒏 = ∑ ��𝒐𝒖𝒕 5.1

• Species conservation:

In addition to mass flowrate, the models also apply species conservation at the component

level. Salt mass fraction, X, is described as the ratio between the mass of salt (NH3-LiNO3) and

total mass (ammonia + NH3+LiNO3). So, for instance, in the condenser where the only specie is

ammonia, X = 0. The specie conservation is expressed mathematically in Equation 5.2:

∑ ��𝒊𝒏 ∙ 𝑿𝒊𝒏 = ∑ ��𝒐𝒖𝒕 ∙ 𝑿𝒐𝒖𝒕 5.2

• Energy Balance:

The models apply energy balance at both component level and overall system level. At the

system level, the energy inputs to the system are in the form of waste heat recovered from the

exhaust stream, heat received by the evaporator and the pump work. Energy output from the system

is in the form of TEG electrical power generation, absorber and condenser heat rejection. A similar

approach is used at the component level where energy balance law is applied at the inlet and outlet

of each component.

Page 97: Integrated Absorption Refrigeration and Thermoelectric

86

5.2.2 Closure Parameters

In developing the cycle models, several closure parameters were used for the in this study.

The closure parameters for cooling delivery temperature and ambient temperature for the ARS

subsystem were based on the guidelines provided by AHRI [72].

A 10% desorber outlet vapor quality was assumed for all three models and reasonable

closest approach temperature (CAT) for heat exchangers, superheat for evaporator and subcooling

for absorber and condenser were also assumed. A reasonable pump efficiency of 50% was also

assumed as a closure parameter. Table 5.1 lists these values:

Component Parameter

Desorber CAT 10 K

DoSH 2 K

SHX CAT 15 K

RPC CAT 10 K

Condenser CAT 10 K

DoSC 2 K

Evaporator CAT 5 K

DoSH 2 K

Condenser CAT 10 K

DoSC 2 K

Absorber CAT 10 K

DoSC 2 K

Table 5.1 Heat Exchanger Closure Parameters for ARS-TEG

In the above table CAT, DoSH and DoSC stand for ‘Closest Approach Temperature’,

‘Degree of Superheat’ and ‘Degree of Subcooling”, respectively. Additionally, the CAT between

the liquid and vapor phase at the desorber outlet was assumed to be 2 K. Moreover, the CAT from

the exhaust stream to the TEG hot side is assumed to be 75 K and CAT form the TEGs to the

desorber is assumed to be 40 K. To model the commercial off-the-shelf (COTS) TEG module [94]

Page 98: Integrated Absorption Refrigeration and Thermoelectric

87

considered for this study, the closure parameters adopted include the assumption that the modules’

electrical resistance, conductivity and Seebeck coefficient, provided by the manufacturer, are

constant with respect to temperature. Reasonable CATs were assumed between TEG array and the

exhaust stream, and between TEG array and desorber.

Based on the conservation laws and closure parameters discussed above, desorber balances

are calculated using the following equations:

��𝟔 = ��𝟏𝟑 + ��𝟕 5.3

��𝟔𝑿𝟔 = ��𝟏𝟑𝑿𝟏𝟑 + ��𝟕𝑿𝟕 5.4

��𝟔𝒉𝟔 + 𝑸𝑫𝑬𝑺 = ��𝟏𝟑𝒉𝟏𝟑 + ��𝟕𝒉𝟕 5.5

Where ṁ6 is the mass flowrate of the concentrated solution (ṁcs) at the inlet of the desorber

and ṁ13 and ṁ7 are the mass flowrates of the refrigerant (ṁref) and dilute solution (ṁds) exiting the

desorber, respectively. The symbol h represents the enthalpy with the respective state point as its

subscript.

For the condenser the heat rejection rate is calculated using the following equation:

𝑸𝒄𝒐𝒏𝒅 = ��𝑟𝑒𝑓(𝒉𝟏𝟑 − 𝒉𝟏𝟒) 5.6

The cooling delivered by the evaporator is calculated using equation 5.7:

𝑸𝒆𝒗𝒂𝒑 = ��𝑟𝑒𝑓(𝒉𝟏𝟖 − 𝒉𝟏𝟕) 5.7

The absorber’s heat rejection rate is calculated using equation 5.8:

𝑸𝒆𝒗𝒂𝒑 + ��𝑐𝑠𝒉𝟏 = ��𝑟𝑒𝑓𝒉𝟏𝟗 + ��𝑑𝑠𝒉𝟗 5.8

The work done by the pump is calculated using equation 5.9:

��𝒑 = ��𝒄𝒔(𝒉12 − 𝒉11) 5.9

To calculate the overall COP of the combined system, Equation 5.10 is used:

Page 99: Integrated Absorption Refrigeration and Thermoelectric

88

𝑪𝑶𝑷 = 𝑸𝑒𝑣𝑎𝑝/𝑸𝑫𝑬𝑺 5.10

Finally, with all the heat rates, and inlet and outlet temperatures calculated for the heat

exchangers, UA, for each heat exchanger can be calculated using Equation 5.11:

𝑼𝑨 = 𝑸/𝑳𝑴𝑻𝑫 5.11

Where Q is the rate of heat transfer in a heat exchanger and LMTD is the log-mean

temperature difference of the heat exchanger in question.

5.2.3 Results

With the closure parameters as input and by using the conservation laws described above,

the model calculates, among other quantities, component wise input and output temperatures and

pressures, quality of vapor, composition of species and, HX overall heat transfer coefficients and

heat rates. A summary of results from the model for the two applications and experimental facility

considered in this study are provided in Table 5.2.

Application Texhaust(°C) QTEG,in(kWth) Qelec (kWe) Qcooling(kWth) ηe(%) COP

Carburizing Furnace 500 50 5.04 26 6.5 0.6

Vehicle Exhaust 500 18 0.564 9.23 6.5 0.53

Experimental Model 500 1.8 56 0.923 3 0.5

Table 5.2 Summary of model results

Table 5.3 shows state point-wise (corresponding to Figure 6.1) results for the carburizing furnace

application:

State Point m [kg s-1] h [J kg-1] P [Pa] T [C] X

6 0.15 70002.00 1643000.00 89.47 0.53

7 0.13 159021.00 1643000.00 115.90 0.62

8 0.13 -30021.00 1643000.00 55.00 0.62

9 0.13 -30021.00 398341.00 55.41 0.62

Page 100: Integrated Absorption Refrigeration and Thermoelectric

89

11 0.15 -93125.00 398341.00 40.00 0.53

12 0.15 -90684.00 1643000.00 40.41 0.53

13 0.02 1509000.00 1643000.00 117.90 0.00

15 0.02 190790.00 1643000.00 40.00 0.00

16 0.02 116352.00 1643000.00 24.64 0.00

17 0.02 116352.00 398341.00 -2.00 0.00

18 0.02 1265000.00 398341.00 0.00 0.00

19 0.02 1340000.00 398341.00 30.00 0.00

Table 5.3 Point-wise results for ARS-TEG cycle

Page 101: Integrated Absorption Refrigeration and Thermoelectric

90

Experimental Results

Page 102: Integrated Absorption Refrigeration and Thermoelectric

91

6.1 Experimental Facility

Figure 6.1 shows the experimental lab facility developed for the current study. The facility

was designed to deliver up to 50 W of electricity and 1 kW of cooling with a 3.7 kW electric hot

air blower simulating a waste heat source. The facility is designed to approximate a 1/10th of the

scale of the proposed refrigerated transport vehicle WHR system, which could provide 500 W of

electrical power and 10 kW of cooling. This experimental facility is operated under a ventilation

hood with curtains that can be closed when the system is charged with ammonia.

THERMAL SYSTEMEXPERIMENT CONTROLS

(a) (b)

HAU

Figure 6.1 Experimental Facility: (a) Experimental Facility highlighting experiment controls

(including DC load) and thermal system with the location of HAU highlighted, (b) Picture shows

the HAU-1 model and simulated exhaust (hot air blower)

The scaled experimental system shown above, was tested according to application specific

test matrix to simulate different operating conditions of the vehicle. The lower exhaust temperatures

and flowrates simulate vehicle’s low RPM operation whereas, the higher exhaust temperatures and

flowrates represent vehicle operating at a higher RPM. This matrix is presented in Chapter 4 –

Table 4.3.

Page 103: Integrated Absorption Refrigeration and Thermoelectric

92

Table 6.1 shows a control matrix that details the different key parameters that will be

controlled, measured or calculated.

CONTROL MEASURE CALCULATE

HAU

Simulated Exhaust Gas mair,in and Tair,in Tair,out Qlost

Coupling Fluid moil,in and Toil,in Toil,out Qoil,in

TEG Modules PTEG ηTEG

Refrigeration

Circuit

Pump mpump ppump,out, Tpump,out Ppump

Refrigerant pevap,in and Tevap,in mref ,Tref,out

Simulated Cooling Load mCL, TCL,in TCL,out Qload, COP

Table 6.1 Control Matrix for Experimental Facility

6.2 Instrumentation

Table 6.2 lists the different types of instruments used to measure temperatures, flowrates

and pressures for the four fluid loops in the WHR system. Also included in the table are the

measurement uncertainties associated with these measurements. Almost all of the temperature

measurement devices were calibrated against a high accuracy reference thermometer, yielding

estimated uncertainties of ±0.25°C. The high temperature thermocouples for the simulated exhaust

gas stream were not calibrated, and therefore are assumed to have ±1°C uncertainty.

For data collection, a data acquisition unit (DAQ) is interfaced with a PC running LabView

[98], and a LabView code was setup to save the collected measurements to a spreadsheet file.

The TEG power output is recorded manually. The readings from the DC resistive load

monitor are recorded twice; once at the start of the experimental run when the system achieves a

relative steady state, and again just before the experimental run is complete. The final reading ,

which is the average of the two readings, has an uncertainty of ±0.2% [99].

Page 104: Integrated Absorption Refrigeration and Thermoelectric

93

Temp (°C) Flowrate Pressure (Pa) Type Unc Type Unc Type Unc

Exhaust K ±1 Diff Pressure ±0.15 L m-1 Diff Pressure ±0.31

Coupling Fluid T ±0.25 Oval Gear ±0.04 L m-1 Transducer ±6 kPa

NH3-LiNO3 T ±0.25 Ultrasonic ±0.02 L m-1 Transducer ±6 kPa

NH3 T ±0.25 Ultrasonic ±0.006 L m-1 Transducer ±6 kPa

Ethelene Glycol T ±0.25 Oval Gear ±0.04 L m-1 - -

Table 6.2 Measurement types and uncertainties for ARS-TEG experimental facility

6.3 Exergetic Efficiency

The theoretical maximum heat available for recovery from an exhaust stream is the amount

of energy the stream must reject to achieve thermal equilibrium with the ambient. The ratio between

this theoretical maximum and the actual waste heat recovered by a WHR system is known as the

exergetic efficiency of a WHR system. For HAU-2, the exergetic efficiency is defined as:

𝜼𝒆𝒙 = (∆𝑬��𝒐𝒊𝒍 + 𝑷𝑻𝑬𝑮) 𝑬��𝒂𝒊𝒓,𝒊𝒏⁄ 6.1

Where 𝜼𝒆𝒙 is the Exergetic Efficiency of the HAU, ∆𝑬��𝒐𝒊𝒍 is the rate of change of

exergy of the oil as it passes through the HAU, in W. PTEG is the total electric power generated by

the TEG array in W and 𝑬��𝒂𝒊𝒓,𝒊𝒏 is the exergy rate of the air at inlet.

Exergetic efficiency, defined this way for a WHR application, indicates the ratio of useful

heat and electrical power recovered from an input hot stream of specific exergy content. In Figure

6.2, the exergetic efficiencies are graphically presented for different input temperatures and mass

flowrates of the hot and cold streams. It can be seen form the graphs that the exergetic efficiencies

are higher for high thermal oil inlet temperatures. This is expected as the smaller the difference

between delivery and source temperature for a WHR system, the higher is its exergetic efficiency.

Page 105: Integrated Absorption Refrigeration and Thermoelectric

94

0.005 0.01 0.005 0.01

0.01 0.02 ṁoil

ṁair

0.0

0.1

0.2

0.3

0.4

0.5

ηex [

-]

Toil,in = 50°C

[kg·s-1]

[kg·s-1]

(a)

0.005 0.01 0.005 0.01

0.01 0.02

0.0

0.1

0.2

0.3

0.4

0.5

ηex [

-]

250[°C]

313[°C]

400[°C]

Toil,in = 75°C

(b)

Exergetic Efficiency

0.005 0.01 0.005 0.01

0.01 0.02 ṁoil

ṁair

0.0

0.1

0.2

0.3

0.4

0.5Toil,in = 100°C

ηex [

-]

[kg·s-1]

[kg·s-1]

(c)

Tair,in

Figure 6.2 Exergetic Efficiency of HAU-2 at different inlet conditions

Another aspect to consider is the mass flowrates of the thermal oil stream and simulated

exhaust (air) stream. From the graphs, it can be observed that, within the margin of error, the

exergetic efficiency is strongly sensitive to the air mass flowrate, with greater exergetic efficiencies

recorded at higher air flowrate when other parameters remain unchanged. Since, air has a much

lower thermal capacity compared to the thermal oil, a higher flowrate of air results in a better

source-use match in terms of heat capacity rates, resulting in higher exergetic efficiencies.

Thus, a thermal oil inlet temperature closer to the hot exhaust stream inlet temperature, and

a higher air mass flow rate result in higher exergetic efficiencies.

Page 106: Integrated Absorption Refrigeration and Thermoelectric

95

6.4 Thermoelectric Power

Owing to the relatively small amount of waste heat recovered in the form of TEG electrical

power generation, exergetic efficiency, as defined previously, does not inform greatly as regards to

TEG performance. For this, the total TEG power generated for each experimental run is analyzed

in this section.

0.005 0.01 0.005 0.01

0.01 0.02 ṁoil

ṁair

0

5

10

15

20

25

PT

EG

,to

t [W

]

[kg·s-1]

[kg·s-1]

Toil,in = 50°C

(a)

0.005 0.01 0.005 0.01

0.01 0.02

0

5

10

15

20

25

PT

EG

,to

t [W

]

250[°C]

313[°C]

400[°C]

Toil,in = 75°C

(b)

Total TEG Power Tair,in

0.005 0.01 0.005 0.01

0.01 0.02 ṁoil

ṁair

0

5

10

15

20

25

PT

EG

,to

t [W

]

[kg·s-1]

[kg·s-1]

Toil,in = 100°C

(c)

Figure 6.3 TEG Power produced by HAU-2 at different inlet conditions (uncertainty of ±0.2%)

.

Page 107: Integrated Absorption Refrigeration and Thermoelectric

96

Graphs in Figure 6.3, for different inlet temperature and mass flowrate conditions, show

the TEG power produced for all 30 experimental runs. It can be observed that the thermoelectric

power output is greatly sensitive to the temperature difference between the hot and cold stream at

the inlet of the HAU-2.

For given inlet temperatures and air mass flowrate, the oil mass flowrate has no significant

impact on the total TEG power generated. But for the cases where the air mass flowrate is reduced,

there is a significant drop in TEG power output. This sensitivity to air mass flowrate is caused by

the much higher thermal capacity of the thermal oil stream relative to the thermal capacity of the

air stream.

Thus, TEG power is greater when the temperature difference between the fluid streams and

the flowrate of air are relatively high. Additionally, for the same hot and cold side temperatures,

the increase in the thermal fluid side flowrate only minimally affects the TEG power production,

instead, TEG power is more sensitive to the air flowrate.

6.5 Comparison of HAU Model Predictions and Experimental Results

The model developed for HAU-2 (Section 4.3) takes as inputs: the time averaged flowrates

of thermal oil stream and simulated exhaust (air) stream, the inlet temperatures of those streams,

and the average TEG power produced for each experimental run. The model calculates the outlet

temperatures, air side pressure drop, heat loss rate, heat transfer rates in each stage, and overall heat

transfers for each stream.

Figure 6.4 and Figure 6.5 present the comparison between the model’s predicted and the

experimentally measured Qoil and Qair, respectively, for all the experimental cases.

Page 108: Integrated Absorption Refrigeration and Thermoelectric

97

The average relative error, and the average absolute relative error in the predicted values

of Qoil is ~0.3, respectively. The average relative error, and the average absolute relative error in

the predicted values of Qair is ~0.2.

The model output of Qoil depends on the Tair,out , as shown by equations 6.2 and 6.3.

However, there is an overestimate error associated with the Tair,out measurement (discussed in the

Appendix).

𝑸𝒐𝒊𝒍 = 𝑸𝑎𝑖𝑟 − 𝑸𝒍𝒐𝒔𝒕 − 𝑷𝑻𝑬𝑮 6.2

𝑸𝒂𝒊𝒓 = ��𝒂𝒊𝒓𝒄𝒑,𝒂𝒊𝒓(𝑻𝒂𝒊𝒓,𝒊𝒏 − 𝑻𝒂𝒊𝒓,𝒐𝒖𝒕) 6.3

Meanwhile, the experimental measurement of Qoil is dependent on the Toil measurements

which have low uncertainties associated to them. In general, the model prediction for both Qoil and

Qair are consistent with the trend of experimental observation of greater heat transfer rates for larger

temperature differences between the hot and cold stream.

Moreover, due to the lower heat capacity rate of air compared to the thermal oil, the heat

transfer rate increase with an increase in air flowrate when other parameters are kept constant.

Page 109: Integrated Absorption Refrigeration and Thermoelectric

98

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02ṁoil

ṁair

Tair,in

0

200

400

600

800

1000

1200

1400

1600

1800Q

oil

,mo

d [

W],

Qo

il,e

xp [

W]

[kg·s-1]

[kg·s-1]

[°C]

Toil,in = 50°C

(a)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

200

400

600

800

1000

1200

1400

1600

1800

Qoil,mod Qoil,exp

Toil,in = 75°C

(b)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

200

400

600

800

1000

1200

1400

1600

1800Toil,in = 100°C

(c)Qoil Comparison

Figure 6.4 Qoil Model vs Experiment Comparison

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02ṁoil

ṁair

Tair,in

0

200

400

600

800

1000

1200

1400

1600

1800

Qai

r,m

od [

W],

Qai

r,ex

p [

W]

[°C]

[kg·s-1]

[kg·s-1]

Toil,in = 50°C

(a)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

200

400

600

800

1000

1200

1400

1600

1800

Qair,mod Qair,exp

Toil,in = 75°C

(b)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

200

400

600

800

1000

1200

1400

1600

1800Toil,in = 100°C

(c)Qair Comparison

Figure 6.5 Qair Model vs Experiment Comparison

Page 110: Integrated Absorption Refrigeration and Thermoelectric

99

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02ṁoil

ṁair

Tair,in

0

20

40

60

80

100

120

140T

oil

,ou

t,m

od [

°C],

To

il,o

ut,

exp [

°C]

[kg·s-1]

[kg·s-1]

[°C]

Toil,in = 50°C(a)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

20

40

60

80

100

120

140

Toil,out,mod Toil,out,exp

Toil,in = 75°C

(b) Toil,out Comparison

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

20

40

60

80

100

120

140 Toil,in = 100°C(c)

Figure 6.6 Toil Model vs Experiment Comparison

0.02

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01ṁoil

ṁair

Tair,in

0

20

40

60

80

100

120

140

160

180

200

Tai

r,o

ut,

mod [

°C],

Tai

r,out,

exp [

°C] Toil,in = 50°C

[°C]

[kg·s-1]

[kg·s-1]

(a)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

20

40

60

80

100

120

140

160

180

200

Tair,out,mod Tair,out,exp

Toil,in = 75°C

(b)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

20

40

60

80

100

120

140

160

180

200Toil,in = 100°C

(c)Tair,out Comparison

Figure 6.7 Tair Model vs Experiment Comparison

Page 111: Integrated Absorption Refrigeration and Thermoelectric

100

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02 ṁoil

ṁair

Tair,in

0

20

40

60

ΔP

mo

d [

Pa]

,ΔP

exp [

Pa]

[°C]

[kg·s-1]

[kg·s-1]

Toil,in = 50°C

(a)

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02

0

20

40

60

ΔP

mo

d [

Pa]

,ΔP

exp [

Pa]

ΔPmod

ΔPexp

Toil,in = 75°C

(b)

ΔP Comparison

0.005 0.01 0.005 0.01

250 313 400 250 313 250 313 400 250 313

0.01 0.02 ṁoil

ṁair

Tair,in

0

20

40

60

ΔP

mo

d [

Pa]

,ΔP

exp [

Pa]

[°C]

[kg·s-1]

[kg·s-1]

Toil,in = 100°C(c)

Figure 6.8 ΔPexp vs ΔPmodel comparison

In Figures 6.6 and 6.7, a comparison between the engineering model’s predicted Toil,out and

Tair,out with their corresponding experimental measurement is presented.

The absolute relative and relative errors in the predicted values of Toil,out are ~0.06 and -

0.06, respectively. Meanwhile, the absolute relative and relative errors in the predicted values of

Tair,out are both ~0.25 in value.

Page 112: Integrated Absorption Refrigeration and Thermoelectric

101

It can be seen that, for all cases, the predicted Toil,out has a much greater degree of qualitative

similarity (within the margin of error) with the measured Toil,out, while the error in the model

predicted Tair,out is greater than for Toil,out.

Another comparison to consider is the pressure drop in the air stream as it goes through the

HAU. From the graphs presented in Figure 6.8, it can be seen that the pressure drop is sensitive to

the air mass flowrate. The model’s prediction of pressure drop in the HAU is consistent with the

experimentally measured value (within the margin of error). The absolute relative and relative

errors in the predicted values of ΔP are ~0.14 and ~-0.13, respectively

6.6 Discussion

Two different designs for an integrated Thermoelectric Generator and Absorption

Refrigeration based WHR heat acquisition units (HAU) were developed. Their performance was

evaluated at different temperatures and flowrates. Through the analysis of the experimental results

from the first design of the HAU (HAU-1), important design factors were identified:

• Spreading Resistance is an increased thermal resistance when there exists an

area mismatch between the TEGs and the air-side heat exchange surface. This

leads a lower TEG hot junction temperature and less efficient TEG operation.

• An undersized oil-block results in higher TEG cold junction temperatures and

a reduction in TEG efficiency. An undersized oil-block also results in lower

thermal oil outlet temperatures.

Page 113: Integrated Absorption Refrigeration and Thermoelectric

102

• Optimal TEG array sizing is important in designing an HAU as an increase in

the number of TEGs results in a larger heat exchange area between the hot and

cold stream, thereby reducing the overall ΔT across the TEG junctions.

Based on these parameters, an improved HAU (HAU-2) was developed and experimentally

tested. The following changes were made to the TEG design:

• To increase thermal energy transfer to the thermal oil, HAU-2 was divided into

two-stages, with the TEGs present in the first stage. The first stage allows the

TEGs to operate at higher ΔT while the second stage recovers more heat

downstream of the TEG which increases the thermal oil outlet temperature

compared to the one-stage design of HAU-1.

• To minimized the effect of spreading resistance, a copper heat spreader was

utilized in the first stage.

• Based on Rattner’s [43] work, the optimum number of TEGs (2) were used in

HAU-2 as opposed to 6 TEGs in HAU-1.

These changes resulted in a HAU system with 4 times more TEG power output than HAU-

1. Moreover, the exergetic efficiencies for the HAU-2 (~40%) were found to be comparable to

ORC based WHR systems [100], [101]. The experimental results also indicate a great degree of

sensitivity of key system variables to the exhaust/air-side flowrate.

An engineering model for HAU-2 was developed and system performance parameters were

compared to the experimental observations. A general congruence of trend between the model’s

predictions and experimental measurements was observed after air-side flowrate corrective

calibration (see Appendix), though some error still exists due air-side temperature outlet

measurements.

Page 114: Integrated Absorption Refrigeration and Thermoelectric

103

Therefore, current study establishes that a cascaded approach to WHR based on TEG and

ARS is a viable pathway for WHR especially for application where cooling and electric power are

the desired modes of WHR.

Page 115: Integrated Absorption Refrigeration and Thermoelectric

104

Conclusions and Recommendations for Future

Research

Page 116: Integrated Absorption Refrigeration and Thermoelectric

105

This chapter provides a discussion of the major findings of this Ph.D. dissertation with

regards to ARS-TEG based integrated cascaded waste heat recovery systems. Thereafter, some

recommended pathways for future research on this topic are also suggested.

7.1 ARS-TEG thermoeconomic studies

Chapter 3 discusses the thermoeconomic studies conducted for the vehicle and carburizing

furnace applications of the proposed ARS-TEG based WHR system. Cycle models were developed

based on AHRI guidelines for closure parameters of heat exchangers and reasonable component

efficiencies were assumed. Another complementary study compared the performance of ORC-

VCC and ARS-TEG based WHR systems for vehicle application and compared the capital

investment cost of the two systems. The key findings for the thermoeconomic studies are:

1. The payback periods for ARS-TEG based cascaded WHR systems are competitive

with conventional ORC based WHR systems.

2. The temperature drop across the TEGs is a significant design parameter, as it affects

the efficiency of both pathways (ARS and TEGs). If the ΔT across the TEG junctions

is too large, the desorber heat delivery temperature can drop below the required

temperature.

3. In light of the real-world data [51] for a refrigerated transport vehicle’s exhaust

temperaure fluctuations, strategies must be employed to ensure a steady operation of

the WHR system. It was proposed in this disseration that electric heaters may be used

to deliver heat to the desorber when the exhaust tempeatures are too low. At higher

than design temperatures, the exhaust can be re-routed to bypass the desorber.

Page 117: Integrated Absorption Refrigeration and Thermoelectric

106

4. By comparing the captial cost of the ARS-TEG system to an ORC-VCC system, it was

surmized that the cost of the ARS-TEG sysytem ($10,464.05), without the TEGs was

significanly lower than the cost of the ORC-VCC system ($15,390.7). But the TEGs

cost another $10,269 for the ARS-TEG system, making it slightly more expensive.

Since, the electrical power requirement of a refrigerated transport vehicle is not the

primary requirement, a reduction in the number of TEGs can result in a more

competitive ARS-TEG based WHR system.

7.2 HAU design

In Chapter 4 of this Ph.D. dissertation a detailed study of the iterative design and

development of HAU for the ARS-TEG based WHR system was presented. Two HAUs were

constructed during the course of this study. Findings from the performance of the first HAU

informed the design of HAU-2. Some of the key findings form this section of study are:

1. Spreading resistance due to the mismatch in contact area of the TEGs and fins on the

air-side results in a large spreading resistance (Rsp = 0.023 K W-1). This results in a

lower TEG hot side temperature and lower TEG efficiency. Using a copper heat

spreader can minimize the spreading resistance (Rsp = 0.013 K W-1).

2. The number of TEGs in a WHR system must be optimized as increasing the number

of TEGs beyond an optimal number can lead to poorer WHR performance as the

average ΔT across the TEGs drops. For HAU-1, using 6 TEGs the maximum power

produced for a specific set of inlet conditions was ~4 W, whereas, for the same inlet

conditions, HAU-2 generated a maximum of ~21 W.

Page 118: Integrated Absorption Refrigeration and Thermoelectric

107

3. Heat exchange surface sizing on both the hot and cold sides of an HAU requires

optimization. Backpressure, especially on the hot exhaust side of the HAU, is a

significant design parameter as an increase in backpressure required additional work.

If the heat exchange surfaces on the thermal oil side are undersized, the TEG cold side

temperature and hot side temperature ‘pinch’ and the efficiency of the TEGs drops.

The thermal fluid’s outlet temperature is also lower than required if the oil side heat

exchange surfaces are undersized.

7.3 HAU Model prediction and experimental results

In Chapter 4 and Chapter 5 a detailed description of the HAU and cycle models is

presented. In Chapter 6 of this dissertation a detailed comparison of the model predication and

experimentally observed values is presented. Some of the key findings are:

1. Due to the smaller heat capacity (cp) of the exhaust side flow, parameters like Qoil,

PTEG and Toil are more sensitive to the inlet conditions on the air side.

2. The exergetic efficiency of the HAU is comparable to that of a typical ORC based

WHR systems (~40%).

7.4 Recommendations for future research

More studies are needed to improve upon the ARS-TEG based WHR system proposed in

this Ph.D. dissertation. My recommendations for future works include:

Page 119: Integrated Absorption Refrigeration and Thermoelectric

108

7.4.1 Advanced HAU design

The HAU developed during this study can be improved upon by eliminated the thermal

fluid coupling fluid between the HAU and desorber. This can be achieved by designing and

developing an integrated desorber where the heat from the exhaust cascades through the TEGs and

is delivered to the ARS working fluid solution directly. This has the potential to significantly

increase the performance of the ARS-TEG waste heat recovery system by:

• Delivering higher-grade waste to the ARS sub-system

• Eliminating parasitic heat losses due to the thermal oil flow circuit

• Reduction in cost by eliminated additional heat exchange surfaces and thermal oil

as working fluid

7.4.2 Experimental Investigations

In addition to the HAU specific experimental work reported in this Ph.D. dissertation

(Chapter 5 and Chapter 6), additional full system (ARS+TEG) experimental data is needed to gain

further insights into the operation of the WHR system at different conditions.

Page 120: Integrated Absorption Refrigeration and Thermoelectric

109

Appendex A: Velocity Measurement

Calibration

Preliminary analysis of the experimental data indicated the presence of an over-estimate

error in measurements related to the air-side flow. It was suspected that the temperature and

flowrate measurement locations at the exit of the HAU were possibly within the Vena Contracta

effect as shown in Figure A.1.

Vena Contracta Effect

Pitot TubeThermocouple

Figure A.1 Stage-2 Airflow, immediately at the exit of the HAU, is contracted due to the shape

of the HAU. This contraction leads to an increase in the velocity at the location of temperature

and velocity measurement

To minimize the effect of this error, a post-experiment, flowrate calibration was performed

as a correction for the measure data by taking flow measurements downstream using a hotwire

anemometer (Figure A.2). However, some error due to Tair,out measurements still persists.

Page 121: Integrated Absorption Refrigeration and Thermoelectric

110

Figure B.2 Stage-2 Vcorr (Velocity measurement downstream of original measurement) vs Vmsrd

(original measurement)

0

0.5

1

1.5

2

2.5

3

3.5

0 0.5 1 1.5 2 2.5 3 3.5 4

Vco

rr[m

/s]

Vmsrd [m/s]

Velocity Correction – Vmsrd vs. Vcorr

Page 122: Integrated Absorption Refrigeration and Thermoelectric

111

Bibliography

[1] P. A. Owusu and S. Asumadu-Sarkodie, “A review of renewable energy sources, sustainability issues and climate change mitigation,” Cogent Eng., vol. 3, no. 1, Apr. 2016,

doi: 10.1080/23311916.2016.1167990.

[2] EPA, “Sources of Greenhouse Gas Emissions,” Clim. Chang., pp. 1–2, 2015.

[3] E. Worrell, L. Bernstein, J. Roy, L. Price, and J. Harnisch, “Industrial energy efficiency

and climate change mitigation,” Energy Effic., vol. 2, no. 2, pp. 109–123, 2009, doi:

10.1007/s12053-008-9032-8.

[4] A. S. Rattner and S. Garimella, “Energy harvesting, reuse and upgrade to reduce primary energy usage in the USA,” Energy, vol. 36, no. 10, pp. 6172–6183, 2011, doi:

10.1016/j.energy.2011.07.047.

[5] H. Jouhara, N. Khordehgah, S. Almahmoud, B. Delpech, A. Chauhan, and S. A. Tassou,

“Waste heat recovery technologies and applications,” Thermal Science and Engineering

Progress, vol. 6. Elsevier Ltd, pp. 268–289, 01-Jun-2018, doi: 10.1016/j.tsep.2018.04.017.

[6] J. Ilona, C. W. T, and D. Amber, “Waste Heat Recovery. Technology and Opportunities in U.S. Industry; Sponsor Org.: USDOE Office of Energy Efficiency and Renewable Energy

(EERE), Advanced Manufacturing Office (EE-5A) (Advanced Manufacturing Office

Corporate),” Mar. 2008.

[7] R. Shen, X. Gou, H. Xu, and K. Qiu, “Dynamic performance analysis of a cascaded thermoelectric generator,” Appl. Energy, vol. 203, pp. 808–815, Oct. 2017, doi:

10.1016/j.apenergy.2017.06.108.

[8] “Module 10: Absorption refrigeration – CIBSE Journal.” [Online]. Available:

https://www.cibsejournal.com/cpd/modules/2009-11/. [Accessed: 29-Sep-2020].

[9] S. O. Enibe, “Solar refrigeration for rural applications,” Renew. Energy, vol. 12, no. 2, pp.

157–167, 1997, doi: 10.1016/S0960-1481(97)00036-0.

[10] M. Liu, W. Saman, and F. Bruno, “Development of a novel refrigeration system for refrigerated trucks incorporating phase change material,” Appl. Energy, vol. 92, pp. 336–

342, 2012, doi: 10.1016/j.apenergy.2011.10.015.

[11] S. Brückner, S. Liu, L. Miró, M. Radspieler, L. F. Cabeza, and E. Lävemann, “Industrial

waste heat recovery technologies: An economic analysis of heat transformation technologies,” Appl. Energy, vol. 151, pp. 157–167, 2015, doi:

10.1016/j.apenergy.2015.01.147.

[12] D. Lu et al., “Modeling and analysis of an ammonia–water absorption refrigeration system

utilizing waste heat with large temperature span,” Int. J. Refrig., vol. 103, pp. 180–190,

Jul. 2019, doi: 10.1016/j.ijrefrig.2019.04.008.

Page 123: Integrated Absorption Refrigeration and Thermoelectric

112

[13] K. E. Herold, R. Radermacher, and S. A. Klein, Absorption Chillers and Heat Pumps.

2016.

[14] H. Kaibe, T. Kajihara, and S. Fujimoto, “Recovery of Plant Waste Heat by a

Thermoelectric Generating System,” Komatsu Tech. Rep., vol. 57, no. 164, pp. 26–30,

2011.

[15] J. Koehler, W. J. Tegethoff, D. Westphalen, and M. Sonnekalb, “Absorption refrigeration system for mobile applications utilizing exhaust gases,” Int. Commun. Heat Mass Transf.,

vol. 32, no. 5, pp. 333–340, 1997.

[16] “Scopus preview - Scopus - Welcome to Scopus.” [Online]. Available:

https://www.scopus.com/. [Accessed: 21-Sep-2020].

[17] N. Zhou, X. Wang, Z. Chen, and Z. Wang, “Experimental study on Organic Rankine

Cycle for waste heat recovery from low-temperature flue gas,” Energy, vol. 55, pp. 216–

225, Jun. 2013, doi: 10.1016/j.energy.2013.03.047.

[18] J. Shon, H. Kim, and K. Lee, “Improved heat storage rate for an automobile coolant waste

heat recovery system using phase-change material in a fin-tube heat exchanger,” Appl.

Energy, vol. 113, pp. 680–689, 2014, doi: 10.1016/j.apenergy.2013.07.049.

[19] A. Mahmoudi, M. Fazli, and M. R. Morad, “A recent review of waste heat recovery by Organic Rankine Cycle,” Appl. Therm. Eng., vol. 143, no. January, pp. 660–675, 2018,

doi: 10.1016/j.applthermaleng.2018.07.136.

[20] G. Bisio, “Energy recovery from molten slag and exploitation of the recovered energy,”

Energy, vol. 22, no. 5, pp. 501–509, 1997, doi: 10.1016/S0360-5442(96)00149-1.

[21] S. Lan, Z. Yang, R. Chen, and R. Stobart, “A dynamic model for thermoelectric generator

applied to vehicle waste heat recovery,” Appl. Energy, vol. 210, pp. 327–338, Jan. 2018,

doi: 10.1016/j.apenergy.2017.11.004.

[22] O. Badr, S. D. Probert, and P. W. O’Callaghan, “Selecting a working fluid for a Rankine-

cycle engine,” Appl. Energy, vol. 21, no. 1, pp. 1–42, 1985, doi: 10.1016/0306-

2619(85)90072-8.

[23] H. Tian, G. Shu, H. Wei, X. Liang, and L. Liu, “Fluids and parameters optimization for

the organic Rankine cycles (ORCs) used in exhaust heat recovery of Internal Combustion Engine (ICE),” Energy, vol. 47, no. 1, pp. 125–136, 2012, doi:

10.1016/j.energy.2012.09.021.

[24] J. Deng, R. Z. Wang, and G. Y. Han, “A review of thermally activated cooling

technologies for combined cooling, heating and power systems,” Prog. Energy Combust.

Sci., vol. 37, no. 2, pp. 172–203, 2011, doi: 10.1016/j.pecs.2010.05.003.

[25] X. Zhang, L. Wu, X. Wang, and G. Ju, “Comparative study of waste heat steam SRC,

ORC and S-ORC power generation systems in medium-low temperature,” Appl. Therm.

Eng., vol. 106, pp. 1427–1439, Aug. 2016, doi: 10.1016/j.applthermaleng.2016.06.108.

[26] T. C. Hung, T. Y. Shai, and S. K. Wang, “A review of organic rankine cycles (ORCs) for

the recovery of low-grade waste heat,” Energy, vol. 22, no. 7, pp. 661–667, Jul. 1997, doi:

10.1016/S0360-5442(96)00165-X.

Page 124: Integrated Absorption Refrigeration and Thermoelectric

113

[27] B. F. Tchanche, G. Lambrinos, A. Frangoudakis, and G. Papadakis, “Low-grade heat

conversion into power using organic Rankine cycles - A review of various applications,”

Renewable and Sustainable Energy Reviews, vol. 15, no. 8. Elsevier Ltd, pp. 3963–3979,

2011, doi: 10.1016/j.rser.2011.07.024.

[28] N. Galanis, E. Cayer, P. Roy, E. S. Denis, and M. Désilets, “Electricity generation from

low temperature sources,” J. Appl. Fluid Mech., vol. 2, no. 2, pp. 55–67, 2009.

[29] T. C. Hung, T. Y. Shai, and S. K. Wang, “A review of organic rankine cycles (ORCs) for

the recovery of low-grade waste heat,” Energy, vol. 22, no. 7, pp. 661–667, 1997, doi:

10.1016/S0360-5442(96)00165-X.

[30] G. Yu, G. Shu, H. Tian, Y. Huo, and W. Zhu, “Experimental investigations on a cascaded

steam-/organic-Rankine-cycle (RC/ORC) system for waste heat recovery (WHR) from diesel engine,” Energy Convers. Manag., vol. 129, pp. 43–51, Dec. 2016, doi:

10.1016/j.enconman.2016.10.010.

[31] H. Wang et al., “Performance of a combined organic Rankine cycle and vapor

compression cycle for heat activated cooling,” Energy, vol. 36, no. 1, pp. 447–458, 2011,

doi: 10.1016/j.energy.2010.10.020.

[32] W. Han, L. Sun, D. Zheng, H. Jin, S. Ma, and X. Jing, “New hybrid absorption-

compression refrigeration system based on cascade use of mid-temperature waste heat,”

Appl. Energy, vol. 106, pp. 383–390, 2013, doi: 10.1016/j.apenergy.2013.01.067.

[33] R. Funahashi, “Waste heat recovery using thermoelectric oxide materials,” Sci. Adv.

Mater., vol. 3, no. 4, pp. 682–686, Aug. 2011, doi: 10.1166/sam.2011.1200.

[34] X. Shi and D. Che, “A combined power cycle utilizing low-temperature waste heat and LNG cold energy,” Energy Convers. Manag., vol. 50, no. 3, pp. 567–575, Mar. 2009, doi:

10.1016/j.enconman.2008.10.015.

[35] M. He, X. Zhang, K. Zeng, and K. Gao, “A combined thermodynamic cycle used for

waste heat recovery of internal combustion engine,” Energy, vol. 36, no. 12, pp. 6821–

6829, 2011, doi: 10.1016/j.energy.2011.10.014.

[36] A. Khaliq, R. Kumar, and I. Dincer, “Exergy analysis of an industrial waste heat recovery

based cogeneration cycle for combined production of power and refrigeration,” J. Energy

Resour. Technol. Trans. ASME, vol. 131, no. 2, pp. 0224021–0224029, Jun. 2009, doi:

10.1115/1.3120381.

[37] I. Mikulic, R. Zhan, and S. Eakle, “Dependence of fuel consumption on engine backpressure generated by a DPF,” in SAE Technical Papers, 2010, doi: 10.4271/2010-01-

0535.

[38] J. Li, S. Xu, and S. Kong, “Investigation of the absorption-compression hybrid

refrigeration applied to the coach air conditioning,” Heat Transf. Res., vol. 46, no. 3, pp.

271–291, 2015, doi: 10.1615/HeatTransRes.2014006867.

[39] A. Sonthalia, S. Reddy, C. R. Kumar, and K. Kamani, “Theoretical Investigation of Waste Heat Recovery from an IC Engine Using Vapor Absorption Refrigeration System and

Thermoelectric Converter,” Heat Transf. - Asian Res., vol. 44, no. 6, pp. 499–514, 2015,

doi: 10.1002/htj.21133.

Page 125: Integrated Absorption Refrigeration and Thermoelectric

114

[40] J. H. Meng, X. D. Wang, and W. H. Chen, “Performance investigation and design

optimization of a thermoelectric generator applied in automobile exhaust waste heat

recovery,” Energy Convers. Manag., vol. 120, pp. 71–80, 2016, doi:

10.1016/j.enconman.2016.04.080.

[41] X. Liu, Y. D. Deng, Z. Li, and C. Q. Su, “Performance analysis of a waste heat recovery thermoelectric generation system for automotive application,” Energy Convers. Manag.,

vol. 90, pp. 121–127, 2015, doi: 10.1016/j.enconman.2014.11.015.

[42] W. He, S. Wang, C. Lu, X. Zhang, and Y. Li, “Influence of different cooling methods on

thermoelectric performance of an engine exhaust gas waste heat recovery system,” Appl.

Energy, vol. 162, pp. 1251–1258, 2016, doi: 10.1016/j.apenergy.2015.03.036.

[43] A. S. Rattner and T. J. Meehan, “Simple analytic model for optimally sizing thermoelectric generator module arrays for waste heat recovery,” Appl. Therm. Eng., vol.

146, pp. 795–804, 2019, doi: 10.1016/j.applthermaleng.2018.10.003.

[44] S. Vasta, V. Palomba, D. La Rosa, and W. Mittelbach, “Adsorption-compression cascade

cycles: An experimental study,” Energy Convers. Manag., vol. 156, pp. 365–375, Jan.

2018, doi: 10.1016/j.enconman.2017.11.061.

[45] K. C. A. Alam, M. Z. I. Khan, A. S. Uyun, Y. Hamamoto, A. Akisawa, and T. Kashiwagi,

“Experimental study of a low temperature heat driven re-heat two-stage adsorption

chiller,” Appl. Therm. Eng., vol. 27, no. 10, pp. 1686–1692, Jul. 2007, doi:

10.1016/j.applthermaleng.2006.07.006.

[46] F. H. Dubberke et al., “Experimental setup of a cascaded two-stage organic Rankine cycle,” Appl. Therm. Eng., vol. 131, pp. 958–964, Feb. 2018, doi:

10.1016/j.applthermaleng.2017.11.137.

[47] X. Wang, A. Christ, K. Regenauer-Lieb, K. Hooman, and H. T. Chua, “Low grade heat

driven multi-effect distillation technology,” Int. J. Heat Mass Transf., vol. 54, no. 25–26,

pp. 5497–5503, Dec. 2011, doi: 10.1016/j.ijheatmasstransfer.2011.07.041.

[48] G. F. Hewitt and S. J. Pugh, “Approximate design and costing methods for heat

exchangers,” Heat Transf. Eng., vol. 28, no. 2, pp. 76–86, 2007, doi:

10.1080/01457630601023229.

[49] A. C. Caputo, P. M. Pelagagge, and P. Salini, “Heat exchanger design based on economic

optimisation,” Appl. Therm. Eng., vol. 28, no. 10, pp. 1151–1159, Jul. 2008, doi:

10.1016/j.applthermaleng.2007.08.010.

[50] “CMO-32-62S TEG POWER Module - Thermoelectric Generator.” [Online]. Available:

https://thermoelectric-generator.com/product/cmo-32-62s/. [Accessed: 05-Oct-2020].

[51] K. Boriboonsomsin et al., “Real-world exhaust temperature profiles of on-road heavy-duty

diesel vehicles equipped with selective catalytic reduction,” Sci. Total Environ., vol. 634,

pp. 909–921, Sep. 2018, doi: 10.1016/j.scitotenv.2018.03.362.

[52] “U.S. Gasoline and Diesel Retail Prices.” [Online]. Available: https://www.eia.gov/dnav/pet/PET_PRI_GND_DCUS_NUS_A.htm. [Accessed: 07-Oct-

2020].

[53] B. Elliston and M. Dennis, “Feasibility of Solar-Assisted Refrigerated Transport in

Page 126: Integrated Absorption Refrigeration and Thermoelectric

115

Australia,” 2009.

[54] D. Bergeron, “SANDIA REPORT Solar Powered Refrigeration for Transport Application-

A Feasibility Study,” 2001.

[55] J. L. S. A. Tassou, G. De-Lille, “Food Transport Refrigeration,” no. 853, pp. 1–25, 2014.

[56] G. F. Nemet, “Beyond the learning curve: factors influencing cost reductions in

photovoltaics,” Energy Policy, vol. 34, no. 17, pp. 3218–3232, Nov. 2006, doi:

10.1016/j.enpol.2005.06.020.

[57] C. F. Yu, W. G. J. H. M. Van Sark, and E. A. Alsema, “Unraveling the photovoltaic technology learning curve by incorporation of input price changes and scale effects,”

Renew. Sustain. Energy Rev., vol. 15, no. 1, pp. 324–337, Jan. 2011, doi:

10.1016/j.rser.2010.09.001.

[58] “Electric Power Monthly - U.S. Energy Information Administration (EIA).” [Online].

Available: https://www.eia.gov/electricity/monthly/epm_table_grapher.php?t=epmt_5_6_a.

[Accessed: 05-Oct-2020].

[59] B. Patel, N. B. Desai, S. S. Kachhwaha, V. Jain, and N. Hadia, “Thermo-economic

analysis of a novel organic Rankine cycle integrated cascaded vapor compression–absorption system,” J. Clean. Prod., vol. 154, pp. 26–40, 2017, doi:

10.1016/j.jclepro.2017.03.220.

[60] G. Shu, G. Yu, H. Tian, H. Wei, X. Liang, and Z. Huang, “Multi-approach evaluations of

a cascade-Organic Rankine Cycle (C-ORC) system driven by diesel engine waste heat:

Part A – Thermodynamic evaluations,” Energy Convers. Manag., vol. 108, pp. 579–595,

Jan. 2016, doi: 10.1016/j.enconman.2015.10.084.

[61] G. Shu, G. Yu, H. Tian, H. Wei, X. Liang, and Z. Huang, “Multi-approach evaluations of

a cascade-Organic Rankine Cycle (C-ORC) system driven by diesel engine waste heat:

Part A – Thermodynamic evaluations,” Energy Convers. Manag., vol. 108, pp. 579–595,

Jan. 2016, doi: 10.1016/j.enconman.2015.10.084.

[62] A. Fontalvo, J. Solano, C. Pedraza, A. Bula, A. G. Quiroga, and R. V. Padilla, “Energy,

exergy and economic evaluation comparison of small-scale single and dual pressure

organic Rankine cycles integrated with low-grade heat sources,” Entropy, vol. 19, no. 10,

p. 476, Sep. 2017, doi: 10.3390/e19100476.

[63] F. Mol es, J. Navarro-Esbrí, B. Peris, A. an Mota-Babiloni, and K. Kontomaris, “Thermodynamic analysis of a combined organic Rankine cycle and vapor compression

cycle system activated with low temperature heat sources using low GWP fluids,” 2015,

doi: 10.1016/j.applthermaleng.2015.04.083.

[64] K. H. Kim and H. Perez-Blanco, “Performance analysis of a combined organic Rankine

cycle and vapor compression cycle for power and refrigeration cogeneration,” Appl. Therm. Eng., vol. 91, no. 140, pp. 964–974, 2015, doi:

10.1016/j.applthermaleng.2015.04.062.

[65] X. Bu, L. Wang, and H. Li, “Performance analysis and working fluid selection for

geothermal energy-powered organic Rankine-vapor compression air conditioning,” 2013.

Page 127: Integrated Absorption Refrigeration and Thermoelectric

116

[66] M. Tauseef Nasir, M. A. Ali, T. S. Khan, E. Al-Hajri, M. B. Kadri, and K. C. Kim,

“Performance assessment and multi objective optimization of an Organic Rankine Cycle

driven cooling air conditioning system,” Energy Build., vol. 191, pp. 13–30, 2019, doi:

10.1016/j.enbuild.2019.03.012.

[67] N. Toujeni, N. Bouaziz, and L. Kairaouani, “Energetic investigation of a new combined ORC-VCC system for cogeneration,” in Energy Procedia, 2017, vol. 139, pp. 670–675,

doi: 10.1016/j.egypro.2017.11.270.

[68] N. Javanshir, S. M. Seyed Mahmoudi, and M. A. Rosen, “Thermodynamic and

Exergoeconomic Analyses of a Novel Combined Cycle Comprised of Vapor-Compression Refrigeration and Organic Rankine Cycles,” Sustainability, vol. 11, no. 12, p. 3374, 2019,

doi: 10.3390/su11123374.

[69] B. Saleh, “Energy and exergy analysis of an integrated organic Rankine cycle-vapor

compression refrigeration system,” 2018, doi: 10.1016/j.applthermaleng.2018.06.018.

[70] B. Saleh, “Parametric and working fluid analysis of a combined organic Rankine-vapor

compression refrigeration system activated by low-grade thermal energy The effect of boiler temperature on the COP S for different candidates in the basic ORC-VCR system.

Production and hosting by Elsevier A R T I C L E I N F O,” J. Adv. Res., vol. 7, pp. 651–

660, 2016, doi: 10.1016/j.jare.2016.06.006.

[71] W. Zhi-qi, Z. Qi-yu, X. Xiao-xia, L. Bin, and Z. Xin, “Performance comparison and

analysis of a combined power and cooling system based on organic Rankine cycle,” J.

Cent. South Univ, vol. 24, pp. 353–359, 2017, doi: 10.1007/s11771-017-3437-5.

[72] AHRI, “AHRI Standard 560-2000 Standard for absorption water chilling and water

heating packages,” 2000.

[73] A. Standard, “Performance Rating of Commercial and Industrial Unitary Air-Conditioning

Condensing Units,” 2011.

[74] J. Bao, L. Zhang, C. Song, N. Zhang, X. Zhang, and G. He, “Comparative study of combined organic Rankine cycle and vapor compression cycle for refrigeration: Single

fluid or dual fluid?,” Sustain. Energy Technol. Assessments, vol. 37, Feb. 2020, doi:

10.1016/j.seta.2019.100595.

[75] B.-S. Park, M. Usman, M. Imran, and A. Pesyridis, “Review of Organic Rankine Cycle

experimental data trends,” 2018, doi: 10.1016/j.enconman.2018.07.097.

[76] S. Karellas and K. Braimakis, “Energy-exergy analysis and economic investigation of a cogeneration and trigeneration ORC-VCC hybrid system utilizing biomass fuel and solar

power,” 2015, doi: 10.1016/j.enconman.2015.06.080.

[77] M. T. Nasir and K. C. Kim, “Working fluids selection and parametric optimization of an

Organic Rankine Cycle coupled Vapor Compression Cycle (ORC-VCC) for air

conditioning using low grade heat,” Energy Build., vol. 129, pp. 378–395, 2016, doi:

10.1016/j.enbuild.2016.07.068.

[78] H. Wang, R. Peterson, and T. Herron, “Design study of configurations on system COP for

a combined ORC (organic Rankine cycle) and VCC (vapor compression cycle),” Energy,

vol. 36, no. 8, pp. 4809–4820, 2011, doi: 10.1016/j.energy.2011.05.015.

Page 128: Integrated Absorption Refrigeration and Thermoelectric

117

[79] J. Hærvig, K. Sørensen, and T. J. Condra, “Guidelines for optimal selection of working

fluid for an organic Rankine cycle in relation to waste heat recovery,” Energy, vol. 96, pp.

592–602, Feb. 2016, doi: 10.1016/j.energy.2015.12.098.

[80] S. Gopalnarayanan, “Choosing the right refrigerant,” Mech. Eng., vol. 120, no. 10, pp. 92–

95, Oct. 1998, doi: 10.1115/1.1998-oct-6.

[81] F-Chart-Software, “EES: Engineering Equation Solver | F-Chart Software : Engineering Software,” 2012. [Online]. Available: http://fchartsoftware.com/ees/. [Accessed: 10-Dec-

2019].

[82] S. F. Mussati, S. Cignitti, S. S. Mansouri, K. V. Gernaey, T. Morosuk, and M. C. Mussati,

“Configuration optimization of series flow double-effect water-lithium bromide

absorption refrigeration systems by cost minimization,” Energy Convers. Manag., vol.

158, pp. 359–372, Feb. 2018, doi: 10.1016/j.enconman.2017.12.079.

[83] S. Sanaye, M. Emadi, and A. Refahi, “Thermal and economic modeling and optimization

of a novel combined ejector refrigeration cycle,” Int. J. Refrig., vol. 98, pp. 480–493,

2019, doi: 10.1016/j.ijrefrig.2018.11.007.

[84] S. Quoilin, S. Declaye, B. F. Tchanche, and V. Lemort, “Thermo-economic optimization of waste heat recovery Organic Rankine Cycles,” Appl. Therm. Eng., vol. 31, no. 14–15,

pp. 2885–2893, 2011, doi: 10.1016/j.applthermaleng.2011.05.014.

[85] C. Zhang, C. Liu, S. Wang, X. Xu, and Q. Li, “Thermo-economic comparison of

subcritical organic Rankine cycle based on different heat exchanger configurations,”

Energy, vol. 123, pp. 728–741, Mar. 2017, doi: 10.1016/j.energy.2017.01.132.

[86] M. P. Bailey, “Chemical engineering plant cost index (cepci),” Chem. Eng., vol. 121, no.

2, pp. 68–69, Feb. 2014.

[87] “2019 Chemical Engineering Plant Cost Index Annual Average - Chemical Engineering |

Page 1.” [Online]. Available: https://www.chemengonline.com/2019-chemical-

engineering-plant-cost-index-annual-average/. [Accessed: 28-Sep-2020].

[88] “22W TEG Module - By TEGpro.” [Online]. Available: https://www.tegmart.com/thermoelectric-modules/22w-teg-module. [Accessed: 28-Sep-

2020].

[89] A. Bejan, S. Lorente, and D. H. Kang, “Constructal design of thermoelectric power

packages,” Int. J. Heat Mass Transf., vol. 79, pp. 291–299, 2014, doi:

10.1016/j.ijheatmasstransfer.2014.08.008.

[90] L. E. Bell, “Cooling, heating, generating power, and recovering waste heat with thermoelectric systems,” Science (80-. )., vol. 321, no. 5895, pp. 1457–1461, 2008, doi:

10.1126/science.1158899.

[91] J. Henderson, “Analysis Of A Heat Exchanger-Thermoelectric Generator System To Be

Presented At The 14th Intersociety Energy Conversion Engineering Conference.”

[92] F. P. Incropera, D. P. DeWitt, T. L. Bergman, and A. S. Lavine, “Fundamentals of Heat

and Mass Tranfser.” Wiley, p. 997, 2006.

[93] S. Lee, S. Song, V. Au, and K. P. Moran, “Constriction/spreading resistance model for

Page 129: Integrated Absorption Refrigeration and Thermoelectric

118

electronics packaging,” 1995.

[94] “Tgpr-22W-7V-56S.” .

[95] “Online Catalog - F - FH80.” [Online]. Available:

https://www.alphanovatech.com/en/c_fh80e.html. [Accessed: 04-Oct-2020].

[96] Z. Yan and J. Chen, “The load matching of a thermoelectric device at maximum power

output,” 2008, pp. 343–344, doi: 10.1063/1.46828.

[97] D. W. Sun, “Comparison of the performances of NH3-H2O, NH3-LiNO3 and NH3-

NaSCN absorption refrigeration systems,” Energy Convers. Manag., vol. 39, no. 5–6, pp.

357–368, Mar. 1998, doi: 10.1016/s0196-8904(97)00027-7.

[98] “What Is Motion Control? - National Instruments.” [Online]. Available:

http://www.ni.com/en-my/innovations/white-papers/10/what-is-motion-control-.html.

[Accessed: 10-Dec-2019].

[99] “RIGOL 4 Current Slew Rate [5].”

[100] Y. Dai, J. Wang, and L. Gao, “Parametric optimization and comparative study of organic

Rankine cycle (ORC) for low grade waste heat recovery,” Energy Convers. Manag., vol.

50, no. 3, pp. 576–582, Mar. 2009, doi: 10.1016/j.enconman.2008.10.018.

[101] D. Wang, X. Ling, H. Peng, L. Liu, and L. L. Tao, “Efficiency and optimal performance

evaluation of organic Rankine cycle for low grade waste heat power generation,” Energy,

vol. 50, no. 1, pp. 343–352, Feb. 2013, doi: 10.1016/j.energy.2012.11.010.

Page 130: Integrated Absorption Refrigeration and Thermoelectric

Vita

Shahzaib B. Abbasi

Education

Ph.D., Mechanical Engineering, The Pennsylvania State University December 2020

M.S., Pakistan Institute of Engineering and Applied Sciences November 2013 B.E., N.E.D University of Engineering and Technology April 2010

Work Experience

Graduate Researcher – University of Twente, Enschede, The Netherlands Feb. 2020 – Present Doctoral Researcher – The Pennsylvania State University, State College, PA Jan. 2017 – Dec. 2020

Graduate Researcher – Mississippi State University, Starkville, MS Aug. 2015 – Dec. 2016

Junior Engineer – National Center for Non-destructive Testing, Islamabad, Pakistan Nov. 2013 – Jun. 2015 Fellow – P.I.E.A.S, Islamabad, Pakistan Nov. 2011 – Nov. 2013

Product Executive (Renewable Energy) – Makkays, Islamabad, Pakistan Aug. 2010 – Feb. 2011

Selected Publications

1. Abbasi, Shahzaib B., and Alexander S. Rattner. "Cascaded Thermoelectric Generation and Absorption

Refrigeration Waste Heat Recovery." ASTFE Digital Library. Begel House Inc., 2018.