You are on page 1of 391

Exergy Analysis of Heating,

Refrigerating, and Air Conditioning


Exergy Analysis of Heating,
Refrigerating, and Air
Conditioning
Methods and Applications

Ibrahim Dincer and Marc A. Rosen

AMSTERDAM • BOSTON • HEIDELBERG • LONDON • NEW YORK • OXFORD


PARIS • SAN DIEGO • SAN FRANCISCO • SINGAPORE • SYDNEY • TOKYO
Elsevier
Radarweg 29, PO Box 211, 1000 AE Amsterdam, Netherlands
The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK
225 Wyman Street, Waltham, MA 02451, USA

© 2015 Elsevier Inc. All rights reserved.

No part of this publication may be reproduced or transmitted in any form or by any means, electronic or
mechanical, including photocopying, recording, or any information storage and retrieval system, without
permission in writing from the publisher. Details on how to seek permission, further information about
the Publisher’s permissions policies and our arrangements with organizations such as the Copyright
Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/
permissions.

This book and the individual contributions contained in it are protected under copyright by the
Publisher (other than as may be noted herein).

Notices
Knowledge and best practice in this field are constantly changing. As new research and experience
broaden our understanding, changes in research methods, professional practices, or medical treatment
may become necessary.

Practitioners and researchers must always rely on their own experience and knowledge in evaluating
and using any information, methods, compounds, or experiments described herein. In using such
information or methods they should be mindful of their own safety and the safety of others, including
parties for whom they have a professional responsibility.

To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume
any liability for any injury and/or damage to persons or property as a matter of products liability,
negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas
contained in the material herein.

ISBN: 978-0-12-417203-6

Library of Congress Cataloging-in-Publication Data


A catalog record for this book is available from the Library of Congress

British Library Cataloguing in Publication Data


A catalogue record for this book is available from the British Library

For information on all Elsevier publications


visit our website at http://store.elsevier.com/
Acknowledgments

The contributions of several graduate students are gratefully acknowledged,


including Canan Acar, Sayantan Ghosh, Monu Malik, Farrukh Khalid, and
Fahad Suleman. In particular, we thank Canan Acar for reviewing and revising
several chapters and checking for consistency.
In addition, some materials coming from Mohammed Al-Ali, Reza Soltani, and
Abdullah Al-Zahrani are acknowledged.
Last but not least, we warmly thank our wives, Gulsen Dincer and Margot
Rosen, and our children, Meliha, Miray, Ibrahim Eren, Zeynep, and Ibrahim
Emir Dincer and Allison and Cassandra Rosen. They have been a great source
of support and motivation.
Ibrahim Dincer and Marc A. Rosen
July 2015

ix
Preface

This book focuses on applications of exergy methods to the heating, refrigerat-


ing, and air conditioning industries and the primary technologies comprising
them, with the aim of providing an enhanced understanding of the behaviors
of such systems and better tools for their improvement. Heating, refrigeration,
and air conditioning processes are very important energy technologies in most
countries and are often responsible for a significant portion of their energy
utilization. Thus, it is beneficial to consider all available tools in efforts to
improve designs, in terms of efficiency, environmental performance, econom-
ics, and other factors.
Exergy analysis represents a relatively recent and exciting innovation in thermo-
dynamics and energy systems. As a method that uses the conservation of mass
and conservation of energy principles together with the second law of thermo-
dynamics for the analysis, exergy analysis helps in the design, optimization,
and improvement of energy systems like HVAC. The exergy method is a useful
tool for furthering the goal of more efficient energy resource use, for it enables
the locations, types, and magnitudes of wastes and losses to be identified and
meaningful efficiencies to be determined. Exergy methods have received nota-
ble attention only over the last few decades. Although that attention has grown
during that period, it has remained somewhat limited and comprehensive
applications of exergy analysis in the heating, refrigeration, and air condition-
ing industries remain needed.
The book seeks to describe comprehensively the application of exergy methods
to heating, refrigerating, and air conditioning systems, so as to aid in their
improvement. In doing so, the book contains eight chapters. The relations of
the material to building energy systems and their management are stressed
throughout.
Chapter 1 describes energy and exergy methods and how they are used to assess,
design, and improve technologies and systems. Fundamental thermodynamic
principles are explained and analysis methodologies based on exergy are cov-
ered in depth. In addition, extensions of these exergy-based methods to xi
xii Preface

environmental, economic, and sustainability assessments are covered. To pro-


vide a broader context, the role of heating, refrigerating, and air conditioning in
regional systems like countries is discussed, including relevant energy, and
exergy perspectives.
In Chapter 2, energy and exergy analyses of basic heating, refrigeration, and air
conditioning technologies and systems are presented. Included are applications
of exergy analysis to the components comprising heating, refrigerating, and air
conditioning, as well as psychrometric processes and overall systems.
Chapter 3, Chapter 4, Chapter 5, and Chapter 6 focus on a range of industrial
systems and applications of heating, refrigeration, and air conditioning.
Chapter 3 focuses on the use of exergy methods in industrial heating and cool-
ing and the diverse range of processes that can be used for those activities. Heat
pump systems are introduced but are examined in much greater depth in
Chapter 4. Cogeneration, trigeneration, and multigeneration, as well as inte-
grated energy systems and district heating and cooling, are covered in
Chapter 5. Chapter 6 describes how exergy methods are applied to heat and
cold storage systems, covering the range of such technologies that find applica-
tions in buildings.
In Chapter 7, building HVAC systems based on renewable energy are described
and several case studies are considered for illustration. The utilization of renew-
able and sustainable energy in place of conventional energy resources like fossil
fuels is an important extension of the material in the preceding chapters since
we feel that renewable and sustainable energy systems represent the future of
energy systems in general and for heating, refrigeration, and air conditioning
in particular.
In Chapter 8, several illustrations are described of exergy-based methods for
improving heating, refrigeration, and air conditioning systems. The methods
covered include design for responsible energy and environment management,
life cycle assessment, energy retrofits, and energy substitution, the strategic inte-
gration of energy systems, and the allocation of environmental emissions.
These are linked to the material in Chapter 7 by linking them to the utilization
of renewable and sustainable energy.
Incorporated throughout are many illustrative examples and case studies,
which provide the reader with a substantial learning experience, especially in
terms of practical applications.
The appendixes contain unit conversion factors and tables and charts of ther-
mophysical properties of various materials in the International System of Units
(SI). A glossary of exergy-related terminology is also provided.
Preface xiii

References are included to direct the reader to sources where more details can be
found and to assist the reader who is simply curious to learn more. The refer-
ences can also help identify information on topics not covered comprehen-
sively in the book.
As a research-oriented textbook, this volume includes theoretical and practical
features often not included in solely academic textbooks. This book is mainly
intended for use by advanced undergraduate or graduate students in several
engineering and nonengineering disciplines and also as an essential tool for
practitioners in HVAC disciplines. Theory and analysis are emphasized
throughout this comprehensive book, reflecting new techniques, models,
and applications, together with complementary materials and recent informa-
tion. Coverage of the material is extensive, and the amount of information and
data presented is sufficient for advanced courses related to heating, cooling, and
air conditioning—and advanced technologies being applied in these areas—or
as a supplement for courses on applied thermodynamics. We believe that this
book will be of interest to students and practitioners and individuals and insti-
tutions who are interested in exergy and its applications to heating, cooling,
and air conditioning as well as the various new technologies and methods that
are increasingly finding use in these areas. This volume is also a valuable and
readable reference for anyone who wishes to learn about exergy methods and/
or advanced heating, cooling, and air conditioning.
We hope this book allows exergy methods to be more widely applied to heat-
ing, refrigerating, and air conditioning industries and both the traditional and
new technologies being applied in them. The book thereby provides an
enhanced understanding of the behaviors of heating, refrigerating, and air con-
ditioning systems and enhanced tools for improving them. By exploiting the
benefits of applying exergy methods to these systems, we believe they can be
made more efficient, clean, and sustainable and help humanity address many
of the challenges it faces.
Ibrahim Dincer and Marc A. Rosen
July 2015
CHAPTER 1

Exergy and its Ties to the Environment,


Economics, and Sustainability

1.1 INTRODUCTION
Heating, refrigeration, and air conditioning processes are treated as important
energy technologies in most countries and are often responsible for a signi-
ficant portion of their energy utilization. Applying exergy methods to techno-
logies for the heating, refrigerating, and air conditioning can provide a
better understanding of their behaviors and enhanced tools for improving
them. This use of exergy analysis not only is advantageous but also is prudent,
since it is useful to consider all available tools in efforts to improve designs, in
terms of efficiency, environmental performance, economics, and other factors.
Exergy analysis represents a recently rediscovered and exciting innovation in
thermodynamics and energy systems. Exergy methods, basic to enhanced
and combined models, have received notable attention only over the last
few decades. Although such attention has grown during that period, it has
remained limited and applications of exergy analysis in the heating, refrigera-
tion, and air conditioning industries, although they have increased notably,
would benefit from an enhanced focus and a consolidation of the information.
The applications of exergy methods to heating, refrigerating, and air
conditioning systems are described in a detailed and comprehensive manner
in this book, with the intent of enhancing understanding and aiding in process
assessments and improvements.
The book starts by describing energy and exergy methods and how they are used
to assess, design, and improve technologies and systems. Fundamental thermo-
dynamic principles are explained and analysis methodologies based on exergy
are covered in depth. In addition, extensions of these exergy-based methods to
environmental assessments and economic evaluations are covered. To provide
a broader context, the role of heating, refrigerating, and air conditioning in
regional systems like countries is discussed, including relevant energy and
exergy perspectives.
1

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00001-6


© 2015 Elsevier Inc. All rights reserved.
2 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Next, energy and exergy analyses of basic heating, refrigeration, and air
conditioning technologies and systems are presented. Included are applications
of exergy analysis to the components comprising heating, refrigerating, and
air conditioning, as well as overall systems. A diverse range of processes are
considered, including industrial heating and cooling, drying, building energy
management, cogeneration and trigeneration, district heating and cooling,
and thermal storage. Measures for improving heating, refrigeration, and air
conditioning systems are also described. Renewable and sustainable energy
systems are covered throughout not only because of their expanding usage
but also because they likely represent the future of energy systems.
The book closes by describing and assessing exergy-based methods for improv-
ing heating, refrigeration, and air conditioning systems. The methods covered
include design for responsible energy and environment management, life cycle
assessment, energy retrofits, and energy substitution. The latter is extended to
the utilization of renewable and sustainable energy in place of conventional
energy resources like fossil fuels.

1.2 WHY EXERGY?


Energy use is pervasive in life, and there is a strong relation between energy and
prosperity. Throughout much of history, the emergence of civilizations has
been characterized by the discovery and effective application of energy to help
meet society’s needs. The desire of people to sustain and improve their well-
being is possibly the biggest driver of the growth in worldwide energy demand.
Therefore, meeting the demand for energy services in a clean, efficient, secure,
and reliable way is an important challenge.
Energy analysis is the traditional method of assessing the way energy is used
in operations (e.g., physical or chemical processing of materials, heat transfer,
and energy conversion). Energy analysis is based on the first law of thermo-
dynamics (FLT) and usually entails performing energy balances and evaluat-
ing energy efficiencies. The FLT embodies the principle of conservation of
energy, which states that, although energy can change form, it can be neither
created nor destroyed. However, this law provides no information about the
direction in which processes can spontaneously occur, that is, it does not
explain reversibility aspects of thermodynamic processes. An energy balance
also cannot explain the degradation of energy or resources during a process
and does not quantify the usefulness or quality of energy and material quan-
tities (e.g., input, product, and waste flows for a system). The FLT provides no
information about the inability of any thermodynamic process to convert
heat fully into mechanical work or any insight into why mixtures cannot
1.2 Why Exergy? 3

spontaneously separate or unmix themselves. Another principle to explain


these phenomena and to characterize the availability of energy is required
to do this.
The exergy method of analysis overcomes many of the limitations of the FLT.
The concept of exergy is based on both the FLT and the second law of thermo-
dynamics (SLT). Exergy analysis clearly indicates the locations, nature, and
causes of energy degradation in a process and therefore can help improve a pro-
cess or technology. Exergy analysis can also quantify the quality of energy dur-
ing heat transfer. The primary aim of exergy analysis is usually to provide
meaningful efficiencies (i.e., exergy efficiencies) and the causes and true mag-
nitudes of exergy losses.
It is important to distinguish between exergy and energy in order to avoid con-
fusion between exergy analysis and traditional energy-based methods of anal-
ysis and design. Table 1.1, which compares energy and exergy in a general
manner, can help in making such a distinction. Energy flows into and out
of a system with mass flows, heat transfers, and work interactions (e.g., work
associated with shafts and piston rods). Energy is conserved, in line with the
FLT. Exergy, although similar in some respects, is different. It loosely repre-
sents a quantitative measure of the usefulness or quality of an energy or mate-
rial substance. More rigorously, exergy is a measure of the ability to do work
(or the work potential) from a quantity or flow (mass, heat, and work), in a
specified environment. A key attribute of exergy is that it permits comparisons
on a common basis of quantities (inputs and outputs) of different types.

Table 1.1 General Comparison of Energy and Exergy


Energy Exergy

• Dependent on properties of quantity (e.g., matter and • Dependent on both properties of quantity (e.g., matter
energy) and independent of properties of a reference and energy) and properties of a reference environment
environment
• Nonzero in value when in equilibrium with the • Zero in value when at the dead state, that is, in complete
reference environment equilibrium with the reference environment
• Conserved for all processes, that is, can be neither • Conserved for reversible processes and nonconserved
destroyed nor produced for real processes, that is, can be neither destroyed nor
produced in a reversible process but is always destroyed
(consumed) in an irreversible process
• Appears in many forms (e.g., kinetic energy, potential • Appears in many forms (e.g., kinetic exergy, potential
energy, work, heat) and is measured in that form exergy, work, thermal exergy) and is measured on the
basis of work equivalent or ability to produce work
• A measure of quantity, but not quality • A measure of quantity and quality
• Based on the FLT • Based on a combination of the FLT and the SLT
4 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Another benefit of exergy analysis is that by accounting for all exergy streams
and quantities for a system, it is possible to determine the extent to which
exergy is destroyed or consumed by the system. The exergy destruction is pro-
portional to the entropy generation due to irreversibilities. Exergy is always
destroyed in real processes, partially or totally, in line with the SLT. Exergy
destruction is usually largely responsible for the less-than-ideal efficiencies
of systems or processes.
Increasing application and recognition of the usefulness of exergy methods by
those in industry, government, and academia across the world has been
observed in recent years. The present authors, for instance, have examined
exergy analysis methodologies and applied them to industrial systems (e.g.,
Rosen et al., 2005; Rosen and Dincer, 2003, 2004; Rosen and Etele, 2004;
Rosen and Scott, 1998; Rosen and Horazak, 1995), thermal energy storage
(Dincer and Rosen, 2002; Rosen et al., 2004), and environmental impact
assessments (Crane et al., 1992; Rosen and Dincer, 1997, 1999; Gunnewiek
and Rosen, 1998; Rosen, 1990).
This chapter covers the energy and exergy, focusing on the relevant portions of
the field of thermodynamics. The necessary background for understanding
energy and exergy concepts and the basic principles, general definitions, funda-
mentals, and practical applications and implications are provided. An illustra-
tive example is provided to highlight the important aspects of energy and
exergy.

1.3 IMPORTANCE OF ENERGY TO INDUSTRY, CULTURE,


AND LIVING STANDARDS
Energy, culture, and standard of life are linked in complex ways that are often
difficult to describe. Throughout history, energy choices have strongly influ-
enced cultural and economic development as well as standards of living. Envi-
ronmental impact is often a significant consequence of energy processes and
also affects culture and standard of life. Environmental issues also affect the sus-
tainability of a country’s development in the longer term.
The environmental impact of human activities has increased significantly dur-
ing recent decades, primarily because of increases in world population, living
standards, resource consumption, and industrial activity. The temporal rela-
tionship between energy consumption and CO2 emissions is shown in
Fig. 1.1, where it can be seen that consumption and emissions exhibit similar
trends, indicating a strong relation.
Environmental concerns associated with energy use range from pollutant
emissions and accidents to the degradation of environmental quality and
1.3 Importance of Energy to Industry, Culture, and Living Standards 5

14,000 40

Primary energy consumption (Mtoe)


12,000 35

CO2 emissions (billion tonnes)


30
10,000
25
8000
20
6000
15
Consumption
4000
Emissions 10
2000 5

0 0
1965
1967
1969
1971
1973
1975
1977
1979
1981
1983
1985
1987
1989
1991
1993
1995
1997
1999
2001
2003
2005
2007
2009
2011
2013
FIGURE 1.1 Relationship between global primary energy consumption and CO2 emissions between 1965 and 2013. Data obtained from
IEA (2014a).

Table 1.2 Selected Pollutants and Some of Their Sources and Risks
Pollutant Source Risks

Carbon monoxide (CO) Incomplete combustion of fuels Urban air pollution


Sulfur dioxide (SO2) Natural processes (e.g., volcanic activity) Hazardous to human health and
environment
Sulfur-containing fuels, oil refining, electricity Respiratory difficulties, acid
generation, pulp and paper industry precipitation
Nitrogen oxides (NOx) Combustion of fuels at high temperatures Respiratory problems, low-level ozone
formation, creation of acids
Volatile organic Petroleum and solvent vapors Impede the formation of ozone
compounds (VOCs)
Particulates (e.g., fly ash) Natural and anthropogenic sources Acid precipitation, toxic effects

ecosystems. Various types of pollutants are listed in Table 1.2, with descriptions
of their sources and effects on environment and human health.
Environmental considerations have received increasing attention from industry,
government, and the public in recent years, especially since such considerations
form an integral component of standard of life. In addition, environmental
issues often affect cultural development. Over the past few decades, energy-
related environmental concerns (such as climate change and ozone depletion)
have affected both local and regional populations as well as national and mul-
tinational government.
6 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Table 1.3 Summary of Major Environmental Concerns Related to Energy use and Their Causes
and Impacts
Environmental Concern Causes Impacts

Global climate change Greenhouse gases (CO2, CH4, CFCs, Increases in Earth surface temperature and
halons, N2O, ozone, etc.) emissions sea level; coastal flooding; fertile zone
Coal mining, deforestation, general displacement; freshwater scarcity
energy-related activities
Stratospheric ozone CFCs, halons, N2O emissions UV radiation increase (skin cancer, eye
depletion damage)
Acid precipitation SO2, NOx, VOC emissions Acidification of lakes, streams, and
Electricity generation, residential groundwaters; damage to forests and
heating, industrial energy use, sour gas agricultural crops; deterioration of materials
treatment, transportation (buildings, metal structures, fabrics)

Table 1.3 summarizes the major areas of environmental concerns related to


energy use. Additional environmental concerns also exist, and many of these
are listed below with some of their main causes/sources:

• Water pollution: Hazardous chemicals from power plants and refineries,


acid drainage from mines, releases of geothermal fluids containing toxic
chemicals, and thermal pollution associated with discharges from
power plant cooling systems
• Maritime pollution: Shipping operations and accidental oil spills
• Solid wastes and their disposal: Chemical, metal, and other industries
• Ambient air quality: Emissions of SO2, NOx, CO, VOCs, and
particulate matter
• Hazardous air pollutants: Lead-based gasoline additives, emissions during
oil and gas extraction, processing and combustion, and mercury,
chlorinated dioxins, and furans from municipal waste incinerators
• Indoor air quality: CO, CO2, and smoke from stoves and fireplaces,
gaseous oxides of nitrogen and sulfur from furnaces, stray natural gas and
heating oil vapors, radon from natural gas-burning appliances and
surrounding soil, cigarette smoke, and formaldehyde from plywood
and glues
• Land use and siting impact: Fuel refining, electricity generation, disposal
sites for solid wastes including radioactive residues, hydroelectric
reservoirs, mining sites, surface needs for biomass production, and
large-scale exploitation of renewable energy
• Radiation and radioactivity: Energy activities (fossil fuel combustion,
uranium mining and milling, etc.), nuclear waste disposal, nuclear facility
decommissioning, etc.
• Major environmental accidents: Explosions and fires at oil/gas refineries,
rigs, tanks, and pipelines, failures at hydroelectric dams causing flooding
and landslides, accidents at nuclear facilities, and explosions in mines
1.3 Importance of Energy to Industry, Culture, and Living Standards 7

Culture is often loosely defined as the form and stage of intellectual develop-
ment or civilization. Energy choices can be dependent on a society’s culture,
and energy factors can contribute to cultural development. Standard of life is
often taken to be the degree of material comfort available to a community,
and this is influenced by the availability of energy resources (of sufficient quan-
tity, quality, and type) to a society and its ability to utilize those resources. Liv-
ing standards also affect energy issues. For example, societies with a high
standard of life usually have good education systems and extensive research
and development activities, both of which foster the development of energy
technologies capable of harnessing energy resources more efficiently and with
less environmental impact.
Energy, culture, standards of life, and industry are linked to each other, and all
affect the ability of a society to develop sustainably, as shown in Fig. 1.2. Over
last few decades, people have become increasingly aware of concerns associ-
ated with energy use, including supply limitations and environmental issues.
Researchers and policy makers have focused on these and related issues, often
by considering factors such as energy, environment, and sustainable develop-
ment and linkages between them. These topics relate to culture and standard
of life. For example, the environmental impact of energy use is usually
reduced by increasing the efficiency of energy resource utilization and/or by
substituting more environmentally benign energy resources for damaging
ones. Such actions can make development more sustainable and improve
standards of life through a cleaner environment. Sustainable development
demands a sustainable supply of energy sources that can be achieved via
the following:

Energy

Sustainable Standard
Industry
development of life

Culture

FIGURE 1.2 Importance to sustainable development of energy, industry, culture, and standard of life.
8 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

• Sustainably available energy resources at reasonable cost that can be


utilized for all required tasks without causing negative societal impacts.
Energy resources such as fossil fuels (coal, oil, and natural gas) and
uranium are generally acknowledged to be finite. Others such as sunlight,
wind, and falling water are generally considered renewable and therefore
sustainable over the relatively long term. Wastes (convertible to useful
energy via waste-to-energy incineration and other processes) and biomass
fuels are sometimes viewed as sustainable energy sources.
• Efficient use of energy resources to enhance the benefits while avoiding
the problems associated with their use. This implication acknowledges
that all energy resources are to some degree finite, so that greater efficiency
allows them to contribute to development over a longer period of
time, thereby making development more sustainable. Even for energy
sources that may eventually become inexpensive and widely available,
increases in efficiency can reduce the resource requirements (energy,
material, etc.) to create and maintain systems and devices to harvest the
energy and also reduce the associated environmental impacts.

1.4 HEATING, REFRIGERATION, AND


AIR CONDITIONING AND THEIR ENERGY USE
Heating, refrigeration, and air conditioning are the major contributors to
energy consumption in residential and commercial buildings in many regions.
Residential energy-utilizing processes include space heating and cooling, water
heating, lighting, cooking, and appliances. Figure 1.3 shows that the residential
sector accounts for about one-quarter of global final energy consumption in
2011 (a total of 8918 Mtoe). This share has not changed significantly over
the last couple of decades and is expected to remain similar in the future. Note
that the International Energy Agency (IEA) is the source of the data in Fig. 1.3
and defines the residential sector, also known as the households sector, as a
collective pool of all the households in a country.
Note that Fig. 1.3 provides a global perspective, but there is a wide difference in
the share of energy use by the residential sector by country, mainly because of
variations in climate, energy resource availability, energy infrastructure, income,
economic structure, and other specific conditions and preferences. For instance,
residential energy needs for developing tropical countries with limited industry
and services differ from those with no heating requirements and with an econ-
omy based on large service and/or industry sectors. The first set of countries
often relies mainly on biomass as the primary source of energy, mostly for
cooking; the share of residential energy demand often exceeds 75% of the
1.4 Heating, Refrigeration, and Air Conditioning and Their Energy Use 9

Nonenergy use
9%
Other*
4%

Industry
Services 29%
8%

Residential
23%

Transport
27%

FIGURE 1.3 Breakdown by sector of total global energy consumption in 2011. *Others include
agriculture/forestry, fishing, and nonspecified (IEA, 2014b).

total energy consumption. For the second set of countries, residential energy
consumption can account for less than 10% of the total final consumption. This
variation can be better seen in Table 1.4, which shows the contribution of the
building sector to the total final energy demand globally and in selected regions
in 2007. This table is based on statistics from the IEA. These data should be used
cautiously as some countries have difficulties separating consumption of the res-
idential sector from the services sector for several end uses and energy forms.

Table 1.4 Contribution of the Building Sector and Other Sectors to the Energy Demand Globally
and in Selected Regions in 2007
Unit Building
Residential Commercial Total Building Energy Demand
Region Sector Share (%) Sector Share (%) Sector Share (%) (MWh/capita/yr)

The United States and 17 13 31 18.6


Canada
European Union 23 11 34 9.64
Former Soviet Union 26 7 33 8.92
Middle East 21 6 27 5.75
China 25 4 29 3.20
Africa 54 3 57 3.19
Latin America 17 5 22 2.32
Asia (excluding China) 36 4 40 2.07
World 23 8 31 4.57
Source: IEA (2014b).
10 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

The percentages of energy use attributable to the residential and commercial


building sectors are shown in Table 1.4 for various world regions. The regions
differ not only geographically but also in terms of factors such as climate and
economic development. Even for region categories in Table 1.4, there can be
variations in energy use breakdowns. For instance, although Canada and the
United States have similar overall residential and commercial energy shares,
the breakdown of energy demand in these buildings differs notably. About
28% of an average residential building’s energy demand is attributable to space
heating in the United States, while this value is about 60% in Canada. In total,
heating, cooling, and refrigeration comprise about 62% of a residential build-
ing’s energy demand in the United States and about 84% in Canada (IEA online
statistics, 2007). Building heating requirements are higher in colder regions,
with buildings in colder climates consuming a greater share of energy for heat-
ing. However, in warmer climates, cooling demands often depend on a coun-
try’s economic development, with building cooling loads lower in developing
countries than in developed ones.
Note that the energy use values listed in Fig. 1.3 and Table 1.4 are often high-
quality energy forms. For example, exergy analysis shows that energy (or heat)
at room temperature is of very low quality, while energy at the temperature of
the atmosphere generally has no exergy. Even though there is a great amount of
energy contained in the atmosphere, for the most part, this energy is at the envi-
ronmental temperature and therefore useless. The part of energy that has no
capacity to perform work is sometimes called anergy and, in such cases, the
sum of anergy and exergy gives energy that is conserved, based on the FLT.
Therefore, an exergy analysis of heating, refrigeration, and air conditioning
gives a better insight on the efficiency of the overall system by identifying
energy quality losses and potential points of efficiency improvement. A combi-
nation of energy and exergy analyses, rather than energy analysis alone, can
increase productivity, likely enabling continued prosperity.

1.5 BENEFITS OF USING EXERGY ANALYSIS FOR


HEATING, REFRIGERATING, AND AIR CONDITIONING
The significant distinctions between energy and exergy (Table 1.1) demonstrate
that it is informative and useful to assess exergy instead of energy since produc-
tive use of exergy is important and can contribute to economic development.
Exergy methods enable better analyses of systems by identifying weaknesses
and strengths that cannot be explained by energy analysis alone. This is because
evaluating the exergy quantities and flows allows us to track the conversion of
high-quality products like shaft power, as well as providing heating and cool-
ing. Exergy analysis provides a better understanding of how to use resources
1.5 Benefits of Using Exergy Analysis for Heating, Refrigerating, and Air Conditioning 11

310
Winter Spring
Summer Fall
305
Exergy output rate (kW)

300

295

290

285

280

275

270

265
)

y)

a)

l)

y)

a)
na

na

ia

sia

pt

ea

an

SA

SA

in
zi
ke

di

an

ad
nd

(U

pa
gy

or
ra
hi

hi

ap
us
In

(U

(U
ur

an
(B
(C

(C

(I

(S
(E
(R

(J
i(

on

er
(T

(C
l(

ity

es
hi

rid
ba

(G
o
ro
ai

ng

nd
w

ou

el
el

ul
ul

ky

o
gh

ai
um

co

ad
iji

ng
D

Lo

nt
Pa
nb

n
Se

To
C

k
an

Be

os

rli

M
A

ro
or
M
ta

o
Sh

Be
M

To
Y

s
Is

Lo
ew
N

FIGURE 1.4 Exergy output rate of a hypothetical 500 kW heater (source temperature of 700 K)
for different locations and seasons.

efficiently. For instance, with energy analysis, it is not possible to identify the
difference between two identical heat sources (with the same heating power
output) in two climates. Yet, the same heating/cooling energy input has differ-
ent exergy values in different locations, times, seasons, etc. Exergy analysis takes
regional, seasonal, daily, and hourly variations into account while assessing the
quality of energy, providing an important advantage over energy methods.
The last point can be seen in Fig. 1.4, which shows the exergy output rate of a
hypothetical 500 kW heater, with a source temperature of 700 K, for various
locations and seasons. From Fig. 1.4, it can be seen that the same heater pro-
vides different exergy outputs in different locations/seasons. This difference
cannot be determined using energy analysis alone. Also, heating, refrigeration,
and air conditioning energy demands are at relatively low temperatures, so the
actual exergy requirements for space heating and cooling are low. Yet, in most
cases, this demand is met by high-grade energy sources (i.e., fossil fuels and
electricity). Energy analysis does not recognize the difference between high-
and low-grade energy utilization for heating, refrigeration, and air conditioning
applications, while exergy analysis does and also points out the potential for
better matching energy supply and demand qualities.
Figure 1.5 shows possible energy sources and uses based on their quality (i.e.,
exergy content). In ideal cases, high-quality sources are used to support high-
quality applications, while low-quality sources are used for low-quality
applications.
12 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Sources Quality Uses

Oil
Coal
Lighting

High
Uranium
Electrical
(fossil fuels)
appliances
Wind energy

High temp waste


Cooking

Me dium
heat, e.g., from
industrial
processes
o Washing
(200 C)
machine

Low te mp. was te


heat, e.g., from DHW

Low
CHP (50-100 oC)

Ground heat Space heating

FIGURE 1.5 Classification of energy sources and uses based on their quality level (i.e., exergy content).

1.6 ENERGY AND EXERGY FUNDAMENTALS


In this section, theoretical and practical aspects of thermodynamics relevant to
energy and exergy analyses are described. Fundamental principles and such
related issues as reference environment selection, efficiency definition, and
material properties’ acquisition are also discussed. General implications of
exergy analyses are examined, and a procedure for energy and exergy analyses
is provided. Although a relatively standard terminology and nomenclature has
evolved for conventional classical thermodynamics, there is at present no gen-
erally agreed upon terminology and nomenclature for exergy analysis, and a
diversity of symbols and names exist for basic and derived quantities (Kotas
et al., 1987; Lucca, 1990). The exergy analysis nomenclature used here follows
that proposed by Kotas et al. (1987) as a standard exergy nomenclature.

1.6.1 First Law of Thermodynamics


For a control mass, the energy interactions for a system may be divided into two
parts: the amount of heat dQ and the amount of work dW. The FLT for a control
mass can be written as follows:
dQ ¼ dE + dW (1.1)
1.6 Energy and Exergy Fundamentals 13

Integrating Eq. (1.1) from an initial state 1 to a final state 2 yields


Q12 ¼ E2  E1 + W12 (1.2)

where E1 and E2 denote the initial and final energies of the control mass, Q1–2 is
the heat transferred to the control mass during the process from state 1 to state
2, and W1–2 is the work done by the control volume during process 1–2. The
energy E generally includes internal energy U, kinetic energy KE, and potential
energy PE terms, as follows:
E ¼ U + KE + PE (1.3)

For a change of state from state 1 to state 2 with a constant gravitational accel-
eration g, Eq. (1.3) becomes
 
E2  E1 ¼ U2  U1 + m V22  V12 =2 + mgðZ2  Z1 Þ (1.4)

where m denotes the mass contained in the system, V the velocity, and Z the
elevation.
The quantities dQ and dW can be specified in terms of the rate laws for heat
transfer and work. For a control volume, an additional term appears for the fluid
flowing across the control surface (entering at state i and exiting at state e). The
FLT for a control volume can be written as
X X
Q_ cv ¼ E_cv + W_ cv + m_ e h^e  m_ i h^i (1.5)

where ṁ is mass flow rate and h^ is total specific energy, equal to the sum of spe-
cific enthalpy, kinetic energy, and potential energy, that is, h^ ¼ h + V 2 =2 + gZ.

1.6.2 Second Law of Thermodynamics


Although a spontaneous process can proceed only in a definite direction, the FLT
gives no information about direction. It merely states that when one form of
energy is converted to another, the quantities of energy involved are conserved
regardless of the feasibility of the process. Thus, processes can be envisioned that
do not violate the FLT but do violate the SLT, for example, transfer of a heat from
a low-temperature body to a high-temperature body, without the input of an ade-
quate external energy form like work. However, such a process is impossible,
emphasizing that the FLT is itself inadequate for explaining energy processes.
The SLT establishes the difference in the quality of different forms of energy and
explains why some processes can spontaneously occur while others cannot. The
SLT is usually expressed as an inequality, stating that the total entropy after a
process is equal to or greater than that before. The equality only holds for ideal
or reversible processes. The SLT has been confirmed experimentally.
14 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

The SLT defines the fundamental quantity entropy as a randomized energy state
unavailable for direct conversion to work. It also states that all spontaneous
processes, both physical and chemical, proceed to maximize entropy, that is,
to become more randomized and to convert energy to a less available form.
A direct consequence of fundamental importance is the implication that at ther-
modynamic equilibrium, the entropy of a system is at a relative maximum; that
is, no further increase in disorder is possible without changing the thermody-
namic state of the system by some external means (such as adding heat). A cor-
ollary of the SLT is the statement that the sum of the entropy changes of a
system and that of its surroundings must always be positive. In other words,
the universe (the sum of all systems and surroundings) is constrained to
become forever more disordered and to proceed towards thermodynamic equi-
librium with some absolute maximum value of entropy. From a biological
standpoint, this is intuitively reasonable since, unless gradients in concentra-
tion and temperature are forcibly maintained by the consumption of energy,
organisms proceed spontaneously towards the biological equivalent of
equilibrium death.
What makes this statement of the SLT valuable as a guide to formulating energy
policy is the relationship between entropy and the usefulness of energy. Energy
is most useful to us when it is of high quality (e.g., available to do work) or we
can get it to do high-quality tasks (e.g., flow from one substance to another to
provide industrial heating). Useful energy thus must have low entropy so that
the SLT will allow transfer or conversions to occur spontaneously.

1.6.3 Exergy
Exergy is a useful quantity that stems from the SLT in combination with the FLT
and helps in analyzing energy and other systems and processes. The exergy of a
system is defined as the maximum shaft work that can be done by the compos-
ite of the system and a specified reference environment. The reference environ-
ment is assumed to be infinite, in equilibrium, and to enclose all other systems.
Typically, the environment is specified by stating its temperature, pressure, and
chemical composition. Exergy is not simply a thermodynamic property, but
rather is a property of both a system and the reference environment. Exergy
is conserved only when all processes occurring in a system and the environment
are reversible, while exergy is destroyed whenever irreversible processes occur.

1.6.3.1 Exergy Analysis


When an exergy analysis is performed, the thermodynamic imperfections can
be quantified as exergy destructions, which represent losses in energy quality or
usefulness (e.g., wasted shaft work or wasted potential for the production of
shaft work). Like energy, exergy can be transferred or transported across the
1.6 Energy and Exergy Fundamentals 15

boundary of a system. For each type of energy transfer or transport, there is a


corresponding exergy transfer or transport.
Exergy analysis takes into account the different thermodynamic values of dif-
ferent energy forms and quantities. The exergy transfer associated with shaft
work is equal to the shaft work. The exergy transfer associated with heat trans-
fer, however, depends on the temperature at which it occurs in relation to the
temperature of the reference environment.
Some important characteristics of exergy are listed as follows:

• A system in complete equilibrium with its environment does not have any
exergy.
• The exergy of a system increases the more it deviates from the
environment.
• When energy loses its quality or is degraded, exergy is destroyed.
• Exergy by definition depends not only on the state of a system or flow but
also on the state of the reference environment.
• Exergy efficiencies are a measure of approach to ideality (or reversibility).
This is not necessarily true for energy efficiencies, which are often
misleading.
• Energy forms with high exergy contents are typically more valued and
useful than energy forms with low exergy.
• A concentrated mineral deposit “contrasts” with the reference
environment and thus has exergy, which increases with the concentration
of the mineral.

Several quantities related to the conceptual exergy balance are described here,
following the presentations by Moran (1989), Kotas (1995), and Dincer and
Rosen (2013).

1.6.3.2 Exergy of a Closed System


The exergy Exnonflow of a closed system of mass m, or the nonflow exergy, can be
expressed as
Exnonflow ¼ Exph + Ex0 + Exkin + Expot (1.6)

where
Expot ¼ PE (1.7)

Exkin ¼ KE (1.8)
X
Ex0 ¼ ðμi0  μi00 ÞNi (1.9)
i

Exnonflow, ph ¼ ðU  U0 Þ + P0 ðV  V0 Þ  T0 ðS  S0 Þ (1.10)
16 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Here, the system has a temperature T, pressure P, chemical potential μi for spe-
cies i, entropy S, energy E, volume V, and number of moles Ni of species i. The
system is within a conceptual environment in an equilibrium state with inten-
sive properties T0, P0, and μi00. The quantity μi0 denotes the value of μ at the
environmental state (i.e., at T0 and P0). The terms on the right side of
Eq. (1.6) represent, respectively, physical, chemical, kinetic, and potential com-
ponents of the nonflow exergy of the system. The exergy Ex is a property of the
system and conceptual environment, combining the extensive properties of the
system with the intensive properties of the environment.
Physical nonflow exergy is the maximum work obtainable from a system as it is
brought to the environmental state (i.e., to thermal and mechanical equilib-
rium with the environment), and chemical nonflow exergy is the maximum
work obtainable from a system as it is brought from the environmental state
to the dead state (i.e., to complete equilibrium with the environment).

1.6.3.3 Exergy of Flows


1.6.3.3.1 Exergy of a Matter Flow
The exergy of a flowing stream of matter Exflow is the sum of nonflow exergy and
the exergy associated with the flow work of the stream (with reference to P0),
that is,
Exflow ¼ Exnonflow + ðP  P0 ÞV (1.11)

Alternatively, Exflow can be expressed following Eq. (1.6) in terms of physical,


chemical, kinetic, and potential components:
Exflow ¼ Exph + Ex0 + Exkin + Expot (1.12)

where Expot, Exkin, and Ex0 are presented in Eqs. (1.7)–(1.9). Exph of a material
flow is calculated as follows:
Exflow, ph ¼ ðH  H0 Þ  T0 ðS  S0 Þ (1.13)

1.6.3.3.2 Exergy of Thermal Energy


Consider a control mass, initially at the dead state, being heated or cooled at
constant volume in an interaction with some other system. A heat transfer Q
is experienced by the control mass. The flow of exergy associated with the heat
transfer Q is denoted by ExQ and can be expressed as
ðf  
T0
ExQ ¼ 1 δQ (1.14)
i T

where δQ is an incremental heat transfer and the integral is from the initial state
(i) to the final state (f). This “thermal exergy” is the minimum work required by
1.6 Energy and Exergy Fundamentals 17

the combined system of the control mass and the environment in bringing the
control mass to the final state from the dead state. If the temperature T of the
control mass is constant, the thermal exergy transfer associated with a heat
transfer is
 
T0
ExQ ¼ 1  Q ¼ τQ (1.15)
T

For heat transfer across a region r on a control surface for which the temperature
may vary, we can write
ð   
T0
ExQ ¼ q 1 dA (1.16)
r T r

where qr is the heat flow per unit area at a region on the control surface at which
the temperature is Tr.

1.6.3.3.3 Exergy of Work


The total work W0 can be separated into two components: Wx and W. That is,
W 0 ¼ W + Wx (1.17)

where W is the work done by a system due to change in its volume and Wx is the
shaft work done by the system. The exergy associated with shaft work ExW is by
definition Wx. The exergy transfer associated with work done by a system due to
volume change is the net usable work due to the volume change and is denoted
by WNET. Thus, for a process in time interval t1 to t2,
ðWNET Þ1, 2 ¼ W1, 2  P0 ðV2  V1 Þ (1.18)

where W1,2 is the work done by the system due to volume change (V2  V1). The
term P0(V2  V1) is the displacement work necessary to change the volume
against the constant pressure P0 exerted by the environment.

1.6.3.3.4 Exergy of Electricity


As for shaft work, the exergy associated with electricity is equal to the energy.

1.6.3.4 Exergy Consumption


For a process occurring in a system, the difference between the total exergy flows
into and out of the system, less the exergy accumulation in the system, is the
exergy consumption I, expressible as
I ¼ T0 Sgen (1.19)

Equation (1.19) points out that exergy consumption is proportional to entropy


creation and is known as the Gouy-Stodola relation.
18 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

1.6.4 Balances
1.6.4.1 Conceptual Balances
A general balance for a quantity in a system may be written as
Input + Generation  Output  Consumption ¼ Accumulation (1.20)

Input and output refer, respectively, to quantities entering and exiting through
system boundaries. Generation and consumption refer, respectively, to quan-
tities produced and consumed within the system. Accumulation refers to the
buildup (either positive or negative) of the quantity within the system.
Versions of the general balance equation may be written for mass, energy,
entropy, and exergy. Mass and energy, being subject to conservation laws
(neglecting nuclear reactions), can be neither generated nor consumed. Conse-
quently, the general balance (Eq. 1.20) written for each of these quantities
becomes
Mass input  Mass output ¼ Mass accumulation (1.21)

Energy input  Energy output ¼ Energy accumulation (1.22)

Entropy input + Entropy generation  Entropy output ¼ Entropy accumulation


(1.23)

Exergy input  Exergy output  Exergy consumption ¼ Exergy accumulation (1.24)

Entropy is created during a process due to irreversibilities, but cannot be con-


sumed. Exergy is consumed due to irreversibilities, and exergy consumption is
proportional to entropy creation. Equations (1.23) and (1.24) demonstrate an
important main difference between energy and exergy: energy is conserved,
while exergy, a measure of energy quality or work potential, can be consumed.
These balances describe what is happening in a system between two instants of
time. For a complete cyclic process where the initial and final states of the sys-
tem are identical, the accumulation terms in all the balances are zero.

1.6.4.2 Detailed Balances


Two types of systems are normally considered: open (flow) and closed (non-
flow). In general, open systems have mass, heat, and work interactions, and
closed systems heat and work interactions. Mass flow into, heat transfer into,
and work transfer out of the system are defined to be positive. Mathematical
formulations of the principles of mass and energy conservation and entropy
nonconservation can be written for any system, following the general physical
interpretations in Eqs. (1.21)–(1.24).
Consider a nonsteady flow process in a time interval t1 to t2. Balances of mass,
energy, entropy, and exergy, respectively, can be written for a control volume as
1.6 Energy and Exergy Fundamentals 19

X X
mi  me ¼ m2  m1 (1.25)
i e
X X X
ðe + PvÞmi  ðe + PvÞme + ðQr Þ1, 2  ðW 0 Þ1, 2 ¼ E2  E1 (1.26)
i e r

X X XQr 
si mi  se me + + Π 1, 2 ¼ S2  S1 (1.27)
i e r
Tr 1, 2
X X X
exi mi  exe me + ðExQr Þ1, 2  ðExW Þ1, 2  ðWNET Þ1, 2  I1, 2 ¼ Ex2  Ex1 (1.28)
i e r

Here, mi and me denote, respectively, the amounts of mass input across port i
and exiting across port e; (Qr)1,2 denotes the amount of heat transferred into the
control volume across region r on the control surface; ExQr denotes the amount
of exergy related to Qr; (W0 )1,2 denotes the amount of work transferred out of
the control volume; ExW denotes the amount of exergy related to W; WNET is
provided in Eq. (1.18); Π 1,2 denotes the amount of entropy created in the con-
trol volume; I denotes the amount of exergy consumption in the control vol-
ume; m1, E1, S1, and Ex1 denote, respectively, the amounts of mass, energy,
entropy, and exergy in the control volume at time t1 and m2, E2, S2, and Ex2
denote, respectively, the same quantities at time t2; and e, s, ex, P, T, and v
denote specific energy, specific entropy, specific exergy, absolute pressure, abso-
lute temperature, and specific volume, respectively. The total work W0 done by a
system excludes flow work, which is provided in Eq. (1.17). The specific energy
e is given by
e ¼ u + ke + pe (1.29)

where u, ke, and pe denote, respectively, specific internal, kinetic, and potential
(due to conservative force fields) energies. For irreversible processes, Π 1,2 > 0,
and for reversible processes, Π 1,2 ¼ 0. I1.2 is defined in Eq. (1.19).
The left sides of Eqs. (1.25)–(1.28) represent the net amounts of mass, energy,
entropy, and exergy transferred into (and in the case of entropy created and
exergy consumed within) the control volume, while the right sides represent
the amounts of these quantities accumulated within the control volume.
For the mass flow mj across port j,
ð t2 "ð #
mj ¼ ðρVn dAÞj dt (1.30)
t1 j

Here, ρ is the density of matter crossing an area element dA on the control sur-
face in time interval t1 to t2 and Vn is the velocity component of the matter flow
normal to dA. The integration is performed over port j on the control surface.
One-dimensional flow (i.e., flow in which the velocity and other intensive
20 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

properties do not vary with position across the port) is often assumed. Then,
the previous equation becomes
ð t2
mj ¼ ðρVn dAÞj dt (1.31)
t1

It has been assumed that heat transfers occur at discrete regions on the control
surface and the temperature across these regions is constant. If the temperature
varies across a region of heat transfer,
ð t2 ð 
ðQr Þ1, 2 ¼ ðqdAÞr dt (1.32)
t1 r

and
  ð t2 ð  
Qr q
¼ dAr dt (1.33)
Tr 1, 2 t1 r T r

ð t2 ð   
T0
ðExQr Þ1, 2 ¼ 1 qr dAr dt (1.34)
t1 r Tr

where Tr is the temperature at the point on the control surface where the heat
flux is qr. The integral is performed over the surface area of region Ar.
The quantities of mass, energy, entropy, and exergy in the control volume
(denoted by m, E, S, and Ex) on the right sides of Eqs. (1.25)–(1.28), respec-
tively, are given more generally by
ð
m ¼ ρdV (1.35)
ð
E ¼ ρedV (1.36)
ð
S ¼ ρsdV (1.37)
ð
Ex ¼ ρξdV (1.38)

where the integrals are over the control volume.


For a closed system, mi ¼ me ¼ 0 and Eqs. (1.25)–(1.28) become
m2  m1 ¼ 0 (1.39)
X
ðQr Þ1, 2  ðW 0 Þ1, 2 ¼ E2  E1 (1.40)
r

XQr 
+ Π 1, 2 ¼ S2  S1 (1.41)
r
Tr 1, 2
1.6 Energy and Exergy Fundamentals 21

X
ðExQr Þ1, 2  ðExW Þ1, 2  ðWNET Þ1, 2  I1, 2 ¼ Ex2  Ex1 (1.42)
r

When volume is fixed, (WNET)1,2 ¼ 0 in Eqs. (1.28) and (1.42). Also, when the
initial and final states are identical as in a complete cycle, the right sides of
Eqs. (1.25)–(1.28) and (1.39)–(1.42) are zero.

1.6.5 Energy and Exergy Efficiencies


Efficiency is an important consideration in decision making regarding resource
utilization. Efficiency is defined as “the ability to produce a desired effect with-
out waste of, or with minimum use of, energy, time, resources, etc.,” and is used
by people to mean the effectiveness with which something is used to produce
something else or the degree to which the ideal is approached in performing
a task.
For general engineering systems, nondimensional ratios of quantities are typ-
ically used to determine efficiencies. Ratios of energy are conventionally used to
determine efficiencies of engineering systems whose primary purpose is the
transformation of energy. These efficiencies are based on the FLT. A process
has maximum efficiency according to the first law if energy input equals recov-
erable energy output (i.e., if no “energy losses” occur). However, efficiencies
determined using energy are misleading because in general, they are not mea-
sures of “an approach to an ideal.” Measures of performance that take into
account limitations imposed by the second law are SLT-based efficiencies.
From energy or exergy viewpoints, a gauge of how effectively the input is con-
verted to the product is the ratio of product to input. Energy (η) and exergy (ψ)
efficiencies are often written for steady-state processes occurring in systems as
Energy in product outputs Energy loss
η¼ ¼1 (1.43)
Energy in inputs Energy in inputs

Exergy in product outputs Exergy loss + Exergy consumption


ψ¼ ¼ 1 (1.44)
Exergy in inputs Exergy in inputs

Two other common exergy-based efficiencies for steady-state devices are as


follows:
Total exergy output Exergy consumption
Rational efficiency ¼ ¼1 (1.45)
Total exergy input Total exergy input

Theoretical minimum exergy input required


Task efficiency ¼ (1.46)
Actual exergy input

Exergy efficiencies often give more illuminating insights into process perfor-
mance than energy efficiencies because (i) they weigh energy flows according
22 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

to their exergy contents and (ii) they separate inefficiencies into those associ-
ated with effluent losses and those due to irreversibilities. In general, exergy effi-
ciencies provide a measure of potential for improvement.

1.7 APPROACHES TO EXERGY AND OTHER SECOND LAW


ANALYSES
As mentioned earlier, second law analysis (SLA) provides better insights on
identifying losses and improvement potentials in a process. Many researchers
and scientists discuss SLA theory, for example, Szargut (1980), Edgerton
(1982), Moran (1989, 1990), Kotas (1995), Szargut et al. (1988), Petit and
Gaggioli (1980), Moran and Sciubba (1994), Rosen (1996), Kestin (1980),
and Dincer and Rosen (2013). Their discussions consider flow processes at
steady state but can easily be extended to other processes.
Table 1.5 shows five SLA approaches (exergy analysis, physical exergy analysis,
exergy consumption analysis, negentropy analysis, and entropy analysis) and
the required reference environment properties for each. The five SLA
approaches are listed in order of comprehensiveness in Table 1.5, starting with
the least comprehensive. Balances of the appropriate quantities for all SLAs
include a consumption or creation term accounting for inefficiencies due to
internal irreversibilities. Consequently, it is meaningless to compare the results
of exergy consumption, negentropy, and entropy analyses to those of energy
analysis because, while exergy and negentropy can be consumed and entropy
created, energy is conserved.
The fact that a reference environment must be subjectively specified does not
detract from the usefulness of SLA. Rather, it provides the ability to choose
an environment such that the analysis appears in its greatest simplicity, the

Table 1.5 Selected Second law Analyses and the Reference Environment Properties Required
for Each
Required Reference
Type of SLA Environment Properties Notes

Entropy analysis – Provides entropy creation due to irreversibilities


Negentropy analysis – Negentropy consumption is equal to entropy creation
due to irreversibilities
Exergy consumption T0 Change in work potentials is investigated
analysis
Physical exergy analysis T0, P0 Useful when stream compositions do not change
Exergy analysis T0, P0, μj00 ( j ¼ 1, 2, …) Most comprehensive SLA

Adapted from Rosen (1999).


1.7 Approaches to Exergy and Other Second Law Analyses 23

Second law of thermodynamics

Entropy analysis

Negentropy analysis

Exergy consumption analysis

Physical exergy analysis

Exergy analysis

FIGURE 1.6 Hierarchy of the various approaches to SLA, indicating all SLAs stem from the SLT and
showing the SLAs in order of increasing comprehensiveness (starting at the top of the diagram).

greatest insight is gained into process performance, and the results are most rel-
evant to the local surroundings. Different types of SLA generally permit the
determination of different types of efficiencies and measures of performance.
For both physical exergy and exergy analyses, the magnitudes of losses can
be related to the magnitudes of inputs and outputs (because measures of work
potential are rigorously defined as physical exergy or exergy). Second law effi-
ciencies that are analogous to first law efficiencies can therefore be evaluated.
This is not the case for exergy consumption, negentropy, and entropy analyses,
which only permit relative comparisons of the performances of different
subprocesses.
As noted above, the SLAs that stem from the SLT can be categorized in order of
decreasing comprehensiveness as follows (see Fig. 1.6): exergy, physical exergy,
exergy consumption, negentropy, and entropy analyses. As the level of compre-
hensiveness decreases, the “usefulness” of a substance becomes less rigorously
defined as a specific work potential, the environment needs to be specified in
less detail, the required calculations become less complex, and the results con-
tain less information.

1.7.1 Illustrative Example


The approaches to SLA are illustrated by considering a boiler, modeled as a
closed heat exchanger and an adiabatic combustor (see Fig. 1.7). Steady-state
operation is assumed and changes in kinetic and potential energies are consid-
ered negligible.
24 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Heat loss
Steam

Hot products
Fuel of combustion Heat
Combustor Flue gas
exchanger
Air

Water
FIGURE 1.7 A boiler, broken down hypothetically into an adiabatic combustor and a closed heat
exchanger.

1.7.1.1 First Law Analysis


Energy rate balances for the heat exchanger (H) and combustor (C) portions,
respectively, of the boiler (B) can be written as
 
m_ h hp  hg ¼ m_ c ðhs  hl Þ + Q_ H (1.47)

m_ f hf + m_ a ha ¼ m_ p hp (1.48)

and can be added to yield the energy rate balance for the overall boiler. Here,
the subscripts f, a, p, g, l, and s denote fuel, air, hot products of combustion, flue
gases, liquid water, and steam, respectively. Also, the mass flow rate through the
hot side of the closed heat exchanger is denoted ṁh and through the cold side
ṁc. Since these flow rates are constant across the heat exchanger, m_ h ¼ m_ p ¼ m_ g
and m_ c ¼ m_ l ¼ m_ s .
Appropriate first law efficiencies can be written as follows for the boiler, the
heat exchanger, and the combustor, respectively:

m_ c ðhs  hl Þ
ηB ¼ (1.49)
m_ f hf

m_ c ðhs  hl Þ
ηH ¼   (1.50)
m_ h hp  hg

m_ p hp
ηC ¼ (1.51)
m_ f hf

The specific enthalpy of the fuel, hf, is evaluated such that it is equal to the
higher heating value. For an adiabatic combustor, the efficiency definition in
Eq. (1.51) always yields 100%.
1.7 Approaches to Exergy and Other Second Law Analyses 25

1.7.1.2 Second Law Analysis


Exergy rate balances for the heat exchanger and combustor portions, respec-
tively, of the boiler can be written as
 
I_H ¼ m_ h exp  exg + m_ c ðexl  exs Þ (1.52)

I_C ¼ m_ f exf + m_ a exa  m_ p exp (1.53)

These rate balances can be added to yield the exergy rate balance for the overall
boiler:

I_B ¼ I_H + I_C (1.54)

Note that the boundary around the heat exchanger has been located so that sur-
face heat losses occur at T0; the exergy flow rate associated with Q_ H is therefore
zero. Exergy efficiencies, analogous to the energy efficiencies in Eqs. (1.49)–
(1.51), respectively, can be written as follows:

m_ c ðexs  exl Þ
ψB ¼ (1.55)
m_ f exf

m_ c ðexs  exl Þ
ψH ¼   (1.56)
m_ h exp  exg

m_ p exp
ψC ¼ (1.57)
m_ f exf

Physical exergy analysis cannot be applied rationally to the overall boiler or the
combustor because the chemical compositions of some streams within these
systems are changing. However, since the chemical compositions for both
the flows remain unaltered through the heat exchanger, a valid physical exergy
balance can be written for that device:
  
I_H ¼ m_ h exph p  exph g + m_ c exph l  exph s (1.58)

and an alternative second law efficiency, ψ H, can be defined based on physical


exergy:
  
m_ c exph s  exph l
ψ ph ¼  (1.59)
H
m_ h exph p  exph g

Since chemical exergy terms in Eq. (1.56) cancel for the present case,
ψ H ¼ (ψ ph)H. Exergy consumption rates for the boiler, the heat exchanger,
and the combustor can be determined using exergy rate balances (Eqs. 1.52
and 1.53) or, for the heat exchanger alone, using the physical exergy rate
balance (Eq. 1.58).
26 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

To provide sample numerical values, data are considered for a coal-fired boiler
(Petit and Gaggioli, 1980). Following the system model in Fig. 1.7, tempera-
tures, pressures, and flow rates for streams are given in Table 1.6, stream exergy
values in Table 1.7, and device efficiencies in Table 1.8. The first and second law
efficiencies in Table 1.8 differ significantly (e.g., the energy and exergy efficien-
cies for the coal-fired boiler are 85% and 34%, respectively).

Table 1.6 Flow Data for a Coal-Fired Boiler Presented in Fig. 1.7
Normalized Mass Flow
Flowa Temperature (K) Pressure (MPa) Rate (kg/kg coal)

Fuel (coal) 298 0.101 1


Air 298 0.101 9.38
Product gas 1844 0.101 10.38
Flue gas 475 0.101 10.38
Liquid water 298 5.84 6.66
Steam 755 5.84 6.66
a
Composition of combustion gases and coal (Illinois No. 6) given by Petit and Gaggioli (1980).

Table 1.7 SLA Results for Flows for a Coal-Fired Boiler (Fig. 1.7)
Normalized Flow Rate (kJ/kg coal)
Flow Physical Exergy Chemical Exergy Exergy

Fuel (coal) 0 26,391 26,391


Air 0 0 0
Product gas 18,458 1031 19.489
Flue gas 424 1031 1455
Liquid water 50 0 50
Steam 8963 0 8963
Source: Rosen (1999).

Table 1.8 Second and First law Analyses Results Relating to Device
Efficiencies for a Coal-Fired Boiler (Fig. 1.7)
Efficiency (%)
Device Energy Physical Exergy Exergy

Combustor 100 – 74
Heat exchanger 98 49 49
Boiler 85 – 34

Source: Rosen (1999).


1.8 Linkages Between Exergy, Economics, The Environment, and Sustainability 27

1.7.2 Implications of Second Law Analysis


SLA permits direct evaluation and improvement of the thermodynamic effi-
ciency and performance of energy processes and systems. In addition, there
exist important implications of SLA in other fields. These are considered by
many to be significant and are discussed in the remainder of this chapter.

1.8 LINKAGES BETWEEN EXERGY, ECONOMICS,


THE ENVIRONMENT, AND SUSTAINABILITY
The linkage between energy and economics was a prime concern in 1970s. At
that time, the linkage between energy and the environment did not receive
much attention. As environmental concerns became major issues in the
1980s, the link between energy utilization and the environment became more
recognized. Since then, there has been increasing attention on this connection,
as it has become clearer that energy production, transformation, transport, and
use all impact the global environment. Simultaneously, concerns have been
expressed about the nonsustainable nature of human activities, and extensive
efforts have begun to be devoted towards developing methods for achieving
sustainable development.
The relation between sustainable development and the use of resources, partic-
ularly energy resources, is of great significance to societies. Attaining sustainable
development requires that sustainable energy resources be used and is assisted
if resources are used efficiently. Exergy methods are important since they are
useful for improving efficiency. The relations between exergy and both energy
and the environment make it clear that exergy is directly related to sustainable
development.
Many suggest that mitigating the environmental impact of energy resource uti-
lization and achieving increased resource utilization efficiency are best
addressed by considering exergy. By extension, since these topics are critical ele-
ments in achieving sustainable development, exergy also appears to provide the
basis for developing comprehensive methodologies for sustainability. The
exergy of an energy form or a substance is a measure of its usefulness or quality
or potential to cause change. The latter point suggests that exergy may be, or
provide the basis for, an effective measure of the potential of a substance or
energy form to impact the environment. In practice, the authors feel that a thor-
ough understanding of exergy and the insights it can provide into the efficiency,
environmental impact, and sustainability of energy systems are required for the
engineer or scientist working in the area of energy systems and the environ-
ment. Further, as energy policies increasingly play an important role in addres-
sing sustainability issues and a broad range of local, regional, and global
28 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

environmental concerns, policy makers also need to appreciate the exergy con-
cept and its ties to these concerns. The need to understand the linkages between
exergy and energy, sustainable development, and environmental impact has
become increasingly significant.
Despite the fact that many studies appeared during the past few decades con-
cerning the close relationship between energy and the environment, there has
only recently been an increasing number of works on the linkage between the
exergy and the environment (e.g., Reistad, 1970; Szargut, 1980; Wepfer and
Gaggioli, 1980; Crane et al., 1992; Rosen and Dincer, 1997, 2001; Dincer
and Rosen, 1999; Sciubba, 1999).
Many researchers (El-Sayed and Gaggioli, 1989; Tsatsaronis, 1987, 1994, 1998;
Torres et al., 1996) have developed methods of performing economic analyses
based on exergy, which are referred to by such names as thermoeconomics, sec-
ond law costing, and exergoeconomics. These analysis techniques recognize
that exergy, not energy, is the commodity of value in a system, and they con-
sequently assign costs and/or prices to exergy-related variables. The objectives
of most of these analysis techniques generally include the determination of the
following items:

• Appropriate allocation of economic resources so as to optimize the design


and/or operation of a system
• Economic feasibility and profitability of a system (by obtaining the actual
costs of products and their appropriate prices)

1.9 RELATIONS BETWEEN EXERGY AND ECONOMICS


In the analysis and design of energy systems, technical disciplines (especially
thermodynamics) are combined with economics to achieve optimum designs.
Economic issues are important in the evaluation of green energy technologies.
For energy-conversion devices, costs are conventionally based on energy. Many
researchers (Rosen and Dincer, 2003; Tsatsaronis, 1987, 1994; El-Sayed and
Gaggioli, 1989; Mazur, 2005), however, have recommended that costs are bet-
ter distributed among outputs based on exergy.
Exergy is a useful concept in economics. In macroeconomics, exergy offers a
way to evaluate resource depletion and environmental destruction by means
of an exergy tax. In microeconomics, exergy has been fruitfully combined with
cost-benefit analysis to improve the design. By minimizing life cycle cost, we
find the best system due to the prevailing economic conditions, and by mini-
mizing the exergy losses, we also minimize environmental effects. Designing
efficient and cost-effective systems, which also meet environmental conditions,
1.9 Relations Between Exergy and Economics 29

is one of the foremost challenges that the engineers face. In the world, with
finite natural resources and large energy demands, it becomes increasingly
important to understand the mechanisms that degrade energy and resources
and to develop systematic approaches for improving systems and thus also
to reduce the impact on the environment. Exergetics combined with econom-
ics, both macro- and microeconomics, represents powerful tools for the system-
atic study and optimization of systems. A number of people working on the
thermoeconomic aspects of energy systems quote Georgescu-Roegen (1971)
as the father of the thermodynamics of economics, who played a pioneering
role in this direction. Exergetics and microeconomics form the basis of thermo-
economics (Evans and Tribus, 1962), which is also named exergoeconomics
(Bejan et al., 1996) or exergonomics (Yantovskii, 1994). The concept of utility
is a central concept in macroeconomics. Utility is also closely related to exergy,
and an exergy tax is an example of how exergy could be introduced into
macroeconomics.
Wall (1993) pointed out that “the concept of exergy is crucial not only to effi-
ciency studies but also to cost accounting and economic analyses. Costs should
reflect value, since the value is not in energy but in exergy, assignment of cost to
energy leads to misappropriations, which are common and often gross. Using
exergy content as a basis for cost accounting is important to management for
pricing products and for their evaluation of profits. It is also useful to engineer-
ing for operating and design decisions, including design optimization. Thus,
exergy is the only rational basis for evaluating: fuels and resources, process,
device, and system efficiencies, dissipations and their costs, and the value
and cost of systems outputs.”
Methods have developed of performing economic analyses based on exergy, which
are referred to as thermoeconomics, second law costing, and exergoeconomics
(Tsatsaronis, 1987, 1994; El-Sayed and Gaggioli, 1989; Mazur, 2005; Jaber
et al., 2004). These methods recognize that exergy, not energy, is the commodity
of value in a system and assign costs and/or prices to exergy-related variables.
These methods usually help determine the appropriate allocation of economic
resources so as to optimize the design and operation of a system and/or the
economic feasibility and profitability of a system (by obtaining actual costs of
products and their appropriate prices).
Tsatsaronis (1987) identified four main types of analysis methodologies,
depending on which of the following forms the basis of the technique:
(i) exergy-economic cost accounting, (ii) exergy-economic calculus analysis,
(iii) exergy-economic similarity number, and (iv) product/cost efficiency
diagrams. These methods are discussed and compared elsewhere (e.g., Moran,
1989; Kotas, 1995; Szargut et al., 1988; Szargut, 1980; Tsatsaronis, 1987;
El-Sayed and Gaggioli, 1989).
30 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

One rationale for the statement that costs are better distributed among outputs
if cost accounting is based on exergy is that exergy often is a consistent measure
of economic value (i.e., a large quantity of exergy is often associated with a valu-
able commodity), while energy is only sometimes a consistent measure of eco-
nomic value. This rationale can be illustrated with results of previous research
by the authors on the coal-fired electricity-generating station considered earlier,
which suggested possible general relations between thermodynamic losses and
capital costs (Rosen and Dincer, 2003). That work examined thermodynamic
and economic data for mature devices and showed that correlations exist
between capital costs and thermodynamic losses for devices. The existence of
such correlations likely implies that designers knowingly or unknowingly
incorporate the recommendations of exergy analysis into process designs indi-
rectly. The results of the analysis of the relations between thermodynamic losses
and capital costs for devices in a modern coal-fired electricity-generating station
led to several observations:
• For the thermodynamic losses considered (energy and exergy loss), a
significant parameter appears to be the ratio of thermodynamic loss rate
to capital cost.
• A systematic correlation appears to exist between exergy loss rate and
capital cost, but not between energy loss rate and capital cost. This finding
is based on the observation that the variation in thermodynamic loss rate/
capital cost ratio values for different devices is large when based on energy
loss and small when based on exergy loss.
• Devices in modern coal-fired electricity-generating stations appear to
conform approximately to a particular value of the thermodynamic loss
rate/capital cost ratio (based on exergy loss), which reflects the
“appropriate” trade-off between exergy losses and capital costs that is
practiced in successful plant designs.
An understanding of the relations between exergy and the environment may
reveal the underlying fundamental patterns and forces affecting changes in
the environment and help researchers to deal better with environmental dam-
age. Rosen (1999) had identified relationships between exergy and environ-
mental impact as resource degradation and waste exergy emissions. These
relationships demonstrate that the potential usefulness of the thermodynamic
property exergy in addressing and solving environmental problems is substan-
tial and that further work in this field is required before this potential can be
properly and fully exploited.
Two of these relationships are as follows:
• Resource degradation: Kestin (1980) defined a resource as a material found
in nature or created artificially, which is in a state of disequilibrium
with the environment. Resources have exergy as a consequence of this
1.10 Relations Between Exergy and Environmental Impact and Ecology 31

disequilibrium. For some resources (e.g., metal ores), it is their


composition that is valued. Many processes exist to increase the value
of such resources by purifying them (i.e., by increasing their exergy). This
is done at the expense of consuming at least an equivalent amount of
exergy elsewhere (e.g., burning coal to produce process heat for metal ore
refining). For other resources (e.g., fuels), it is normally their reactivity
that is valued (i.e., their potential to cause change or “drive” a task or
process). By preserving exergy through increased efficiency (i.e., degrading
as little exergy as necessary for a process), environmental damage is
reduced. Increased efficiency also has the effect of reducing exergy
emissions, which, as discussed in the next subsection, also play a role in
environmental damage.
• Waste exergy emissions: The exergy associated with process wastes emitted
to the environment can be viewed as a potential for environmental
damage. Typical process wastes have exergy, a potential to cause change,
as a consequence of not being in stable equilibrium with the
environment. When emitted to the environment, this exergy represents
a potential to change the environment. In some cases, this exergy may
cause a change perceived to be beneficial (e.g., the increased rate of growth
of fish and plants near the cooling water outlets from thermal power
plants). More often, however, emitted exergy causes a change, which is
damaging to the environment (e.g., the deaths of fish and plants in some
lakes due to the release of specific substances in stack gases as they
react and come to equilibrium with the environment).

1.10 RELATIONS BETWEEN EXERGY AND


ENVIRONMENTAL IMPACT AND ECOLOGY
People have long been intrigued by the implications of the laws of thermody-
namics on the environment. One myth speaks of ouroboros, a serpent-like
creature that survived and regenerated itself by eating only its own tail. By nei-
ther taking from nor adding to its environment, this creature was said to be
completely environmentally benign and self-sufficient. It is useful to examine
this creature in light of the thermodynamic principles recognized today.
Assuming that ouroboros was an isolated system (i.e., it received no energy
from the sun or the environment and emitted no energy during any process),
ouroboros’ existence would have violated neither the conservation law for mass
nor the FLT (which states energy is conserved). However, unless it was a revers-
ible creature, ouroboros’ existence would have violated the second law (which
states that exergy is reduced for all real processes), since ouroboros would have
had to obtain exergy externally to regenerate the tail it ate into an equally
ordered part of its body (or it would ultimately have dissipated itself to an
32 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

unordered lump of mass). Thus, ouroboros would have to have had an impact
on its environment.
Besides demonstrating that, within the limits imposed by the laws of thermo-
dynamics, all real processes must have some impact on the environment, this
example is intended to illustrate the following key point: the second law is instru-
mental in providing insights into environmental impact (e.g., Hafele, 1981;
Edgerton, 1982; Rosen and Dincer, 1997). Today, the principles demonstrated
through this example remain relevant, and technologies are sought having
ouroboros’ characteristics of being environmentally benign and self-sufficient
(e.g., the University of Minnesota researchers built an “energy-conserving” house
called ouroboros (Markovich, 1978)). The importance of the second law in
understanding environmental impact implies that exergy, which is based on
the second law, has an important role to play in this field.
The most appropriate link between the second law and environmental impact
has been suggested to be exergy (Rosen and Dincer, 1997), in part because it is a
measure of the departure of the state of a system from that of the environment.
The magnitude of the exergy of a system depends on the states of both the sys-
tem and the environment. This departure is zero only when the system is in
equilibrium with its environment. The concept of exergy analysis as it applies
to the environment is discussed in detail elsewhere (Rosen and Dincer, 1997).
An understanding of the relations between exergy and the environment may
reveal the underlying fundamental patterns and forces affecting changes in
the environment and help researchers more effectively address the causes of
environmental damage. Tribus and McIrivne (1971) suggested that performing
exergy analyses of the natural processes occurring on the Earth could form a
foundation for ecologically sound planning because it would indicate the dis-
turbance caused by large-scale changes. Three relationships between exergy and
environmental impact (Rosen and Dincer, 1997) are discussed below:
Order destruction and chaos creation: The destruction of order, or the creation of
chaos, is a form of environmental damage. Entropy is fundamentally a measure
of chaos, and exergy of order. A system of high entropy is more chaotic or dis-
ordered than one of low entropy, and relative to the same environment, the
exergy of an ordered system is greater than that of a chaotic one. For example,
a field with papers scattered about has higher entropy and lower exergy than the
field with the papers neatly piled. The exergy difference of the two systems is a
measure of (i) the exergy (and order) destroyed when the wind scatters the stack
of papers and (ii) the minimum work required to convert the chaotic system to
the ordered one (i.e., to collect the scattered papers). In reality, more than this
minimum work, which only applies if a reversible cleanup process is employed,
is required. The observation that people are bothered by a landscape polluted
with papers chaotically scattered about, but value the order of a clean field with
1.10 Relations Between Exergy and Environmental Impact and Ecology 33

the papers neatly piled at the side, suggests that, on a more abstract level, ideas
relating exergy and order in the environment may involve human values
(Hafele, 1981) and that human values may in part be based on exergy
and order.
Resource degradation: The degradation of resources found in nature is a form of
environmental damage. Kestin (1980) defined a resource as a material found in
nature or created artificially, which is in a state of disequilibrium with the envi-
ronment, and noted that resources have exergy as a consequence of this disequi-
librium. Two main characteristics of resources are valued:

• Composition (e.g., metal ores): Many processes exist to increase the value
of such resources by purifying them, which increases their exergy. Note
that purification is accomplished at the expense of consuming at least an
equivalent amount of exergy elsewhere (e.g., using coal to drive metal
ore refining).
• Reactivity (e.g., fuels): That is, their potential to cause change or “drive” a
task or process.

Two principal general approaches exist to reduce the environmental impact


associated with resource degradation:

• Increased efficiency. Increased efficiency preserves exergy by reducing the


exergy necessary for a process and therefore reduces environmental
damage. Increased efficiency also usually reduces exergy emissions, which,
as discussed in the next section, also play a role in environmental damage.
• Using external exergy resources (e.g., solar energy). The Earth is an open
system subject to a net influx of exergy from the sun. It is the exergy
(or order states) delivered with solar radiation that is valued; all the energy
received from the sun is ultimately radiated out to the universe.
Environmental damage can be reduced by taking advantage of the
openness of the Earth and utilizing solar radiation (instead of degrading
resources found in nature to supply exergy demands). This would not
be possible if the Earth was a closed system, as it would eventually become
more and more degraded or “entropic.”

Waste exergy emissions: The exergy associated with waste emissions can be
viewed as a potential for environmental damage in that the exergy of the wastes,
as a consequence of not being in stable equilibrium with the environment, rep-
resents a potential to cause change. When emitted to the environment, this
exergy represents a potential to change the environment. Usually, emitted
exergy causes a change that is damaging to the environment, such as the deaths
of fish and plants in some lakes due to the release of specific substances in stack
gases as they react and come to equilibrium with the environment, although in
some cases, the change may be perceived to be beneficial (e.g., the increased
34 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

growth rate of fish and plants near the cooling water outlets from thermal
power plants). Further, exergy emissions to the environment can interfere with
the net input of exergy via solar radiation to the Earth (e.g., emissions of CO2
and other greenhouse gases from many processes appear to cause changes to
the atmospheric CO2 concentration, affecting the receiving and reradiating
of solar radiation by the Earth). The relation between waste exergy emissions
and environmental damage has been recognized by several researchers (e.g.,
Reistad, 1970). By considering the economic value of exergy in fuels, Reistad
developed an air pollution rating that he felt was preferable to the mainly
empirical ratings then in use, in which the air pollution cost for a fuel was esti-
mated as either (i) the cost to remove the pollutant or (ii) the cost to society of
the pollution in the form of a tax that should be levied if pollutants are not
removed from effluent streams.
Although the previous two points indicate simultaneously that exergy in the
environment in the form of resources is of value while exergy in the environ-
ment in the form of emissions is harmful due to its potential to cause environ-
mental damage, confusion can be avoided by considering whether or not the
exergy is constrained (see Fig. 1.8). Most resources found in the environment
are constrained and by virtue of their exergy are of value, while unconstrained
emissions of exergy are free to impact in an uncontrolled manner on the envi-
ronment. To elaborate further on this point, consider a scenario in which emis-
sions to the environment are constrained (e.g., by separating sulfur from stack
gases). This action yields two potential benefits: the potential for environmen-
tal damage is restrained from entering the environment, and the now con-
strained emission potentially becomes a valued commodity, that is, a source
of exergy.

Unconstrained exergy
(a potential to cause a change in the
environment)

Exergy
emissions to the
environment

Constrained exergy
(a potential to cause a change)

FIGURE 1.8 Comparison of constrained and unconstrained exergy illustrating that exergy constrained
in a system represents a resource, while exergy emitted to the environment becomes unconstrained
and represents a driving potential for environmental damage.
1.11 Relations Between Exergy and Sustainability 35

Process exergy efficiency

• Order destruction and chaos creation


• Resource degradation
• Waste exergy emissions

FIGURE 1.9 Qualitative depiction of relation between environmental impact (in terms of order
destruction and chaos creation, or resource degradation, or waste exergy emissions) of a process
and its exergy efficiency.

The decrease in the environmental impact of a process, in terms of several mea-


sures, as the process exergy efficiency increases is illustrated approximately in
Fig. 1.9.

1.11 RELATIONS BETWEEN EXERGY AND


SUSTAINABILITY
Exergy can be considered as the confluence of energy, environment, and sus-
tainable development, as shown in Fig. 1.10, which illustrates the interdisci-
plinary character of exergy and its central focus among these disciplines.

Sustainable
development

Exergy
Energy Environment

FIGURE 1.10 Interdisciplinary coverage of exergy analysis.


36 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Exergy methods can be used to improve sustainability. For example, Cornelissen


(1997) pointed out that one important element in obtaining sustainable
development is the use of exergy analysis. By noting that energy can never be
“lost”, as it is conserved according to the FLT, while exergy can be lost due to
internal irreversibilities, that study suggests that exergy losses, particularly
due to the use of nonrenewable energy forms, should be minimized to obtain
sustainable development. Further, the study shows that environmental effects
associated with emissions and resource depletion can be expressed in terms
of one exergy-based indicator, which is founded on physical principles.
Sustainable development also includes economic viability. Thus, the methods
relating exergy and economics also reinforce the link between exergy and
sustainable development. The objectives of most existing analysis techniques
integrating exergy and economics include the determination of (i) the app-
ropriate allocation of economic resources so as to optimize the design and
operation of a system and/or (ii) the economic feasibility and profitability
of a system. Exergy-based economic analysis methods are referred to by
such names as thermoeconomics, second law costing, cost accounting, and
exergoeconomics.
Figure 1.11 illustratively presents the relation between exergy and sustain-
ability and environmental impact. There, sustainability is seen to increase
and environmental impact to decrease as the process exergy efficiency increases.
The two limiting efficiency cases are significant. First, as exergy efficiency
approaches 100%, environmental impact approaches zero, since exergy is only
converted from one form to another without loss, through either internal con-
sumptions or waste emissions. Also sustainability approaches infinity because
the process approaches reversibility. Second, as exergy efficiency approaches
0%, sustainability approaches zero because exergy-containing resources are
Environmental impact

Sustainability

0 Exergy efficiency (%) 100

FIGURE 1.11 Qualitative depiction of the relation between the environmental impact and sustainability
of a process and its exergy efficiency.
1.12 Closing Remarks 37

Nonsustainable development

Increased exergy efficiency

Reduction of exergy related


environmental degradation

Use of sustainable energy


resources

Sustainable development

FIGURE 1.12 Some key contributions of exergy methods to increasing the sustainability of
nonsustainable systems and processes.

used but nothing is accomplished. Also, environmental impact approaches


infinity because, to provide a fixed service, an ever-increasing quantity of
resources must be used and a correspondingly increasing amount of exergy-
containing wastes is emitted.
Some important contributions that can be derived from exergy methods, for
increasing the sustainability of development that is nonsustainable, are pre-
sented in Fig. 1.12. Development typical of most modern processes, which
are generally nonsustainable, is shown at the bottom of the figure. A future
in which development is sustainable is shown at the top of the figure, while
some key exergy-based contributions towards making development more sus-
tainable are shown and include increased exergy efficiency, reduction of
exergy-based environmental degradation, and use of sustainable exergy
resources.

1.12 CLOSING REMARKS


In this chapter, the benefits of using exergy for assessing efficiency, environ-
mental impact, and sustainability have been demonstrated, with an introduc-
tion to heating, refrigeration, and air conditioning. It is believed that the
concepts encompassing exergy have a significant role to play in evaluating
and increasing the use of sustainable energy and technologies. Although deci-
sions regarding the design and modification of energy systems are normally
concerned not only with efficiency but also with economics, environmental
impact, safety, and other issues, exergy should prove useful in design and
improvement activities to engineers and scientists, as well as decision and pol-
icy makers.
38 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Several important concluding remarks can be drawn from this chapter indicat-
ing that exergy

• provides a suitable technique for furthering the goal of more efficient


energy resource use and hence energy conservation,
• is the basis of an effective method using the conservation of mass and
conservation of energy principles together with the SLT for the design and
analysis of energy systems,
• embodies an indicator of energy quality,
• provides an effective measure of whether or not and by how much it is
possible to design more efficient energy systems by reducing the
inefficiencies in existing systems,
• can be integrated beneficially with economic concepts,
• provides a useful optic for addressing the impact of energy resource use on
the environment,
• provides a loose measure of environmental degradation and hence one of
the potential techniques to reduce environmental impact,
• can contribute to moving towards sustainable development.

An enhanced understanding of the environmental problems relating to energy


presents a significant challenge, both to allow problems to be addressed and to
ensure that the solutions are beneficial for society and the energy policy mak-
ing. The potential usefulness of exergy analysis in addressing and solving
energy-related challenges is substantial and that exergy can play a role in deci-
sion and policy making related to energy activities.

Nomenclature
A area
e specific energy
E energy
ex specific exergy
Ex exergy
g gravitational acceleration
h specific enthalpy
H enthalpy
I exergy consumption
KE kinetic energy
ke specific kinetic energy
m mass
PE potential energy
pe specific potential energy
P pressure
Q heat
s specific entropy
1.12 Closing Remarks 39

S entropy
t time
T temperature
u specific internal energy
U internal energy
v specific volume
V velocity, volume
W work
Z elevation

Greek symbols
η energy efficiency
ρ density
μ chemical potential
Π entropy production
ξ specific exergy
ψ exergy efficiency

Subscripts
0 environmental state, chemical exergy
00 dead state
a air
B boiler
c cold side of heat exchanger
C combustor
cv control volume
e exit
f fuel
g flue gases
gen generation
h hot side of heat exchanger
H heat exchanger
i inlet
I irreversibility
j jth constituent
kin kinetic component
l liquid water
p products of combustion
ph physical component
pot potential component
Q heat
r region of heat interaction
s steam
W work
40 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Superscripts
˙ rate with respect to time

Acronyms
CFC chlorofluorocarbon
FLT first law of thermodynamics
IEA International Energy Agency
Mtoe million tons of oil equivalent
SLA second law analysis
SLT second law of thermodynamics
VOC volatile organic compound

References
Bejan, A., Tsatsaronis, G., Moran, M., 1996. Thermal Design and Optimization. Wiley, New York.
Cornelissen, R.L., 1997. Thermodynamics and Sustainable Development (Ph.D. thesis). University
of Twente, The Netherlands.
Crane, P., Scott, D.S., Rosen, M.A., 1992. Comparison of exergy of emissions from two energy con-
version technologies, considering potential for environmental impact. Int. J. Hydrog. Energy
17, 345–350.
Dincer, I., Rosen, M.A., 1999. Energy, environment and sustainable development. Appl. Energy
64 (1–4), 427–440.
Dincer, I., Rosen, M.A., 2002. Thermal Energy Storage: Systems and Applications. Wiley, London.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development. 2nd ed.,
Elsevier Science, Oxford.
Edgerton, R.H., 1982. Available Energy and Environmental Economics. D.C. Heath, Toronto.
El-Sayed, Y.M., Gaggioli, R.A., 1989. A critical review of second law costing methods. Parts I and II.
Trans. ASME: J. Energy Resour. Technol. 111, 1–15.
Evans, R.B., Tribus, M., 1962. A Contribution to the Theory of Thermoeconomics. Report No. 62-
63, UCLA Department of Energy, Los Angeles, CA.
Georgescu-Roegen, N., 1971. The Entropy Law and Economic Process. Harvard University Press,
Cambridge.
Gunnewiek, L.H., Rosen, M.A., 1998. Relation between the exergy of waste emissions and measures
of environmental impact. Int. J. Environ. Pollut. 10 (2), 261–272.
Hafele, W., 1981. Energy in a Finite World: A Global Systems Analysis. Ballinger, Toronto.
IEA Online Statistics, 2007. Statistics and Balances. International Energy Agency (IEA) of the Orga-
nisation for Economic Cooperation and Development (OECD), Paris, France, www.iea.org/
stats/index.asp.
International Energy Agency Technical Report, 2014a. 2014 key world energy statistics. http://www.
iea.org/publications/freepublications/publication/KeyWorld2014.pdf.
International Energy Agency Technical Report, 2014b. Energy efficiency indicators: fundamentals
on statistics technical report. http://www.iea.org/publications/freepublications/publication/
IEA_EnergyEfficiencyIndicatorsFundamentalsonStatistics.pdf.
References 41

Jaber, M.Y., Nuwayhid, R.Y., Rosen, M.A., 2004. Price-driven economic order systems from a ther-
modynamic point of view. Int. J. Prod. Res. 42, 5167–5184.
Kestin, J., 1980. Availability: the concept and associated terminology. Energy 5, 679–692.
Kotas, T.J., 1995. The Exergy Method of Thermal Plant Analysis, reprint ed. Krieger, Malabar, FL.
Kotas, T.J., Raichura, R.C., Mayhew, Y.R., 1987. Nomenclature for exergy analysis. In: Second Law
Analysis of Thermal Systems. ASME, New York, pp. 171–176.
Lucca, G., 1990. The exergy analysis: role and didactic importance of a standard use of basic con-
cepts, terms and symbols. In: A Future for Energy: Proceedings of the Florence World Energy
Research Symposium, pp. 295–308.
Markovich, S.J., 1978. Autonomous living in the Ouroboros house. In: Solar Energy Handbook,
Popular Science, pp. 46–48.
Mazur, V.A., 2005. Fuzzy thermoeconomic optimization. Int. J. Exergy 2, 1–13.
Moran, M.J., 1989. Availability Analysis: A Guide to Efficient Energy Use, revised ed. American Soci-
ety of Mechanical Engineers, New York.
Moran, M.J., 1990. Second law analysis: what is the state of the art? In: Proceedings of Florence
World Energy Research Symposium (FLOWERS 90), Firenze, Italy, pp. 249–260.
Moran, M.J., Sciubba, E., 1994. Exergy analysis: principles and practice. Trans. ASME: J. Eng. Gas
Turbines Power 116, 285–290.
Petit, P.J., Gaggioli, R.A., 1980. Second law procedures for evaluating processes. In: Gaggioli, R.A.
(Ed.), Thermodynamics: Second Law Analysis. In: ACS Symposium Series, vol. 122. American
Chemical Society, Washington, DC, pp. 15–37.
Reistad, G.M., 1970. Availability: Concepts and Applications. (Ph.D. dissertation). University of
Wisconsin, Madison.
Rosen, M.A., 1990. Comparison based on energy and exergy analyses of the potential cogeneration
efficiencies for fuel cells and other electricity generation devices. Int. J. Hydrog. Energy 15, 267–274.
Rosen, M.A., 1996. Thermodynamic investigation and comparison of selected production processes
for hydrogen and hydrogen derived fuels. Energy Int. J. 21, 1079–1094.
Rosen, M.A., Dincer, I., 1997. On exergy and environmental impact. Int. J. Energy Res. 21, 643–654.
Rosen, M.A., 1999. Second-Law Analysis: Approaches and Implications. Int. J. Energy Res.
23, 415–429.
Rosen, M.A., Dincer, I., 1999. Exergy analysis of waste emissions. Int. J. Energy Res. 23 (13),
1153–1163.
Rosen, M.A., Dincer, I., 2001. Exergy as the confluence of energy, environment and sustainable
development. Exergy Int. J. 1 (1), 3–13.
Rosen, M.A., Dincer, I., 2003. Thermoeconomic analysis of power plants: an application to a coal-
fired electrical generating station. Energy Convers. Manag. 44, 2743–2761.
Rosen, M.A., Dincer, I., 2004. Effect of varying dead-state properties on energy and exergy analyses
of thermal systems. Int. J. Therm. Sci. 43 (2), 121–133.
Rosen, M.A., Etele, J., 2004. Aerospace systems and exergy analysis: applications and methodology
development needs. Int. J. Exergy 1 (4), 411–425.
Rosen, M.A., Horazak, D.A., 1995. Energy and exergy analyses of PFBC power plants. In: Alvarez
Cuenca, M., Anthony, E.J. (Eds.), Pressurized Fluidized Bed Combustion. Chapman and Hall,
London, pp. 419–448 (Chapter 11).
Rosen, M.A., Scott, D.S., 1998. Comparative efficiency assessments for a range of hydrogen produc-
tion processes. Int. J. Hydrog. Energy 23, 653–659.
Rosen, M.A., Tang, R., Dincer, I., 2004. Effect of stratification on energy and exergy capacities in
thermal storage systems. Int. J. Energy Res. 28, 177–193.
42 C HA PT E R 1 : Exergy and its Ties to the Environment, Economics, and Sustainability

Rosen, M.A., Le, M.N., Dincer, I., 2005. Efficiency analysis of a cogeneration and district energy sys-
tem. Appl. Therm. Eng. 25 (1), 147–159.
Sciubba, E., 1999. Exergy as a direct measure of environmental impact. In: Proceedings of the ASME
Advanced Energy Systems Division, vol. 39. American Society of Mechanical Engineers,
New York.
Szargut, J., 1980. International progress in second law analysis. Energy Int. J. 5, 709–718.
Szargut, J., Morris, D.R., Steward, F.R., 1988. Exergy Analysis of Thermal, Chemical and Metallur-
gical Processes. Hemisphere, New York.
Torres, C., Serra, L., Valero, A., Lozano, M.A., 1996. The productive structure and thermoeconomic
theories of system optimization. In: Proceedings of ASME Advanced Energy Systems Division,
vol. 36. American Society of Mechanical Engineers, New York.
Tribus, M., McIrivne, E.C., 1971. Energy and information. Sci. Am. 225 (3), 179–188.
Tsatsaronis, G., 1987. A review of exergoeconomic methodologies. In: Moran, M.J., Sciubba, E.
(Eds.), Second Law Analysis of Thermal Systems. American Society of Mechanical Engineers,
New York, pp. 81–87.
Tsatsaronis, G., 1994. Invited papers on exergoeconomics. Energy Int. J. 19, 279–381.
Tsatsaronis, G., 1998. Design optimization using exergoeconomics. In: Thermodynamics and the
Optimization of Complex Energy Systems. NATO Advanced Study Institute, Neptun,
Romania, pp. 394–410.
Wall, G., 1993. Exergy, ecology and democracy-concepts of a vital society. In: Szargut, J., et al. (Ed.),
ENSEC’93 International Conference on Energy Systems and Ecology, July 5–9, Cracow, Poland,
pp. 111–121.
Wepfer, W.J., Gaggioli, R.A., 1980. Reference datums for available energy. In: Thermodynamics: Sec-
ond Law Analysis. ACS Symposium Series, vol. 122. American Chemical Society, Washington,
DC, pp. 77–92.
Yantovskii, E.I., 1994. Energy and Exergy Currents. Nova, New York.
CHAPTER 2

Energy and Exergy Assessments

2.1 INTRODUCTION
Psychrometrics involves the use of thermodynamics to analyze conditions and
processes involving moist air. A thorough understanding of psychrometrics
is important in the heating, ventilating, air conditioning, and refrigeration
(HVACR) community. Psychrometrics is used not only in assessing and design-
ing heating and cooling processes and ensuring the comfort of building occu-
pants but also in constructing building materials (e.g., insulation and roofing)
and in assessing their stability and fire resistance (Dincer and Rosen, 2013).
Numerous researchers in their related publications and books (e.g., Dincer
et al., 2007; Wepfer et al., 1979; Stecco and Manfrida, 1986; Dincer and
Rosen, 2011; Dincer and Rosen, 2013; Kanoglu et al., 2007; Ratlamwala and
Dincer, 2012) illustrate the application of exergy analysis to a variety of heating,
ventilating, and air conditioning (HVAC) processes.
This chapter describes energy and exergy assessments of the components and
psychrometric processes in HVAC systems and illustrates this material by asses-
sing a novel integrated system for HVACR applications. The basic components
in HVACR systems include heat exchangers, pumps, compressors, throttles,
and turbines, and these are introduced, classified, and thermodynamically
analyzed. This chapter also describes the energy and exergy assessments of psy-
chrometric processes. Mass, energy, entropy, and exergy balances for all com-
ponents and processes are provided.
In this chapter, kinetic and potential energy changes are considered to be neg-
ligible and all processes are assumed to be steady-flow and steady-state. Of
course, transient processes can be assessed if required.
For a proposed integrated system involving psychrometric processes, thermo-
dynamic analyses are performed. The energy and exergy efficiencies for individ-
ual components and the integrated system are calculated and parametric
studies are performed that determine the impact on system performance of
varying dead-state properties and system operating conditions. 43

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00002-8


© 2015 Elsevier Inc. All rights reserved.
44 C HA PT E R 2 : Energy and Exergy Assessments

2.2 HEAT EXCHANGERS (HEATING/COOLING)


Closed heat exchangers (see Fig. 2.1) transfer heat from one fluid to another
without the fluids coming in direct contact with each other. Heat transfer in
a heat exchanger can occur without the fluid undergoing phase change or with
phase change (e.g., from a liquid to a vapor, as in an evaporator, or from a vapor
to a liquid, as in a condenser). The transfer of heat is driven by a temperature
difference. In most HVACR applications, heat exchangers are selected to trans-
fer either sensible or latent heat. Sensible heat applications involve heat transfer
that results in a temperature change without phase change. Latent heat transfer
involves a phase change of one of the liquids, for example, transferring heat to a
liquid by condensing steam.
Heat exchanger performance is commonly evaluated with one of two methods,
which are described in the next two subsections.

2.2.1 Log Mean Temperature Difference Method


One method of evaluating heat exchanger performance is the logarithmic mean
temperature difference method. When heat is exchanged between two fluids
flowing through a heat exchanger, the rate of heat transfer may be expressed as
Q ¼ UAΔtm (2.1)

where U is the overall heat transfer coefficient from fluid to fluid, A is the heat
transfer surface area of the heat exchanger associated with U, and △tm is the log
mean temperature difference (LMTD or △tm).
For a heat exchanger with a constant U, the LMTD can be calculated as
ðT1  T4 Þ  ðT2  T3 Þ
Δtm ¼ Cf (2.2)
ln ðT1  T4 Þ=ðT2  T3 Þ

where Cf is a correction factor (less than 1.0) that is applied to heat exchanger
configurations that are not truly counterflow. Figure 2.1 illustrates a temperature
cross, where the outlet temperature of the heating fluid is less than the outlet
temperature of the fluid while heated (T2 < T4). A temperature cross is only

FIGURE 2.1 Closed heat exchanger.


2.2 Heat Exchangers (Heating/Cooling) 45

possible with a heat exchanger with a counterflow arrangement. The physical


arrangement of the surface area affects the overall coefficient UA. Not every heat
exchanger with identical surface area carry out equally for a given load. Hence-
forth, for specific applications, defining load conditions when selecting a heat
exchanger is critical.
The load for each fluid stream can be calculated as

Q_ ¼ mc
_ p ðTin  Tout Þ (2.3)

value of △tm is an significant factor in selection of heat exchanger. For a


given load, if △tm has a high value, a comparatively minor heat exchanger sur-
face area is necessary. The commercial effect is that the design of the heat
exchanger must accommodate the forces and actions convoying with huge
difference in temperatures. When the approach temperature is small i.e. the
change in T2 and T4 is minor, △tm is also insignificant and a fairly large A is
obligatory.

2.2.2 ε-NTU (Effectiveness Analysis)


An substitute method of assessing heat exchanger performance includes the
calculation of exchanger heat transfer effectiveness ε and number of exch-
anger transfer units (NTU). This method is grounded on the identical assump-
tions as the logarithmic mean temperature difference technique designated
earlier.
Equations (2.1) and (2.2) for △tm are more conveniently applied when inlet
and outlet temperatures are known for both fluids. Though during most times,
the temperatures of fluids leaving the heat exchanger are unidentified. To avoid
trial-and-error calculations, the ε-NTU method uses three dimensionless param-
eters: effectiveness ε, number of transfer units (NTU), and capacity rate ratio cr.
The mean temperature difference in Eq. (2.2) is not needed.
Heat exchanger effectiveness ε is the ratio of actual heat transfer rate to the max-
imum possible heat transfer rate in a counterflow heat exchanger of infinite sur-
face area with the same mass flow rates and inlet temperatures. The maximum
possible heat transfer rate for hot fluid entering at Thi and cold fluid entering at
Tci is
qmax ¼ Cmin ðThi  Tci Þ (2.4)

where Cmin is the smaller of the hot and cold fluid capacity rates in W/K.
The actual heat transfer rate can be written as
q ¼ εqmax (2.5)
46 C HA PT E R 2 : Energy and Exergy Assessments

Heat exchanger effectiveness can generally be expressed as a function of the


NTU and the capacity ratio cr:
ε ¼ 1  expðNTUÞ (2.6)

The mean temperature difference in then given by


ðThi  Tci Þε
Δtm ¼ (2.7)
NTU

After finding the heat transfer rate q, the exit temperature for constant-density
fluids can be found from
q
jTe  Ti j ¼ (2.8)
_ p
mc

The mass, energy, entropy, and exergy rate balance equations can be written for
a closed heat exchanger.
Mass rate balance
For the hot fluid : m_ h, i ¼ m_ h, o ¼ m_ h (2.9)

For cold fluid : m_ c, i ¼ m_ c, o ¼ m_ c (2.10)

Energy rate balance


m_ h, i h1 + m_ c, i h3 ¼ m_ h, o h2 + m_ c, o h4 + Q_ surr (2.11)

If we consider each flow independently, we have

Q_ c ¼ m_ c ðh4  h4 Þ (2.12)

Q_ h ¼ m_ h ðh2  h1 Þ (2.13)

Note that Q_ c is positive while Q_ h is negative, because the cold fluid gains heat
and the hot fluid loses it. The energy loss rate to the surroundings Q_ surr is
therefore

Q_ surr ¼ Q_ c + Q_ h (2.14)

Entropy rate balance


Q_ surr
m_ h, i s1 + m_ c, i s3 + S_gen ¼ m_ h, o s2 + m_ c, o s4 + (2.15)
T

Exergy rate balance  


_ D + Q_ surr 1  To
m_ h, i ex1 + m_ c, i ex3 ¼ m_ h, o ex2 + m_ c, o ex4 + Ex (2.16)
T

2.2.3 Efficiencies
Heat exchangers are a significant component in many engineering systems.
However, various methods are used to evaluate their thermodynamic
2.2 Heat Exchangers (Heating/Cooling) 47

efficiencies, and sometimes, these do not reflect the quality of the heat
exchanger from an energy point of view.
For example, Mikheev (1956) suggested that the efficiency of a heat exchanger η
is the fraction q1 of the available heat Qavail from a hot fluid that is used to pro-
vide heat Q1 to a cold one:
m_ 2 ðh3  h4 Þ
η¼ (2.17)
m_ 1 ðh1  h0 Þ

where h0 is the specific enthalpy of the hot fluid at the ambient temperature, ṁ1
and ṁ2 are the mass flow rates of the hot and cold fluids, h1 is the specific
enthalpy of the hot fluid at the inlet to the heat exchanger, and h3 and h4
are the specific enthalpies of the cold fluid at the inlet and outlet.
Andreev and Kostenko (1965) examined the exergy efficiency of a heat
exchanger, defined as the ratio of the exergy changes of cold and hot flows
through the exchanger:
_ 3  Ex
Ex _ 4
ηex ¼ (2.18)
_ 1  Ex
Ex _ 2

where Ėx3 and Ėx4 are the exergy flow rates of the cold fluid and Ėx1 and Ėx2 are
the exergy flow rates of the hot fluid at the inlet and outlet, respectively.
We can express the exergy efficiency using the exergy rate balance:
X
_ out
Ex _ D
Ex _ D
Ex
ηex ¼ X ¼1X ¼1 (2.19)
_
Exin Exin _
m h, i 1 + m
ex _ c, i ex3

2.2.4 Illustrative Example


In large steam power plants, feedwater is frequently heated in closed feedwater
heaters by steam extracted from the turbine at some stage. Steam enters the
feedwater heater at 1 MPa and 200 °C and leaves as saturated liquid at the same
pressure, as shown in Fig. 2.2. Feedwater enters the heater at 2.5 MPa and 50 °C
and leaves 10 °C below the exit temperature of the steam.

T0
Steam from
turbine
1.6 MPa 1 2
250 °C Sat. liquid

4 3
Feedwater
4 MPa
30 °C
FIGURE 2.2 Feedwater heater.
48 C HA PT E R 2 : Energy and Exergy Assessments

The following assumptions are used in the analysis:

• Heat loss from the device to the surroundings is negligible, and thus, heat
transfer from the hot fluid is equal to the heat transfer to the cold fluid.
• The surrounding temperature is at 25 °C.

2.2.4.1 Results and Discussion


Based on the data provided and assumptions, details about the thermodynamic
properties at each state point in the feedwater heater are determined (see
Table 2.1). Also, the following parameters are calculated using EES software:

ηex ¼ 0:5632 ηen ¼ 1 S_gen ¼ 0:3755kJ=kgK Ex


_ d ¼ 111:9kW

2.2.4.2 Parametric Study


The effect of varying selected parameters (ambient temperature and inlet
feedwater temperature) on feedwater heater efficiencies is investigated. Figure 2.3
demonstrates that ambient temperature variations do not affect heat exchanger

Table 2.1 Thermodynamic Properties at Each State Point in the Feedwater Heater
State No. P (kPa) T (K) h (kJ/kg) s (kJ/kg K) ex (kJ/kg) m_ (kg/s)

1 1600 523 2918 6.671 934.5 1


2 1600 474.6 858.7 2.344 164.7 3.005
3 4000 303 128.7 0.4332 4.081
4 1600 464.6 814 2.249 148.4

1 1

0.9 0.9

0.8 0.8
hen

hex

0.7 0.7

0.6 0.6

0.5 0.5
hen
hex
0.4 0.4
273 283 293 303 313 323
To (K)

FIGURE 2.3 Effect of ambient temperature on feedwater heater energy and exergy efficiencies.
2.3 Pumps 49

1 1

0.9 0.9

0.8 0.8
hen

hex
0.7 0.7

0.6 0.6
hen
hex
0.5 0.5
373 393 413 433 453 473
T1 (K)

FIGURE 2.4 Effect of inlet steam temperature on feedwater heater energy and exergy efficiencies.

energy efficiency, but that exergy efficiency decreases with increasing ambient tem-
perature. Figure 2.4 shows that energy efficiency also does not change as inlet feed-
water temperature varies and that exergy efficiency decreases with increasing inlet
steam temperature.

2.3 PUMPS
A pump transfers mechanical energy to a fluid by raising its pressure. It is a
device that moves fluids (liquids or gases) by mechanical action. Pumps can
be classified into three major groups according to the method they use to move
the fluid: direct lift, displacement, and gravity.
The mass, energy, entropy, and exergy rate balances for an adiabatic pump can
be written as follows:
Mass balance : m_ in ¼ m_ out (2.20)

Energy balance : m_ in hin + W_ in ¼ m_ out hout (2.21)

Entropy balance : m_ in sin + S_gen ¼ m_ out sout (2.22)

Exergy balance : m_ in exin + W_ in ¼ m_ out exout + Ex


_ d (2.23)

2.3.1 Energy Efficiency


Pump energy efficiency is defined as the ratio of the mechanical energy increase of
the fluid as it flows through the pump to the mechanical energy input to the pump:
50 C HA PT E R 2 : Energy and Exergy Assessments

m_ in ðhout  hin Þ
ηen ¼ (2.24)
W_ in

2.3.2 Exergy Efficiency


The exergy efficiency for a pump is defined as the ratio of the mechanical exergy
increase of the fluid as it flows through the pump to the mechanical energy
input to the pump:
m_ in ðexout  exin Þ
ηex ¼ (2.25)
W_ in

2.3.3 Illustrative Example


Water enters a pump at 100 kPa and 30 °C at a rate of 1.35 kg/s and leaves at
4 MPa, as shown in Fig. 2.5. The pump has an isentropic efficiency of 70%. The
following parameters are calculated: (a) the actual power input, (b) the rate of
frictional heating, (c) the exergy destruction, and (d) the exergy efficiency for an
environment temperature of 20 °C, respectively.

2.3.3.1 Results and Discussion


For the given data, the following parameters are calculated using EES software:

ηs ¼ 0:7 ηen ¼ 1 Q_ loss ¼ 2:266kW W_ rev ¼ 5:224kW


Here, ηex ¼ 0:6944 _ d ¼ 2:308kW
Ex W_ a ¼ 7:552kW W_ s ¼ 5:286kW

Note that Q_ loss is the frictional heat loss rate for the pump, which is the differ-
ence between the actual and the reversible work rates.

2.3.3.2 Parametric Study


The effect of varying selected parameters (ambient and inlet temperatures) on
pump efficiencies is investigated. Figure 2.6 demonstrates that the energy effi-
ciency and the isentropic efficiency of the pump do not change with ambient
temperature, while the exergy efficiency decreases with increasing ambient tem-
perature. Figure 2.7 shows the effect of inlet temperature on efficiencies. It can be
seen that, while energy efficiency and isentropic efficiency of the pump do not
change, exergy efficiency increases as the pump inlet temperature with increases.

1 2
Water
100 kPa
30°C
1.35 kg/s 4 MPa
FIGURE 2.5 Pump.
2.4 Compressors 51

1 1
hen hs hex

0.9 0.9

hex, hs
hen

0.8 0.8

0.7 0.7

0.6 0.6
273 280 287 294 301 308 315 322
T0 (K)

FIGURE 2.6 Effect of ambient temperature on pump isentropic, energy, and exergy efficiencies.

1 1
hen hex hs

0.9 0.9 hex, hs


hen

0.8 0.8

0.7 0.7

0.6 0.6
274 281 288 295 302 309 316 323
T1 (K)

FIGURE 2.7 Effect of inlet temperature on pump isentropic, energy, and exergy efficiencies.

2.4 COMPRESSORS
Compressors are commonly employed. For example, a compressor (Fig. 2.8) is
one of the four main components of the basic vapor compression refrigeration
system (the others are the condenser, the evaporator, and the expander). In that
application, the compressor circulates refrigerant through the system and
52 C HA PT E R 2 : Energy and Exergy Assessments

FIGURE 2.8 Compressor.

increases refrigerant vapor pressure to create a pressure difference between the


condenser and the evaporator.
The mass, energy, entropy, and exergy rate balances for a compressor can be
written as follows:
Mass balance : m_ in ¼ m_ out (2.26)

Energy balance : m_ in hin + W_ in ¼ m_ out hout (2.27)

Entropy balance : m_ in sin + S_gen ¼ m_ out sout (2.28)

Exergy balance : m_ in exin + W_ in ¼ m_ out exout + Ex


_ d (2.29)

2.4.1 Efficiencies
Compressor-related efficiencies can be grouped by category. There are multiple
efficiencies because deviations from ideal performance can be evaluated in various
ways, to measure different parameters. Various compressor efficiencies follow:
Volumetric efficiency (ηv) is the ratio of actual volumetric flow to the ideal vol-
umetric flow (i.e., the geometric compressor displacement).
Compression isentropic efficiency (ηs) considers only what occurs within the
compression volume and is a measure of the deviation of actual compression
from isentropic compression. This efficiency is defined as the ratio of work
required for isentropic compression of the gas (ws) to work delivered to the
gas within the compression volume (wa):
2.4 Compressors 53

ws
ηs ¼ (2.30)
wa

For a multicylinder or multistage compressor, this equation applies only for


each individual cylinder or stage.
Mechanical efficiency (ηm) is the ratio of work delivered to the gas to work input
to the compressor shaft (wm):
wa
ηm ¼ (2.31)
wm

Isentropic (reversible adiabatic) efficiency (ηi) is the ratio of work required


for isentropic compression of a gas (ws) to work input to the compressor
shaft (wm):
ws
ηi ¼ (2.32)
wm

Motor efficiency (ηe) is the ratio of work input to the compressor shaft (wm) to
work input to the motor (we):
wm
ηe ¼ (2.33)
we

Total compressor efficiency (ηcomp) is the ratio of work required for isentropic
compression (ws) to actual work input to the motor (we):
ws
ηcomp ¼ (2.34)
we

The energy efficiency of a compressor can be defined as the ratio of the total
energy change of the fluid passing through the compressor to the net input
compressor work:
m_ ðhout  hin Þ
ηen ¼ (2.35)
W_ c

The exergy efficiency of a compressor can be expressed as the ratio of the revers-
ible work to the net input compressor work:

W_ rev
ηex ¼ (2.36)
W_ c

2.4.2 Illustrative Example


Air is compressed steadily by an 8 kW compressor from 100 kPa and 17 °C to
600 kPa and 167 °C at a rate of 2.1 kg/min, as shown in Fig. 2.9. The following
54 C HA PT E R 2 : Energy and Exergy Assessments

600 kPa
167 °C 2

.
Win = 8 kW

Air

100 kPa
1
17 °C

FIGURE 2.9 Compressor example.

parameters are calculated: (a) the increase in the exergy of the air and (b) the
rate exergy destroyed during this process. The surroundings are assumed to be
at 17 °C.

2.4.2.1 Results and Discussion


Based on the data provided and assumptions, the following parameters are
determined using EES software:

ηen ¼ 0:6629 ηex ¼ 0:7819 _ dest ¼ 1:745kW


Ex
ηs ¼ 0:7778 S_gen ¼ 0:006014kW=K W_ rev ¼ 6:255kW

2.4.2.2 Parametric Study


The effect of varying selected parameters (ambient and compressor outlet tem-
peratures) on compressor efficiencies is investigated. Figure 2.10 demonstrates
that the energy and isentropic efficiencies of the compressor do not vary with
changing ambient temperature but that the exergy efficiency increases as ambi-
ent temperature rises. Figure 2.11 shows the effect of varying compressor outlet
temperature on efficiencies. It can be seen that all the three efficiencies consid-
ered increase with increasing compressor outlet temperature.

2.5 FANS
A fan creates a pressure difference and causes flow of a gas, often air. The
impeller does work on the gas, imparting to it both static and kinetic
2.5 Fans 55

0.8 0.8

0.76 0.76

0.72 0.72
hen, hs

hex
0.68 0.68

0.64 0.64

hen hex hs
0.6 0.6
0 10 20 30 40 50
T0 (C)

FIGURE 2.10 Effect of ambient temperature on compressor isentropic, energy, and exergy efficiencies.

1 1
hen
hex
0.9 0.9
hs

0.8 0.8
hen, hs

hex

0.7 0.7

0.6 0.6

0.5 0.5
150 160 170 180 190 200
T2 (C)

FIGURE 2.11 Effect of outlet temperature on compressor isentropic, energy, and exergy efficiencies.

energies, in different proportions depending on the fan type. Fan efficiency


ratings are usually based on ideal conditions, with some fans having ratings
of more than 90% total efficiency. However, actual connections normally
cause the actual efficiencies in the field to be lower than the corresponding
ideal efficiencies.
56 C HA PT E R 2 : Energy and Exergy Assessments

The mass, energy, entropy, and exergy rate balances for a fan are as follows:
Mass balance : m_ in ¼ m_ out (2.37)
2
v
Energy balance : m_ in hin + W_ in ¼ m_ out hout + m_ out exit (2.38)
2
Entropy balance : m_ in sin + S_gen ¼ m_ out sout (2.39)

v2exit
Exergy balance : m_ in exin + W_ in ¼ m_ out exout + Ex
_ d + m_ out (2.40)
2

2.5.1 Efficiencies
The energy (and exergy) efficiency of a fan is the ratio of the kinetic energy flow
rate of the gas at the fan exit to the mechanical power input:
v2exit
ΔE_mech m_ 2
η¼ ¼ (2.41)
W_ in W_ in

2.5.2 Illustrative Example


Figure 2.12 shows a fan increasing the air velocity from 0 to 12 m/s with a work
input rate of 50 kW at constant pressure and mass flow rate. Using Eq. (2.41),
the efficiency of the fan is calculated as
v2exit
ΔE_mech m_ 2
η¼ ¼ ¼ 72%
W_ in W_ in

2.6 THROTTLING VALVES


In a throttling (or expansion) valve, the pressure of the fluid passing through
the valve is reduced suddenly, and the temperature decreases simultaneously.

FIGURE 2.12 Fan.


2.6 Throttling Valves 57

The pressure reduction when a fluid passes through the orifice (i.e., small open-
ing) of a throttling valve is due to sudden expansion into a larger space and
frictional effects. A throttling valve is another common component in refriger-
ation, air conditioning, and other systems.
In refrigeration applications, for instance, the refrigerant exiting the compres-
sor at high pressure and temperature passes through the condenser and leaves
at medium temperature and high pressure and passes through a throttling
valve. There, the pressure and the temperature of the refrigerant drop suffi-
ciently so the refrigerant is then able to produce the cooling effect in the evap-
orator of the refrigerator. A similar operation is observed for the cooling coil
of an air conditioner. The throttling valve also controls the amount of the
refrigerant entering the evaporator, which depends on the refrigeration load.
Specifically, the refrigerant flow rate through the throttling valve depends on
cross-sectional area of the orifice and the pressure difference across the
throttling valve.

2.6.1 Functions Performed by Throttling Devices


in Refrigeration Systems
When the high-pressure refrigerant from the condenser enters the throttling
valve in a refrigeration system, the pressure of the refrigerant decreases sud-
denly, and due to this, the temperature of the refrigerant drops significantly.
The two main functions performed by the throttling (or expansion) valve are
to reduce refrigerant pressure and meeting the refrigerant load.
The mass, energy, entropy, and exergy rate balances for a throttling valve can be
written as follows:
Mass balance : m_ in ¼ m_ out (2.42)

Energy balance : m_ in hin ¼ m_ in hin (2.43)

Entropy balance : m_ in sin + S_gen ¼ m_ out sout (2.44)

_ d
Exergy balance : m_ in exin ¼ m_ out exout + Ex (2.45)

2.6.2 Types of Throttling Devices


Some of the most commonly used types of throttling valves are listed as follows:

1. Capillary tube
2. Constant pressure or automatic throttling valve
3. Thermostatic expansion valve
4. Float valve
58 C HA PT E R 2 : Energy and Exergy Assessments

2.6.3 Throttle Efficiencies


Efficiencies are not straightforwardly defined for throttling devices as they are
dissipative in nature. Nonetheless, they are sometimes defined for convenience
or to allow component comparisons.
The energy efficiency of a throttle valve can be defined as the ratio of the total
output energy to the total input energy:

m_ ðhout Þ
ηen ¼ (2.46)
m_ ðhin Þ

Correspondingly, the exergy efficiency can be defined as the ratio of the useful
exergy output to the total exergy input:

m_ ðexout Þ Exd
ηex ¼ ¼1 (2.47)
m_ ðexin Þ m_ ðexin Þ

For an adiabatic throttling operation, which is often the case, the energy effi-
ciency is often 100%. The exergy efficiency is lower than 100% due to irrevers-
ibilities associated with unconstrained expansion.

2.6.4 Illustrative Example


R-134a at 1 MPa and 100 °C is throttled to a pressure of 0.8 MPa, as shown in
Fig. 2.13. The reversible work and exergy destroyed during this throttling pro-
cess are determined. The surroundings are assumed to be at 30 °C.

2.6.4.1 Results and Discussion


Based on the data provided and assumptions, the following parameters are cal-
culated for the throttling valve using EES software:

ηen ¼ 1 S_gen ¼ 0:03176kW=K Ex


_ d ¼ 9:464kW

2.6.4.2 Parametric Study


The effect of varying selected parameters (ambient temperature and inlet pres-
sure) on throttling valve efficiencies is investigated. Figure 2.14 shows that

1 2
T1 T1 = T2
h1 h1 = h2
FIGURE 2.13 Throttling valve.
2.6 Throttling Valves 59

1 1
hen hex

0.98 0.98

0.96 0.96
hen

hex
0.94 0.94

0.92 0.92

0.9 0.9
273 283 293 303 313 323
T0 (K)

FIGURE 2.14 Effect of ambient temperature on throttling valve energy and exergy efficiencies.

1 1
hen hex

0.96 0.96

0.92 0.92
hen

hex

0.88 0.88

0.84 0.84

0.8 0.8
1000 2000 3000 4000
P1 (kPa)

FIGURE 2.15 Effect of inlet pressure on throttling valve energy and exergy efficiencies.

ambient temperature has almost no effect on the energy efficiency of a throt-


tling valve but that increasing the ambient temperature decreases the exergy
efficiency. Figure 2.15 demonstrates that varying the throttling valve inlet pres-
sure does not affect the energy efficiency, while increasing the throttling valve
inlet pressure decreases the device’s exergy efficiency.
60 C HA PT E R 2 : Energy and Exergy Assessments

2.7 TURBINES
A turbine is a device that extracts energy from a pressurized fluid as it expands
and yields mechanical work, often in the form of a rotating shaft. Turbines are
used in many systems for electrical power generation (Fig. 2.16).
In an ideal Rankine cycle, for instance, superheated vapor from a boiler
enters a vapor turbine at high temperature and pressure. The vapor passes
through a nozzle, where some of its pressure is converted into kinetic energy
as the velocity increases. The high velocity vapor enters the turbine and
flows over the turbine blades. A force is created on the blades due to kinetic
energy of the vapor and the expansion of the vapor as it flows over the
blades, causing them to move. An electrical generator or another device
is attached to the shaft. The fluid often exits the turbine as a saturated vapor
at a reduced temperature and pressure and is conveyed to the condenser
where it is cooled.
The mass, energy, entropy, and exergy rate balance equations can be written for
a turbine as follows:
Mass balance : m_ in ¼ m_ out (2.48)

Energy balance : m_ in hin ¼ m_ in hin + W_ out (2.49)

Entropy balance : m_ in sin + S_gen ¼ m_ out sout (2.50)

_ d + W_ out
Exergy balance : m_ in exin ¼ m_ out exout + Ex (2.51)

FIGURE 2.16 Turbine.


2.7 Turbines 61

2.7.1 Turbine Efficiencies


The energy efficiency of a turbine can be defined as the ratio of the work output
to the total change in energy between states 1 and 2:

W_ out
ηen ¼ (2.52)
m_ ðhin  hout Þ

The isentropic efficiency of a turbine is defined as ratio of the work output from
the turbine to the work output if the process were isentropic:
h1  h2 T1  T2
ηs ¼ ¼ (2.53)
h1  h2s T1  T2s

The exergy efficiency of a compressor can be defined in several ways, one being
the ratio of work output to the reversible work:

W_ out W_ out
ηex ¼ ¼ (2.54)
W_ rev m_ ðexin  exout Þ

2.7.2 Illustrative Example


Air expands in an adiabatic turbine from one specified state to another, as
shown in Fig. 2.17. The exergy efficiency is to be determined.
Assumptions:

1. The working fluid, air, behaves like an ideal gas with constant
specific heats.
2. Kinetic and potential energy changes are negligible.

FIGURE 2.17 Turbine example.


62 C HA PT E R 2 : Energy and Exergy Assessments

2.7.2.1 Results and Discussion


Based on the data provided and assumptions, the following parameters are
determined using EES software:

ηs ¼ 0:7584 ηen ¼ 1 W_ a ¼ 101:1kW


ηex ¼ 0:6372 _ d ¼ 56:95kW
Ex W_ rev ¼ 158:7kW

2.7.2.2 Parametric Study


The effect of varying selected parameters (ambient and inlet temperatures) on
turbine efficiencies is investigated. Figure 2.18 shows that ambient temperature
has no effect on the energy and isentropic efficiencies but that increasing the
ambient temperature decreases slightly the exergy efficiency of a turbine.
Figure 2.19 demonstrates that varying the turbine inlet temperature does not
affect the energy efficiency. But raising the turbine inlet temperature increases
the isentropic and exergy efficiencies of the turbine.

2.8 ENERGY AND EXERGY ASSESSMENTS OF


PSYCHROMETRIC PROCESSES
In this section, we consider several psychrometric processes that are commonly
encountered in HVAC systems, including the following:

• Sensible cooling
• Sensible heating

1 1
hen hex hs

0.9 0.9
hen, hs

hex

0.8 0.8

0.7 0.7

0.6 0.6
273 283 293 303 313 323
T0 (K)

FIGURE 2.18 Effect of ambient temperature on turbine isentropic, energy, and exergy efficiencies.
2.8 Energy and Exergy Assessments of Psychrometric Processes 63

1 1
hen hex hs

0.9 0.9
hen, hs

hex
0.8 0.8

0.7 0.7

0.6 0.6
430 440 450 460 470
T1 (K)

FIGURE 2.19 Effect of inlet temperature on turbine isentropic, energy, and exergy efficiencies.

• Heating with humidification


• Cooling with dehumidification
• Evaporative cooling
• Adiabatic mixing of air streams

All of these can be treated as steady-state, steady-flow processes. For computa-


tional purposes in the examples, several air and water vapor properties are taken
to be constant. Values for these properties are listed in Table 2.2, and the values
of the reference environment (i.e., dead-state) properties are given in Table 2.3.

Table 2.2 Material Properties


cp,a 1.004 kJ/kg K
cp,v 1.872 kJ/kg K
Ra 0.287 kJ/kg K
Rv 0.4615 kJ/kg K
Tav 323.15 K
(x)3 0.055
(xv)0 0.024

Source: Dincer and Rosen (2013).

Table 2.3 Dead-State Properties


T0 298 K
P0 101.325 kPa
ϕ0 0.7
64 C HA PT E R 2 : Energy and Exergy Assessments

FIGURE 2.20 Sensible heating and cooling processes: schematic (left) and representation on
psychrometric chart (right).

The rate balances for various commodities can be written in a general form for
common air conditioning processes (see Fig. 2.20) in which one or more moist
air flows enter and exit. These are given below:
X X
Mass balance f or dry air : m_ a ¼ m_ a (2.55)
in out
X X X X
Mass balance f or water : m_ w ¼ m_ w or m_ a ω ¼ m_ a ω or (2.56)
in out in out

where
m_ w ¼ m_ a ðωout  ωin Þ (2.57)
X X
Energy balance : Q_ in + _ ¼ Q_ out +
mh _
mh (2.58)
in out

Entropy balance : S_in  S_out + S_gen ¼ 0 (2.59)


X X X X
or S_Q_ + _ 
ms S_Q_  _ + S_gen ¼ 0
ms (2.60)
in in out out

X Q_ X X Q_ X
or + _ 
ms  _ + S_gen ¼ 0
ms (2.61)
in
T in out
T out
X X X X
Exergy balance : _ _+
Ex _ 
mex _ _
Ex mex _ dest ¼ 0
_ + Ex (2.62)
Q Q
in in out out

X  T0
 X X  T0
 X
or Q_ 1  + _ 
mex Q_ 1   mex _ dest ¼ 0
_ + Ex (2.63)
in
T in in
T out

Here, the specific flow exergy for a stream is given by


ex ¼ h  h0  T0 ðs  s0 Þ (2.64)
2.9 Sensible Cooling (ω1 ¼ω2 ) 65

The exergy destruction rate is directly proportional to the entropy generation


rate due to irreversibilities:
_ dest ¼ T0 S_gen
Ex (2.65)

When considering dry air and water vapor as an ideal gas, the flow exergy for a
stream can be defined as
 
  T T P
ex ¼ cp, a + ωcp, v T0  1  ln e ÞRa T0 ln
+ ð1 + ω
T0 T0 P0
 
1+ωe e
ω
e Þ ln
+ Ra T0 ð1 + ω e ln
+ω (2.66)
e
1+ ω e0
ω
  
1+ωe e
ω
e Þ ln
Note that the last term Ra T0 ð1 + ω e ln
+ω is the specific chem-
e
1+ ω e0
ω
ical exergy. Here,

ω
e ¼ 1:608ω (2.67)

The humidity ratio ω is defined as

ω ¼ mv =ma (2.68)

For any process or system, the exergy efficiency is defined as the ratio of exergy
of the products to the input exergy as follows:

_ products
Ex _ dest
Ex
ηex ¼ ¼1 (2.69)
_ in
Ex _ in
Ex

2.9 SENSIBLE COOLING (ω15ω2)


Cooling of air is one a common psychrometric process in air conditioning sys-
tems. The basic function of the air conditioners is to cool the air in the room,
which often includes some air from the atmosphere, and sometimes to dehumid-
ify the air. Sensible cooling of air is the process in which only the sensible heat of
the air is removed so as to reduce its temperature, and there is no change in the
moisture content of the air (as measured by the specific or absolute humidity, in
units of kg/kg of dry air). During a sensible cooling process, the dry bulb (DB)
temperature and wet bulb (WB) temperature of the air are reduced, while the
absolute humidity and the dew point (DP) temperature of the air remain con-
stant. There is an overall reduction in the specific enthalpy of the air.
In an ordinary window or the air conditioner, sensible cooling of air is carried
out by passing it over an evaporator coil, also called the cooling coil. The coil
66 C HA PT E R 2 : Energy and Exergy Assessments

carries the refrigerant at a relatively low temperature. The lower-temperature air


passes through the space, which is to be maintained at desired conditions, usu-
ally for human comfort.
In central air conditioners, the cooling coils are cooled by the chilled water,
which is chilled by flowing over the evaporator of the refrigeration unit accom-
panying large air conditioning systems. In certain cases, the cooling coil uses a
gas as a working fluid that passes through it.
The sensible cooling process is represented by a straight horizontal line on a
psychrometric chart. The line starts from the initial DB temperature of the
air and ends at the final DB temperature, at the left. The cooling line is also
the constant DP temperature line since the moisture content of the air remains
constant. All properties of the moist air, including the initial and final points,
can be read off of the psychrometric chart.
The mass, energy, entropy, and exergy rate balance equations can be written for
a sensible cooling process:
Dry air mass balance : m_ a1 ¼ m_ a2 (2.70)

Water mass balance : m_ w1 ¼ m_ w2 (2.71)

Energy balance : m_ a1 h1 ¼ m_ a2 h2 + Q_ c (2.72)

Q_
Entropy balance : m_ a1 s1 + S_gen, c ¼ m_ a2 s2 + c (2.73)
Ts
 
T0
Exergy balance : m_ a1 ex1 ¼ m_ a2 ex2 + Q_ c 1  _ dest, c
+ Ex (2.74)
Ts
 _ 
_ dest, c ¼ T0 S_gen, c ¼ T0 m_ a2 s2  m_ a1 s1 + Qc
Ex (2.75)
Ts

2.9.1 Efficiencies
Various efficiencies can be defined for sensible cooling. Ratlamwala and Dincer
(2012) performed a comprehensive study to investigate all possible efficiency
options for psychrometric processes and comparatively assessed them for some
selected applications. The first defines the efficiency as the ratio of change in
energy/exergy of the air to the energy/exergy input to the system. In this defi-
nition, change in energy/exergy means the difference between the energy/
exergy of the stream entering the system and the energy/exergy of the stream
exiting the process. Moreover, this definition presumes that the energy/exergy
output for this process is heat rejected from the system. Then, we can write

m_ a2 h2 + Q_ c
ηen, 1 ¼ (2.76)
m_ a1 h1
2.9 Sensible Cooling (ω1 ¼ω2 ) 67

 
T0
m_ a2 ex2 + Q_ c 1 
Ts
ηex, 1 ¼ (2.77)
m_ a1 ex1

Another efficiency definition is based on the concept that the required output
of the system is the energy/exergy of the stream leaving the system and that the
required input to the system is the amount of heat rejected from the system to
attain the desired output. In this case,
m_ a2 h2
ηen, 2 ¼ (2.78)
Q_ c

m_ ex
ηex, 2 ¼  a2 2  (2.79)
T0
Q_ c 1 
Ts

A third efficiency definition defines the efficiency of the process as the energy/
exergy of output stream plus the amount of heat rejected from the process,
divided by the energy/exergy of the input stream. For this case,

m_ a2 h2 + m_ w2 hw2 + Q_ c
ηen, 3 ¼ (2.80)
m_ a1 h1
 
T0
m_ a2 ex2 + m_ w2 exw2 + Q_ c 1 
Ts
ηex, 3 ¼ (2.81)
m_ a1 ex1

2.9.2 Illustrative Example


Humid air at 1 atm, 50 °C DB, and 80% relative humidity is sensibly cooled at
constant pressure to 32 °C DB and 100% relative humidity. The energy and
exergy efficiencies of the process are to be determined.
The following assumptions are made:

• Steady-flow, steady-state process.


• Dry air and water vapor behave as ideal gases.
• Kinetic and potential energy changes are negligible.

2.9.2.1 Results and Discussion


Based on the data provided and assumptions, the following parameters (shown
in Table 2.4) are determined using EES software:
The rate of exergy destroyed, the rate of cooling, and the rate of entropy gener-
ation are calculated as follows:
_ dest ¼ 1:478kW Q_ c ¼ 70:06kW S_gen ¼ 0:00496kJ=kgK
Ex
68 C HA PT E R 2 : Energy and Exergy Assessments

Table 2.4 Thermodynamic Properties at all State Points in a Sensible Cooling Process
State Point P (kPa) T (K) h (kJ/kg) RH s (kJ/kg K) ex (kJ/kg) ω m_ (kg/s)

1 101.3 323 223.1 0.8 6.350 5.162 0.066 0.618


2 101.3 305 109.8 1.0 5.987 0.174 0.066 0.0040
3 101.3 323 2546 – 8.559 – – –
4 101.3 – 2559 – 8.414 56.000 – –

FIGURE 2.21 Sensible cooling example.

Using the efficiency expressions in Eqs. (2.76)–(2.81), the energy and exergy
efficiencies for sensible cooling in the example (Fig. 2.21) become as follows:
ηen, 1 ¼ 1 ηen, 2 ¼ 0:97 ηen, 3 ¼ 0:567 ηex, 1 ¼ 0:537 ηex, 2 ¼ 0:067 ηex, 2 ¼ 0:105

2.9.2.2 Parametric Studies


The effect of varying selected parameters (ambient temperature and relative
humidity) on the various efficiencies for sensible cooling discussed earlier is
investigated. Figure 2.22 demonstrates that ambient temperature has no effect
on the various energy efficiencies considered, while increasing the ambient
temperature decreases the various exergy efficiencies considered, for a sensible
cooling process. Figure 2.23 demonstrates that ambient relative humidity also
has no effect on the energy efficiencies but that increasing the ambient relative
humidity increases the three exergy efficiencies for a sensible cooling process.

2.10 SENSIBLE HEATING (ω15ω2)


Sensible heating is essentially the opposite of sensible cooling. In a sensible
heating process, the temperature of air is increased without changing its mois-
ture content. During this process, the DB and WB temperatures of the air
increase while the absolute humidity and DP point temperature of the air
remain constant.
In general, heating of air is often carried out by passing it over a heating coil.
This coil may be heated by passing a heated fluid (e.g., hot water or steam)
through it or by using an electric resistance heating coil. Hot water and steam
are often used as heating media in industrial applications.
2.10 Sensible Heating (ω1 ¼ω2 ) 69

1 1

0.8 0.8

0.6 0.6
hen

hex
0.4 0.4

hen,1 hex,1
0.2 0.2
hen,2 hex,2
hen,3 hex,3
0 0
275 280 285 290 295 300
T0 (K)

FIGURE 2.22 Effect of ambient temperature on sensible cooling of various energy and exergy efficiencies.

1 1

hen,1 hex,1

0.8 hen,2 hex,2 0.8


hen,3 hex,3

0.6 0.6
hen

hex

0.4 0.4

0.2 0.2

0 0
0.5 0.6 0.7 0.8 0.9 1
RH0

FIGURE 2.23 Effect of ambient relative humidity on sensible cooling of various energy and
exergy efficiencies.

Like sensible cooling, the sensible heating process is represented by a straight


horizontal line on the psychrometric chart, as shown in Fig. 2.20. The line starts
from the initial DB temperature of air and ends at the final temperature,
towards the right. The sensible heating line is also the constant DP
temperature line.
70 C HA PT E R 2 : Energy and Exergy Assessments

Note that heating of air is also important in heat pumps that provide space
heating. In a heat pump, the air is heated by passing it over a condenser coil
(or the heating coil) that contains a high-temperature working fluid (often a
refrigerant). In some cases, the heating of air is also done to suit different indus-
trial and comfort air conditioning applications where large air conditioning
systems are used.

2.10.1 Rate Balance Equations


The mass, energy, entropy, and exergy rate balance equations can be written for
a sensible heating process:
Dry air mass balance : m_ a1 ¼ m_ a2 (2.82)

Water maas balance : m_ w1 ¼ m_ w2 (2.83)

Energy balance : Q_ h + m_ a1 h1 ¼ m_ a2 h2 (2.84)

Q_ h _
Entropy balance : m_ a1 s1 + + Sgen, h ¼ m_ a2 s2 (2.85)
T0
 
T0
Exergy balance : m_ a1 ex1 + Q_ h 1  _ dest, h
¼ m_ a2 ex2 + Ex (2.86)
Ts
 
Q_
_ dest,
Ex h ¼ T0 S_gen, h ¼ T0 m_ a2 s2  m_ a1 s1  h (2.87)
T0

2.10.2 Efficiencies
Various efficiencies can be defined for sensible heating. The first defines the effi-
ciency as the ratio of the change in energy/exergy of the air to the energy/exergy
input to the process. In this definition, change in energy/exergy denotes the dif-
ference between the energy/exergy of the stream entering and exiting the pro-
cess. Moreover, this definition states that energy/exergy input to this system is
heat provided to the process to drive the heating operation. That is,
 
m_ a2 h2  m_ a1 h1
ηen, 1 ¼ (2.88)
Q_ h
m_ a2 ex2  m_ a1 ex1
ηex, 1 ¼   (2.89)
T0 _
1 Qh
T

The second definition of efficiency is based on the concept that the required
output of the system is the energy/exergy of the stream leaving the process
and that the required input by the process is the amount of heat added to
the process. Then,
2.10 Sensible Heating (ω1 ¼ω2 ) 71

 
m_ a2 h2
ηen, 2 ¼ (2.90)
Q_ h
m_ a2 ex2
ηex, 2 ¼   (2.91)
T0 _
1 Qh
T

The third efficiency defines the efficiency of the system as the energy/exergy of
output stream divided by the amount of heat added to the process plus the
energy/exergy of the input stream. This definition states
 
m_ a2 h2
ηen, 3 ¼ (2.92)
Q_ h + m_ a1 h1
m_ a2 ex2
ηex, 3 ¼   (2.93)
T0
m_ a1 ex1 + Q_ h 1 
Ts

2.10.3 Illustrative Example


The saturated humid air at 15 °C DB is heated to a higher temperature. The
energy and exergy efficiencies of the process are to be determined.
The following assumptions are made:

• Steady-flow, steady-state process.


• Mass flow rate of dry air remains constant throughout the process.
• Dry air and water vapor behave like ideal gases.
• Kinetic and potential energy changes are negligible.

In the analysis, the amount of moisture in the air remains constant ðω1 ¼ ω2 Þ as
it flows through the heating section since the process involves no humidifica-
tion or dehumidification. The inlet state of the air is completely specified, and
the total pressure is 101.325 kPa. The properties of the air at the inlet and exit
states are determined by EES.

2.10.3.1 Results and Discussion


Based on the data provided and assumptions, the following parameters are
calculated using EES software. Table 2.5 gives details about the thermody-
namic properties at each state point in the heating with humidification
process.
The rate of exergy destruction, the rate of cooling, and the rate of entropy gen-
eration are calculated as follows:
_ dest ¼ 2:466kW Q_ h ¼ 59:91kW S_gen ¼ 0:008277kJ=kgK
Ex
72 C HA PT E R 2 : Energy and Exergy Assessments

Table 2.5 Thermodynamic Properties at Each State Point for Heating with Humidification Process
State Point P (kPa) T (K) h (kJ/kg) RH s (kJ/kg K) ex (kJ/kg) ω m_ (kg/s)

1 101.3 288 41.61 1 5.758 0.1122 0.0105 0.6188


2 101.3 313 138.4 0.8 6.081 0.7664 0.0105 0.004068
3 101.3 313 2546 – 8.559 – – –
4 101.3 – 2573 – 8.258 116.8 – –

FIGURE 2.24 Sensible heating example.

Using the efficiency expressions in Eqs. (2.88)–(2.93), the energy and exergy
efficiencies for sensible heating in the example (Fig. 2.24) are found to be as
follows:
ηen, 1 ¼ 1 ηen, 2 ¼ 1 ηen, 3 ¼ 0:699 ηex, 1 ¼ 0:161 ηex, 2 ¼ 0:141 ηex, 2 ¼ 0:165

2.10.3.2 Parametric Studies


The effect of varying selected parameters (ambient temperature and relative
humidity as well as outlet temperature) on the various efficiencies for sensible
heating discussed earlier is investigated. Figure 2.25 demonstrates that ambient
temperature has no effect on all the energy efficiencies and increasing the ambi-
ent temperature decreases the exergy efficiency of a sensible heating process.
Figure 2.26 demonstrates that ambient relative humidity also has no effect
on all the energy efficiencies and increasing the ambient relative humidity
increases the three exergy efficiencies of a sensible heating process. Figure 2.27
shows the effect of outlet temperature on sensible heating process efficiencies.
It is seen that increasing the outlet temperature increases the exergy efficiency
of the heating process.

2.11 HEATING WITH HUMIDIFICATION


In heating and humidification processes for air, the DB temperature and rela-
tive humidity of the air increase. Heating and humidification are carried out by
passing air over a spray of water, which is maintained at a temperature higher
than the DB temperature of air or by mixing air and steam.
2.11 Heating with Humidification 73

1 1

0.8 0.8

0.6 0.6
hen

hex
0.4 0.4

hen,1 hex,1
0.2 0.2
hen,2 hex,2
hen,3 hex,3

0 0
275 280 285 290 295 300
T0 (K)

FIGURE 2.25 Effect of ambient temperature on sensible heating energy and exergy efficiencies.

1 1
hen,1 hex,1
hen,2 hex,2
0.8 hen,3 hex,3 0.8

0.6 0.6
hen

hex

0.4 0.4

0.2 0.2

0 0
0.5 0.6 0.7 0.8 0.9 1
RH0

FIGURE 2.26 Effect of ambient relative humidity on sensible heating energy and exergy efficiencies.

When ordinary air is passed over a warm spray of water, moisture particles from
the spray evaporate partially and the vapor is added to the air, increasing its
moisture content. Also, since the temperature of the moisture is greater than
the DB temperature of the air, there is overall increase in temperature.
During heating and humidification processes, the DB, WB, and DP tempera-
tures of the air increase, along with its relative humidity. The heating and
74 C HA PT E R 2 : Energy and Exergy Assessments

1 1

0.8 0.8

0.6 0.6
hen,1 hex,1
hen

hex
hen,2 hex,2
0.4 0.4
hen,3 hex,3

0.2 0.2

0 0
310 315 320 325 330
T2 (K)

FIGURE 2.27 Effect of outlet temperature on sensible heating energy and exergy efficiencies.

humidification process is represented on the psychrometric chart by a line start-


ing from the initial condition and extending up and right.

2.11.1 Rate Balance Equations


The mass, energy, entropy, and exergy rate balance equations can be written for
heating with humidification (Fig. 2.28):
Dry air mass balance : m_ a1 ¼ m_ a2 ¼ m_ a3 (2.94)

Water mass balance : m_ w1 ¼ m_ w2 (2.95)

m_ w2 + m_ w ¼ m_ w3 ! m_ a2 ω2 + m_ w ¼ m_ a3 ω3

Energy balance : Q_ in, 12 + m_ a1 h1 ¼ m_ a2 h2 ðprocess 1  2Þ (2.96)


Humidity ratio

Sat. vapor
Heating
coils
T1
T3 3
RH1 Air P = 1 atm RH3
V1 1 2
1 2 3
Dry bulb temperature
FIGURE 2.28 Heating with humidification process: schematic (left) and representation on psychrometric chart (right).
2.11 Heating with Humidification 75

m_ a2 h2 + m_ w hw ¼ m_ a3 h3 ðprocess 2  3Þ (2.97)

Q_ in, 13 + m_ a1 h1 + m_ w hw ¼ m_ a3 h3 ðprocess 1  3Þ (2.98)

Q_ in, 13 _
Entropy balance : m_ a1 s1 + m_ w sw + + Sgen, 13 ¼ m_ a3 s3 ðprocess 1  3Þ (2.99)
T0
 
T0
Exergy balance : Q_ in, 12 1  _ dest, 12
+ m_ a1 ex1 ¼ m_ a2 ex2 + Ex (2.100)
Ts

_ dest, 23 ðprocess 2  3Þ


m_ a2 ex2 + m_ w exw ¼ m_ a3 ex3 + Ex (2.101)
 
T0
Q_ in, 13 1  _ dest, 13 ðprocess 1  3Þ
+ m_ a1 ex1 + m_ w exw ¼ m_ a3 ex3 + Ex (2.102)
Ts
!
Q_ in, 13
_ dest, 13 ¼ T0 S_gen, 13 ¼ T0 m_ a3 s3  m_ a1 s1  m_ w sw 
Ex ðprocess 1  3Þ
T0
(2.103)

2.11.2 Efficiencies
Various efficiencies can be defined for heating with humidification. The first is
based on taking the desired output of the system to be the amount of energy
gained by the system and the required input to the system to be the energies
added to the system via heat and hot water. The desired output is the difference
between the energy rates of the stream exiting and entering the system, while
the required input to the system is the sum of the heat rate added to the system
and the energy rate input via the hot water entering the system. In this case,
 
m_ a3 h3  m_ a1 h1
ηen, 1 ¼ (2.104)
Q_ in + m_ w hw
m_ a2 ex3
ηex, 1 ¼   (2.105)
T0
m_ a1 ex1 + m_ w exw + Q_ in, 13 1 
Ts

The second definition of the efficiency treats the desired output of the system as
the energy rate of the exiting stream and the required input as the energy rate
added to the system through heat and hot water. Then,
 
m_ a3 h3
ηen, 2 ¼ (2.106)
Q_ in + m_ w hw
0 1
B m_ a3 ðexÞ3  m_ a1 ðexÞ1 C
ηex, 2 ¼ B
@  C
A (2.107)
T0 _
1 Qin + m_ w hw
T
76 C HA PT E R 2 : Energy and Exergy Assessments

The third efficiency is defined as the ratio of the energy rate of the exiting stream
to the heat addition rate to the system, carried by the inlet stream and energy of
the hot water:
 
m_ a3 h3
ηen, 3 ¼ (2.108)
Q_ in + m_ a1 h1 + m_ w hw
0 1
B m_ a3 ðexÞ3 C
ηex, 3 ¼ B
@  C
A (2.109)
T0 _
1 Qin + m_ w ðexÞw
T

2.11.3 Illustrative Example


Stated air at 1 atm, 10 °C DB, and 70% relative humidity is heated and then
humidified at constant pressure to 25 °C DB and 60% relative humidity. The
energy and exergy efficiencies of the process are to be determined.

2.11.3.1 Results and Discussion


Table 2.6 gives thermodynamic properties at each state in the heating with
humidification example (Fig. 2.29).
The rate of exergy destruction, the rate of cooling, and the rate of entropy gen-
eration are calculated as follows:
_ dest ¼ 12:1kW Q_ h ¼ 41:84kW S_gen ¼ 0:04276kJ=kg K
Ex

Table 2.6 Thermodynamic Properties at State Points for Heating with Humidification Example
State Point P (kPa) T (K) h (kJ/kg) RH s (kJ/kg K) ex (kJ/kg) m_ (kg/s)

101.3 283 27.02 0.9 5.707 – –


1 101.3 265 4.925 0.7 5.59 1.06 0.6188
2 101.3 373.1 2676 0.5 7.355 595 0.004068
3 101.3 308 80.28 – 5.889 1.742 –
4 101.3 373.1 2519 – 8.902 – –

FIGURE 2.29 Heating with humidification example.


2.12 Cooling with Dehumidification 77

Using the efficiencies in Eqs. (2.88)–(2.93), the energy and exergy efficiencies
for heating with humidification in the example are as follows:
ηen, 1 ¼ 1 ηen, 2 ¼ 0:942 ηen, 3 ¼ 1 ηex, 1 ¼ 0:033 ηex, 2 ¼ 0:069 ηex, 2 ¼ 0:081

2.11.3.2 Parametric Studies


The effect of varying selected parameters (ambient temperature and relative
humidity) on the various efficiencies for heating with humidification discussed
earlier is investigated. Figure 2.30 demonstrates that ambient temperature has
no effect on the energy efficiencies, but increasing the ambient temperature
decreases the exergy efficiencies. Figure 2.31 shows that ambient relative
humidity also has no effect on all the energy efficiencies of the heating and
humidification process but that increasing the ambient relative humidity raises
the three corresponding exergy efficiencies.

2.12 COOLING WITH DEHUMIDIFICATION


Cooling with dehumidification is a process in which air is cooled sensibly and,
at the same time, the moisture is removed. Cooling and dehumidification
occurs when the air at a given DB and DP temperature is cooled below the
DP temperature.

1
0.2

0.8
0.16

0.6 0.12
hen

hex

0.4 0.08

hen,1 hex,1
0.2 0.04
hen,2 hex,2
hen,3 hex,3
0 0
273 276 279 282 285
T0 [K]

FIGURE 2.30 Effect of ambient temperature on heating with humidification energy and exergy
efficiencies.
78 C HA PT E R 2 : Energy and Exergy Assessments

1 0.1

0.8 0.08

0.6 0.06

hen

hex
0.4 0.04

hen,1 hex,1
0.2 0.02
hen,2 hex,2
hen,3 hex,3
0 0
0.2 0.4 0.6 0.8 1
RH0

FIGURE 2.31 Effect of ambient relative humidity on heating with humidification energy and exergy
efficiencies.

It is instructive to consider the cooling and dehumidification process in more


detail. When the air comes in contact with a cooling coil that is maintained at a
temperature below the DP temperature, the air’s DB temperature decreases. The
cooling process continues until it reaches the DP temperature of the air. At this
point, water vapor in the air starts to condense and dew particles form on the
surface of the cooling coils. The moisture content of the air declines, reducing
its relative humidity. Thus, when air is cooled below its DP temperature, there,
cooling and dehumidification occur.
The cooling and dehumidification process is most widely applied in air
conditioning. It is used in all types of window, split, packaged, and central air’s
DB temperature decreases after the air comes in interaction with a cooling coil
that is upheld at a temperature below the DP temperature. The cooling proce-
dure remains up until it hits the DP temperature of the air. At this point, water
vapor in the air begins to condense and dew particles appear on the surface of
the cooling coils. The moisture content of the air drops, dipping its relative
humidity. Therefore, once air is refrigerated lower than its DP temperature,
at that point, cooling and dehumidification take place. The cooling and dehu-
midification process is most extensively useful in air conditioning. It is applied
in all types of packaged, split, window, and central air conditioning systems for
creating the comfort situations preferred inside a space. In window and split air
conditioners, the evaporator (or cooling) coil is preserved at a temperature
lower the DP temperature of the room air by passing cool refrigerant over it.
When the room air moves above this coil, its DB temperature decreases and
2.12 Cooling with Dehumidification 79

moisture is removed since the air is cooled below its DP temperature. The dew
formed on the cooling coil is removed with tubing. In central air conditioning
systems, the cooling coil is cooled by a refrigerant or chilled water.
In a general cooling and dehumidification process, air passes over a coil
through which a cool refrigerant, chilled water, or cooled gas passes. During
the process, the DB, WB, and DP temperatures of air are reduced. As both sen-
sible and latent heats are removed from the air, a reduction is observed in the
enthalpy of the air. The cooling and dehumidification process is represented on
a psychrometric chart as shown in Fig. 2.32 (right side) by a line extending from
the initial condition down and to the left.

2.12.1 Rate Balance Equations


The mass, energy, entropy, and exergy rate balances can be written for cooling
with dehumidification:
Dry air mass balance : m_ a1 ¼ m_ a2 (2.110)

Water mass balance : m_ w1 ¼ m_ w2 + m_ w ! m_ a1 ω1 ¼ m_ a2 ω2 + m_ w (2.111)

Energy balance : m_ a1 h1 ¼ Q_ out + m_ a2 h2 + m_ w hw (2.112)

Q_ out
Entropy balance : m_ a1 s1 + S_gen ¼ m_ a2 s2 + m_ w sw + (2.113)
Ts
 
T0
Exergy balance : m_ a1 ex1 ¼ m_ a3 ex3 + m_ w exw + Q_ out 1  _ dest
+ Ex (2.114)
Ts
 _ 
_ dest ¼ T0 S_gen ¼ T0 m_ a3 s3  m_ a1 s1  m_ w sw  Qout
Ex (2.115)
Ts
Humidity ratio

Cooling
coils
x 1
1 T1 T
2 RH2 2
RH1
Air P 2

Condensate
Dry bulb temperature

FIGURE 2.32 Cooling with dehumidification: schematic (left) and representation on psychrometric
chart (right).
80 C HA PT E R 2 : Energy and Exergy Assessments

2.12.2 Efficiencies
Various efficiencies can be defined for cooling with dehumidification. For the
first definition, the energy efficiency is defined as the ratio of thermal energy
released by the system plus the energy of the exiting stream to the energy carried
by the incoming stream. The corresponding exergy efficiency is defined simi-
larly, but with exergy quantities. That is,

Q_ out + m_ a2 h2
ηen, 1 ¼ (2.116)
m_ a1 h1
 
T0 _
1 Qout + m_ a2 ex2
T
ηex, 1 ¼ (2.117)
m_ a1 ex1

The second efficiency definition is based on the ratio of energy of the exiting
stream to the heat rejected by the system. This definition presumes that the pur-
pose of the system is to cool the incoming stream, so the heat rejected by the
system is the actual input to the system. Then,
m_ a2 h2
ηen, 2 ¼ (2.118)
Q_ out

m_ a2 ex2
ηex, 2 ¼   (2.119)
T0 _
1 Qout
T

The third definition of efficiency states that the desired output of the system is
the sum of the heat released by the system, the energy of the exiting stream, and
the energy carried by the water, following the approach of Ratlamwala and
Dincer (2012). However, the required input is the energy of the entering
stream. For this case,

Q_ out + m_ a2 h2 + m_ w hw
ηen, 3 ¼ (2.120)
m_ a1 h1
 
T0
m_ a2 ex2 + m_ w exw + Q_ out 1 
Ts
ηex, 3 ¼ (2.121)
m_ a1 ex1

2.12.3 Illustrative Example


Air at 1 atm, 28 °C DB, and 25 °C DP is cooled and dehumidified at constant
pressure to 20 °C DB and 60% relative humidity. The energy and exergy effi-
ciencies of the process are to be determined.
2.12 Cooling with Dehumidification 81

Table 2.7 Thermodynamic Properties at State Points for Cooling with Dehumidification Example
State Point P (kPa) T (K) h (kJ/kg) RH s (kJ/kg K) ex (kJ/kg) m_ (kg/s)

0 101.3 292 51.84 0.95 5.793 – –


1 101.3 301.2 86.79 0.95 5.911 0.4952 1.105
2 101.3 293.2 42.3 0.6 5.76 0.1428 0.01989
3 101.3 301.2 117.3 – 0.4088 0.5891 –
4 101.3 292 51.84 0.95 5.793 – –

2.12.3.1 Results and Discussion


Table 2.7 gives the thermodynamic properties at each state in the cooling with
dehumidification example.
The rate of exergy destruction, the rate of cooling, and the rate of entropy gen-
eration are calculated as follows:
_ dest ¼ 0:43kW Q_ h ¼ 46:34kW S_gen ¼ 0:0015kJ=kg K
Ex

Using the efficiencies in Eqs. (2.116)–(2.121), the energy and exergy efficien-
cies for cooling with dehumidification example (Fig. 2.33) are found to be as
follows:
ηen, 1 ¼ 0:976 ηen, 2 ¼ 0:999 ηen, 3 ¼ 1 ηex, 1 ¼ 0:624 ηex, 2 ¼ 0:859 ηex, 2 ¼ 0:645

2.12.3.2 Parametric Study


The effect of varying selected parameters (ambient temperature and relative
humidity) on the various efficiencies for heating with humidification discussed
earlier is investigated. Figure 2.34 illustrates that ambient temperature does not
affect the energy efficiencies, while increasing the ambient temperature
decreases the exergy efficiency of the cooling with dehumidification process.

FIGURE 2.33 Cooling with dehumidification example.


82 C HA PT E R 2 : Energy and Exergy Assessments

1 1

0.8 0.8

0.6 hen,1 hex,1 0.6


hen,2 hex,2

hex
hen 0.4
hen,3 hex,3
0.4

0.2 0.2

0 0
273 276 279 282 285 288
T0 [K]

FIGURE 2.34 Effect of ambient temperature on cooling with dehumidification energy and exergy
efficiencies.

Similarly, Fig. 2.35 demonstrates that ambient relative humidity also has no
effect on all the energy efficiencies, while increasing the ambient relative
humidity causes one of the exergy efficiencies to increase and the other two
to decrease, for the cooling and dehumidification process.

1 1

0.9 0.9
hen

hex

0.8 0.8

hen,1 hex,1
hen,2 hex,2
hen,3 hex,3
0.7 0.7
0.8 0.82 0.84 0.86 0.88 0.9
RH0

FIGURE 2.35 Effect of ambient relative humidity on cooling with dehumidification energy and exergy
efficiencies.
2.13 Adiabatic Mixing of Air Streams 83

2.13 ADIABATIC MIXING OF AIR STREAMS


Many air conditioning applications require the mixing of two air streams.
This is particularly true for large buildings, production and process plants,
and hospitals, which require that the conditioned air be mixed with a cer-
tain fraction of fresh outdoor air before it is processed and routed to the
building spaces. Mixing is accomplished by merging the two air streams,
as shown in Fig. 2.36. The heat interaction with the surroundings is usually
small during the process, and thus, the mixing processes can normally be
assumed to be adiabatic. Mixing processes normally involve no work inter-
actions, and changes in kinetic and potential energies, if any, are usually
negligible.

2.13.1 Rate Balance Equations


Then, the mass, energy, entropy, and exergy rate balances for the adiabatic mix-
ing of two air streams are written as follows:
Dry air mass balance : m_ a1 ¼ m_ a2 ¼ ma3 (2.122)

Water mass balance : m_ w1 + m_ w2 ¼ m_ w3 ! m_ a1 ω1 + m_ a2 ω2 ¼ m_ a3 ω3 (2.123)

Energy balance : m_ a1 h1 + m_ a2 h3 ¼ m_ a3 h3 (2.124)

Entropy balance : m_ a1 s1 + m_ a2 s2 + S_gen ¼ m_ a3 s3 (2.125)

_ dest
Exergy balance : m_ a1 ex1 + m_ a2 ex2 ¼ m_ a3 ex3 + Ex (2.126)

_ dest ¼ T0 S_gen ¼ T0 ðm_ a3 s3  m_ a1 s1  m_ a2 s2 Þ


Ex (2.127)

For this process, only one efficiency definition is possible based on either
energy or exergy. The efficiency of this process is defined as the ratio of the
energy rate of the exiting stream to the energy rate of the two entering streams.

1 36 °C
8 kg/s
Twb1 = 30 °C
w3
P = 1 atm f3 3
Air T3

10 kg/s
12 °C
2 100%

FIGURE 2.36 Adiabatic mixing of air streams.


84 C HA PT E R 2 : Energy and Exergy Assessments

Thus, energy and exergy efficiencies, respectively, for adiabatic mixing of two air
streams can be written as follows:
 
m_ a3 h3
ηen ¼ (2.128)
m_ a1 h1 + m_ a2 h2

m_ a3 ex3
ηex ¼ (2.129)
m_ a1 ex1 + m_ a2 ex2

2.13.2 Illustrative Example


A stream of warm air is mixed with a stream of saturated cool air. The temper-
ature, the specific humidity, and the relative humidity of the mixture are to be
determined.
The assumptions made are listed as follows:
• Steady-flow, steady-state operating conditions exist.
• Dry air and water vapor behave like ideal gases.
• Kinetic and potential energy changes are negligible.
• The mixing section is adiabatic.

2.13.2.1 Results and Discussion


Table 2.8 gives the details of the thermodynamic properties at each state in the
cooling with dehumidification example.
The rate of exergy destroyed and the rate of entropy generation are calculated as
follows:
_ dest ¼ 8:82kW S_gen ¼ 0:0312kJ=kg K
Ex

Using the efficiencies in Eqs. (2.128)–(2.129), the energy and exergy


efficiencies for heating with humidification example are found to be as follows:
ηen ¼ 1 ηex ¼ 0:602

Table 2.8 Thermodynamic Properties at State Points for Adiabatic Mixing Example
State ex
Point h (kJ/kg) m_ (kg/s) P (kPa) RH Tdb (K) Twb (K) ω s (kJ/kg K) (kJ/kg)

1 98.65 8 101.3 0.647 309 303 0.024 5.95 2.726


2 33.73 10 101.3 1 285 285 0.0085 5.73 0.03912
3 62.58 18 101.3 0.899 295.8 294.6 0.0153 5.83 0.7433
2.13 Adiabatic Mixing of Air Streams 85

2.13.2.2 Parametric Study


The effect of varying selected parameters (ambient temperature and relative
humidity) on the energy and exergy efficiencies for adiabatic mixing is inves-
tigated. Figure 2.37 demonstrates that ambient temperature has no effect on
the energy efficiency of the adiabatic mixing process but that the exergy effi-
ciency increases with decreasing ambient temperature. Figure 2.38 shows

1 1
hen,1 hex,1

0.8 0.8

0.6 0.6
hen

hex
0.4 0.4

0.2 0.2

0 0
273 278 283 288 293
T0 [K]

FIGURE 2.37 Effect of ambient temperature on adiabatic mixing energy and exergy efficiencies.

1 1
hen
hex
0.8 0.8

0.6 0.6
hen

hex

0.4 0.4

0.2 0.2

0 0
0.2 0.4 0.6 0.8 1
RH0

FIGURE 2.38 Effect of ambient relative on adiabatic mixing energy and exergy efficiencies.
86 C HA PT E R 2 : Energy and Exergy Assessments

that ambient relative humidity also does not affect the energy efficiency,
while the exergy efficiency of adiabatic mixing increases with ambient rela-
tive humidity.

2.14 EVAPORATIVE COOLING


A common application of cooling with humidification is the evaporative
cooling, also called desert cooling. An evaporative cooler contains a water
tank, a small water pump, and a fan. Water from the tank is circulated
by the pump and sprayed into the airflow. The fan blows air over the water
spray, simultaneously humidifying and cooling the air. Cooling with
humidification can be used for space cooling only in dry and hot climates
like desert areas. Although this cooling process cannot operate in hot cli-
mate and high humidity, it is effective for cooling in hot, dry environments
and has very low initial and operating costs compared to unitary air
conditioners.

2.14.1 Rate Balance Equations


The mass, energy, entropy, and exergy rate balances for evaporative cooling can
be written as follows:
Dry air mass balance : m_ a1 ¼ m_ a2 (2.130)

Water mass balance : m_ w1 + m_ w ¼ m_ w2 ! m_ a1 ω1 + m_ w ¼ m_ a2 ω2 (2.131)

Energy balance : m_ a1 h1 ¼ m_ a2 h2 ! h1 ¼ h2 (2.132)

Entropy balance : m_ a1 s1 + m_ w sw + S_gen ¼ m_ a2 s2 (2.133)

_ dest
Exergy balance : m_ a1 ex1 + m_ w exw ¼ m_ a2 ex2 + Ex (2.134)

_ dest ¼ T0 S_gen ¼ T0 ðm_ a2 s2  m_ a1 s1  m_ w sw Þ


Ex (2.135)

2.14.2 Efficiencies
Various efficiencies can be defined for evaporative cooling (Fig. 2.39), that is,
cooling with humidification. The first defines the efficiency as the ratio of the
energy rate gained by the system to the energy rate provided to the system. In
this definition, the rate energy gained by the system is calculated by subtracting
the energy rate of the exiting stream from the energy rate of the entering stream.
However, the energy rate provided to the system is the energy rate of the input
stream and the energy rate of the water. Hence, the first energy and exergy effi-
ciencies can be written as follows:
2.15 Integrated System 87

FIGURE 2.39 Evaporative cooling: schematic (left) and representation on psychrometric chart (right).

m_ a2 h2 + m_ a1 h1
ηen, 1 ¼ (2.136)
m_ a1 h1 + m_ w hw
m_ a2 ex2
ηex, 1 ¼ (2.137)
m_ a1 ex1 + m_ w exw

The second efficiency definition considers the ratio of the energy rate of the exit-
ing stream to the energy rate of the water for the energy efficiency, and the cor-
responding exergy efficiencies then becomes
m_ a2 h2
ηen, 2 ¼ (2.138)
m_ w hw
m_ a2 ðexÞ2
ηex, 2 ¼ (2.139)
m_ w ðexÞw

The third efficiency definition is based on the ratio of energy rate of the exiting
stream to the energy rate of the inlet stream and water, based on the approach of
Ratlamwala and Dincer (2012). In this case,
m_ a2 h2
ηen, 3 ¼ (2.140)
m_ a1 h1 + m_ w hw
m_ a2 ðexÞ2 + m_ a1 ðexÞ1
ηex, 3 ¼ (2.141)
m_ a1 ðexÞ1 + mw ðexÞw

2.15 INTEGRATED SYSTEM


An integrated system, proposed by Ghosh and Dincer (2015), consists of two
open cycles, that is, a heating and cooling cycle and a closed refrigerant cycle.
The heating and cooling cycle uses atmospheric air and a water mixture as the
working fluid, while the refrigerant cycle uses R-134a.
88 C HA PT E R 2 : Energy and Exergy Assessments

The following psychrometric processes are utilized in the integrated system:

• Cooling with dehumidification


• Space cooling
• Evaporative cooling (in a cooling tower)
• Heating with humidification
• Space heating
• Ventilation

The integrated system produces multiple outputs depending on the load con-
ditions. On a hot day, atmospheric air is first cooled in the evaporator of the
vapor-compression cycle. The heat rejected by the atmospheric air is utilized
by the evaporator to increase the temperature of R-134a. The cooled air is then
used for space cooling and further utilized to cool hot water for residential pur-
poses. But on a cold day, atmospheric air is first heated and humidified using
the heat rejected by the condenser of the vapor compression cycle. The air is
then used for space heating and further used for ventilation purposes.
On a hot day, atmospheric air at 32 °C is first cooled and dehumidified to 20 °C
at state point 2. This air is used for space cooling for residential applications,
leaving at 28 °C at state point 4. This air is further used to cool water from
40 to 32 °C in a cooling tower for residential purposes. The air leaves the cool-
ing tower at an atmosphere at 33 °C.
On a cold day, atmospheric air at 10 °C is first heated and humidified to 48 °C,
from state points 12-15. The heated air is then used for space heating for res-
idential and/or industrial purposes, leaving the building at 32 °C.

2.15.1 Rate Balance Equations


The mass, energy, entropy, and exergy rate balances for each component in the
integrated system are written as follows:
Cooling with humidification (A in Fig. 2.40)
Dry air mass balance : m_ a1 ¼ m_ a2 (2.142)

Water mass balance : m_ w1 ¼ m_ w2 + m_ w ! m_ a1 ω1 ¼ m_ a2 ω2 + m_ w (2.143)

Energy balance : m_ a1 h1 ¼ Q_ A + m_ a2 h2 + m_ w3 hw3 (2.144)

Q_
Entropy balance : m_ a1 s1 + S_gen ¼ m_ a2 s2 + m_ w3 sw3 + A (2.145)
Ts
 
T0
Exergy balance : m_ a1 ex1 ¼ m_ a2 ex2 + m_ w3 exw3 + Q_ A 1  _ dest, A
+ Ex (2.146)
Ts

Space cooling (B in Fig. 2.40)


2.15 Integrated System 89

FIGURE 2.40 Integrated system process flow diagram.

Dry air mass balance : m_ a2 ¼ m_ a4 (2.147)

Water mass balance : m_ w2 ¼ m_ w4 (2.148)

Energy balance : Q_ B + m_ a2 h2 ¼ m_ a4 h4 (2.149)

Q_ B _
Entropy balance : m_ a2 s2 + + Sgen, B ¼ m_ a2 s4 (2.150)
T0
 
T0
Exergy balance : m_ a2 ex2 + Q_ B 1  _ dest, B
¼ m_ a4 ex4 + Ex (2.151)
Ts
 
_ _ Q_ B
_ _
Exdest, B ¼ T0 Sgen, B ¼ T0 ma4 s4  ma2 s2  (2.152)
T0

Evaporative cooling (C in Fig. 2.40)


Dry air mass balance : m_ a4 ¼ m_ a5 (2.153)

Water mass balance : m_ w4 + m_ w6 ¼ m_ 52 ! m_ a4 ω4 + m_ w6 ¼ m_ a5 ω5 (2.154)

Energy balance : m_ a4 h4 ¼ m_ a5 h5 ! h4 ¼ h5 (2.155)

Entropy balance : m_ a4 s4 + m_ w4 sw4 + S_gen, C ¼ m_ a4 s4 (2.156)

_ dest, C
Exergy balance : m_ a4 ex4 + m_ w4 exw4 ¼ m_ a4 ex5 + Ex (2.157)

_ dest, C ¼ T0 S_gen, C ¼ T0 ðm_ a5 s5  m_ a4 s4  m_ w4 sw4 Þ


Ex (2.158)

Heating with humidification (D in Fig. 2.40)

Dry air mass balance : m_ a12 ¼ m_ a13 ¼ m_ a15 (2.159)


90 C HA PT E R 2 : Energy and Exergy Assessments

Water mass balance : m_ w12 ¼ m_ w13 (2.160)

m_ w12 + m_ 14 ¼ m_ w15 ! m_ a12 ω12 + m_ w14 ¼ m_ a15 ω15

Energy balance : Q_ in, 1232 + m_ a12 h12 ¼ m_ a13 h13 ðprocess 1  2Þ

m_ a12 h12 + m_ w14 h14 ¼ m_ a15 h15 ðprocess 2  3Þ

Q_ in, 1215 + m_ a12 h12 + m_ w14 h14 ¼ m_ a15 h15 ðprocess 1  3Þ (2.161)

Q_ in, 1215 _
Entropy balance : m_ a12 s12 + m_ w14 s14 + + Sgen, 1215 ¼ m_ a3 s3 ðprocess 1  3Þ
T0
(2.162)
 
T0
Exergy balance : Q_ in, 1215 1  _ dest, 1215
+ m_ a12 ex12 ¼ m_ a15 ex15 + Ex
Ts

m_ a12 ex15 + m_ w14 ex14 ¼ m_ a15 ex15 + Ex_ dest, 1315 ðprocess 2  3Þ
 
T0
Q_ in, 1215 1  _ dest, 1215 ðprocess 1  3Þ
+ m_ a12 ex12 + m_ w14 ex14 ¼ m_ a15 ex15 + Ex
Ts
!
Q_ in, 1215
_ dest, 1215 ¼ T0 S_gen, 1215 ¼ T0 m_ a15 s15  m_ a12 s12  m_ w14 s14 
Ex (2.163)
T0

Space heating (E in Fig. 2.40)


Dry air mass balance : m_ a15 ¼ m_ a16 (2.164)

Water mass balance : m_ w15 ¼ m_ w16 (2.165)

Energy balance : Q_ E + m_ a15 h15 ¼ m_ a15 h15 (2.166)

Q_ E _
Entropy balance : m_ a15 s15 + + Sgen, E ¼ m_ a16 s16 (2.167)
T0
 
T0
Exergy balance : m_ a15 ex15 + Q_ E 1  _ dest, E
¼ m_ a16 ex16 + Ex (2.168)
Ts
 _ 
_ dest, E ¼ T0 S_gen, E ¼ T0 m_ a15 s15  m_ a16 s16  QE
Ex (2.169)
T0

Compressor (I in Fig. 2.40)


Mass balance : m_ 8 ¼ m_ 9 (2.170)

Energy balance : m_ 8 h8 + W_ comp ¼ m_ 9 h9 (2.171)

Entropy balance : m_ 8 s8 + S_gen, comp ¼ m_ 9 s9 (2.172)

Exergy balance : m_ 8 ex8 + W_ comp ¼ m_ 9 ex9 + Ex


_ d, comp (2.173)
2.15 Integrated System 91

Condenser (II in Fig. 2.40)


Mass balance : m_ 9 ¼ m_ 10 (2.174)

Energy balance : m_ 9 h10 ¼ m_ 10 h10 + Q_ cond (2.175)

Q_
Entropy balance : m_ 9 s9 + S_gen, cond ¼ m_ 9 s9 + cond (2.176)
T0
 
_ d, cond + Q_ cond 1  T0
Exergy balance : m_ 9 ex9 ¼ m_ 10 ex10 + Ex (2.177)
Ts

Throttling valve (III in Fig. 2.40)


Mass balance : m_ 10 ¼ m_ 11 (2.178)

Energy balance : m_ 10 h10 ¼ m_ 11 h11 (2.179)

Entropy balance : m_ 10 s10 + S_gen, val ¼ m_ 10 s10 (2.180)

_ d, val
Exergy balance : m_ 10 ex10 ¼ m_ 10 ex10 + Ex (2.181)

Evaporator (IV in Fig. 2.40)


Mass balance : m_ 11 ¼ m_ 8 (2.182)

Energy balance : m_ 11 h11 + Q_ evap ¼ m_ 8 h8 (2.183)

Q_ evap
Entropy balance : m_ 11 s11 + S_gen, evap + ¼ m_ 8 s8 (2.184)
T0
 
T0
Exergy balance : m_ 11 ex11 + Q_ evap 1  _ d, comp
¼ m_ 8 ex8 + Ex (2.185)
Ts

2.15.2 Results and Discussion


2.15.2.1 The Results of Exergy Destruction Rates
Figure 2.41 presents the ratio of exergy destruction rate for each component in the
integrated system to the total exergy destruction rate of the system. It is seen that the
condenser has the greatest exergy destruction rate while the compressor has the
lowest exergy destruction, for the components present. This is because the com-
pressor is assumed to be isentropic, precluding it from exhibiting entropy gener-
ation or exergy destruction. Space heating also has a high exergy destruction rate
due to the large change in exergy rates between the inlet and outlet streams.
Table 2.9 presents the exergy destruction rate for each component as well as the
total exergy destruction rate for the integrated system, that is, the summation of
the exergy destruction rates for each component (see last column).
92 C HA PT E R 2 : Energy and Exergy Assessments

0.40

0.35

Exergy destruction ratio


0.30

0.25

0.20

0.15

0.10

0.05

0.00
Cooling Compressor Condenser Space Evaporator Space Throttling
tower cooling heating valve

FIGURE 2.41 Ratio of exergy destruction rate of component to the total exergy destruction rate.

Table 2.9 Exergy Destruction Rates for the Overall Integrated System
and its Components
Component Exergy Destruction Rate (kW)

Cooling tower 77
Compressor 0
Condenser 6425
Space cooler 109
Evaporator 4600
Space heater 5929
Throttling valve 1819
Integrated system 18,959

2.15.2.2 Efficiency Results


The efficiencies of the integrated system can be defined under different loading
conditions. For instance, on a hot summer day with ambient temperature of
32 °C, the requirement would be of 100% cooling load and 0% heating load.
The energy efficiency for this loading condition would be the total useful energy
output rate (Q_ c for space cooling, the change in energy rate between hot and
cold waters in the cooling tower and Q_ cond , the useful heat rate from the con-
denser) divided by the total input energy rate (input air and compressor work
input rates). Then,

Q_ c + m_ w, 6 ðh7  h6 Þ + Q_ cond
ηen, c ¼ (2.186)
m_ a, 1 ha, 1 + W_ comp
2.15 Integrated System 93

Correspondingly, the exergy efficiency for a 100% cooling load can be defined as
   
_ T0 _ T0
Qc 1  + m_ w, 6 ðex7  ex6 Þ + Qcond 1 
T2 T9
ηex, c ¼ (2.187)
m_ a, 1 exa, 1 + W_ comp

On a cold winter day with an ambient temperature of 10 °C, the requirement


would be of 0% cooling load and 100% heating load. The energy efficiency for
this loading condition would be the total useful energy output rate (Q_ h for
space heating) divided by the total input energy rate (input air, compressor
work input, and the evaporator input energy rates). Then,

Q_ h
ηen, h ¼ (2.188)
m_ a, 1 ha, 12 + W_ comp + Q_ evap

Correspondingly, the exergy efficiency for 100% heating load is


 
T12
Q_ h 1 
T15
ηex, h ¼   (2.189)
T12
m_ a, 1 exa, 12 + W_ comp + Q_ evap 1 
T15

The overall energy and exergy efficiencies of the integrated system on an average
day with an ambient temperature of 24 °C can be defined according to a
requirement of 50% cooling load and 50% heating load:

Q_ c + m_ w, 6 ðh7  h6 Þ + Q_ h
ηen, sys ¼ (2.190)
m_ a, 1 ha, 1 + W_ comp

   
T0 T0
Q_ c 1  + m_ w, 6 ðex7  ex6 Þ + Q_ h 1 
T2 T15
ηex, sys ¼ (2.191)
_ _ _
ma, 1 exa, 1 + ma, 1 exa, 12 + W comp

The energy and exergy efficiencies for the integrated system can be evaluated
with the above expressions as follows:
ηen, c ¼ 18:4% ηex, c ¼ 18:1% ηen, h ¼ 77:8%
ηex, h ¼ 75:4% ηen, sys ¼ 18:6% ηex, sys ¼ 33:3%

2.15.2.3 Parametric Study


The effect of varying selected parameters (ambient temperature and relative
humidity) on the energy and exergy efficiencies for integrated system and its
94 C HA PT E R 2 : Energy and Exergy Assessments

FIGURE 2.42 Effect of ambient temperature on energy and exergy efficiencies of integrated system for
100% cooling load.

components is investigated. Figure 2.42 demonstrates that for a 100% cooling


load, an increase in ambient temperature increases the energy efficiency of the
system while decreasing the exergy efficiency of the system. Figure 2.43 shows
that for a 100% heating load, an increase in ambient temperature increases the
energy and exergy efficiencies of the system. Figure 2.44 shows the for a 50%

FIGURE 2.43 Effect of ambient temperature on energy and exergy efficiencies of integrated system for
100% heating load.
2.15 Integrated System 95

FIGURE 2.44 Effect of ambient temperature on energy and exergy efficiencies of integrated system for
50% cooling load and 50% heating load.

heating load and 50% cooling load, an increase in the ambient temperature
increases the energy and exergy efficiencies of the system. Figure 2.45 demon-
strates the effect of ambient relative humidity on energy and exergy efficiencies
of the system under different loading conditions. In general, increasing the
ambient relative temperature increases the energy and exergy efficiencies of
the system.

FIGURE 2.45 Effect of ambient relative humidity on energy and exergy efficiencies of integrated system
under different loading conditions.
96 C HA PT E R 2 : Energy and Exergy Assessments

2.16 CLOSING REMARKS


In this chapter, energy and exergy assessments of basic components, psychro-
metric processes, and an integrated system for HVACR applications are
described and illustrated. The basic components in HVACR systems such as
heat exchangers, pumps, compressors, throttles, and turbines are introduced,
classified, and thermodynamically analyzed. Energy and exergy assessments
of psychrometric processes are considered, including sensible heating, sensible
cooling, heating with humidification, cooling with dehumidification, evapora-
tive cooling, and adiabatic mixing of air streams. Mass, energy, entropy, and
exergy rate balances for all components and psychrometric processes are writ-
ten. The integrated system presented in this chapter produces multiple outputs
for HVACR applications such as space heating, space cooling, and cooling of
water. The energy and exergy efficiencies for individual components and the
overall integrated system are determined and parametric analyses of the effects
of varying dead-state properties and operating conditions are provided. The
energy and exergy efficiencies of the system are calculated to be 18.6% and
33.3%, respectively. It is observed that increasing the ambient relative humidity
and the ambient temperature increases the energy and exergy efficiencies of the
system.

Nomenclature
Ėx exergy rate (kW)
ex specific exergy (kJ/kg)
h specific enthalpy (kJ/kg)
ṁ mass flow rate (kg/s)
P pressure (kPa)
Q heat transfer (kJ)
Q̇ heat rate (kW)
s specific entropy (kJ/kg K)
S entropy rate
T temperature (K or °C)
ρ density (kg/m3)
v specific volume (m3/kg)
Ẇ work rate (kW)
V volume (m3)

Greek symbols
η efficiency
ρ density (kg/m3)
ω specific humidity or humidity ratio (kg/kg)
References 97

Subscripts
a air
c cooling
cd cooling with dehumidification
cond condenser
comp compressor
ct cooling tower
d/dest destruction
en energy
evap evaporator
ex exergy
gen generation
h heating
hh heating with humidification
in input
ref refrigerant
s source
sc space cooling
sh space heating
sys system
val valve
w water
0–17 state points

References
Andreev, L.P., Kostenko, G.N., 1965. Exergetic characteristics of the efficiency of heat exchangers.
Izv. Vuzov Ser. Energetika 3, 53–60.
Dincer, I., Rosen, M.A., 2011. Thermal Energy Storage: Systems and Applications, second ed. Wiley,
London.
Dincer, I., Rosen, M.A., 2013. Exergy-Energy, Environment and Sustainable Development, second
ed. Elsevier, New York.
Dincer, I., Kanoglu, M., Rosen, M.A., 2007. Exergy analysis of psychrometric processes for HVAC&R
applications. ASHRAE Trans. 113, 172–180.
Ghosh, S., Dincer, I., 2015. Development and performance assessment of a new integrated system
for HVAC&R applications. Energy 80, 159–167.
Kanoglu, M., Dincer, I., Rosen, M.A., 2007. Exergy analysis of psychometric processes for HVAC&R
applications. ASHRAE Trans. 113, 172–180.
Mikheev, M.A., 1956. Fundamentals of Heat Transfer. Gosénergoizdat, Moscow.
Ratlamwala, T.A.H., Dincer, I., 2012. Efficiency assessment of key psychrometric processes. Int. J.
Refrig. 36 (3), 1142–1153.
Stecco, S.S., Manfrida, G., 1986. Exergy analysis of compression and expansion processes. Energy
11 (6), 573–577.
Wepfer, W.J., Gaggioli, R.A., Obert, E.F., 1979. Proper evaluation of available energy for HVAC.
ASHRAE Trans. 85 (I), 214–230.
CHAPTER 3

Industrial Heating and Cooling Systems

3.1 INTRODUCTION
Energy is needed for industrial processes throughout the world and is integral
to technology improvement, economic growth, and modernization in the
industrial sector. In fact, the industrial sector is one of the largest energy con-
sumers in many countries, and about 35% of world’s total energy use is
reported to be in the industrial sector (Suleman et al., 2014a,b). This of course
depends on factors related to each country, such as the level of technological
and economic development.
Heating and cooling are often large energy users in the industrial sector. In indus-
trial heating, process heating, in which energy is transferred to a process or to
a material for treatment, is widespread. Various materials are used in the indus-
trial sector, including metals, plastics, rubber, concrete, glass, and ceramics.
Similarly, industrial cooling is common, in applications such as cooling or
freezing of perishables, cryogenic processes like air liquefaction, and others.
In addition, the industrial sector requires heating and cooling of the facilities
in which industrial operations are located.
There are various methods for industrial heating, such as direct or indirect heat-
ing. In direct heating systems, heat is generated directly within a material by
passing electric current through it, or by causing controlled exothermic reac-
tions, or by exciting atoms or molecules inside the material by electromagnetic
radiation (e.g., microwaves). In indirect heating systems, energy is transferred
to materials from heat sources by heat transfer in the form of conduction, con-
vection, radiation, or a combination of these. The range of industrial heating
systems is categorized in Fig. 3.1.
Industrial heating and cooling operations often use fossil fuels, directly through
or indirectly via other energy conversion processes. But over the last few
decades, energy issues, such as supply security and environmental impact, have
led many countries to offer incentives for renewable energy exploitation, in the
industrial and other sectors of the economy. The aim of such incentives is often 99

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00003-X


© 2015 Elsevier Inc. All rights reserved.
100 C HA PT E R 3 : Industrial Heating and Cooling Systems

Industrial
heating

Conventional Non-conventional
heating heating

Combustion Renewable Waste heat


Electric based
based energy recovery

Heating
medium

Heat transfer
Hot water Hot air Steam
fluid

FIGURE 3.1 Classifications of industrial heating systems.

substituting or reducing fossil fuel use, offsetting in whole or in part their emis-
sions of pollutants and greenhouse gases like carbon dioxide (Suleman et al.,
2014a,b). Natural energy resources, such as sunlight, wind, rain, tides, waves,
geothermal heat, and biomass, are renewable and thus can contribute to
these aims.
Concerns about energy resources have also fostered interest in better under-
standing and improving the efficiency of energy conversion devices. According
to the first law of thermodynamics, which deals with energy quantities and the
principle of conservation of energy, some energy efficiency improvements can
be identified. But the range of improvements demonstrated with energy
methods is limited since energy analysis often does not evaluate thermody-
namic losses and efficiencies meaningfully. As a consequence, energy methods
often do not indicate means for the effective use of fuels and resources. But
exergy methods, which are based on the first and second laws of thermodynam-
ics, help overcome these weaknesses. Exergy conservation and exergy efficiency
improvement are logical and meaningful thermodynamic objectives, if the
objective is better use of resources. As pointed our earlier, the benefits of exergy
methods stem in large part from its foundation in the second law, which
addresses the quality of energy. As pointed out earlier, exergy is defined as
the maximum useful work that could be obtained from a system at a given state
in a specified environment. Exergy is not conserved, except in reversible (ideal)
processes, but reducing the use of exergy can be accomplished with appropriate
measures.
3.3 Renewable Heating And Cooling 101

Exergy analysis is, consequently, an effective method for improving the efficien-
cies and performance of industrial heating and cooling systems. It does this by
using the conservation of mass and conservation of energy principles together
with the second law in the design and analysis of energy systems. Applying
exergy analysis to an industrial processes helps determine the usefulness, qual-
ity, and potential to cause change for any material or energy flow in a process
stream. Hence, exergy analysis is an important tool for determining the loca-
tions, types, and magnitudes of efficiencies and losses in the industrial sector
generally and in industrial heating and cooling in particular.
This chapter focuses on the use of exergy methods in assessing and improving
industrial heating and cooling systems and covers both renewable and tradi-
tional (i.e., fossil fuel and electric) processes.

3.2 INDUSTRIAL PROCESS HEATING TEMPERATURES


Industrial process heating can be categorized on the basis of heating tempera-
ture. Brown et al. (1985) divided the industrial heating sector into low-,
medium-, and high-temperature ranges. Table 3.1 lists process heating temper-
ature ranges for the industrial sector and the corresponding energy efficiencies
for heating with electricity or fuel. Table 3.2 shows an example of the temper-
ature ranges along with the efficiencies, for a variety of industries.

3.3 RENEWABLE HEATING AND COOLING


Heating with renewable energy is a relatively new mode of heating for industry.
Heat is generated from renewable sources instead of fossil fuels (e.g., replacing a
fossil fuel furnace/boiler using concentrating solar thermal to feed radiators).
The use of renewable energy sources for heating and cooling has received much
less attention than electricity generation from renewable sources. This is

Table 3.1 Process Heating Temperatures and Efficiencies for


the Industrial Sector
Energy Efficiency (%)
Heating Process Heating Temp. Electrical Fuel
Category Range, Tp (°C) Heating, ηe,h Heating, ηe,h

Low <121 100 65.5


Medium 121-399 90 60
High >399 70 50

Compiled from Dincer and Rosen (2013)


102 C HA PT E R 3 : Industrial Heating and Cooling Systems

Table 3.2 Process Heating Data and Energy and Exergy Efficiencies for All Categories of Product
Heat Temperature Tp for Various Industries in the Industrial Sector
Breakdown of Energy and
Breakdown of Energy Used for Exergy Efficiencies for Each
Each Tp Tp Category, by Type

Electrical
Heating Fuel Heating
Industry Tp Range Mean Tp (°C) Electricity Fuel
ηe,h ψ e,h ηf,h ψ f,h

Iron and steel Low 45 4.2 0 100.0 6.3 65.0 4.1


Medium 0 0 90.0 – 60.0 –
High 983 95.8 100 70.0 53.4 50.0 38.1
Chemical and petrochemical Low 42 62.5 0 100.0 5.4 65.0 3.5
Medium 141 37.5 100 90.0 25.2 60.0 16.8
High 494 0 0 70.0 42.8 50.0 30.6
Petrochemical feedstock Low 57 0 0 100.0 9.7 65.0 6.3
Medium 227 0 0 90.0 36.4 60.0 24.2
High 494 0 100 70.0 42.8 50.0 30.6
Fertilizer Low 57 10 30 100.0 9.7 65.0 6.3
Medium 350 80 30 90.0 47.0 60.0 31.3
High 900 10 40 70.0 52.2 50.0 37.3
Cement Low 42 91.7 0.9 100.0 5.4 65.0 3.5
Medium 141 0 9 90.0 25.2 60.0 16.8
High 586 8.3 90.1 70.0 45.7 50.0 32.7
Sugar Low 83 100 59 100.0 16.3 65.0 10.6
Medium 315 0 9 90.0 44.4 60.0 29.6
High 400 0 32 70.0 39.0 50.0 27.9
Noniron metals Low 61 10 13.8 100.0 10.8 65.0 7.0
Medium 132 9.4 22.6 90.0 23.8 60.0 15.9
High 401 80.4 63.6 70.0 39.0 50.0 27.9
Other industries Low 57 10.6 13.8 100.0 9.7 65.0 6.3
Medium 132 89.4 86.2 90.0 23.8 60.0 15.9
High 400 0.1 0.1 70.0 39.0 50.0 27.9
Compiled from Dincer and Rosen (2013), Utlu et al. (2007)

observed despite the fact that meeting thermal energy demands consumes the
largest share of primary energy supply in many countries and renewable energy
sources in some circumstances are practical alternatives to fossil fuels.
The main renewable heat resources are as follows:

• Solar
• Geothermal
• Biomass
3.3 Renewable Heating And Cooling 103

Input Output

Renewable energy Direct useful heat


Energy conversion
sources

Electricity
generation or other
energy carrier

Useful heat from


CHP

FIGURE 3.2 Renewable energy heat production. Modified from IEA (2008).

The use of solar, geothermal, and biomass energy for heating and cooling is
growing (Seyboth et al., 2008). The main facets of renewable energy heat
production, from input to output, are illustrated in Fig. 3.2 in a general
sense.
Two of the main renewable heat resources, solar and geothermal, are discussed
further in the next two subsections.

3.3.1 Solar Energy


Solar energy technology is well proven. Solar energy has little negative effects on
the environment compared to conventional energy resources like fossil fuels,
which emit greenhouse gases and other pollutants. Solar energy is also one
of the most abundant forms of energy. The Earth’s surface receives 108 kWh
of solar energy daily (Suleman et al., 2014a,b), which is equivalent to
500,000 billion barrels of oil. This value of solar energy is 1000 times greater
than known oil reserves (Mittal et al., 2005).
Solar energy can be used to generate electricity indirectly using various types of
solar thermal systems, including parabolic trough solar collectors, solar dishes,
and solar towers. These solar thermal collection systems can also convert solar
energy into useful thermal outputs, allowing solar thermal energy to be applied
in industrial process heating as well as space heating and space cooling (using
heat-driven absorption chillers).
In addition to the active solar energy systems described above, passive solar
energy systems can be useful. They can be used for space heating and cooling
and natural lighting of dwellings or commercial buildings.
The free availability of solar energy makes the operating costs to exploit it low,
although the initial costs can be high.
104 C HA PT E R 3 : Industrial Heating and Cooling Systems

3.3.2 Geothermal Energy


Heat contained within the Earth’s crust can be discharged by heat conduction as
well as in the form of hot water or steam at particular locations. This thermal
energy can be utilized for power generation, or directly for heating (e.g., indus-
trial process heating, district heating, swimming pool heating, greenhouse heat-
ing, aquaculture pond heating, or agricultural drying), and to operate thermally
driven heat pumps. The energy provided by using geothermal energy is usually
measured as the difference of the energy of the fluid (liquid or vapor) extracted
and energy of the fluid reinjected (if reinjection occurs).
Geothermal energy is usually considered a renewable and environmentally
benign energy resource. It is reliable and relatively inexpensive. Since the dis-
covery of geothermal energy, its use has increased more than two times
(Ratlamwala and Dincer, 2013).
A distinction is usually made between deep geothermal and shallow geother-
mal energy resources (IEA, 2008).
Geothermal energy resources are generally divided into three categories, based
on the resource temperature: low temperature (below 90°C), moderate temper-
ature (90-150°C), and high temperature (above 150°C) (Kanoglu and
Bolatturk, 2008). Efficient utilization of lower geothermal wells (about 70-
100 °C) is the subject of much investigation, especially since wells with these
temperatures are available in many places in the world.
Industrial applications of geothermal energy usually require geothermal fluids
at low to medium temperatures and encompass a rather wide range of indus-
trial activities:

• Process heating
• Industrial space air conditioning
• Food processing
• Food drying
• Fish drying
• Pulp and paper processing
• Washing and dying of textiles
• Leather and fur treatment
• Fuel production and oil enhancement
• Chemicals production

3.4 REQUIREMENTS AND SYSTEMS FOR LOW- TO


MEDIUM-TEMPERATURE INDUSTRIAL HEATING AND
COOLING
Many industries require systems for low- to medium-temperature heating and cool-
ing. Some of the important processes and the range of the temperatures required for
3.4 Requirements And Systems For Low- To Medium-Temperature Industrial Heating And Cooling 105

Table 3.3 Heating Temperature Ranges for Various Industries and Their
Main Processes
Industry Process Temperature Range (°C)

Food industries
Dairy Pressurization 60-80
Sterilization 100-120
Drying 120-180
Concentrates 60-80
Boiler feed water 60-90
Tinned food Sterilization 110-120
Pasteurization 60-80
Cooking 60-90
Bleaching 60-90
Meat Washing sterilization 60-90
Cooking 90-100
Beverages Washing sterilization 60-80
Pasteurization 60-70
Flours and by-products Sterilization 60-80

Non-food industries
Textile Bleaching, dying 60-90
Drying, degreasing 100-130
Dyeing 70-90
Fixing 160-180
Pressing 80-100
Paper Cooking, drying 60-80
Boiler feed water 60-90
Bleaching 130-150
Timber by-products Thermodiffusion beams 80-100
Drying 60-100
Preheating water 60-90
Preparation pulp 120-170
Chemical Soaps 200-260
Synthetic rubber 150-200
Processing heat 120-180
Preheating water 60-90
Plastics Preparation 120-140
Distillation 140-150
Separation 200-220
Extension 140-160
Drying 180-200
Compiled from Kalogirou (2003)
106 C HA PT E R 3 : Industrial Heating and Cooling Systems

each are outlined in Table 3.3. The main industrial processes utilizing heat at a low-
to medium-temperature range are sterilizing, pasteurizing, drying, hydrolyzing,
distillation and evaporation, washing and cleaning, and polymerization.
Studies of industrial heat demands reveal that several industries have heating
and cooling temperature requirements that can be met with solar thermal
energy. Figure 3.3 shows a solar energy-driven process heating system.
Large-scale solar applications for process heat often benefit from the effects of
scale. To improve economics, the initial investment costs need to be low,
although this is often challenging since solar collector costs are relatively high.
One way to enhance the economics of such systems is to design them so the
solar heat is fed directly to a suitable process without heat storage. In this case,
the greatest amount of solar energy that the system delivers cannot appreciably
exceed the rate at which the process uses energy.
It can be seen in Table 3.3 that most of the energy utilized is in the food industry
and the manufacture of nonmetallic mineral products. The types of food indus-
tries that can employ solar thermal energy are the milk, tinned foods, and meat
(sausage, salami, etc.) industries and breweries. Most of the low- to medium-
temperature process heat used in the food and textile industry is for a diverse
array of applications, including drying, cooking, cleaning, extraction, and
others. Particularly favorable conditions exist in food industry, since food treat-
ment and storage processes exhibit high energy consumption rates and high
running times. In the textile industry, thermal energy is mainly consumed
for heating of liquid baths close to 100°C, washing, bleaching, and dying. Dry-
ing processes normally use hot air or gases at 140-220°C.

Central steam
supply
Solar collector
used for steam
generation

Boiler
Process Process Process

Solar collector
applied to a
Pump Return water particular
process

Make-up water
FIGURE 3.3 An industrial heating system using solar thermal energy (also adaptable to be used with
conventional heat supplies). Modified from Kalogirou (2003).
3.5 Industrial Heat Pumps 107

3.5 INDUSTRIAL HEAT PUMPS


Heat pumps can be used for heating and/or cooling and find application in the
building, industrial, and other sectors.
In heating applications, heat pumps can transfer heat from a lower to a higher tem-
perature, and the heat delivered at the higher temperature is the product. That is, the
higher-temperature reservoir is normally the item or region requiring heating. In
such heating processes, the lower-temperature reservoir is often a natural resource
such as the surroundings (e.g., air or ground or water) from which heat is extracted.
But the lower-temperature reservoir can also be as anthropogenic, such as waste heat
from industrial or domestic operations, building systems, or industrial processes.
In cooling applications, heat pumps also transfer heat from a lower to a higher
temperature, but the heat extracted at the lower temperature is the product.
Then, the lower-temperature reservoir is normally the item or region requiring
cooling. In essence, heat pumps operate similarly when they provide cooling
instead of heating, but they achieve a reverse purpose. That is, heat is transferred
from the application requiring cooling to a higher-temperature region, nor-
mally the surroundings. Sometimes, the rejected heat from cooling is used to
meet a simultaneous heat demand.
Industrial heat pumps are a class of active heat recovery device that raises the
temperature of a waste heat stream so it is more useful. Consequently, heat
pumps can lead to energy savings, and they often find application when con-
ventional passive heat recovery is not possible. A large-scale application of heat
pumps for industrial processes is shown in Fig. 3.4. The heat pump can provide
approximately 390 PJ/year, which is 16% of the total energy consumption of
German industry in 2005. Table 3.4 provides data regarding the industrial heat

140
Hot water
120
Heating (up to 100 °C)
Annual heat use (PJ)

100
Process heat 100 °C
80

60

40

20

0
r

ile
al

s
ile

al

er
od

pe

tic
oo

ic

et

th
ob
xt
Fo

pa

as
m

M
W

O
Te

om
he

Pl
d
an

ut
A
l p
Pu

FIGURE 3.4 Potential large-scale applications of heat pumps for industrial processes. Modified from EHPA (2012).
108 C HA PT E R 3 : Industrial Heating and Cooling Systems

Table 3.4 Heat Pump Applications in Industrial Processes


Industry Activity Process

Petroleum refining Distillation of petroleum and petrochemical Separation of propane/propylene, butane/


and petrochemicals products butylene and ethane/ethylene
Chemicals Inorganic salt manufacture, including salt, Concentration of product salt solutions
sodium sulfate, sodium carbonate, boric acid
Heat recovery Compression of low-pressure steam or
vapor for use as a heating medium
Pharmaceuticals Process water heating
Treatment of process effluent Concentration of waste streams to reduce
hydraulic load on waste treatment facilities
Wood products Pulp manufacturing Concentration of black liquor
Paper manufacturing Process water heating
Paper manufacturing Flash-steam recovery
Lumber manufacturing Product drying
Food and beverage Manufacturing of alcohol Concentration of waste liquids
Beer brewing Concentration of waste beer
Wet corn milling/corn syrup Concentration of steep water
Sugar refining Concentration of sugar solution
Dairy products Concentration of milk and of whey
Juice manufacturing Juice concentration
General food products Heating of process and cleaning water
Soft drink manufacturing Concentration of effluent
Utilities Nuclear power generation Concentration of radioactive waste
Concentration of cooling tower blowdown
Miscellaneous Manufacturing of drinking water Desalination of seawater
Steam stripping of waste water or process Flash-steam recovery
streams
Electroplating industries Heating of process solutions
Concentration of effluent
Textiles Process and wash-water heating
Space heating
Concentration of dilute dope stream
General manufacturing Process and wash-water heating
Space heating
District heating Large-scale space heating
Solvent recovery Removal of solvent from air streams
Compiled from Chua et al. (2010)
3.6 Combustion-Based Process Heating 109

pump application and Table 3.5 lists the heat pump selection criteria. Some
typical applications of industrial heat pumps follow:

• Space heating
• Heating and cooling of process streams
• Water heating for washing, sanitation, and cleaning
• Steam production
• Drying/dehumidification
• Evaporation
• Distillation

Table 3.5 Parameters Affecting Heat Pump Selection


Temperature
Lift (°C) Heat Source Heat Sink Heat Pump Type

<38 Sensible cooling of liquid Sensible heating of gas Closed-cycled mechanical


or liquid
Boiling liquid Absorption (LiBr-water)
Partial condensation of liquid Sensible heating of gas Closed-cycled mechanical
from vapor stream or liquid
Boiling liquid Absorption (LiBr-water)
Condensing steam Evaporation of water Open-cycle mechanical (single-stage
compressor)
Thermocompression
Condensing vapor (steam or Boiling liquid Semi-open-cycle mechanical
others) Sensible heating of gas (single-stage compressor)
or liquid
>38 All heat sources (except All heat sink (except Absorption (with high lift working fluid)
steam) steam) Multistage mechanical compression
Low pressure steam Higher-pressure steam Open-cycle mechanical
header Absorption (with high lift working fluid)
Multistage mechanical compression
Compiled from Chua et al. (2010)

3.6 COMBUSTION-BASED PROCESS HEATING


In combustion-based process heating systems, heat is produced during the
combustion of fuel and transferred to a process or material where it is needed.
Fossil fuels are commonly used in combustion-based process heating, but bio-
mass fuels are also employed. In the case of direct heating, combustion prod-
ucts (i.e., flue gases) directly contact the materials requiring heating, while in
the case of indirect heating, the flue gases do not come into direct contact with
the material but heat is nonetheless transferred (e.g., via a radiant panel)
(Lawrence Berkeley National Laboratory, 2007).
110 C HA PT E R 3 : Industrial Heating and Cooling Systems

Raw material input

Kiln hood Flame

Burner pipe
FIGURE 3.5 Rotary kiln structure.

As an example, consider cement production. This process utilizes combustion-


based process heating, in the step where clinker is produced by heat treatment
in a rotary kiln machine. The raw material is heated to around 1500°C in the
kiln. Calcium carbonate dissociates at this temperature to form cement clinker,
which contains hydraulic calcium silicates. To heat the materials to this very
high temperature, a flame is required at a temperature of about 2000°C, which
is producible using fossil fuels. The clinker is subsequently cooled and stored
until it is ready for grinding, after which it is used to produce cement. The struc-
ture, a simple rotary kiln, is shown in Fig. 3.5.
Considering the rotary burner as an example of combustion-based heating pro-
cess, an energy analysis is described. The energy analysis is used to assess the
system and to aid efforts to reduce heat losses and improve heat recovery.
Steady-state behavior is assumed. Then, a mass rate balance for the rotary kiln
can be written as follows:
X X
m_ in ¼ m_ out (3.1)

where
X
m_ in ¼ m_ air + m_ fuel + m_ raw (3.2)
X
m_ out ¼ m_ clinker + m_ dust + m_ gases (3.3)

An energy rate balance for the rotary kiln expresses the principle of conservation
of energy and can be written as
X X X
E_in ¼ E_out + Q_ L (3.4)

Alternatively, we can express the exergy rate balance, by substituting terms, as


follows:

m_ airin hairin + m_ fuelin hfuelin + m_ rawin hrawin ¼ m_ clinker hclinker + m_ gases hgases + Q_ L + m_ dust hdust
(3.5)
3.7 Electric Process Heating 111

The use of exergy analysis is uncommon in industrial processes, despite its ben-
efits, which suggest that it may be used more regularly in industry in the future.
For the kiln, the exergy inputs to the system are the coal, raw material, and air.
The exergy outputs of the kiln machine are exergy of clinker product, kiln
exhaust, and heat loss across kiln systems. An exergy rate balance for the kiln,
accounting for exergy input, output, and destruction rates, can be written as
X X X X
_ in ¼
Ex _ out +
Ex _ Q+
Ex _ dest
Ex (3.6)

or, by substituting terms, as


m_ airin exairin + m_ fuelin exfuelin + m_ rawin exrawin ¼ m_ clinker exclinker + m_ gases exgases
 
_ T0 _ dest
+ QL 1  + Ex (3.7)
Ts

The energy efficiency of the rotary kiln can be expressed in general terms as
Energy in products output Energy loss
η¼ ¼1 (3.8)
Energy inputs Energy inputs

More specifically, this energy efficiency can be expressed as follows:


m_ clinker hclinker
η¼ (3.9)
m_ fuel hfuel

Similarly, the exergy efficiency of the rotary kiln can be expressed in general and
specific terms, respectively, as follows:
Exergy in useful products output
ψ¼ (3.10)
Exergy inputs

m_ clinker exclinker
ψ¼ (3.11)
m_ fuel exfuel

3.7 ELECTRIC PROCESS HEATING


Electric process heating involves the use of an electric current or electromag-
netic field to produce heat for processes or heating materials. In this direct elec-
tric heating system, heat is generated within the workpiece, using one of several
methods:

(a) Passing electric current through the materials


(b) Inducing electric current into the materials
(c) Exciting atoms in the materials by using electromagnetic radiation

Electric process heating is utilized in many industries, including the iron and
steel industry, which is one of the largest industrial energy consumers globally.
112 C HA PT E R 3 : Industrial Heating and Cooling Systems

For instance, Camdali and Tunc (2003) examined an electric arc furnace having
55 tons of casting capacity. In electric arc furnaces, electric arcs pass from the
electrodes to a metal. This electric current through the metal charge causes heat
generation due to electric resistance of the metal.
The exergy rate balance for an electric arc furnace can be written as follows:
X X X X
_ in ¼
Ex _ out +
Ex _ Q+
Ex _ dest
Ex (3.12)

or
_ elect + Ex
m_ scrap exscrap + m_ cwin excwin + Ex _ chr ¼ m_ ls exls + m_ cwout excwout + m_ dst exdst
(3.13)
_ dest
+ m_ slg exslg + m_ st exst + m_ sg exsg + Ex

The exergy efficiency can be expressed as the ratio of the exergy rates exiting the
electric arc furnace as products to the exergy rates entering. The exergy efficiency
of the electric arc furnace can be expressed in general and specific terms, respec-
tively, as follows:
Exergy in useful products output
ψ¼ (3.14)
Exergy inputs

m_ st exst
ψ ¼X (3.15)
Ex_ in

As an example, it is pointed out that, for the above-mentioned electric arc fur-
nace with 55 tons of casting capacity, Camdali and Tunc (2003) determined the
exergy efficiency to be 55%.

3.8 STEAM-BASED PROCESS HEATING SYSTEMS


In steam-based heating systems, steam supplies process heat. In direct steam-
based heating systems, steam is injected into the liquids or gases in the process.
In indirect steam-based heating systems, a heat exchanger is used. Then, the
steam is cooled and condensed in the heat exchanger tubes as heat is transferred
to the liquids and gases on the other side of the heat exchanger. Steam process
heating has many advantages and can be used in conjunction with various by-
product fuels.
Rosen and Dincer (2004) applied exergy analysis to process heating applica-
tions in various plants. For instance, they investigated the factors affecting
steam production for industrial heating using nuclear energy at the Bruce
Energy Centre in Ontario, Canada. The authors concluded that industries with
process heat requirements can benefit from using steam to supply some or all of
their heating requirements.
3.9 Case Study 113

Nuclear generating station

1 2
Steam transformer plant

4 3

5 Energy center 6

FIGURE 3.6 Steam supply system at the Bruce Energy Centre. Rosen and Dincer, 2004.

As the case of the Bruce Energy Centre is quite instructive, it is examined in


greater detail. Figure 3.6 shows the Bruce Energy Centre steam supply system.
The heat supply needs to satisfy the energy and exergy requirements of the pro-
cess simultaneously. The required heat temperature can be determined as the
ratio of exergy requirement to energy requirement. A high ratio specifies
high-temperature heat is required, while a low ratio indicates the opposite.
The exergy supplied is in theory equal to the exergy required, but in practice,
this ideal condition is never met due to the required minimum temperature dif-
ferences for heat transfer and losses. However, keeping temperature differences
as small as possible lowers the exergy destruction, raising the efficiency towards
the optimum (Rosen and Dincer, 2004).

3.9 CASE STUDY


As a case study, we consider an integrated system composed of the following:
a solar thermal system that raises the temperature of a heat transfer fluid
(HTF), a heat pump with a heat recovery unit for industrial heating, and var-
ious textile processes where hot water is used for dying, cleaning, and ironing/
pressing. The system is illustrated in Fig. 3.7. To have an effective solar heating
system, its working fluid requires a high critical temperature (Al-Sulaiman
et al., 2010), as this permits an increased amount of heat to be transferred
from the solar system to the HTF. Therminol VP-1 is used as the HTF in
the case study. This fluid has several benefits over other fluids. For instance,
it can withstand high temperatures and has good corrosion resistance quali-
ties. In the integrated system, the temperature of the HTF can be raised to
114 C HA PT E R 3 : Industrial Heating and Cooling Systems

I
3r
V

1a 2r
Solar
IV II collector
5a used for
Process steam
4r III 1r Process Process Boiler
generation

4a V

`
I Pump
3a 2a

VII

I Condenser
II Compressor
III Evaporator
Make-up
IV Expansion valve
water
V Supply fan
VI Suction fan
VII Heat recovery unit

FIGURE 3.7 Integrated heating system incorporating solar thermal energy and a heat pump.

150 °C by the solar collector (Rowshanzadeh, 2010), and this is used as an


input to the textile ironing process through steam. In the ironing process, a
proper amount of heat is required to avoid damaging clothes and causing tex-
ture shining. After the ironing process, the hot water passes to other processes
such as dying and cleaning, which require heat at temperatures of around 60-
90°C. In another part of the system, a heat pump is also used for the low-
temperature process of cleaning and dying, transferring heat through warm
air. The heat pump has a heat recovery unit that allows returned air to be uti-
lized for preheating in the evaporator. In this heat recovery system, energy is
recovered to raise efficiency.

3.9.1 Analysis
In the analysis of the system’s performance and behavior, various quantities are
determined, including inlet and outlet specific enthalpies, specific exergies,
mass flow rates, pressures, and temperatures. Also, exergy destructions are cal-
culated to pinpoint the system irreversibilities.
To model the system and simplify the analysis, several typical assumptions are
made:

• The reference (dead) state for the system has an environmental


temperature T0 ¼ 20°C and a pressure P0 ¼ 101.325 kPa.
• The system operates at steady-state conditions.
• The changes in the kinetic and potential energy terms are negligibly small
compared to other terms in the energy and exergy balances.
• The isentropic efficiency of compressors and pumps is taken as 90%.
3.9 Case Study 115

• The compressors and pumps are adiabatic.


• At the exit of the condenser and the evaporator of the heat pump, the
refrigerant (water) is in a saturated state.
• Pressure losses in all heat exchangers and the pipelines are negligible.
• Air behaves like an ideal gas.
Energy and exergy rate balances for the important sections of the integrated sys-
tem are expressed and described in the following sections. These are based on
the general balances for mass, energy, and exergy, respectively, which can be
written for a steady-state process as follows:
X X
m_ i ¼ m_ e (3.16)
i e

X X
Ei + Q_ ¼ E_e + W_ (3.17)
i e

X X 
T0 _ X _
_ i+
Ex 1 Q ¼ Exe + W_ + Ex
_ dest (3.18)
i i
Ti i e

Various types of heating efficiencies can be defined. Rosen and Dincer (2004)
gave electrical heating efficiencies as follows:
Energy efficiency ¼ ðEnergy output with productsÞ=ðTotal energy inputÞ

Q_ p
ηe ¼ (3.19)
W_ e

Exergy efficiency ¼ ðExergy output with productsÞ=ðTotal exergy inputÞ


 
_ T0
Qp 1 
Tp
ψe ¼ (3.20)
W_ e

Similarly, for fuel heating,

Q_ p
ηf ¼ (3.21)
_
mf HHV
 
_ T0
Qp 1 
Tp
ψf ¼ (3.22)
m_ f exf

Here, Tp is the temperature associated with the product heat rate, Q_ p ; T0 is


the temperature of the environment; Ẇe is the electrical energy rate; and ṁf,
HHV, and exf are the mass flow rate, specific higher heating value, and specific
exergy, respectively, for the fuel.
116 C HA PT E R 3 : Industrial Heating and Cooling Systems

3.9.2 Heating Devices


In a building environment, heating is normally done using hot water, warm
air, steam, radiant heat, or electric heat. To produce the high-temperature
water, air, and steam needed in many industrial operations, fuels such as nat-
ural gas, oil, and coal or electric energy can be used. According to ASHRAE
(2005), various types of heating equipment exist and are suitable for different
applications. The main types of heating equipment according to ASHRAE
follow:

• Burners
• Furnaces
• Boilers
• Heat exchangers
• Combustors
• Terminal units
• Electric heaters

3.9.2.1 Burner
In a gas burner, a fuel such as gas and a mixture of gas and air are supplied to the
combustion zone. Fuel gas is injected from a high-pressure source through an
injector nozzle so that it forms a gas jet. Separately, primary air is provided to
the burner throat by venture action. Fuel gas and air are mixed in a mixing tube
and form a stoichiometric mixture. The rate balance equations for burner are
given below:

Mass balance equation : m_ f + m_ a ¼ m_ p (3.23)

Energy balance equation : m_ f hf + m_ a ha ¼ m_ p hp + Q_ loss (3.24)

Entropybalance equation : m_ f sf + m_ a sa + S_gen ¼ m_ p sp + Q_ loss (3.25)

 
T0
Exergy balance equation : m_ f exf + m_ a exa ¼ m_ p exp + Q_ l 1  _ dest
+ Ex (3.26)
Ti

3.9.2.2 Furnace
The furnace is another device for residential and industrial heating. For large
industrial heating applications, several types of furnaces are available, includ-
ing kilns for the cement industry, blast furnaces for the metal industry, and
electric arc furnaces for steelmaking. For residential and industrial HVAC
applications, fuel burning furnaces are common in which combustion occurs
in a combustion chamber. The air passes over the outside surface of the furnace
3.9 Case Study 117

Gases

1
Air
3
Combustion
Furnace gases
Fuel
2
FIGURE 3.8 Furnace.

heat exchanger and does not directly contact the combustion products. Resi-
dential and industrial furnaces come in various capacities. The input and out-
put flow over the furnace is shown in Fig. 3.8.
The mass, energy, entropy, and exergy rate balance equations, respectively, for a
furnace are as follows:

m_ f + m_ a ¼ m_ cg (3.27)

m_ f hf + m_ a ha + Q_ in ¼ m_ cg hcg (3.28)

Q_ in _
m_ f sf + m_ a sa + + Sgen ¼ m_ cg scg (3.29)
Ts
 
T0
m_ f exf + m_ a exa + Q_ in 1  _ dest
¼ m_ cg excg + Ex (3.30)
Ts

3.9.2.3 Boiler
Boilers and furnaces are commonly used heating devices in the industrial sec-
tor. Most industries in which steam is generated for heating use boilers (Saidur
et al., 2010). A boiler is basically a pressure vessel that is designed to heat and
vaporize a fluid, usually water. The respective mass, energy, entropy, and exergy
rate balance equations for a boiler can be written as follows:

m_ in ¼ m_ out (3.31)

m_ in hin + Q_ in ¼ m_ out hout (3.32)

Q_ in _
m_ in sin + + Sgen ¼ m_ out sout (3.33)
Ts
 
T0
m_ in exin + Q_ in 1  _ dest
¼ m_ out exout + Ex (3.34)
Ts
118 C HA PT E R 3 : Industrial Heating and Cooling Systems

⋅ ⋅
1 2

⋅ ⋅
1 2

FIGURE 3.9 Heat exchanger.

3.9.2.4 Heat Exchangers


Mass, energy, entropy, and exergy rate balance equations, respectively, for a
heat exchanger can be expressed as follows and energy flow of heat exchanger
is shown in Fig. 3.9:
m_ 1in + m_ 2in ¼ m_ 1out + m_ 2out (3.35)

m_ 1in h1in + m_ 2in h2in ¼ m_ 1out h1out + m_ 2out h2out (3.36)

m_ 1in s1in + m_ 2in s2in + S_gen ¼ m_ 1out s1out + m_ 2out s2out (3.37)

_ dest
m_ 1in ex1in + m_ 2in ex2in ¼ m_ 1out s1out + m_ 2out s2out + Ex (3.38)

3.9.3 Rate Balances Equation for Remainder of System


3.9.3.1 Condenser
Two mass rate balances can be written for the condenser:
m_ RðinÞ ¼ m_ RðoutÞ (3.39)

m_ airðinÞ ¼ m_ airðoutÞ (3.40)

Similarly, energy, entropy, and exergy rate balances for the condenser can be
written as follows:

m_ airðinÞ hairðinÞ + m_ RðinÞ hRðinÞ ¼ m_ airðoutÞ hairðoutÞ + m_ RðoutÞ hRðoutÞ + Q_ Heating (3.41)

Q_ Heating
m_ airðinÞ sairðinÞ + m_ RðinÞ sRðinÞ + S_gen ¼ m_ airðoutÞ sairðoutÞ + m_ RðoutÞ sRðoutÞ + (3.42)
T0
 
T0
m_ airðinÞ exairðinÞ + m_ RðinÞ exRðinÞ ¼ m_ airðoutÞ exairðoutÞ + m_ RðoutÞ exRðoutÞ + Q_ Heating 1  _ dest
+ Ex
Ts
(3.43)
3.9 Case Study 119

3.9.3.2 Expansion Valve


The mass, energy, entropy, and exergy rate balance equations, respectively, for
an expansion valve can be written as follows:

m_ RðinÞ ¼ m_ RðoutÞ (3.44)

hRðinÞ ¼ hRðoutÞ (3.45)

m_ RðinÞ sRðinÞ + S_gen ¼ m_ RðoutÞ sRðoutÞ (3.46)

_ dest
m_ RðoutÞ exRðoutÞ ¼ m_ RðoutÞ exRðoutÞ + Ex (3.47)

3.9.3.3 Evaporator
Two mass rate balances can be written for the evaporator:

m_ RðinÞ ¼ m_ RðoutÞ (3.48)

m_ airðinÞ ¼ m_ airðoutÞ (3.49)

Similarly, the energy, entropy, and exergy rate balance equations for the con-
denser can be written as follows:

m_ airðinÞ hairðinÞ + m_ RðinÞ hRðinÞ ¼ m_ airðoutÞ hairðoutÞ + m_ RðoutÞ hRðoutÞ + Q_ cooling (3.50)

Q_ cooling
m_ airðinÞ sairðinÞ + m_ RðinÞ sRðinÞ + S_gen ¼ m_ airðoutÞ sairðoutÞ + m_ RðoutÞ sRðoutÞ + (3.51)
T0
 
T0
m_ airðinÞ exairðinÞ + m_ RðinÞ exRðinÞ ¼ m_ airðoutÞ exairðoutÞ + m_ RðoutÞ exRðoutÞ + Q_ cooling 1  _ dest
+ Ex
Ts
(3.52)

3.9.3.4 Compressor
The mass, energy, entropy, and exergy rate balance equations, respectively, for a
compressor can be written as follows:

m_ Rðcomp_inÞ ¼ m_ Rðcomp_outÞ (3.53)

m_ RðinÞ hRðinÞ + W_ in ¼ m_ RðoutÞ hRðoutÞ (3.54)

m_ ref ðinÞ hRðinÞ + W_ in + S_gen, comp ¼ m_ RðoutÞ sRðoutÞ (3.55)

m_ RðinÞ exRðinÞ + W_ in ¼ m_ RðoutÞ exRðoutÞ + Ex


_ dest (3.56)
120 C HA PT E R 3 : Industrial Heating and Cooling Systems

3.9.3.5 Fan
The mass, energy, entropy, and exergy rate balance equations, respectively, for a
fan can be written as follows:
m_ airðinÞ ¼ m_ airðoutÞ (3.57)

v2
m_ airðinÞ hairðinÞ + Wfan ¼ m_ airðoutÞ hairðoutÞ + m_ airðoutÞ (3.58)
2
2
v
m_ airðinÞ sairðinÞ + Wfan + S_gen, fan ¼ m_ airðoutÞ sairðoutÞ + m_ airðoutÞ (3.59)
2
_ dest
m_ airðinÞ exairðinÞ + Wfan ¼ m_ airðoutÞ exairðoutÞ + Ex (3.60)

3.10 RESULTS AND DISCUSSION


Comprehensive energy and exergy analyses of the integrated system are per-
formed and results are presented in Tables 3.6 and 3.7. The assumptions stated
above are invoked and the reference-environment conditions are taken to be
T0 ¼ 20°C and P0 ¼ 101 kPa.

Table 3.6 Input and Calculated Process Data for the System in Fig. 3.7
State Fluid Type P (kPa) ṁ (kg/s) T (K) h (kJ/kg) ex (kJ/kg)

1a Air 101.325 4.2 333 333.6 2.4


2a Air 101.325 4.2 312 311.3 1.2
3a Air 101.325 4.2 302 302.6 0.1
4a Air 101.325 4.2 285 285.4 0.1
5a Air 101.325 4.2 308 308.5 0.3
1r R407C 490 1.5 283 418.9 42.7
2r R407C 1700 1.5 357 472.4 77.9
3r R407C 1600 1.5 314 261.7 60.9
4r R407C 500 1.5 271.2 271.7 124.4
6a Air 101.325 4.2 335 336 2.7
1VP1 Therminol VP1 1000 3.5 423 238 40.9
2VP1 Therminol VP1 700 3.5 343 93.7 6.8
3VP1 Therminol VP1 1000 3.5 348 102.5 –
1w Water 1800 2.0 343 294 17.7
2w Water 1750 2.0 403 547 71.5
3w Water 1750 2.0 403 547 71.5
4w Water 800 2.0 313 168 –
3.10 Results And Discussion 121

Table 3.7 Measures of Merit for Major Components of System


Measure of Merit and Component Value

Energy efficiency of condenser 66.2%


Exergy efficiency of condenser 60.3%
Energy efficiency of evaporator 35.4%
Exergy efficiency of evaporator 34.0%
Energetic COP of cycle 3.54
Exergy efficiency of cycle 42.5%
Energetic COP of system 2.97
Exergy efficiency of system 35.7%
Energy efficiency of process 36.7%
Exergy efficiency of process 75.0%

3.10.1 Efficiencies and Other Measures of Merit


The exergy efficiencies of the heat pump cycle and overall system are found to
be 42.5% and 35.7%, respectively. These values may seem low because there are
no heat losses because the system is assumed adiabatic, losses due to exergy
destructions make up the difference. Moreover, the energetic coefficients of per-
formance (COPs) of the heat pump cycle and overall system are 3.54 and 2.97,
respectively. The exergy efficiency of the textile process is found to be 75.0%,
which is much greater than the corresponding energy efficiency of 36.7%.
The individual energy and exergy efficiencies of condenser are found to be
66.2% and 60.3%, respectively. The energy and exergy efficiencies of evaporator
are found to be 35.4% and 34.0%, respectively, which are comparatively lower
than the condenser efficiencies because of the high exergy destruction rate in
the evaporator. Table 3.6 provides the thermodynamic properties of each state
in the system, as determined using Engineering Equation Solver. Table 3.7 lists
the measures of merit for the main components and overall system, including
efficiencies and energetic COPs of the heat pump cycle.

3.10.2 Exergy Destruction Rates


The exergy analyses reveal the components of the system with the largest exergy
destruction rates (see Figs. 3.10 and 3.11). The exergy destruction rate in the
ironing process is greatest for all processes considered here due to the large tem-
perature difference between the two fluids in the heat exchanger. The second
highest value occurs in the evaporator, for which the exergy destruction rate
is around 110 kW. This exergy destruction rate can be reduced by implementing
suitable temperature and pressure parameters for different components in the
system. The largest exergy destruction rates are observed in heat transfer devices.
Selecting a more suitable working fluid also has a significant effect on the
122 C HA PT E R 3 : Industrial Heating and Cooling Systems

160

Exergy destruction rate (kW)


140
120
100
80
60
40
20
0

or

or

s
so

to

es
lv
Fa

m
ns

at

oc
va

ra
es

pu
or
de
pr

ne

pr
n
ap

r
on
om

io

ge

la

ile
Ev

ns
C

So
m

xt
C

pa

ea

Te
Ex

St
FIGURE 3.10 Exergy destruction rates of main components of the system.

0.45
0.40
Exergy destruction ratio

0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
e

or

or

s
to

so

es
lv

Fa

m
ns

at
va

oc
ra

es
pu

or
de

pr
ne

pr
n

ap
r
on

om
io

ge

la

ile
Ev
ns

So
m

xt
C
pa

ea

Te
Ex

St

FIGURE 3.11 Exergy destruction ratio of main components of system.

energy and exergy efficiencies of the system. The condenser, the fan, and the
expansion valve exhibit the lowest exergy destruction rates.

3.10.3 Parametric Analysis


We now describe how the performance of the system in terms of energy and
exergy efficiencies is affected as selected design parameters vary, such as ambi-
ent temperature and inlet pressure and temperature for several components.
The design parameters selected for consideration are those that play a signifi-
cant role in system performance.
3.10 Results And Discussion 123

3.10.3.1 Effect of Ambient Temperature


The reference-environment temperature often has an impact on the perfor-
mance of a system (IEA, 2008). The effect of varying the reference-environment
temperature on the performance of the heat pump cycle considered here is
shown in Fig. 3.12. As the reference-environment temperature increases from
18 °C to 30 °C, the exergy efficiency decreases from 44% to 35%. However, this
variation in reference-environment temperature does not affect the energetic
COP because it is independent of reference-environment conditions. Similarly,
it can be seen in Fig. 3.13 that an increase in the reference-environment

0.46 4

3.5
0.44
3
0.42
2.5

COPen
hex

0.4 2

1.5
0.38
hex
1
0.36 COPen
0.5

0.34 0
292 294 296 298 300
T0 (K)

FIGURE 3.12 Effect of variation of ambient temperature on COP and exergy efficiency of heat pump cycle.

0.4 3.5

0.38 3

0.36 2.5

0.34 2
hex,sys

COPen

0.32 1.5
hex
0.3 1
COPen
0.28 0.5

0.26 0
291 294 297 300 303
T0(K)

FIGURE 3.13 Effect of variation of ambient temperature on COP and exergy efficiency of overall system.
124 C HA PT E R 3 : Industrial Heating and Cooling Systems

0.55 120

0.5 hex
Exd,comp 96
Exd,cond
0.45
Exd,eva

Exd (kW)
72
Exd,fan

hex
0.4
48
0.35

24
0.3

0.25 0
285 290 295 300 305
T0 (K)

FIGURE 3.14 Effect of variation of ambient temperature on exergy destruction rate and efficiency of
various components of the system.

temperature reduces the exergy efficiency from 39% to 26%, while the ener-
getic COP is unaffected. The decrease in the exergy efficiency with respect to
reference-environment temperature is due to the fact that the exergy destruction
rises when the reference-environment temperature rises. The increase in
exergy destruction, and the corresponding decrease in exergy efficiency, as
reference-environment temperature rises is shown for various components in
Fig. 3.14.

3.10.3.2 Effect of Pressure and Temperature of Condenser


The condenser pressure and temperature affect the performance of the heat
pump cycle and the overall system. Figure 3.15 shows the effect of increasing
condenser inlet pressure on the exergy efficiency and energetic COP of the heat
pump cycle. When the inlet pressure is raised from 1600 kPa to 1800 kPa, the
exergy efficiency of the heat pump cycle rises from 41% to 43% and energetic
COP rises from 3.4 to 3.6. But the outlet temperature from the condenser
exhibits a negative effect on the performance of the system, as shown in
Fig. 3.16. Increasing the outlet temperature from 77 °C to 97 °C causes both
the exergy efficiency and the energetic COP of the system to decrease.

3.10.3.3 Effect of Varying Refrigerant and Water Temperature in Heat


Pump Cycle
The system performance is observed in Fig. 3.17 to be negatively affected by
increasing the refrigerant temperature in the heat pump cycle, with the exergy
efficiency decreasing from 48.0% to 36.0% and the energetic COP decreasing
3.10 Results And Discussion 125

0.435 3.65
COPen
hex
0.43 3.6

COPen
hex

0.425 3.55

0.42 3.5

0.415 3.45
1600 1650 1700 1750 1800
Inlet pressure of condenser (kPa)
FIGURE 3.15 Effect of variation of inlet pressure of condenser on COP and exergy efficiency of heat
pump cycle.

0.4 3.4
heex
COPen
0.38 3.2

0.36 3
COPen,sys
hex,sys

0.34 2.8

0.32 2.6

0.3 2.4
350 355 360 365 370
Outlet temperature of condenser (K)
FIGURE 3.16 Effect of variation of outlet temperature of condenser on COP and exergy efficiency of
overall system.

from 4.0 to 3.0. The textile process efficiencies are positively affected by increas-
ing the outlet temperature of the water from the steam generator at point 2 w
(see Fig. 3.18). This water is heated only by the solar source via the solar col-
lector and receives no other input. The energy efficiency of the heating process
through solar system rises from 36.0% to 42.0% and the exergy efficiency
increases from 74.0% to 83.0%.
126 C HA PT E R 3 : Industrial Heating and Cooling Systems

0.48 4
COPen
0.46
hex 3.8

0.44
3.6
0.42

COPen
hex
3.4
0.4
3.2
0.38

0.36 3

0.34 2.8
350 355 360 365 370
Tref (K)

FIGURE 3.17 Effect of variation of refrigeration temperature of compressor on COP and exergetic
efficiency of overall system.

0.46 0.86
hex
0.44 0.84
hen
0.82
0.42
0.8
hen

hex
0.4
0.78
0.38
0.76

0.36 0.74

0.34 0.72
400 410 420 430
Outlet temperature of water (K)
FIGURE 3.18 Effect of varying supply temperature of water on efficiencies of heating process.

3.11 FURTHER DISCUSSION


For the integrated solar and heat pump-based system for industrial heating sys-
tem considered, energy and exergy analyses are performed. The system com-
prises a heat pump cycle for process heating of water and solar energy for
another industrial heating process. These options are intended to be more
3.12 Closing Remarks 127

environmentally benign than conventional systems. The energy efficiency of


the heating process is seen to be 37% and the exergy efficiency 75%. The ener-
getic COP of the heat pump cycle is 3.54, whereas its exergy efficiency is 42.5%.
Moreover, the energetic COP of the overall system is 2.97, while the exergy effi-
ciency is 35.7%. Further analyses could provide additional insights, such as
investigating the impact on performance and efficiency of such additional
parameters as solar radiation levels and working fluid choice.

3.12 CLOSING REMARKS


This chapter covers the energy and exergy analyses of industrial heating and
cooling. It demonstrates how exergy analysis depicts the actual performance
of a process clearly and meaningfully. Exergy methods are being increasingly
applied in industrial processes, for determining the exergy destruction rates
and efficiencies of components. The case study in this chapter considers an
industrial textile heating process and the exergy destruction rate of each com-
ponent is determined. These exergy results are useful as they help to improve
the efficiency of an overall system and its components.

Nomenclature
Ė exergy rate (kW)
ex specific exergy (kJ/kg)
h specific enthalpy (kJ/kg)
COP coefficient of performance
ṁ mass flow rate (kg/s)
P pressure (kPa)
Q heat transfer (kJ)
Q_ heat rate (kW)
s specific entropy (kJ/kgK)
Ṡ entropy generation
T temperature (K or °C)
t time (s)
ρ density (kg/m3)
v specific volume (m3/kg)
Ẇ work rate (kW)
V volume (m3)

Greek symbols
η energy efficiency
ψ exergy efficiency
ω specific humidity or humidity ratio (kg/kg)
128 C HA PT E R 3 : Industrial Heating and Cooling Systems

Subscripts
a air
cond condenser
comp compressor
eva evaporator
ref refrigerant
d destruction
en energy
ex exergy
geo geothermal
w water
prod product
avg average

Acronyms
HEX heat exchanger
HTF heat transfer fluid
PTSC parabolic trough solar collector

References
Al-Sulaiman, F.A., Dincer, I., Hamdullahpur, F., 2010. Exergy analysis of an integrated solid oxide
fuel cell and organic Rankine cycle for cooling, heating and power production. J. Power Sources
195, 2346–2354.
American Society of Heating, Refrigerating and Air-Conditioning Engineers 2005 “Fundamentals”
ASHRAE, Inc. Atlanta, GA.
Brown, H.L., Hamel, B.B., Edman, B.A., Koluch, M., Gajanana, B.C., Troy, P., 1985. Analysis of 10
Industrial Processes. Library of congress, Cataloging-in-publication data, Fairmount Press,
Philadelphia.
Camdali, U., Tunc, M., 2003. Exergy analysis and efficiency in an industrial AC electric arc furnace.
Appl. Therm. Eng. 23, 2255–2267.
Chua, K.J., Chou, S.K., Yang, W.M., 2010. Advances in heat pump systems: a review. Appl. Energy
87, 3611–3624.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, second
ed. Elsevier, Oxford, UK.
International Energy Agency, 2008. Combined Heat and Power. OECD/IEA, Paris. http://www.iea.
org/publications/freepublications/publication/chp_report.pdf.
Kalogirou, S., 2003. The potential of solar industrial process heat applications. Appl. Energy
76, 337–361.
Kanoglu, M., Bolatturk, A., 2008. Performance and parametric investigation of a binary geothermal
power plant by exergy. Renew. Energy 33, 2366–2374.
Lawrence Berkeley National Laboratory, 2007. Improving Process Heating System Performance:
A Source Book for Industry. U.S. Department of Energy, Berkeley, CA.
Mittal, V., Kasana, K.S., Thakur, N.S., 2005. The study of solar absorption air-conditioning systems.
J. Energy South Africa 16, 59–66.
References 129

Ratlamwala, T.A.H., Dincer, I., 2013. Development of a geothermal based integrated system for
building multigenerational needs. Energy Build. 62, 496–506.
Rosen, M.A., Dincer, I., 2004. A study of industrial steam process heating through exergy analysis.
Int. J. Energy Res. 28, 917–930.
Rowshanzadeh, R., 2010. Performance and Cost Evaluation of Organic Rankine Cycle at Different
Technologies. (Master thesis). Department of Energy Technology, Kungliga Tekniska Hogsko-
lan (KTH), Sweden.
Saidur, R., Ahamed, J.U., Masjuki, H.H., 2010. Energy, exergy and economic analysis of industrial
boilers. Energy Policy 38, 2188–2197.
Seyboth, K., Beurskens, L., Langniss, O., Ralph, E.H., 2008. Recognising the potential for renewable
energy heating and cooling. Energy Policy 36, 2460–2463.
Suleman, F., Dincer, I., Agelin-Chaab, M., 2014a. Development of an integrated renewable energy
system for multigeneration. Energy 78, 1–9.
Suleman, F., Dincer, I., Agelin-Chaab, M., 2014b. Energy and exergy analyses of an integrated solar
heat pump system. Appl. Therm. Eng. 73, 557–564.
Utlu, Z., Hepbasli, A., 2007. A review and assessment of the energy utilization efficiency in the Turk-
ish industrial sector using energy and exergy analysis method. Renew. Sust. Energ. Rev.
11, 1438–1459.
CHAPTER 4

Heat Pump Systems

4.1 INTRODUCTION
The principles governing the operation of the heat pump have been recognized
since the 1800s and are the basis of most refrigeration. The idea of using a heat
engine in reverse mode, as a heat pump, was proposed by Lord Kelvin, but it
was only in the twentieth century that practical machines began to be used,
mainly for refrigeration. Beginning in the 1970s, air source heat pumps started
to come into common use. They have the advantage of being combustion-free
and thus do not generate indoor pollutants like carbon monoxide. Heat pumps
are also competitive in terms of installation cost with central combustion
furnace/central air conditioner combinations. Hence, heat pumps now rou-
tinely provide central air conditioning and heating for buildings.
Today, heat pumps are widely used not only for space cooling and heating but
also for chilling materials, producing hot water, and preheating feedwater in
various types of facilities, including office and institutional buildings, com-
puter centers, restaurants, hotels, district heating and cooling systems, and
industrial plants (Zamfirescu and Dincer, 2009).
Efficient energy use, facilitated by such measures such as waste heat recovery,
and the application of renewable energy can mitigate carbon dioxide emissions
and global warming. A heat pump system can contribute to this objective, by
providing effective and efficient cooling and heating. It is noted in the latter
case that a heat pump normally delivers more thermal energy than the electrical
energy required to operate it.
A significant portion of global energy consumption is attributable to domestic
heating and cooling. Heat pumps are advantageous and widely used in many
applications due to their high utilization efficiencies compared with conven-
tional heating and cooling systems. Utilization of heat pump systems often
leads to three significant benefits: environmental, economic, and technologi-
cal. Preservation of the natural environment can be assisted by replacing heat-
ing and cooling devices that are driven directly or indirectly by fossil fuel 131

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00004-1


© 2015 Elsevier Inc. All rights reserved.
132 C HA PT E R 4 : Heat Pump Systems

combustion with heat pumps due to the higher efficiencies. The benefits
are even more pronounced when the electricity to drive the heat pump is
derived from renewable energy sources (e.g., hydro, wind, and solar). In terms
of the economy, the use of heat pumps can facilitate reductions in oil
import and transport costs and heating costs, with a parallel increase in private
purchasing power.
Heat pump technology is sufficiently simple and mature that it enables
installers to apply standardized designs that help keep initial costs relatively
low for both new construction and retrofits of existing systems. Good reliability
and efficiency are additional beneficial characteristics of heat pumps, permit-
ting them to play an important role in reducing energy use in societies. This
partly explains the sudden increase in heat pump use often observed during
“energy crises.”
One of the most common heat sources for a heat pump is air, although water
and ground are also used. Recently, there has been increasing interest in using
the ground (rock and soil) as a heat source for heating and cooling applications.
Ground source heat pumps (GSHPs) have achieved a notable and growing mar-
ket share (Lubis et al., 2011). Note that by utilizing low-temperature resources
like the ground, exergy efficiencies for heating and cooling usually increase sig-
nificantly. The energy (thermal) efficiency of a heat pump, usually reflected
through a coefficient of performance (COP), is also higher with heat pumps than
simple heaters, but the higher exergy efficiency is particularly instructive.
In this chapter, heat pump cycles, systems, and applications are described, and
the technical, operational, thermodynamic, and environmental aspects are
detailed. Some illustrative examples are presented to highlight the importance
of heat pumps.

4.2 HEAT PUMP EFFICIENCIES


Various measures are used to assess the efficiency of heat pumps. For all mea-
sures, higher values imply higher efficiency. The heat pump efficiency is con-
ceptually determined by comparing the product energy delivered by the heat
pump with the amount of energy it consumes to drive it. The efficiency mea-
sures can be based on laboratory tests in controlled environments on actual
use in the field (Dincer and Rosen, 2013).
There are many performance assessment criteria for heat pump systems. Some
commonly used categories of efficiency measures for heat pumps are illustrated
in Fig. 4.1 and described below. We focus here on heat pumps used in a heating
mode, noting that corresponding information for heat pumps used in a cooling
mode is also available.
4.2 Heat Pump Efficiencies 133

Heat pump efficiencies

Coefficient of Primary energy Energy efficiency Heating season Seasonal energy


performance ratio ratio performance efficiency ratio
(COP) (PER) (EER) factor (HSPF) (SEER)

FIGURE 4.1 Summary of heat pump efficiency measures.

4.2.1 Coefficient of Performance


The COP is the most common measure of heat pump efficiency and can be
expressed as the ratio of the product heat output of a heat pump to its electrical
energy input:
COP ¼ Product heat output=Electrical energy input (4.1)

Air source heat pumps generally have COPs ranging from 2 to 4, indicating that
they deliver 2-4 times more thermal energy than they consume in terms of elec-
trical energy. Water and ground source heat pumps normally have COPs of 3-5
(Soltani et al., 2015). The COP of an air source heat pump decreases as the out-
side temperature drops. Therefore, two COP ratings are usually given for a sys-
tem: one at 8.3 °C (47 °F) and the other at 9.4 °C (17 °F). When comparing
COPs, one must be sure the ratings are based on the same outside air temper-
ature to avoid inconsistencies. COPs for ground and water source heat pumps
do not vary as widely because ground and water temperatures are more con-
stant than air temperatures (Caliskan et al., 2011).
While comparing COPs can be informative, it does not provide a complete pic-
ture. When the outside temperature drops below 4.4 °C (40 °F), the outdoor
coils of a heat pump must be defrosted periodically. The outdoor coil temper-
ature can be below the freezing temperature for water when a heat pump is in
the heating cycle. Under these conditions, moisture in the air freezes on the sur-
face of the cold coil. Eventually, enough frost can build up to keep air from
passing over the coil and to inhibit heat transfer, causing it to lose efficiency.
When the coil efficiency is reduced sufficiently to appreciably affect system
capacity, the frost must be eliminated. To defrost the coils, the heat pump cycle
normally is reversed, and heat is transferred from the house to the outdoor coil
to melt the ice. This process reduces the average COP significantly.
Another factor that lowers the overall efficiency of air-to-air heat pumps is their
inability to provide sufficient heat during the coldest days of winter. This weak-
ness usually causes a backup heating system to be required. The backup is often
provided by electric resistance heating, which has a COP of only one. When the
ambient temperature drops to the 3.8 °C to 1.1 °C range, or a different
134 C HA PT E R 4 : Heat Pump Systems

system-specific balance point, this electric resistance heating engages, and the
overall system efficiency decreases.

4.2.2 Primary Energy Ratio


Heat pumps may be driven electrically or by engines (e.g., internal combustion
engines or gas motors). Unless electricity is derived from an alternative source
(e.g., hydro, wind, and solar), heat pumps also utilize primary energy sources
upstream or on-site, as in the case of a natural gas motor. When comparing heat
pump systems driven by different energy sources, it is appropriate to use the pri-
mary energy ratio (PER), defined as the ratio of useful heat delivered to primary
energy input (Holland et al., 1982). The PER is related to the COP as follows:
PER ¼ ηCOP (4.2)

where η is the efficiency with which the primary energy input is converted to
compressor shaft work.
Due to the high COP of heat pumps, their PER values can be high relative to
those for conventional fossil fuel-fired systems. In the case of an electrically
driven compressor where the electricity is generated in a coal power plant,
the efficiency η may be as low as 25%. The PER expression indicates that gas
engine-driven heat pumps are very attractive from a PER point of view since
values for η up to 75% can be obtained. However, heat recovery systems tend
to be judged on their potential financial savings, rather than their potential
energy savings.

4.2.3 Energy Efficiency Ratio


The energy efficiency ratio (EER) is used for evaluating the efficiency of a heat
pump in the cooling cycle. The EER is defined as the ratio of cooling capacity
provided to electricity consumed:
EER ¼ Cooling capacity=Electrical energy input (4.3)

The same rating system is used for air conditioners, allowing for straightforward
comparisons of different units. In practice, EER ratings higher than 10, expres-
sed in units of BTU/h per watt of total electrical input rate, are desirable.

4.2.4 Heating Season Performance Factor


A heat pump’s performance varies depending on the weather and how much
supplementary heat is required. Therefore, a more realistic performance mea-
sure, especially for air-to-air heat pumps, is evaluated on a seasonal basis. One
such measure is referred to as the heating season performance factor (HSPF) for
the heating cycle. An industry standard test for overall heating efficiency
4.3 Classification of Heat Pump Systems 135

provides an HSPF rating. Such laboratory testing accounts for the reductions in
efficiency caused by defrosting, temperature fluctuations, supplemental heat-
ing, fan operation, and on/off cycling. The HSPF is estimated as the seasonal
heating output divided by the seasonal power consumption:
HSPF ¼ Total seasonal heating output=Total electrical energy input (4.4)

The higher the HSPF, the more efficient the heat pump on a seasonal basis is.
The HSPF can be thought of as the “average COP” for the entire heating system.
To estimate the average COP, one divides the HSPF by 3.4. Hence, an HSPF
of 6.8 corresponds roughly with an average COP of 2. HSPFs of 5-7 are
considered good.
Most utility-sponsored heat pump incentive programs require that heat pumps
have an HSPF of at least 6.8. Many heat pumps meet this requirement, and
some have HSPF ratings above 9. More efficient heat pumps are generally more
expensive, so financial assessments must also account for the annual energy
savings along with the added cost.

4.2.5 Seasonal Energy Efficiency Ratio


A heat pump’s performance varies depending on the weather and the amount
of supplementary heat required, so a more realistic efficiency measure can be
obtained on a seasonal basis. The seasonal energy efficiency ratio (SEER) for
the cooling cycle is such a measure. The SEER is the ratio of the total cooling
of the heat pump to the total electrical energy input during the same period,
that is,
SEER ¼ Total seasonal cooling output=Total electrical energy input (4.5)

The SEER rates the seasonal cooling performance of the heat pump. The SEER
for a unit varies depending on where it is located. SEER values of 8-10 are con-
sidered good. The higher the SEER, the more efficiently the heat pump cools.
The SEER compares the heat removed from a house or structure being cooled
and the energy used by the heat pump, including fans. The SEER is usually
noticeably higher than the HSPF since defrosting is not needed, and there usu-
ally is no need for expensive supplemental heat during conditions when air
conditioning is used.

4.3 CLASSIFICATION OF HEAT PUMP SYSTEMS


A systematic classification of the different types of heat pumps is difficult
because the classification can be made from numerous points of view, for
example, type of heat source, heat source/heat sink configuration, application,
136 C HA PT E R 4 : Heat Pump Systems

output, and type of heat pump process. If the heat is distributed via a mass flow
(e.g., warm air or warm water), this mass flow is called the heat carrier.
In the United States, heat pumps are generally classified for the heating of
buildings according to the type of heat source (firstly) and type of heat carrier
(secondly). A distinction can be made between the terms

• heat pump, covering only the refrigeration machine aspect, and


• heat pump plant, which, besides the heat pump itself, also contains the
heat source.

This differentiation is due to heat from the heat source being transferred to the
cold side of the heat pump by an intermediate circuit, the cold carrier. Another
usual classification differentiates between

• primary heat pumps that utilize a natural heat source present in the
environment, such as external air, ground, groundwater, and
surface water;
• secondary heat pumps that reuse waste heat as heat source, that is, already
used heat, such as extracted air, wastewater, and waste heat from rooms
to be cooled; and
• tertiary heat pumps that are in series with a primary or secondary heat
pump in order to raise the achieved, but still relatively low, temperature
further (e.g., for hot water preparation).

In this chapter, heat pumps are classified based on their (i) heat sources and
(ii) heat source/heat sink configurations (Fig. 4.2). Air, water, ground (rock
and soil) and geothermal, and solar are considered as heat sources. The classi-
fication of heat pumps based on source/sink configuration includes (i) water-
to-water, (ii) water-to-air, (iii) air-to-air, (iv) air-to-water, (v) ground-to-water,
and (vi) ground-to-air heat pump systems.

Heat pumps

Source Source / sink configuration


• Air • Water to water
• Water • Water to air
• Ground and geothermal • Air to air
• Solar • Air to water
• Ground to water
• Ground to air

FIGURE 4.2 Classification of heat pump systems based on heat source and heat source/heat sink
configuration.
4.3 Classification of Heat Pump Systems 137

4.3.1 Type of Heat Source


In heat pumps, air, water, and ground are considered as primary heat sources.
In practice, air is the most common source for heat pumps, while water and
ground source systems are less commonly applicable. In general, air, ground,
and groundwater are considered practicable as heat sources for small heat
pump systems, while surface water, seawater, and geothermal systems are more
suited to larger heat pump systems (Koohi-Fayegh and Rosen, 2014). As far as
low-temperature sources are concerned, groundwater or surface water, air, and
ground are most commonly used.
The characteristics of the heat source strongly affect the technical and economic
performance of a heat pump. An ideal heat source for heat pumps in buildings
has a high and stable temperature during the heating season, is abundantly avail-
able, is not corrosive or polluted, has favorable thermophysical properties, and
requires low investment and operational costs. In most cases, however, the avail-
ability of the heat source is the key factor in determining its use. Table 4.1 lists
commonly used heat sources. Ambient and exhaust air, ground, and groundwater
are practical heat sources for small heat pump systems, while sea/lake/river water,
rock (geothermal), and wastewater are used for large heat pump systems.

4.3.1.1 Air
Air is commonly used to meet the heating, ventilation, and air conditioning
(HVAC) needs of light residential and commercial applications. While ambient
air is free and widely available, there are a number of problems associated with
its use as a heat source. The performance of air source heat pumps depends
greatly on ambient temperature. In cooler and more humid climates, some
residual frost tends to accumulate on the outdoor heat transfer coil as the tem-
perature falls below 2-5 °C, leading to a reduction in the capacity of the heat
pump. Coil defrosting can be achieved by reversing the heat pump cycle or
by other less energy-efficient means. This results in a small energy penalty
because during the defrost cycles, cool air is circulated in the building. Provided
the defrost cycle is of short duration, this is not significant. In addition, for ther-
modynamic reasons, the capacity and performance of the heat pump decline
with decreasing temperature. As the heating load is greatest at this time, a sup-
plementary heating source is required. This device is commonly an existing oil,
gas, or electric furnace or an electric resistance heater; the latter is usually part of
the heat pump system. The alternative to the provision of a supplementary
heating device is to ensure that the capacity of the heat pump is adequate to
cope with the most extreme weather conditions. This can result in oversizing
of the unit at a high additional capital cost and is not cost-effective compared
with the cost of supplementary heating devices.
Exhaust (ventilation) air is a common heat source for heat pumps in residential
and commercial buildings. The heat pump recovers heat from the ventilation
138 C HA PT E R 4 : Heat Pump Systems

Table 4.1 Summary of Commonly Used Heat Sources for Heat Pumps
Heat Temperature
Sources Range (°C) Limitations Example

Air
Outdoor 10 to 15 Low heating efficiency and capacity at low Ambient air
ambient temperatures
Low cooling efficiency and capacity at high
ambient temperatures
Defrosting and backup energy as
supplemental heat requirement
Exhaust 15 to 25 Insufficient capacity for typical heating/ Ventilation
cooling loads

Water
Well 4 to 10 Water disposal and permit requirements; Groundwater well
fouling problems
Surface 0 to 10 Often regulated or prohibited; may clog, foul, Lake, river, etc.
or scale
Deep 3 to 8 Often regulated or prohibited; may clog, foul, Sea, ocean, etc.
or scale
Waste >10 Usually regulated; may clog, foul, scale, or Raw or treated sewage,
corrode gray water, etc.

Ground and geothermal


Ground- 0 to 10 Poor performance if ground is not moist Buried fluid loops (brine
coupled High initial cost for ground loop systems)
Direct 0 to 10 Performance depends on ground conditions Refrigerant circulated in
expansion Leak repair is very expensive; requires large ground coil
refrigerant quantities

Solar
Direct – Poor; usually unacceptable performance Refrigerant circulated in
Supplemental source or storage is required solar panels
Indirect – Poor; usually unacceptable performance Water or air circulated in
Supplemental source or storage is required collectors

Source: Dincer and Kanoglu (2010) and ASHRAE (2008)

air and provides water and/or space heating. Continuous operation of the
ventilation system is required during the heating season or throughout the year.
Some units are also designed to utilize both exhaust air and ambient air. For
large buildings, exhaust air heat pumps are often used in combination with
air-to-air heat recovery units.
Outside ambient air is the most common heat source as far as availability is
concerned. Unfortunately, when the space heating load is the highest, the air
4.3 Classification of Heat Pump Systems 139

temperature is the lowest. Further, ambient temperatures are not stable. The
COPs of vapor compression heat pumps decrease with decreasing cold
source temperature. In addition, at evaporator temperatures below 5 °C, air
humidity is deposited on the evaporator surface in the form of ice. This hin-
ders heat transfer and leads to lower working fluid temperatures and,
therefore, lower COP values, depending on the temperature of the air flowing
over the evaporator. If ice formation occurs, periodic deicing of the evapo-
rator surface is required, leading to decreased values of the overall system
COP (by 5-10%).

4.3.1.2 Water
Water source heat pumps are common in installations where internal heat
sources or heat or cold reclamation is possible. In addition, solar or off-peak
thermal storage systems can be used. These sources have a more stable temper-
ature, compared with ambient air. The combination of a high first-cost solar
device with a heat pump is generally not attractive economically on either a
first-cost or a life-cycle cost basis.
Groundwater is available at stable temperatures between 4 and 10 °C in
many regions. Open or closed systems can be used to tap into this heat source.
In open systems, the groundwater is pumped up, cooled, and then reinjected
into a separate reinjection well or discharged to surface water. Open systems
need to be carefully designed to avoid problems such as freezing, corrosion,
and fouling. Closed systems can be either direct expansion systems, with the
working fluid evaporating in underground heat exchanger pipes, or brine
loop systems. Due to the extra internal temperature difference, heat pump
brine systems generally have a lower performance but are easier to maintain.
A major disadvantage of groundwater heat pumps is the cost of installing
the heat source. Additionally, local regulations may impose severe constraints
regarding interference with the water table and the possibility of ground and
water pollution.
Most groundwater at depths more than 10 m is available throughout the year at
temperatures high enough (e.g., 10 °C) to be used as a low-temperature source
for heat pumps. Its temperature remains practically constant over the year and
permits high seasonal heating COPs (3 and more). The energy necessary to
pump up this water has a considerable effect on COP (typically a 10% reduc-
tion per 20 m pumping height). It is necessary to pump the evaporator water
back into the ground to avoid depletion of groundwater layers.
The groundwater needs to be of a purity almost up to the level of drinking water
to be usable directly in the evaporator. The large consumption of water of high
purity limits the number of heat pump systems that can make use of this source.
Also, surface waters constitute a heat source that can be used only for a limited
number of applications.
140 C HA PT E R 4 : Heat Pump Systems

Groundwater at considerable depth (aquifers) offers possibilities for direct


heating or for heating with heat pump systems. The drilling and operating costs
involved usually require large-scale applications of this heat source. The quality
of these waters (e.g., they can contain corrosive salt) often presents serious lim-
itations to their use.
Groundwater (i.e., water at depths of up to 80 m) is available in most areas with
temperatures generally in the 5-18 °C range. One of the main difficulties with
these sources is that often, the water has a high dissolved solid content, produc-
ing fouling or corrosion problems with heat exchangers. In addition, the flow
rate required for a single-family house is high, and groundwater systems are dif-
ficult to use widely in densely populated areas. The inclusion of the cost of pro-
viding the heat source has a significant impact on the economic attractiveness
of these systems. A rule of thumb seems to be that such systems are economic if
both the supply and the reinjection sources are available, marginally economic
if one is available, and not cost-competitive if neither source is available
(Dincer and Kanoglu, 2010). In addition, if a well has to be sunk, the necessity
for drilling teams to act in coordination with heating and ventilation contrac-
tors can pose problems. Also, many local governments impose constraints lim-
iting interference with the water table, sometimes posing difficulties for
reinjection wells.
Surface waters such as rivers and lakes constitute in principle good heat sources,
but they suffer from the major disadvantage that either the source freezes in
winter or the temperature can be very close to 0 °C (typically 2-4 °C). As a
result, great care is needed to avoid freezing on the evaporator. Where the water
is warmed by industry or by power station effluents, the situation is somewhat
improved.
Seawater appears to be an excellent heat source under certain conditions and is
mainly used for medium- and large-sized heat pump installations. At a depth of
25-50 m, the sea temperature is constant (5-8 °C), and ice formation is gener-
ally not a problem (freezing point of 1 °C to 2 °C). Both direct expansion
systems and brine systems can be used. It is important to use corrosion-resistant
heat exchangers and pumps and to minimize organic fouling in seawater pipe-
lines, heat exchangers, evaporators, etc. Where salinity is low, however, the
freezing point may be near 0 °C, and the situation can be similar to that for riv-
ers and lakes with respect to freezing.
Wastewater and effluent are characterized by a relatively high and constant tem-
perature throughout the year. Examples of possible heat sources in this category
are effluents from public sewers (treated and untreated sewage water) at a tem-
perature range of 10-20 °C throughout the year, industrial effluent, cooling
water from industrial processes or electricity generation, and condenser heat
from refrigeration plants. Condenser cooling water from electricity generation
4.3 Classification of Heat Pump Systems 141

or industrial effluents can also be used as heat sources. The main constraints for
use in residential and commercial buildings are distance to the user and vari-
able availability of the waste heat flow. However, wastewaters and effluents
often serve as good heat sources for industrial heat pumps, providing energy
savings in industry.
Apart from surface water systems, which may be prone to freezing, water source
systems generally do not suffer from the low-temperature problems of air
source heat pumps because of the higher annual average temperature. This
ensures that the temperature difference between the source and sink is smaller
and results in improved heat pump performance. The evaporator should be
cleaned regularly, as heat transfer across the evaporator can drop by as much
as 75% within 5 months without proper cleaning (Dincer and Kanoglu,
2010). The costs of cleaning become relatively low for larger projects, making
the use of this heat source more likely to be economic.

4.3.1.3 Ground and Geothermal


Ground or subground (ground source) systems are used for residential and
commercial applications and have similar advantages to water source systems,
because of the relatively high and constant annual temperatures that result in
high COPs. Generally, the heat can be extracted from pipes laid horizontally or
sunk vertically in the ground. The latter system appears to be suitable for larger
heat pump systems. In the former case, adequate spacing between the coils is
necessary, and the availability of suitably large areas (about double the area to
be heated) may restrict the number of applications. For vertical systems, vari-
able or unknown geologic structures and ground thermal properties can cause
difficulties. Owing to the removal of heat from the ground, the ground temper-
ature may fall during the heating season. Depending on the depth of the coils,
recharging may be necessary during the warm months to raise the ground tem-
perature to its normal levels. This can be achieved by passive (e.g., solar irradi-
ation) or active means. In the latter case, this can increase the overall cost of the
system. Leakage from the coils may also pose problems. Both horizontal and
vertical systems tend to be expensive to design and install, often involving
various types of expertise (heating and cooling, laying the pipework, etc.).
Rock (geothermal heat) can be used in regions with no or little groundwater.
Typical borehole depths range from 100 to 200 m, but depths of several hun-
dred meters also are used. When large thermal capacity is needed, the drilled
holes can be inclined to reach a large rock volume. This type of heat pump
is normally connected to a brine system with welded plastic pipes extracting
heat from the rock. Some rock-coupled systems in commercial buildings use
the rock for heat and cold storage. Because of the relatively high cost of drilling,
rock is seldom economically attractive for domestic use.
142 C HA PT E R 4 : Heat Pump Systems

The ground constitutes a suitable heat source for a heat pump in many countries.
At small depths, temperatures remain above freezing, and seasonal temperature
fluctuations are much smaller than those of ambient air. Heat is extracted from
the ground by means of a glycol solution flowing through tubing embedded
in the ground. If a horizontal grid of tubing is utilized, several hundred square
meters of surface area are needed to heat a single-family building. In urban
areas, such space is rarely available. In addition, considerable costs are involved.
For these reasons, vertical ground heat exchangers are preferred presently.
Geothermal heat sources for heat pumps are currently utilized in various coun-
tries, for example, the United States, Canada, and France. These resources are
generally localized and do not usually coincide with areas of high population
density. In addition, the water often has a high salt content that leads to diffi-
culties with the heat exchangers. Due to the high and constant temperatures of
these resources, the performance is generally high.

4.3.1.4 Solar
Solar energy, as either direct or diffuse radiation, behaves similar to air in terms
of some of its characteristics. A solar source heat pump or a combined solar/
heat pump heating system exhibits the disadvantages of the air source heat
pump, such as low efficiency and extreme variability, with the additional dis-
advantage of high capital cost, particularly because in all cases, a heat storage or
backup system is required. In areas with high daily irradiation levels, this may
not be the case (Suleman et al., 2014).
Each of the aforementioned heat sources for heat pumps presents some draw-
backs. Presently, considerable research is being devoted to resolving the tech-
nical problems and utilizing alternative heat sources. Solar energy may provide
a suitable heat source, but solar systems presently are costly, and the intermit-
tent character of solar energy requires the use of large and expensive storage
volumes.

4.3.2 Heat Source/Heat Sink Configuration


Heat pumps are generally classified by their heat sources and sinks. Depending
on cooling requirements, various heat source and heat sink arrangements are
possible in practical applications. The six basic types of heat pumps are as
follows:
• Water-to-water
• Water-to-air
• Air-to-air
• Air-to-water
• Ground-to-water
• Ground-to-air
4.3 Classification of Heat Pump Systems 143

Water B Air
A Heat pump B A Heat pump
Water Water

(a) (b)

Air A B Air Air A Water


Heat pump Heat pump B

(c) (d)

Water B Air
A Heat pump B A Heat pump
Ground Ground
(e) (f)
FIGURE 4.3 Some types of heat pumps (A: heat source and B: heat sink): (a) Water-to-water,
(b) water-to-air, (c) air-to-air, (d) air-to-water, (e) ground-to-water, and (f) ground-to-air.

In each of these, the first term represents a heat source for heating applications
or a heat sink for cooling applications. The common types of heat pumps are
shown in Fig. 4.3.

4.3.2.1 Water-to-Water Heat Pumps


In these heat pump systems (Fig. 4.3a), the heat source and the heat sink are
water. The heat pump system takes heat from a water source (by coil A) while
simultaneously rejecting it to a water heat sink (by coil B) and either heats or
cools a space or a process. In practice, there are many sources of water, for exam-
ple, wastewater, single or double well, lake, pool, and cooling tower. These heat
pumps use less electricity than other heat pumps when properly maintained,
but operating costs increase dramatically without proper maintenance.

4.3.2.2 Water-to-Air and Air-to-Water Heat Pumps


Some heat pumps have been designed to operate utilizing a water source,
instead of an air source, by designing the outdoor heat exchanger to operate
between the heat pump working fluid and water, instead of between the work-
ing fluid and air. These water-to-air heat pumps (Fig. 4.3b) have advantages
over the air-to-air type if a relatively warm source of water is available that does
not require excessive pumping power. In particular, industrial waste heat can
be used.
The difference with the water-to-water heat pump is in the method of treating
the air. Water-to-air and air-to-water (Fig. 4.3d) heat pump systems provide the
heating and cooling of air with water as the heat sink or source. The same
144 C HA PT E R 4 : Heat Pump Systems

sources of water can be used in these systems. They are less efficient than water-
to-water systems because of the lower heat transfer coefficient of air. These
systems are commonly used in large buildings and sometimes in industrial
applications to provide hot or cold water.
Air-to-water systems operate opposite to water-to-air heat pumps: they extract
heat from ambient or exhaust air to heat or preheat water used for space heating
or process heating. Heat is extracted from the air inside the home and trans-
ferred to water and returned to the ground. Households select the desired
indoor temperature (Dincer and Naterer, 2010).

4.3.2.3 Air-to-Air Heat Pumps


These systems (Fig. 4.3c) use air on both sides (on coils) and provide heating or
cooling. In the cooling mode, heat is removed from the air in the space and
discharged to the outside air. In the heating mode, heat is removed from the
outside air and discharged to air in the space. In these units, it is necessary
to provide defrost controls and periods to maintain maximum efficiency. These
are the most popular systems for residential and commercial applications,
mainly because of their easy and economical installation and low maintenance
cost. The most popular heat pump is the air source type (air-to-air) that operates
in two basic modes:

• As an air conditioner: The heat pump’s indoor coil (heat exchanger)


extracts heat from the interior of a structure and pumps it to the coil in
the unit outside where it is discharged to the air outside (hence the term
air-to-air heat pump).
• As a heating device: The heat pump’s outdoor coil (heat exchanger)
extracts heat from the air outside and pumps it indoors where it is
discharged to the air inside.

Depending on the climate, air source heat pumps (including their supplemen-
tary resistance heaters) are about 1.5-3 times more efficient than resistance
heating alone. Operating efficiency has improved since the 1970s, making their
operating costs generally competitive with combustion-based systems, depend-
ing on local fuel prices. With their outdoor unit subject to weathering, some
maintenance is necessary.

4.3.2.4 Ground-to-Water and Ground-to-Air Heat Pumps


In these systems, coil A in Fig. 4.3e and f is buried underground and heat is
extracted from the ground. These heat pump systems do not find widespread
use. Practical applications are limited to space heating where the total heating
or cooling effect is small, and the ground coil size is equally small. This system
requires the burial of several meters of pipe per ton of refrigeration, thus requir-
ing a large amount of land.
4.4 Assessment of Basic Heat Pump 145

4.4 ASSESSMENT OF BASIC HEAT PUMP: ENERGY


AND EXERGY ANALYSES OF VAPOR COMPRESSION
HEAT PUMP CYCLE
Figure 4.4a is a schematic of a simple, idealized vapor compression heat pump
cycle operating between a low-temperature (TL) reservoir and high-temperature
(TH) reservoir. A basic vapor compression heat pump cycle has four main com-
ponents: compressor, condenser, expansion valve, and evaporator. A
temperature-entropy diagram for the simple vapor compression heat pump
cycle is given in Fig. 4.4b.
In the cycle, the refrigerant enters the compressor as a saturated vapor. It is com-
pressed isentropically in a compressor; it is cooled and condensed at constant
pressure by rejecting heat to a high-temperature medium and exits the con-
denser as a saturated vapor. The refrigerant flows through an expansion valve,
at constant enthalpy. The working fluid evaporates at constant pressure in the
evaporator by absorbing heat from the refrigerated space and exits as a
saturated vapor.
In the present analysis, we consider steady-flow, steady-state processes and treat
kinetic and potential energy changes as negligible.
The analysis of a heat pump normally begins with the application of the
conservation of mass principle. Mass rate balances can be expressed for
the compressor, condenser, expansion valve, and evaporator as follows:
m_ ¼ m_ 1 ¼ m_ 2 ¼ m_ 3 ¼ m_ 4 (4.6)

FIGURE 4.4 (a) Schematic of an ideal vapor compression heat pump system and
(b) its temperature-entropy diagram.
146 C HA PT E R 4 : Heat Pump Systems

where ṁ denotes the mass flow rate (kg/s) and subscripts refer to the stream
numbers in Fig. 4.4.
The conservation of energy principle is usually applied next. Energy rate bal-
ances can be expressed for each of the processes of the cycle in Fig. 4.4a for
steady-flow, steady-state operation with negligible kinetic and potential energy
changes as follows:

_ 1 + W_ ¼ mh
Compressor : mh _ 2 (4.7)

Condenser : mh _ 3 + Q_ H
_ 2 ¼ mh (4.8)

Expansion valve : h3 ¼ h4 (4.9)

_ 4 + Q_ L ¼ mh
Evaporator : mh _ 1 (4.10)

where h denotes specific enthalpy (kJ/kg), Ẇ denotes compressor work input


rate, and Q_ H and Q_ L denote condenser and evaporator heat loads, respectively.
An energy rate balance for the entire system gives

W_ + Q_ L ¼ Q_ H (4.11)

Then, we usually move on to the nonconservation of entropy principle.


Entropy rate balances for the system components in Fig. 4.4a can be written as

_ 1 + S_gen, comp ¼ ms
Compressor : ms _ 2 (4.12)

Q_ H
_ 2 + S_gen, cond ¼ ms
Condenser : ms _ 3+ (4.13)
TH

_ 3 + S_gen, ev ¼ ms
Expansion valve : ms _ 4 (4.14)

Q_ L _
_ 4+
Evaporator : ms _ 1
+ Sgen, e ¼ ms (4.15)
TL

where s denotes specific entropy (kJ/kg K) and Ṡgen denotes entropy generation
rate (which is specified for each component).
In many cases, we now apply exergy analysis. Two important aims are to deter-
mine the exergy destruction rates and exergy efficiencies for the system and each
of its components. The components with greater exergy destruction rates are also
those with greater potential for improvement. The exergy destruction rate in a
component can be determined from an exergy rate balance for the component.
Entropy rate balances for the system components in Fig. 4.4a can be written as

_ 1 + W_ ¼ mex
Compressor : mex _ 2 + Ex_ dest, comp (4.16)
 
T
_ 3 + Q_ H 1  0 + Ex
_ 2 ¼ mex
Condenser : mex _ dest, cond (4.17)
TH
4.4 Assessment of Basic Heat Pump 147

_ 3 ¼ mex
Expansion valve : mex _ dest, ev
_ 4 + Ex (4.18)
 
T
_ 4 + Q_ L 1  0 ¼ mex
Evaporator : mex _ 1 + Ex_ dest, e (4.19)
TL

where ex denotes specific exergy (kJ/kg) and Ėxdest denotes exergy destruction
rate, which is specified for each component. T0 implies the dead-state temper-
ature or the reference environment temperature. In a heat pump analysis when
the device operates as a heater, T0 is usually set to the temperature of the low-
temperature medium TL. The exergy destruction rate in the overall cycle can be
determined by analyzing the overall system separately or by summing the
exergy destruction rates of all of the components:

Ex _ dest, comp + Ex
_ dest, total ¼ Ex _ dest, cond + Ex
_ dest, ev + Ex
_ dest, e (4.20)

The total exergy destruction rate of the cycle can be expressed as the difference
between the exergy input rate (power input) and the exergy output rate (the
exergy rate of the heat transferred to the high-temperature medium):
 
_ dest, total ¼ W_  Q_ H 1  T0
Ex (4.21)
TH

In the case where the total exergy destruction rate is zero, the mechanical power
input is a minimum and is equal to the product of the required heating load Q_ H
(in the heating mode) and the temperature-related term in parentheses in the
above equation. That is,
 
T0
W_ min ¼ Q_ H 1  (4.22)
TH

The maximum COP of a heat pump cycle operating between temperature limits
of TL and TH based on the Carnot heat pump can be written as
TH
COPCarnot ¼ (4.23)
TH  TL

The energetic COP of the heat pump cycle during heating mode is defined as

Q_ H
COPen ¼ (4.24)
W_

The exergetic COP, which is defined as exergy efficiency (ψ) of the cycle during
heating mode, can be written as
 
_ T0
QH 1 
TH W_ min COPen
ψ ¼ COPex ¼ ¼ ¼ (4.25)
W_ W_ COPCarnot
148 C HA PT E R 4 : Heat Pump Systems

With exergy analysis, it is possible to assess the system based on the exergy
destruction rates of each component. Higher exergy destruction rates indicate
where larger irreversibilities occur, as can be seen by considering the relative
irreversibility (RI). The RI determines the contribution of each component
to the overall system irreversibility as follows:
_ dest, i
Ex
RIi ¼ (4.26)
_ dest, total
Ex

where the subscript “i” denotes the ith device.

EXAMPLE 4.4.1
A heat pump is used as a heater to keep a room at 25 °C by extracting heat from an environment at
5 °C. The heat pump operates on the idealized simple vapor compression cycle. The total heat loss
rate from the room to the environment is estimated to be 45,000 kJ/h, and the power input to the
compressor is 4.5 kW. Determine (a) the rate of heat extraction from the environment (in kW),
(b) the energetic COP of the heat pump, (c) the maximum rate of heat supply to the room for
the given power input, (d) the exergetic COP of the heat pump, (e) the minimum power input
for the same heating load, and (f) the exergy destruction rate of the cycle.

Solution
(a) The rate of heat extraction from the environment can be written as
 
1kW
Q_ L ¼ Q_ H  W_ ¼ 45, 000kJ=h  4:5kW ¼ 8:0kW
3600kJ=h

(b) The energetic COP of the heat pump is


Q_ H 12:5kW
COPen ¼ ¼ ¼ 2:78
W_ 4:5kW
(c) The COP of the Carnot cycle operating between the same temperature limits is
TH 298
COPCarnot ¼ ¼ ¼ 14:9
TH  TL 298  278
Hence, the maximum rate of heat supply to the room for the given power input is

Q_ H, max ¼ W_ COPCarnot ¼ 4:5kW  14:9 ¼ 67:1kW

(d) The exergetic COP of the heat pump is


COPen 2:78
COPex ¼ ¼ ¼ 0:186
COPCarnot 14:9
(e) The minimum power input for the same heating load is
   
T0 278
W_ min ¼ Q_ H 1  ¼ 12:5kW 1  ¼ 0:84kW
TH 298

(f) The exergy destruction rate of the cycle is


_ dest ¼ W_  W_ min ¼ 4:5kW  0:84kW ¼ 3:66kW
Ex
4.4 Assessment of Basic Heat Pump 149

EXAMPLE 4.4.2
An air source heat pump operates as a heater on the idealized simple vapor compression refrig-
eration cycle with refrigerant-134a as the working fluid. The refrigerant evaporates at 20 °C and
condenses at 1200 kPa. The refrigerant absorbs heat from ambient air at 4 °C and transfers it to a
space at 24 °C. Determine (a) the specific work input and the energetic COP, (b) the specific exergy
destruction in each component of the cycle and the total specific exergy destruction of the cycle,
and (c) the minimum specific work input and the exergetic COP of the cycle. (d) Determine the
energetic COP, the minimum power input, the total exergy destruction, and the exergetic COP
of the cycle if a GSHP is used with a ground temperature of 18 °C. The evaporating temperature
in this case is 6 °C. Everything else remains the same.

Solution
(a) Table 4.2 is constructed by using the data provided in the problem statement and the prop-
erties of refrigerant-134a. Stream numbers are based on Fig. 4.4a.
Following Fig. 4.4, we can write:

qL ¼ h1  h4 ¼ 238:41  117:77 ¼ 120:64kJ=kg


qH ¼ h2  h3 ¼ 284:43  117:77 ¼ 166:66kJ=kg

Then, we can determine the specific work input and the energetic COP, respectively:

w ¼ h2  h1 ¼ 284:43  238:41 ¼ 46:02kJ=kg


qH 166:66kJ=kg
COPen ¼ ¼ ¼ 3:62
w 46:02kJ=kg

(b) The specific exergy destruction in each component of the cycle is determined as follows:

Compressor : sgen, comp ¼ s2  s1 ¼ 0:9456  0:9456 ¼ 0


exdest, comp ¼ T0 sgen, comp ¼ 0kJ=kg
qH 166:66
Condenser : sgen, cond ¼ s3  s2 + ¼ 0:4244  0:9456 + ¼ 0:03996kJ=kgK
TH 297
exdest, cond ¼ T0 sgen, cond ¼ 277  0:03996 ¼ 11:07kJ=kg

Expansion valve : sgen, ev ¼ s4  s3 ¼ 0:4691  0:4244 ¼ 0:0447kJ=kgK


exdest, ev ¼ T0 sgen, ev ¼ 277  0:0447 ¼ 12:38kJ=kg

Table 4.2 Process Data of Streams Used in the Heat Pump Cycle in Example 4.4.2
Stream T (°C) P (kPa) xa h (kJ/kg) s (kJ/kg K)

1 20 1 238.41 0.9456


2 1200 284.43 0.9456
3 1200 0 117.77 0.4244
4 20 117.77 0.4691
a
x: quality.
150 C HA PT E R 4 : Heat Pump Systems

qL 120:64
Evaporator : sgen, evap ¼ s1  s4  ¼ 0:9456  0:4691 + ¼ 0:0410kJ=kgK
TL 277
exdest, evap ¼ T0 sgen, evap ¼ 277  0:0410 ¼ 11:36kJ=kg

The total specific exergy destruction can be determined by adding specific exergy
destructions for all components:

exdest, total ¼ exdest, comp + exdest, cond + exdest, ev + exdest, evap ¼ 34:81kJ=kg

(c) The specific exergy of the heat transferred to the high-temperature medium is
   
T0 277
exqH ¼ qH 1  ¼ 166:66 1  ¼ 11:22kJ=kg
TH 297

The minimum specific work input is

wmin ¼ exqH ¼ 11:22kJ=kg

The exergetic COP is

exqH 11:22kJ=kg
COPex ¼ ¼ ¼ 0:244
w 46:02kJ=kg

(d) Repeating calculations for TL ¼ 18 °C and T1 ¼ 6 °C, we obtain


qH 163:2kJ=kg
COPen ¼ ¼ ¼ 4:80
w 34kJ=kg
wmin ¼ exqH ¼ 3:30kJ=kg
exqH 3:30kJ=kg
COPex ¼ ¼ ¼ 0:097
w 34kJ=kg
exdest, total ¼ w  exqH ¼ 34  3:30 ¼ 30:7kJ=kg

4.5 HEAT PUMP APPLICATIONS


Heat pumps have significant potential for saving energy. They can be used to
recover waste heat and to raise its temperature to more useful levels.
Recent research and development has indicated that heat pump performance is
likely to improve over the coming years. Improvements in component design
and in the use of waste heat sources will raise heat pump performance. More-
over, new ideas and equipment appearing in the last decade have simplified the
construction of the heat pump heating and cooling systems.
Heat pumps appear and operate very much like forced-air air conditioners, with
the notable exception that they can provide both heating and cooling. While
heat pumps and air conditioners require the use of some different components,
they both operate on the same basic principles.
4.5 Heat Pump Applications 151

Heat naturally flows from a higher to a lower temperature. Heat pumps “pump”
heat in the other direction, using a relatively small amount of high-quality drive
energy (electricity, fuel, or high-temperature waste heat). Thus, heat pumps can
transfer heat from natural heat sources in the surroundings, such as the air,
ground, or water, or from man-made heat sources such as industrial or domes-
tic waste to a building or an industrial application. Heat pumps can also be used
for cooling. Heat is then transferred in the opposite direction, from the appli-
cation that is cooled to surroundings at a higher temperature. Sometimes, the
excess heat from cooling is used to meet a simultaneous heat demand.
When operated to provide heat (e.g., for space heating or water heating), the
heat pump is said to operate in the heating mode; when operated to remove
heat (e.g., for air conditioning), it operates in the cooling mode. In both cases,
additional energy has to be provided to drive the pump. Overall, this operation
becomes energetically attractive if the total energy output is greater than the
energy used to drive the heat pump and economically attractive if the total
life-cycle cost (including installation, maintenance, and operating costs) is
lower than that for competing devices.
The most common heat source for a heat pump is air, although water is also
used in many applications. During the past decade, ground or geothermal
resources have received increasing attention as a heat source, particularly in res-
idential, commercial, and institutional applications. From the utilization point
of view, air is considered the most common distribution medium where the
heat pump provides both heating and cooling. For heating only, air is also a
common medium, except in regions where many water distribution systems
are installed in the residential sector. The energy needed to drive a heat pump
is normally provided by electricity or fossil fuels, such as oil or gas.
The general characteristics of some typical commercially available heat pump
systems are listed in Table 4.3 for the residential, commercial, and industrial

Table 4.3 Typical Heat Pump Characteristics and Applications


Commercial and
Residentiala Industrialb District

Primary energy source Electricity Electricity, natural gas, Electricity, natural gas, or oil
or oil
Heat source Air, ground, or water Air, ground, or water Sewage water, waste heat, water,
or ground
Heat sink Air or water Air or water Water
End use Heating and/or cooling Heating and/or cooling Heating and/or cooling
a
Single-family or two-family houses.
b
Multifamily residences, industrial space heating, commercial and institutional sectors, etc.
Source: Dincer and Kanoglu (2010)
152 C HA PT E R 4 : Heat Pump Systems

sectors. For the commercial sector, all the basic characteristics are similar to
those in the residential sector except for the fuel drive. In the former sector,
a greater variety of fuels can be used because of the larger-scale operation, which
suits fossil engine systems. In industry, large-scale uses also result in greater fuel
flexibility, and the heat source is usually waste hot water, steam, or humid air.
The type of heat sink employed depends on the particular industrial process.

4.5.1 Residential Applications


The heating and cooling of single-family and multifamily homes have proved
to be a successful and popular application of heat pumps. A large variety of
systems exist, depending in part on whether they are intended for both heating
and cooling or only heating and on the nature of the low-temperature
heat source and the medium for distributing heat (cold) to the building (air,
water, etc.).
Heat pumps for residential heating and cooling can be classified into four main
categories depending on their operational function:

• Heating-only heat pumps for space heating and/or water heating


applications
• Heating and cooling heat pumps for both space heating and cooling
applications
• Integrated heat pump systems for space heating and cooling, water
heating, and sometimes exhaust air heat recovery
• Heat pump water heaters for water heating

The range of heat pump applications for HVAC purposes is presented and clas-
sified in Fig. 4.5.

4.5.2 Industrial Applications


Heat pumps are available for many industrial processes ranging from opera-
tions in the petrochemical and pulp and paper sectors to those in the food
industry. The uses of the applications include space heating/cooling and pro-
cess water heating/cooling, steam production, drying, dehumidification, evap-
oration and distillation, and concentration processes.
Table 4.4 summarizes the heat pump applications in various industrial pro-
cesses. The table is not comprehensive but highlights the most common indus-
trial applications and heat pump types.
Some novel applications of heat pump systems in various energy-intensive
industries are presented and classified in Fig. 4.6.
4.5 Heat Pump Applications 153

Wall-mounted,
ductless, split, etc.
Decentralized
Domestic with air,
water, solar, etc.

Small scale
(residential single
Central with back-up
family, light commercial
system
and institutional, etc.)
Unitary ground

Ground source, ground


Centralized
water, brine, etc.

Water source type

Adsorption
HVAC
applications

Water source type

Ground coupled brine,


Residential multi-family, etc.
commercial, etc.
Ground source, ground
water, etc.

Absorption
Large scale
Cooling
(absorption chiller)

District
Waste water source
Heating
Heat-driven rankine

FIGURE 4.5 Classification of heat pump applications for heating, cooling, and air conditioning purposes.

Table 4.4 Summary of Selected Heat-Pump Applications in Industrial and Manufacturing Activities
Industry Activity Process Heat Pump Type

Petroleum Petroleum refining and Separation (e.g., propane/propylene


petrochemical product and butane/butylene)
distillation
Inorganic salt production Waste stream concentration to
(e.g., salt and sodium reduce hydraulic load on waste
carbonate) treatment facilities
Process effluent treatment Mechanical vapor
Chemicals compression, open cycle
Heat recovery Low-pressure steam or vapor
compression for use as a heating
medium
Pharmaceuticals Process water heating
Wood products Pulp manufacturing Concentration of black liquor
Paper manufacturing Process water heating Mechanical compression
Flash steam recovery Thermocompression

Continued
154 C HA PT E R 4 : Heat Pump Systems

Table 4.4 Summary of Selected Heat-Pump Applications in Industrial and Manufacturing


Activities Continued
Industry Activity Process Heat Pump Type
Lumber manufacturing Product drying Mechanical vapor
Food and Soft drink/alcohol Effluent concentration compression, open cycle
beverage production
Wet corn milling/corn syrup Deep water and syrup concentration Mechanical vapor
production compression, open cycle,
Sugar refining Sugar solution concentration thermocompression
Dairy products Milk and whey concentration
Juice production Juice concentration
General food product Process heating and cleaning water Mechanical vapor
manufacturing compression, open cycle
Utilities Nuclear power Radioactive waste concentration
Cooling tower blow down
concentration
Miscellaneous Drinking water processing Seawater desalination
Steam stripping Flash steam recovery Thermocompression, open
(wastewater or process cycle
streams)
Electroplating industries Process solution heating
Effluent concentration
Textiles Process and wash water heating
Space heating
Dilute dope stream concentration Mechanical vapor
General manufacturing Process and wash water heating compression, open cycle
Space heating
District heating Large-scale space heating
Solvent recovery Removal of solvent from air streams
Source: U.S. DOE (2009)

Novel solutions or applications

Geothermal heating/ Heating/cooling in


Drying Desalination Cogeneration
cooling vehicles/buildings

FIGURE 4.6 A generalized classification of the novel applications of heat pump technologies. Adapted
from Chua et al. (2010).
4.6 Case Studies 155

4.6 CASE STUDIES


In this section, four systems using vapor compression heat pump cycles are
investigated, and an air sink heat pump is modeled and analyzed with different
sources. The systems selected for consideration are the following:

• System 1: Air source heat pump (with circulating air)


• System 2: Air source heat pump with ventilation
• System 3: Submerged water source heat pump (direct expansion)
• System 4: Open-loop water source heat pump

The performance of each of these systems is assessed based on comprehensive


mass, energy, and exergy analyses. For comparison purposes, system heat loads
are kept constant for all systems considered, although it is more suitable to use
water or ground-source heat pumps for higher loads. Table 4.5 presents the
design parameters used in the assessment of all four systems.

4.6.1 General Assumptions and Simplifications


In the analysis, numerous assumptions and simplifications are applied. The
heat transfer and pressure losses for piping systems, compressor, expansion
valve, and fans are neglected. Potential energy changes within the system are
neglected. Except for the evaporator and condenser fans, kinetic energy changes
are neglected. The temperature drop of the air heat source through the evapo-
rator is considered to be 3 °C. The saturation temperature in the evaporator is
taken to be 10-20 °C below the source temperature (ASHRAE, 2008). The sur-
face area of the outdoor evaporator coil is assumed to be 50-100% larger than
that of the inside coil (ASHRAE, 2008). The condenser and evaporator fan areas
are taken to be 0.35 m2 and 0.70 m2, respectively. The mechanical and electri-
cal efficiencies are assumed to be 68% and 69% for the compressor, 82% and
88% for the pump, and 40% and 80% for the fan, respectively.

Table 4.5 Design Parameters Used in Case Studies


Parameter Value

Required heat load 3 kW


Ambient temperature 5 °C
Ambient pressure 101.3 kPa
Evaporator superheating level 5 °C
Condenser subcooling level 5 °C
Evaporator saturation temperature 10 °C below ambient temperature
Compression ratio 5
156 C HA PT E R 4 : Heat Pump Systems

Mass, energy, entropy, and exergy rate balance equations for the compressor,
the condenser, the expansion valve, and the evaporator and energetic and exer-
getic COPs (exergy efficiency, ψ) are determined for the systems using
Eqs. (4.6)–(4.26). The rate balances for the condenser and evaporator fans
can be written as
m_ ¼ m_ in ¼ m_ out (4.27)

_ in + W_ ¼ mh
mh _ 2 (4.28)
 2 
V V2
W_ ¼ m_ out  in (4.29)
2 2

W_
W_ fan ¼ (4.30)
ηfan, mech  ηfan, elec

_ in + S_gen, fan ¼ ms
ms _ out (4.31)

_ in + W_ fan ¼ mex
mex _ dest, fan
_ in + Ex (4.32)

In each system, R-134a is used as refrigerant and air is assumed to behave as an


ideal gas. The system is investigated based on steady-state, steady-flow operat-
ing conditions, and the heat transfer and pressure losses through piping sys-
tems, compressor, expansion valve, and fans are neglected. Potential energy
changes within the system are neglected. Except evaporator and condenser fans,
kinetic energy changes are neglected. The temperature drop of air source
through the evaporator is considered to be 3 °C. The saturation temperature
in the evaporator is taken to be 10-20 °C below the source temperature
(ASHRAE, 2008). The surface of the outdoor coil of the evaporator is assumed
to be 50-100% larger than the inside coil (ASHRAE, 2008). The condenser and
evaporator fan areas are taken as 0.35 m2 and 0.70 m2, respectively. Compres-
sor mechanical and electrical efficiencies are assumed to be 68% and 69%,
respectively. Pump mechanical and electrical efficiencies are assumed to be
82% and 88%, respectively. Fan mechanical and electrical efficiencies are
assumed to be 40% and 80%, respectively.

4.6.2 System 1
The first system considered, as presented in Fig. 4.7, is an air-to-air heat pump. It
includes condenser and evaporator fans to circulate the air within the system.
R-134a is used as the refrigerant and air is assumed to behave like an ideal
gas. The system is investigated based on steady-state, steady-flow operating
conditions.
Table 4.6 summarizes the state properties (phase, temperature, and pressure)
and the specific enthalpy, specific entropy, and specific exergy of each stream
in System 1 (see Fig. 4.7).
4.6 Case Studies 157

1
2

9 8 5 6 7
10

3
4
A Compressor
B Condenser
C Expansion valve
D Evaporator
E Condenser fan
F Evaporator fan

FIGURE 4.7 Schematic of System 1 layout.

An exergy analysis of System 1 shows that the overall system has an exergy effi-
ciency of 17.6% and an exergy destruction rate of 2.10 kW. Of this exergy
destruction rate, 1.31 kW is caused by the heat pump unit (RIheat pump is
62.2%), and the rest is mainly associated with the evaporator and condenser
fans (RIfans is 37.8%). Among the heat pump components, the condenser
and the evaporator have the largest exergy destruction rates and, therefore, rel-
ative irreversibilities. The exergy destruction rate associated with heat transfer is
the main cause of these irreversibilities. Relative irreversibilities for all compo-
nents in System 1 are presented in Fig. 4.8.
Exergy efficiencies of the overall system, the heat pump unit, and the system
components are presented in Fig. 4.9 for System 1. Since heat losses in the
expansion valve and compressor are neglected, these units have the highest
exergy efficiencies. The heat pump unit (without the fans) has an exergy effi-
ciency of 45.1%.

4.6.3 System 2
System 2 is essentially similar to System 1, but ventilation is added to permit
some of the building air to be replaced by fresh outdoor air. Central heat pump
systems generally provide ventilation when applied in residential applications.
In this case, the heat load is taken to be 3 kW per resident, and six people are
assumed to be living in each unit. Also, each unit is assumed to require 9.4 L/s
(20 cfm) of fresh air, based on data reported by Sugarman (2005). Figure 4.10
shows the air-handling duct system of System 2. The volumetric flow rate of
ventilation air 56.6 L/s (120 cfm) is fixed based on the design and amount
of return air from the room, which is calculated considering the assumed
heat load.
158
C HA PT E R 4 :
Heat Pump Systems
Table 4.6 State Properties and Thermodynamic Data for the Heat Pump System 1, Shown in Fig. 4.7
Specific Specific Specific
Temperature Pressure Enthalpy, Entropy, Exergy,
State Description Fluid Phase (°C) (kPa) h (kJ/kg) s (kJ/kg K) ex (kJ/kg)

1 Evaporator 5.093 200 248.7 0.953 14.97


outlet
2,s Compressor Superheated 50.35 283.1 0.953 49.38
outlet vapor
2 Compressor 57.45 290.7 0.976 50.52
outlet Refrigerant 1000
3 Condenser Subcooled 34.37 99.92 0.368 29.01
outlet
4 Evaporator Mixture 10.09 200 99.92 0.388 23.37
inlet
5 Condenser fan 19
inlet 292.5 5.675
6 Condenser fan 19
outlet
7 Condenser air 30 303.6 5.712
outlet
Dry air Gas 101.3 1254
8 Evaporator fan 5
inlet
9 Evaporator fan 5 278.5 5.62
outlet
10 Evaporator air 2 275.5 5.615
outlet
4.6 Case Studies 159

Compressor
9%

Evaporator fan
30%

Condenser
29%

Condenser fan
8%

Evaporator
20% Expansion valve
4%

FIGURE 4.8 Relative irreversibilities for components in System 1.

90
80.55
80
69.45
70
Exergy efficiency (%)

60

50 47.35 46.19 45.1


40 40
40

30
17.64
20

10

0
Compressor Condenser Expansion Evaporator Condenser Evaporator Heat pump Overall
valve fan fan unit system

FIGURE 4.9 Exergy efficiencies of the overall system, heat pump unit, and system components for
System 1.

Mass, energy, entropy, and exergy balance equations for the compressor,
condenser, expansion valve, evaporator, and fans and energetic and exergetic
COP (exergy efficiency, ψ) calculations are presented in Eqs. (4.6)–(4.32).
The exergy analysis of System 2 shows that the overall system has an exergy effi-
ciency of 19.9% and an exergy destruction rate of 1.83 kW. 1.16 kW of this
exergy destruction rate is caused by the heat pump unit (RIheat pump is
63.7%); the rest is due to evaporator and condenser fans (RIfans is 36.3%).
160 C HA PT E R 4 : Heat Pump Systems

Fresh ventilation
air

Return air from room

Indoor coil
condenser

Room air supply

FIGURE 4.10 Schematic of the air-handling duct system of System 2.

Among heat pump components, the condenser and evaporator have the largest
extraction rates, therefore, relative irreversibilities. Exergy loss associated with
heat transfer is the main cause of these irreversibilities. Relative irreversibilities
of each component are presented in Fig. 4.11.
Exergy efficiencies of the overall system, the heat pump unit, and the system
components are presented in Fig. 4.12 for System 2. As for System 1, due to

Compressor,
10%

Evaporator fan,
35%
Condenser, 25%

Condenser fan,
Expansion
2%
Evaporator, valve, 5%
23%

FIGURE 4.11 Relative irreversibilities for components in System 2.


4.6 Case Studies 161

90
80.55
80
69.45
70
Exergy efficiency (%)

60

50 46.19 45.1
40 40
40 35.14
30
19.91
20

10

0
Compressor Condenser Expansion Evaporator Condenser Evaporator Heat pump Overall
valve fan fan unit system

FIGURE 4.12 Exergy efficiencies of the overall system, heat pump unit, and system components for
System 2.

neglected heat losses in the expansion valve and the compressor, these units
exhibit the highest exergy efficiencies, and the heat pump unit (without the
fans) exhibits an exergy efficiency of 45.1%.

4.6.4 System 3
In System 3, the evaporator is submerged in surface water, as is common prac-
tice where a pond or lake is located near the system. System 3, unlike Systems 1
and 2, does not have an evaporator fan. Figure 4.13 depicts the evaporator
system.

Heat pump unit


To condenser

From expansion valve

FIGURE 4.13 Schematic of System 3.


162 C HA PT E R 4 : Heat Pump Systems

Mass, energy, entropy, and exergy balance equations for the compressor, con-
denser, expansion valve, evaporator, and fans and energetic and exergetic COP
(exergy efficiency, ψ) calculations are presented in Eqs. (4.6)–(4.32).
The exergy analysis of System 3 shows that the overall system has an exergy effi-
ciency of 34.23% and an exergy destruction rate of 2.062 kW. 1.90 kW of
this exergy destruction rate is attributable to the heat pump unit, for which
RIheat pump ¼ 92.14%. The remainder is linked to the condenser fan (RIfans is
7.86%). Among the heat pump components, the condenser and the evaporator
exhibit the largest exergy destruction rates and, as a consequence, the largest
relative irreversibilities. The exergy destruction associated with heat transfer
is the main cause of these irreversibilities. The relative irreversibilities of the sys-
tem components are presented in Fig. 4.14.
Exergy efficiencies of the overall system, the heat pump unit, and the system
components are presented in Fig. 4.15 for System 3. Due to neglected losses
in the expansion valve and compressor, these units have the highest exergy effi-
ciencies. The heat pump unit (without the fans) is 45.1% exergy-efficient.

4.6.5 System 4
System 4 contains an open-loop water source heat pump. Water is transported
to the evaporator from a surface water source such as a lake or pond. After heat
exchange in the evaporator, the water is returned to the same or another loca-
tion. Figure 4.16 illustrates the System 4.

Condenser Compressor
fan 9%
8%

Condenser
30%
Evaporator
49%

Expansion valve
4%
FIGURE 4.14 Relative irreversibilities of components in System 3.
4.6 Case Studies 163

90
80.55
80
69.45
70
Exergy efficiency (%)

60

50 47.35 45.1
40
40 34.23
30

20 15.05

10

0
Compressor Condenser Expansion Evaporator Condenser Heat pump Overall
valve fan unit system

FIGURE 4.15 Exergy efficiencies of the overall system, heat pump unit, and system components for
System 3.

Heat pump unit


To evaporator

From evaporator

FIGURE 4.16 Schematic of System 4.

Mass, energy, entropy, and exergy balance equations for the compressor, con-
denser, expansion valve, evaporator, and fans and energetic and exergetic COP
(exergy efficiency, ψ) calculations are presented in Eqs. (4.6)–(4.32).
Exergy analysis of System 4 shows that the overall system has an exergy effi-
ciency of 33.35% and an exergy destruction rate of 1.240 kW. 1.06 kW of this
exergy destruction rate is caused by the heat pump unit (RIheat pump is 85.6%);
the rest is due to the pump and condenser fan (RIfans is 14.4%). Among heat
pump components, the condenser and evaporator have the largest extraction
rates, therefore, relative irreversibilities. Exergy loss associated with heat transfer
164 C HA PT E R 4 : Heat Pump Systems

Pump Condenser
1% fan Compressor
13% 16%

Evaporator
14%

Expansion valve
7%
Condenser
49%

FIGURE 4.17 Relative irreversibilities for components in System 4.

is the main cause of these irreversibilities. Relative irreversibilities of each com-


ponent are presented in Fig. 4.17.
Exergy efficiencies of the overall system, the heat pump unit, and the system
components are presented in Fig. 4.18 for System 4. Due to neglected losses
in the expansion valve and compressor, these units have the highest exergy effi-
ciencies. The heat pump unit (without the fans) is 45.1% exergy-efficient.

90

80

70
Exergy efficiency (%)

60

50

40

30

20

10

0
Compressor Condenser Expansion Evaporator Pump Condenser Heat pump Overall
valve fan unit system

FIGURE 4.18 Exergy efficiencies of the overall system, heat pump unit, and system components for
System 4.
4.7 Closing Remarks 165

Table 4.7 Performance Comparison of Investigated Systems


System
Performance Criteria
1 2 3 4

Carnot COP 4.89 4.89 4.89 4.89


Energetic COP (heat pump) 4.54 4.54 4.54 4.54
Energetic COP (overall system) 2.9 3.11 4.07 4.02
Exergetic COP (heat pump) 0.51 0.51 0.51 0.51
Exergetic COP (overall system) 0.33 0.35 0.46 0.45

4.6.6 Overall Comparison of Systems


Table 4.7 presents the Carnot, energetic, and exergetic COPs of all four systems
considered in this case study. The Carnot efficiency depends on heat sink/
source temperatures, and in the four systems, the sink and source temperatures
do not vary. Therefore, the Carnot efficiency is the same for all four systems.
Similarly, when taken alone, the energetic and exergetic COPs of the heat
pumps do not vary among the four systems. This is because of the constant
sink/source temperature and heating requirements assumed in each system.
When the overall systems are compared, System 3 (submerged water source
heat pump with direct expansion) is seen to have the highest overall system
energetic and exergetic COPs, at 4.07 and 0.46, respectively.

4.7 CLOSING REMARKS


In this chapter, heat pump systems along with their various performance cri-
teria are described. Heat pump systems are classified based on heat source
and heat source/sink configuration. For illustrative purposes, a basic heat pump
system is analyzed both energetically and exergetically. Numerical examples are
also provided. Residential and industrial applications of heat pumps are dis-
cussed, and four heat pump systems are comparatively assessed in a case study.
Several important observations can be drawn from the chapter:

• Although the most common measure of heat pump efficiency from an


energy perspective is energy-based COP, exergy-based COP or exergy
efficiency is utilized to assess the actual performance of the heat pump.
• The largest irreversibilities in the heat pump unit are typically associated
with the condenser, followed by the compressor, the evaporator, and
the expansion valve.
166 C HA PT E R 4 : Heat Pump Systems

• Exergy analyses of heat pump systems focus attention on components


where the greatest potential is destroyed and quantify the extent to which
modifications affect, favorably or unfavorably, the performance of the
system and its components.

Four heat pump systems are considered as case studies. These systems are (i) air
source heat pump with circulating air, (ii) air source heat pump with ventila-
tion, (iii) submerged water source heat pump (direct expansion), and
(iv) open-loop water source heat pump. The energetic and exergetic COPs of
these systems are determined, demonstrating that the submerged water source
heat pump (direct expansion) system has the highest energetic and exergetic
COPs. The next highest COPs are exhibited by the open-loop water source heat
pump, the air source heat pump with ventilation, and the air source heat pump
with circulating air.

Nomenclature
ex specific exergy
Ex exergy
h specific enthalpy
m mass
q specific heat
Q heat
s specific entropy
S entropy
T temperature
V velocity
w specific work
W work

Greek symbols
η energy efficiency
ψ exergy efficiency

Subscripts
0 reference environment state
comp compressor
cond condenser
dest destruction
e evaporator
elec electrical
en energy
ev expansion valve
References 167

ex exergy
gen generation
H high-temperature
i ith constituent
in inlet
L low-temperature
max maximum
mech mechanical
min minimum
out outlet
Q heat
W work

Superscripts
˙ rate with respect to time

Acronyms
COP coefficient of performance
EER energy efficiency ratio
GSHP ground source heat pump
HSPF heating season performance factor
HVAC heating, ventilation, and air conditioning
PER primary energy ratio
RI relative irreversibility
SEER seasonal energy efficiency ratio

References
ASHRAE, 2008. HVAC Systems and Equipment. American Society of Heating, Refrigerating, and Air
Conditioning Engineers, Atlanta, GA.
Caliskan, H., Hepbasli, A., Dincer, I., 2011. Exergy analysis and sustainability assessment of a solar-
ground based heat pump with thermal energy storage. J. Solar Energy Eng. 133, 11–25.
Chua, K.J., Chou, S.K., Yang, W.M., 2010. Advances in heat pump systems: a review. Appl. Energy
87, 3611–3624.
Dincer, I., Kanoglu, M., 2010. Refrigeration Systems and Applications. John Wiley & Sons, Oxford.
Dincer, I., Naterer, G.F., 2010. Assessment of exergy efficiency and Sustainability Index of an air-
water heat pump. Int. J. Exergy 7, 37–50.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, 2nd ed.
Elsevier Science, Oxford.
Holland, F.A., Watson, F.A., Devotta, S., 1982. Thermodynamic Design Date for Pump Systems.
Pergamon Press, Oxford.
Koohi-Fayegh, S., Rosen, M.A., 2014. An analytical approach to evaluating the effect of thermal
interaction of geothermal heat exchangers on ground heat pump efficiency. Energy Convers.
Manag. 78, 184–192.
168 C HA PT E R 4 : Heat Pump Systems

Lubis, L., Kanoglu, M., Dincer, I., Rosen, M.A., 2011. Thermodynamic analysis of a hybrid geother-
mal heat pump system. Geothermics 40, 233–238.
Soltani, R., Dincer, I., Rosen, M.A., 2015. Comparative performance evaluation of cascaded air-
source hydronic heat pumps. Energy Convers. Manag. 89, 577–587.
Sugarman, S.C., 2005. HVAC Fundamentals. CRC Press, Boca Raton, FL.
Suleman, F., Dincer, I., Agelin-Chaab, M., 2014. Energy and exergy analyses of an integrated solar
heat pump system. Appl. Therm. Eng. 73, 559–566.
U.S. DOE (Department of Energy), 2009. Industrial heat pumps for steam and fuel savings. http://
www1.eere.energy.gov/manufacturing/tech_assistance/pdfs/heatpump.pdf.
Zamfirescu, C., Dincer, I., 2009. Performance investigation of high-temperature heat pumps with
various BZT working fluids. Thermochim. Acta 488, 66–77.
CHAPTER 5

Cogeneration, Multigeneration,
and Integrated Energy Systems

5.1 INTRODUCTION
Issues surrounding fossil fuel prices, resources, and environmental impacts
have led in recent decades to increased efforts to develop more efficient systems.
One way to do this is by producing multiple outputs in a single system. Cogen-
eration, or combined heat and power (CHP), is a technique for producing heat
and electricity in a single process that is very efficient and thereby able to reduce
energy use considerably. Cogeneration is often associated with the combustion
of fossil fuels but can also be carried out using other sources of thermal energy
(e.g., some renewable energy resources, nuclear energy, and burning wastes).
The trend recently has been to use cleaner fuels such as natural gas for cogen-
eration. Cogeneration appears to have significant long-term prospects in global
energy markets, primarily due to its numerous operational, environmental, and
economic benefits.
Cogeneration often reduces energy use cost-effectively and improves security of
energy supply. In addition, since cogeneration installations are usually located
close to consumers, electrical grid losses can be reduced when cogeneration is
applied, and cogeneration is often well suited for use in isolated or remote
areas. Cogeneration can offer an attractive option for facilities with high electric
rates and buildings that consume large amounts of hot water and electricity.
Usually, the higher the electric rates, the greater the savings with cogeneration
and the lower the payback period (i.e., the savings pay for the initial capital
investment faster). The thermal energy product from cogeneration can be used
for domestic hot water heating, space heating, pool and spa heating, laundry
processes, and absorption cooling. The more the product heat from cogenera-
tion used year round in existing systems, the more financially attractive cogen-
eration is in most instances. Facilities that use large amounts of thermal energy
during all months of the year include

• apartments and condominiums;


• colleges, universities, and other educational institutions;
169

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00005-3


© 2015 Elsevier Inc. All rights reserved.
170 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

• hospitals;
• hotels;
• athletic clubs;
• assisted living facilities, nursing homes, and senior housing;
• industrial and waste treatment facilities; and
• laundries.

Cogeneration helps overcome one of the main drawbacks of conventional elec-


trical systems: the significant heat losses that detract greatly from efficiency.
Heat losses are reduced and efficiency is increased when cogeneration is used
to supply heat to various applications and facilities. The overall energy effi-
ciency of a cogeneration system is the percent of the fuel converted to both elec-
tricity and useful thermal energy. Typical cogeneration systems have overall
efficiencies ranging from 65% to 90%.
Trigeneration, also known as combined cooling, heat, and power (CCHP), is
the simultaneous generation of electric power, heating, and cooling, either
from the combustion of fossil fuels or from other thermal energy resources
(e.g., renewable energy thermal sources such as geothermal and solar energies).
One relatively common trigeneration system uses high-temperature heat to
drive a gas or steam turbine, and then, the residual low-temperature heat
(which may be waste heat) is recovered for heating and/or to produce cooling.
Trigeneration systems differ from cogeneration systems in that both heating
and cooling are simultaneously produced. In trigeneration systems, an absorp-
tion cooling system is often used to provide cooling from some of the thermal
energy, although electric chillers can also be used.
The potential improvement in the overall system efficiency via trigeneration
can in some circumstances be enhanced by extending trigeneration to multige-
neration, typically taken to be the simultaneous production of more than three
products. In multigeneration systems, part of the electricity or cooling or heat-
ing is often utilized to produce an additional product such as hydrogen or dry-
ing or hot water.
Multigeneration systems can sometimes help satisfy global energy needs while
reducing environmental impacts and costs. Reductions in fuel usage and emis-
sions of CO2 and wastes and increased efficiencies are some of the potential
benefits of multigeneration systems over conventional single-generation pro-
cesses. Figure 5.1 shows how the overall system energy efficiency is evaluated
as the number of outputs increases. Systems with multiple outputs can at times
exhibit other benefits over conventional energy/heat and cooling/fuel genera-
tion processes, for example, improved reliability.
District energy systems can utilize numerous energy resources, ranging from
fossil fuels to renewable energy to waste heat, and are sometimes called
5.1 Introduction 171

Power
heating
2
Single generation:
Cooling
O1 hot water Cogeneration:
4
O2
Hydrogen Trigeneration:
I Process O3 fresh water
6 Quadgeneration:
Drying ¼
N-generation:
refrigeration
On 8

Less sustainable More sustainable

FIGURE 5.1 Illustration of multigeneration energy efficiencies as the number of outputs


increases (Dincer and Acar, 2015).

“community energy systems.” By linking a community’s energy users, district


energy systems can help improve efficiency and provide opportunities to con-
nect generators of waste energy (e.g., electric power plants or industrial facili-
ties) with consumers who may be able to use the waste energy. The heat in a
district energy system can be used for heating or can be converted to cooling
using absorption chillers or steam turbine-driven chillers.
District energy systems generally include both district heating and district cooling
systems. They distribute steam, hot water, and chilled water from a central plant
to individual buildings through a network of pipes. District energy systems can
provide space heating, air conditioning, domestic hot water, and/or industrial
process energy and often are linked with electricity generation via cogeneration.
With district energy, boilers and chillers in individual buildings are not required.
District energy is often an attractive, efficient, and environmentally benign way to
reduce energy consumption. The Intergovernmental Panel on Climate Change
(IPCC) identified cogeneration and district energy as key measures for green-
house gas (GHG) reduction, and the European Commission has been develop-
ing cogeneration and district energy systems for the European Union. District
energy systems can provide other environmental and economic benefits,
including
• reduced local/regional air pollution,
• increased opportunities to use ozone-friendly cooling technologies,
• infrastructure upgrades and development that provide new jobs,
• enhanced opportunities for electric peak demand reduction through
chilled water or ice storage,
• increased fuel flexibility, and
• improved energy security.
Integrated energy systems integrate multiple processes in ways that typically
allow for one or more benefits. Integrated energy systems often are taken to
include those systems described thus far in this chapter, like multigeneration
172 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

and district energy systems, but they can be broader. Rather, for instance, a dif-
ferent type of integrated energy system is one that integrates multiple systems in
a way that permits cascading of waste heat at high temperatures so it becomes
an input to users needing heat at lower temperatures. Locating industrial plants
requiring waste heat or materials near other industries so they may produce
those wastes, thereby potentially making them by-products instead, is a way
of integrating energy systems.
Energy and exergy analyses of cogeneration, trigeneration, district heating and
cooling (DHC), and integrated energy systems are described in this chapter.
Relative to conventional systems, such systems can be complex, particularly
because they simultaneously provide electrical, heating, and cooling services.
This chapter also describes the benefits of applying exergy analysis to such sys-
tems and explains key exergy-based performance measures for such systems.
A specific case is considered to illustrate the topics covered including the deter-
mination of system and component efficiencies and their improvement.
This chapter reveals insights that can aid in the design of cogeneration and inte-
grated energy systems and related optimization activities and in the selection of
the proper types of systems for various applications and situations. This knowl-
edge can help energy utilities improve existing plants where appropriate and
develop better designs. Another key point covered relates to difficulties associ-
ated with the types of analysis tools used for cogeneration, trigeneration, dis-
trict energy, and integrated energy systems. In general, energy technologies
are normally examined thermodynamically using energy analysis, although a
better understanding is attained when a more complete thermodynamic view
is taken. Exergy analysis provides an additional thermodynamic perspective
and, in conjunction with energy analysis, permits more complete thermody-
namic analyses.
Applications of exergy analysis to cogeneration, trigeneration, district energy,
and integrated energy systems have increased in recent years (Rosen et al.,
2005; Dincer and Rosen, 2013; Dincer and Zamfirescu, 2012) and have yielded
useful and meaningful insights into performance that assist in achieving opti-
mal designs.

5.2 COGENERATION
Cogeneration, or CHP, is the simultaneous production of electrical and ther-
mal energy from a single-energy source. The technology is proved and reliable,
mostly used in large-scale centralized power plants, and it has been applied for
more than 100 years. The waste heat from electricity generation is recovered,
sometimes at the expense of some electrical output, and used for applications
such as space heating and cooling, water heating, and industrial process heat.
5.2 Cogeneration 173

By making use of the waste from one process as an input to another, substantial
gains in energy efficiency can be realized. Most of the world’s electricity is gen-
erated by rotating machinery that is driven by the combustion of fuels.
As a relatively clean, efficient, and cost-effective technology, cogeneration can
help address global energy, environmental, and economic concerns. In conven-
tional separate methods for electricity generation and heating, electricity is pro-
duced centrally in many ways and transported to users via the electrical grid
while heat is produced by burning fuels in a combustor, usually on site.
CHP systems take advantage of the by-product heat that can be as high as
60–80% of total primary energy in combustion-based electricity generation.
CHP systems combine the production of electrical (or mechanical) energy
and useful thermal energy from the same primary energy source in one oper-
ation. The advantages of CHP include (i) high efficiency, (ii) low emissions
of GHGs like CO2 and other pollutants, (iii) cost savings, (iv) wide geographic
applicability, and (v) enhanced energy security.
The fact that CHP plants recover a share of the waste heat that is otherwise
released by power plants that generate only electricity is significant. The global
average energy efficiency of fossil-fueled electric power plants is 37%, whereas
the global average efficiency of CHP units, accounting for both coproducts, is
58%, and state-of-the-art CHP plants have energy efficiencies that can exceed
85% (International Energy Agency, 2014). For instance, a low-temperature
heat-driven CHP system proposed as cost-effective for small-scale applications
exhibited energy and exergy efficiencies of 87% and 35%, respectively
(Hogerwaard et al., 2013). In electricity generation mode, the corresponding
efficiencies are 17% and 5%, respectively. The usefulness of decentralized
cogeneration units is discussed by Pehnt (2008). Further emission reductions
from fossil fuel systems are possible through carbon dioxide capture and stor-
age (CCS)/sequestration.
Cogeneration can be implemented at a range of scales, from large-scale systems
serving communities or large industrial complexes to independent small systems
for hospitals, universities, or residential buildings. Since heat losses occur during
transport, CHP facilities are normally located near users of thermal energy,
although the transport distances for some advanced CHP-district energy systems
can be on the order of tens of kilometers. System efficiency and heat output char-
acteristics are important attributes of a cogeneration system. Cogeneration
energy efficiency is the percent of input energy converted to electricity and useful
thermal energy. Most cogeneration systems have overall efficiencies between
65% and 85%. The nature of the heat output from cogeneration systems varies
depending on system type. The thermal output can be of high quality (e.g., high-
pressure, high-temperature steam) for industrial process needs or low quality
(e.g., hot water) for limited thermal applications such as space and domestic
hot water heating.
174 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

A cogeneration facility is composed of two basic parts: a power generator and a


heat recovery system. A range of technologies can be used to achieve cogenera-
tion, including steam turbines, gas turbines, reciprocating engines, microtur-
bines, fuel cells, and Stirling engines. Table 5.1 lists selected performance
criteria of diesel and natural gas engines, stream/gas/microturbines, and fuel
cells as CHP technologies. Diesel and natural gas engines and gas turbines have
the advantages of lower capital costs, quick start-up times, high efficiencies, and
reliability. However, they require regular maintenance and their NOx emissions
are high. Steam turbines are flexible in terms of fuel input, but they have lower
electric efficiencies and longer start-up times. Microturbines have flexible fuel
input capabilities and compact size, fewer moving parts, and lower noise. But
they also have high capital costs and low electric efficiencies. Microturbines are
beneficial when energy systems are distributed with micro- to small-scale pro-
duction needs. Fuel cells operate quietly with high reliability and efficiency and
extremely low emissions and are actively undergoing research to improve

Table 5.1 Performance Summary for CHP Prime Movers


Diesel Natural Steam Gas Fuel
Engine Gas Engine Turbine Turbine Microturbine Cell

Electrical energy efficiency 30–50 25–45 30–42 25–60 20–30 40–70


(LHV, %)
Total capacity (MW) 0.05–5 0.05–5 0.05–250 3–200 0.025–0.25 0.2–2
Land area footprint (m2/kW) 0.02 0.02–0.03 <0.01 0.002– 0.01–0.14 0.02–
0.06 0.2
Installment cost (USD/kW)a 800–1500 800–1500 800–1000 700–900 500–1300 >3000
Operation and maintenance 0.5–0.8 0.7–1.5 0.4 0.2–0.8 0.2–1 0.3–
cost (USD/kWh)a 1.5
Start-up time 10 s 10 s 1–24 h 10– 1 min 3–48 h
60 min
Fuel pressure (bar) 0.3 0.1–3 N/A 8–35 3–7 0.1–3
NOx emissions (kg/MWh) 1.4–15 1–13 0.8 0.1–1.8 0.2–1 <0.01
CHP output (total kJ output/ 3400 1000–5000 N/A 3400– 4000–15,000 500–
kWh input) 12,000 3700
Usable temperature of 80–500 150–260 N/A 260–600 200–350 60–
thermal coproduct (°C) 400

Uses for heat recovery


Hot water heating + + + + +
Space heating + +
District heating + + + +
LP steam + + + + + +
HP steam + + +
a
Monetary values are based on year 2012.
Data sources: IEA (2008) and US DOE (2012).
5.2 Cogeneration 175

performance (e.g., reduce energy consumption) (Liu et al., 2014). Selecting the
most appropriate prime mover for a CHP system depends on local resources,
size and cost constraints, and GHG emission requirements. The advantages
and disadvantages of CHP prime movers are summarized in Table 5.2.
Operational flexibility of CHP plants may be constrained by heat loads,
although thermal storages and complementary heat sources can mitigate this
effect (Blarke, 2012; Christidis et al., 2012; Nuytten et al., 2013). Obtaining
flexibility from fossil generation has a cost and can affect the overall GHG
reduction potential of various renewable energy sources (Pehnt, 2008; Ludig
et al., 2011). Demand response and energy storage can potentially offer addi-
tional flexibility. Demand response can include intentional modifications to
consumption patterns of energy by inducing consumers to alter their timing
of use and/or level of instantaneous demand and/or total consumption.
Demand response is of increasing interest due to its potentially low cost
(Depuru et al., 2011; Cook et al., 2012; Joung and Kim, 2013; Procter,
2013), albeit some emphasize its limitation compared to flexible conventional
supply technologies (Cutter et al., 2012). Smart meters and remote controls are
key components of smart grids, in which information technology is used to
improve the operation of power systems, especially with resources located
closer to distribution end points. The development of intelligent DHC net-
works in combination with heat storage allows for more flexibility and diversity
of energy sources and facilitates additional opportunities for low-carbon tech-
nologies (CHP, waste heat use, heat pumps, and solar heating and cooling). In
addition, excess renewable electricity can be converted to heat to replace what
otherwise would have been produced by fossil fuels (Meibom et al., 2007).
By using outputs and wastes from one process as inputs to other processes,
cogeneration systems have the potential to increase efficiency and reduce energy
costs, GHG and other pollutant emissions, and releases of ozone-depleting chlo-
rofluorocarbons from air conditioning and refrigeration units. If the thermal out-
put is greater than necessary, excess thermal energy is produced that could have
been used to generate more electricity. Locating a cogeneration facility near the
loads for heat means electricity is produced closer to the load than often occurs
with centralized power generation. This “distributed” energy approach allows for
geographically dispersed generating plants, reduces transmission losses, and pro-
vides process heating/cooling for buildings. In addition, the reduced need for
energy resulting from cogeneration can help reduce dependency on fuel imports.
Cogeneration provides an opportunity to increase the output diversity of a gen-
eration plant and provides competition in energy generation. The benefits of
cogeneration systems suggest it has significant potential for growth.
Siting of cogeneration can be challenging as facilities often must be located near
their thermal users. In addition, integrating distributed energy sources into the
176 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.2 Summary of Advantages and Disadvantages of CHP Prime Movers


Prime mover Advantages Disadvantages

Reciprocating Low capital cost per kW High emissions


engines Readily available in various sizes Noisy
Good electric efficiency Frequent maintenance
Quick start-up
Fuel flexibility
Proved high reliability
Low natural gas pressure requirement
Gas turbine Low capital cost per kW for large systems High emissions
generators Readily available in various sizes Reduced efficiency at partial load
High electric efficiency Sensitive to ambient air conditions
High heat recovery (can produce high-temperature High-pressure requirements
steam)
Quick start-up High cost and low efficiency for smaller
engines
Fuel flexibility
Proved high reliability
Microturbines High efficiency High capital cost per kW
Low maintenance (few moving parts) Low electric efficiency
Small size, lightweight Reduced efficiency at partial load
Low noise Sensitive to ambient air conditions
Commercially available Fuel gas compression requirements
Fuel flexibility Reliability not yet proved
Quick start-up
Low emissions
Stirling engines High efficiency Higher capital cost per kW
Low maintenance (few moving parts) Few commercial units in operation
Low noise
Commercially available Limited authorized service providers
Fuel flexibility Reliability not yet proved
Low emissions Reliability not yet proved
Fuel cells High efficiency High capital cost per kW
Low maintenance Limited fuel options
Quiet operation Reliability not yet proved
Quick start-up
Low emissions
Not sensitive to ambient air conditions
Constant efficiency at partial load
Source: CH2MHILL (2007).
5.2 Cogeneration 177

electricity grid may require transmission and distribution system upgrades. Many
stakeholders must work together in the development and operation of a cogen-
eration facility, including the electricity generator, the utility that distributes the
electricity (and possibly thermal energy), and the thermal users. The distribution
utility must be willing to purchase power from the generator and may put restric-
tions or costs on connecting to the grid. In addition, thermal users may have fluc-
tuating heat requirements that can be difficult for the system to follow.

5.2.1 Case Study


Three cases are considered in this case study to demonstrate several benefits of
cogeneration. These case studies are based directly on the ones reported by
Dincer and Rosen (2013). In each case, the demands for thermal and electrical
energy are specified. Then, three methods are assessed for satisfying the
demands, two using cogeneration and one based on separate processes for heat
and electricity generation. Device efficiencies are specified and minor losses
such as those associated with distribution are neglected. The case study high-
lights the reduction in energy consumption and environmental emissions
and the increase in energy efficiency, when cogeneration is substituted for sep-
arate electrical and heat generation processes.

5.2.1.1 Case 1: Fuel Cogeneration vs. Fuel Electricity Generation and


Fuel Heating
This case considers the substitution of fuel-driven cogeneration for fuel-driven
electricity generation and fuel heating (see Fig. 5.2). A demand for 20 units of

Separate processes for Cogeneration of


heat and electricity heat and electricity

Losses 34
Elec.
Fuel Elec. gen. Elec.
h = 37% 20
54 20 Fuel Cogen.
100 h = 92% Product
Product heat
Fuel Heat gen. heat 72
85 h = 85% 72
Losses 8
Losses 13

139 47 92 100 8 92

FIGURE 5.2 Case 1: fuel cogeneration vs. fuel electricity generation and fuel heating (Dincer and
Rosen, 2013).
178 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.3 Data for Sample Coal and Nuclear Cogeneration Options
Product Energy
Cogeneration Rate (MW) Steam Data Efficiencies (%)

Flow Rate Temperature Pressure


Option Electricity Heat (kg/s) (°C) (MPa) Electricity Heat Total

Coal 267 984 367 361 1.03 20 72 92


Nuclear 199 1423 603 238 0.45 11 81 92

Source: Dincer and Rosen (2013).

electricity and 72 units of product heat is considered. The cogeneration unit is


assumed to operate like the coal cogeneration option described in Table 5.3,
which has electric and thermal efficiencies of 20% and 72%, respectively,
and an overall energy efficiency of 92%. The separate electricity generation pro-
cess is assumed to have the same energy efficiency as the coal-fired Nanticoke
Generating Station (37%), which used to operate in Ontario, Canada, and the
efficiency of the separate heating process is assumed to be 85%, which is typical
of a fuel-fired boiler (MacRae, 1992).
It can be seen in Fig. 5.2 that the cogeneration system consumes 100 units of fuel
energy and losses 8 units of energy, while the two separate processes together
consume 139 units of fuel energy (54 for electricity generation and 85 for heat
production) and lose 47 units of energy (34 from electricity generation and 13
from heat production). Thus, cogeneration substitution decreases fuel energy
consumption here by [(139  100)/139]  100% ¼ 28%. Also, the cogeneration
system efficiency (92%) exceeds that for the electricity generation system (37%),
the heating system (85%), and the combined system containing the separate
electricity generation and heating processes ([92/139]  100% ¼ 66%).

5.2.1.2 Case 2: Nuclear Cogeneration vs. Nuclear Electricity


Generation and Fuel Heating
This case considers the substitution of nuclear cogeneration for nuclear electric-
ity generation and fuel heating (see Fig. 5.3). A demand for 11 units of electric-
ity and 81 units of product heat is considered. The cogeneration unit is assumed
to operate like the nuclear cogeneration option described in Table 5.3, which
has electric and thermal efficiencies of 11% and 81%, respectively, and an over-
all efficiency of 92%. The separate electricity generation process is assumed to
have the same efficiency as Pickering Nuclear Generating Station, which oper-
ates in Ontario, Canada, (30%) and the efficiency of the separate heating pro-
cess is assumed to be 85%.
5.2 Cogeneration 179

Separate processes for Cogeneration of


heat and electricity heat and electricity

Losses 26
Nuclear Elec.
heat Elec. gen. Elec.
Nuclear 11
37 h = 30% 11 heat Cogen.
100 h = 92% Product
Fossil Product heat
fuel Heat gen. heat 81
95 h = 85% 81
Losses 8
Losses 14

132 40 92 100 8 92

FIGURE 5.3 Case 2: nuclear cogeneration vs. nuclear electricity generation and fuel heating (Dincer and
Rosen, 2013).

It is seen in Fig. 5.3 that the cogeneration unit consumes 100 units of nuclear
energy and losses 8 units of energy, while the two separate processes together
consume 132 units of energy (37 units of nuclear energy for electricity gener-
ation and 95 units of fuel energy for heat production) and lose 40 units
of energy (26 from electricity generation and 14 for heat production).
Thus, cogeneration substitution decreases fuel energy consumption here by
[(132  100)/132]  100% ¼ 24% and eliminates fossil fuel consumption.
Also, the cogeneration system efficiency (92%) exceeds that for the electricity
generation system (30%), the heating system (85%), and the combined system
containing the separate electricity generation and heating processes
([92/132]  100% ¼ 70%).

5.2.1.3 Case 3: Fuel Cogeneration vs. Fuel Electricity Generation


and Electrical Heating
This case considers the substitution of fuel-driven cogeneration for fuel-driven
electricity generation and electrical heating (see Fig. 5.4). This case is identical
to case 1 (Fig. 5.2), except that the separate fuel heating process is replaced by
heating (at 95% efficiency) with electricity produced by the separate
electrical plant.
As in case 1, the cogeneration unit consumes 100 units of fuel energy and losses
8 units of energy. However, the two separate processes together consume 259
units of fuel energy (all during electricity generation, which subsequently
supplies 76 units of electricity to the heating process) and lose 167 units of
energy (163 from electricity generation and 4 from heat production). Thus,
180 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Separate processes for Cogeneration of


heat and electricity heat and electricity

Losses 163
Elec.
Fuel Elec. gen. Elec.
h = 37% 20
259 20 Fuel Cogen.
Elec. 76 100 h = 92% Product
Elec. Product heat
heat`g heat 72
h = 95% 72
Losses 8
Losses 4

259 167 92 100 8 92

FIGURE 5.4 Case 3: fuel cogeneration vs. fuel electricity generation and electrical heating (Dincer and
Rosen, 2013).

cogeneration substitution decreases fuel energy consumption here by


[(259  100)/259]  100% ¼ 61%. Also, the cogeneration system efficiency
(92%) exceeds that of both the electricity generation system (37%) and the
combined system containing the separate electricity generation and heating
processes ([92/259]  100% ¼ 36%).

5.2.1.4 Energy and Exergy Efficiencies


For the three cases, energy and exergy efficiencies for the separate processes for
electricity and heat generation and for the cogeneration process are listed in
Table 5.4. In evaluating the efficiencies, it is assumed that the reference-
environment temperature T0 is 15 °C (288 K), the thermal product is delivered
for all cases at an effective temperature T of 150 °C (423 K), the energy and exergy
of the fuel are identical, and the energy and exergy of “nuclear heat” are identical.
The exergetic temperature factor τ in this example is constant for all cases at

Table 5.4 Efficiencies for the Three Cases


Energy Efficiency (%) Exergy Efficiency (%)
Ratio of Efficiencies for
Case Separate Separate Cogeneration and Separate
Study Processes Cogeneration Processes Cogeneration Processes

1 66.2 92 30.9 43.0 1.39


2 70.0 92 27.9 36.9 1.32
3 35.5 92 16.6 43.0 2.59

Source: Dincer and Rosen (2013).


5.2 Cogeneration 181

T0 288K
τ ¼ 1 ¼1 ¼ 0:3191
T 423K

For the first case, the energy and exergy efficiencies, respectively, for the separate
processes taken as a whole are
20 + 72
η¼  ð100%Þ ¼ 66:2%
54 + 85
20 + 72ð0:3191Þ
ψ¼  ð100%Þ ¼ 30:9%
54 + 85

and for the cogeneration process are


20 + 72
η¼  ð100%Þ ¼ 92:0%
100
20 + 72ð0:3191Þ
ψ¼  ð100%Þ ¼ 43:0%
100

Similarly, for the second case, the energy and exergy efficiencies, respectively,
for the separate processes are
11 + 82
η¼  ð100%Þ ¼ 70:0%
37 + 95
11 + 81ð0:3191Þ
ψ¼  ð100%Þ ¼ 27:9%
37 + 95

and for the cogeneration process are


11 + 82
η¼  ð100%Þ ¼ 92:0%
100
11 + 81ð0:3191Þ
ψ¼  ð100%Þ ¼ 36:9%
100

For the third case, the energy and exergy efficiencies, respectively, for the sep-
arate processes are
20 + 72
η¼  ð100%Þ ¼ 35:5%
259
20 + 72ð0:3191Þ
ψ¼  ð100%Þ ¼ 16:6%
259

and for the cogeneration process are


20 + 72
η¼  ð100%Þ ¼ 92:0%
100
20 + 72ð0:3191Þ
ψ¼  ð100%Þ ¼ 43:0%
100
182 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

The ratio of energy efficiencies for the cogeneration and separate processes is
also shown in Table 5.4, along with the corresponding ratio of exergy efficien-
cies. These ratios are the same when based on energy or exergy, mainly due to
the assumptions used in the analysis. It is seen that cogeneration increases both
the energy and exergy efficiencies by 39% for case 1, 32% for case 2, and 159%
for case 3. Two main points are illustrated in Table 5.4:
• Cogeneration increases significantly the energy and exergy efficiencies
compared to separate processes for the same electrical and heating
services.
• The exergy efficiency is markedly lower than the corresponding energy
efficiency for all cogeneration and noncogeneration cases considered.
This is because the thermal product, which is significantly larger than the
electrical product, is delivered at a relatively low temperature (150 °C)
compared to the temperatures potentially achievable.
• As a consequence of the prior point, although cogeneration improves
efficiencies greatly compared to separate processes for each product, there
remains a great margin for further improvement in thermodynamic
efficiencies of the electricity generation and heating processes.

5.2.1.5 Impact of Cogeneration on Environmental Emissions


Four key points related to environmental emissions are demonstrated in the
case study. First, the substitution of cogeneration for separate electrical and heat
generation processes for all cases considered leads to significant reductions in
fuel energy consumption (24–61%), which approximately leads to propor-
tional reductions in emissions. Second, the elimination of fossil fuel consump-
tion in case study 2 eliminates fossil fuel emissions. Third, additional emission
reductions may occur (although these are not evaluated here) when central
cogeneration replaces many small heat producers. Controllable emissions
are reduced at large central stations relative to small plants since better emission
control technologies and stricter emission limits and limit-verification mecha-
nisms often exist at central stations. Fourth, energy losses, which relate to such
environmental impacts as thermal pollution, are reduced significantly with
cogeneration (by 83% for case 1, 80% for case 2, and 95% for case 3).

5.2.1.6 Further Discussion


Several other important points regarding the case study results can be noted as
follows:
• An energy-consumption decrease similar to that for case 3 (61%) would
occur if nuclear cogeneration were substituted for nuclear electricity
generation and electrical heating.
• Sufficient markets must exist for cogenerated heat before cogeneration can
be implemented. Potential markets, and the impacts of the various
5.3 Trigeneration 183

degrees of implementation possible for cogeneration, have been


examined by Dincer and Rosen (2013).
• More detailed and comprehensive assessments are needed regarding how
cogeneration can be integrated into regions before all effects of
utility-based cogeneration can be fully understood. Other investigations
directed towards obtaining a more complete understanding of these
effects have been performed (Dincer and Rosen, 2013).
• The numerical values used in this case study are approximate and only
intended for illustration. However, despite the approximate nature of
these values and the other assumptions and simplifications introduced,
the general findings remain valid when reasonably realistic alternative
values are used.

The key points demonstrated by the case study can be summarized as follows:

• Better efficiencies and emission-reduction technologies for central


cogeneration lead to reduced energy utilization and environmental
emissions when cogeneration replaces separate electricity generation and
heating processes.
• Fuel consumptions and emissions are eliminated when
nuclear-cogenerated heat is substituted for fuel heat.
• Large decreases in energy utilization and emissions occur when
cogenerated heat offsets electrical heat.

5.3 TRIGENERATION
Trigeneration is the simultaneous production of three products in a process.
Often, trigeneration yields electricity, heating, and cooling, in which case, it
is also referred to as CCHP. Trigeneration as CCHP is an extension of CHP.
In addition to heat and electric power produced simultaneously by CHP, CCHP
systems further exploit energy to provide space or process cooling capacity.
CCHP systems in building applications often have “seasonal operation” since
there is little or no cooling load during the winter months in most locations.
Some recent progress in CCHP technologies has been linked to research on dis-
tributed/decentralized energy sources as they can be efficiently implemented in
small distributed scales to meet multiple energy demands of various end users.
CCHPs can also be used to support large-scale applications.
The main advantage of CCHP systems is the higher fuel energy utilization effi-
ciency they exhibit (typically 70–90%) compared to single-generation systems
(around 30–45%). Therefore, they require less input to generate the same
amount of electrical/mechanical and thermal energy. This reduces operating
costs and often leads to reduced lifetime costs. CCHP systems also reduce
184 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

transmission and distribution losses and emissions by consuming less fuel to


meet the same demand. Reliability is sometimes an additional advantage of
small-scale CCHP compared with large-scale centralized plants that can be
more vulnerable to changing environments (i.e., varying customer and market
needs) (Wu and Wang, 2006). CCHP systems are sometimes grid-independent,
which can protect them during electricity blackouts, although that indepen-
dence can also leave them without the grid as a backup in case of system break-
downs. This issue was examined by Alanne and Saari (2006), who compared
the reliability of distributed and centralized energy systems, with an emphasis
on Finland and Sweden.
Typical CCHP system components are (i) an electric power generation unit and
(ii) heating, ventilation, and air conditioning (HVAC) components such as vapor
compression of absorption chillers, cooling towers, heat exchangers, and air
handling units (AHUs). The power generation unit includes a prime mover
(Table 5.1) and electricity generator. Al-Sulaiman et al. (2011) compared CCHP
prime movers using various criteria. The heat recovery unit plays a significant role
in CCHP systems in providing the heat by-product from the prime mover. The
absorption chiller is the most commonly used thermally driven cooling technol-
ogy in CHP/CCHP systems. Several characteristics of absorption technologies,
including operating temperatures, working fluids, cooling capacities, and coeffi-
cient of performances (COPs), are listed in Table 5.5.
Two types of single-effect absorption cooling cycles are included in Table 5.5.
LiBr/water cycles are the simplest option and are widely used. However, since
water is the working fluid, such cycles cannot provide cooling at lower than 0 °C
and they require a water-cooled absorber to prevent crystallization at high
concentrations. In water/NH3 absorption cooling systems, cooling below the

Table 5.5 Technical Characteristics of Available Absorption


Cooling Systems
Operating
Cooling
Temperature (°C)
Capacity
Working (Tons of
Cycle
Heat Input Cooling Fluid Refrigeration) COP

Single-effect 80–110 5–10 LiBr/water 10–100 >0.7


120–150 <0 Water/NH3 3–25 0.5
Double-effect 120–150 5–10 LiBr/water <1000 >1.2
(series)
Double-effect 120–150 <0 Water/NH3 <1000 0.8–1.2
(parallel)
Triple-effect 200–230 5–10 LiBr/water N/A 1.4–1.5

Data source: Srikhirin et al. (2001).


5.3 Trigeneration 185

freezing temperature of water is possible. Water/NH3 systems also avoid the


crystallization problem and have wide operating temperature ranges.
Double-effect absorption cooling technologies with series flow have high
performances and are commercially available. Since the steps are in series,
the output of one step is used as the input to another to increase efficiency.
Double-effect absorption cooling is more efficient than single-effect cooling.
Triple-effect absorption cooling cycles also exist, but they are complex and
require advanced control systems. Since their operating temperatures are higher
than the other options, they require increased maintenance to avoid corrosion.
Each absorption cooling technology in Table 5.5 has advantages and disadvan-
tages. The selection of cooling and heating units depends on the design of the
HVAC components of a CCHP system.
Numerous thermodynamic analyses of CCHP systems have been reported,
often aimed at reducing losses and raising efficiencies. Ahmadi et al. (2012)
compared an integrated organic Rankine cycle (ORC) CCHP system with sim-
pler alternatives and concluded that exergy efficiency of a gas turbine-ORC
CCHP system is higher than that of a CHP system or gas turbine system alone.
Al-Sulaiman et al. (2012) showed that a CCHP system combined with para-
bolic trough solar collectors and an ORC has high energy efficiencies in trigen-
eration mode. The highest energy efficiency they obtained is 94%, which
exceeds that for solar single-generation or cooling/heating cogeneration.
Ozcan and Dincer (2013) thermodynamically analyzed a CCHP system pow-
ered by a solid oxide fuel cell (SOFC) fueled by syngas, integrated with an ORC
driven by the heat of the fuel cell stack exhaust gases, and a LiBr absorption
chiller also driven by SOFC exhaust gases. The energy efficiency of this system
arrangement was found to exceed 50%, significantly higher than that of an
SOFC system operating alone. Suleman et al. (2014) applied energy and exergy
analyses to an integrated solar- and geothermal-based CCHP system and
showed that the integration permitted efficiencies to reach approximately 80%.

5.3.1 Case Study


In this case study, an integrated system for trigeneration of cooling, electric
power, and liquefied hydrogen is examined. The system utilizes a combination
of a quadruple effect absorption system (QEAS), a Linde-Hampson (LH) gas
liquefaction cycle, and a binary power plant. The case study draws on prior
energy and exergy analyses (Ratlamwala et al., 2012).

5.3.1.1 System Description


The trigeneration system considered is shown in Fig. 5.5. Geothermal heat
passes through the very-high-temperature generator (VHTG) before reaching
the binary isobutane power plant, as shown in Fig. 5.6.
186 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Geothermal water Cooling


from source

Hydrogen Liquefied
Precooled hydrogen
gas
QEAS gas Linde-Hampson
cycle

Power Building

Geothermal water
from QEAS
Binary power plant

Return to
underground well

FIGURE 5.5 Schematic of an integrated trigeneration system. Adapted from Dincer and Rosen (2013).

Condenser
8 (con)

39
37 28 5 6
4 9
31 27 Expansion
valve
10
Very-high- High- Medium- Low- 6
temperature temperature temperature temperature Hydrogen
generator generator generator generator Condenser gas
(VHTG) (HTG) (MTG) (LTG)
and heat Evaporator
exchanger (eva) Building
22 23 (CHX)
34 30 26 12 18
36 Exansion 17 11 Precooled
38 valve hydrogen
40 41 32 33 13 14 15 16 gas
Absorber
To binary (abs)
power 34 26 21 Pump
plant
35 29 25 24 20 3 19 2 1

Very-high- High- Medium- Low-


temperature heat temperature heat temperature heat temperature heat
exchanger exchanger exchanger exchanger
(VHHX) (HHX) (MHX) (LHX)
Geothermal water source

FIGURE 5.6 Schematic of the quadruple effect absorption system (QEAS) shown in Fig. 5.5. Adapted
from Dincer and Rosen (2013).

In the VHTG, a strong solution entering at state 36 is heated and exits as a con-
centrated ammonia-water vapor at state 37 and as a weak solution at state 38.
The stream exiting the VHTG at state 38 enters the very-high-temperature heat
exchanger (VHHX) and transfers heat to the stream from the absorber at state 35.
After transferring heat in the VHHX, the stream at state 40 mixes with the stream
from the high-temperature generator (HTG) at state 30 and exits at state 41.
5.3 Trigeneration 187

The process of releasing heat to the strong solution from the weak solution occurs
in the heat exchangers before the weak solution enters the expansion valve at
state 15. The pressure of the weak solution at state 15 is then reduced by throttling
in the expansion valve. The low-pressure (LP) weak solution exits the expansion
valve at state 16 and enters the absorber. The concentrated ammonia-water vapor
at state 37 then enters the HTG, where heat is transferred from the stream at state
37 to that at state 28. The transferred heat from the concentrated ammonia-water
vapor at state 37 heats the strong solution entering HTG at state 34, resulting in a
concentrated ammonia-water vapor exiting at state 31 and a weak solution exit-
ing at state 30. Then, the concentrated ammonia-water vapors at states 28 and 31
mix and enter the medium-temperature generator (MTG) at state 39. The process
continues until the concentrated ammonia-water vapor exiting the low-
temperature generator (LTG) at state 7 enters the condenser and the other
ammonia-water vapor stream exiting the LTG at state 6 is input to the condenser
and heat exchanger, where the concentrated ammonia-water vapor at state 6
rejects heats and passes to the condenser at state 8. In the condenser, concen-
trated ammonia-water vapor flows at states 7 and 8 reject heat to the environ-
ment and are conveyed to the expansion valve, where the pressure decreases
and exiting concentrated ammonia-water vapor at state 10 enters the evaporator.
There, the concentrated ammonia-water vapor gains heat from the hydrogen and
the return air from the building envelope, before exiting at a higher temperature
at state 11. That ammonia-water mixture and the weak solution at state 16 enter
the absorber and reject heat and enter the pump as a strong solution in liquid
form at state 1.
The geothermal water leaving the VHTG is supplied to the binary isobutane
cycle to produce power as seen in Fig. 5.7. The isobutane is heated and

High- 4 Low-
Geothermal 5
pressure pressure
from QEAS turbine
turbine
Heat
exchanger Binary multistage
isobutane cycle Condenser

Reinjection

2 1
Pump

FIGURE 5.7 Schematic of the multistage binary power plant shown in Fig. 5.5. Adapted from Dincer and
Rosen (2013).
188 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

FIGURE 5.8 Schematic of the Linde-Hampson cycle shown in Fig. 5.5. Adapted from Dincer and Rosen
(2013).

vaporized using the geothermal water from the HTG. This vaporized isobutane
flows through the turbine generating power and then is condensed, pumped,
and returned to the heat exchanger. Part of power produced by the binary cycle
is utilized to compress hydrogen gas from state 1 to state 2 in a LH cycle, as
shown in Fig. 5.8. The exiting geothermal water is injected back into the under-
ground water well.

5.3.1.2 Analysis
The mass and energy balances are written for the components of the integrated
system: the QEAS, the isobutane binary multistage cycle, and the LH cycle. Sev-
eral assumptions are made in the analysis and they, along with various input
data values, are listed as follows:

• About 90% of cooling produced by the QEAS is used to precool the


hydrogen from 25 to 16.4 °C in the evaporator, at a constant pressure of
101.3 kPa.
• The remaining 10% of the cooling is supplied to the building.
• In the LH cycle, hydrogen is compressed from 101 kPa to 10 MPa.
• The heat needed to operate the isobutane cycle is provided by geothermal
water at a temperature 20 °C below that of the geothermal source.
• The pressures at the inlet and exit of the stage 1 turbine are 3600 and
1000 kPa, respectively.
• The temperature at the inlet of the stage 1 turbine is 15 °C lower than the
geothermal temperature at the inlet of the binary power plant heat
exchanger.
5.3 Trigeneration 189

• The condenser pressure of the binary multistage power plant is


400 kPa.
• Ninety percent of the electricity produced by the binary multistage plant
is supplied to the LH cycle for compressing hydrogen gas and the
remaining 10% is supplied to the building.
• The concentrations of the ammonia-water strong solution,  [36], and of
the ammonia-water weak solution,  [38], are 0.6 and 0.4, respectively.
• The turbine, compressor, and pump isentropic efficiencies are 80%, 70%,
and 65%, respectively.
The heat input rate to the VHTG of the absorption system is provided using the
geothermal water source as follows:

Q_ VHTG ¼ Q_ geo (5.1)

where

Q_ geo ¼ m_ geo hgeo, source  hgeo, exit

The cooling produced by the QEAS can be calculated from its energy balance as

m_ 10 h10 + Q_ eva ¼ m_ 11 h11 (5.2)

The energetic and exergetic COPs, respectively, of the absorption cooling sys-
tem can be expressed as

Q_ eva
COPen ¼ (5.3)
Q_ geo + W_ p

_ eva
Ex
COPex ¼ (5.4)
_ geo + W_ p
Ex

The actual net power obtained from the binary isobutane plant driven by the
geothermal energy source is expressed as

W_ net, geo ¼ W_ turb  W_ P, iso  W_ parasitic (5.5)

The ideal specific work required to compress hydrogen from P1 ¼ 101 kPa to
P2 ¼ 10 MPa is given as
P2
wcomp, ideal ¼ RT0 ln (5.6)
P1

However, the actual specific work input to liquefaction cycle per unit mass of
hydrogen is expressible as
wcomp, ideal
wcomp, actual ¼ (5.7)
ηcomp
190 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

5.3.1.3 Results and Discussion


Varying the mass flow rate of the geothermal water ṁgeo significantly affects the
performance of the integrated system. An increase in the mass flow rate of geo-
thermal water increases the energy input to the QEAS. The power and cooling
loads provided to the building are observed to vary as the mass flow rate of geo-
thermal water and condenser load change. As seen in Fig. 5.9, the power varies
from 16 to 32 kW and cooling load from 31.6 to 20.6 kW. The increase in QEAS
energy input and the decrease in cooling load with increasing ṁgeo lower the
energetic and exergetic COPs of the QEAS (see Fig. 5.10). These COPs vary,
respectively, from 5.0 to 1.7 and from 1.2 to 0.4 as ṁgeo increases. The decreases
in the COPs directly result from the higher energy input to the QEAS and the
lower output from the QEAS in the form of cooling load. The QEAS perfor-
mance decrease with increasing ṁgeo also affects the rate of hydrogen precool-
ing by the QEAS and liquefaction by the LH system. Figure 5.11 shows that,
with increasing ṁgeo, the amounts of hydrogen gas precooled and liquefied
decrease from 0.53 to 0.34 kg/s and from 0.09 to 0.06 kg/s, respectively.
It is thus observed that an increase in ṁgeo reduces the performance of the
system in terms of cooling production and hydrogen liquefaction. In the case
of power production, however, an increase in ṁgeo enhances the system power
production.
The geothermal water source temperature Tgeo plays an important role in
designing and assessing the system. That parameter also determines if the sys-
tem is capable of producing power or limited to low-grade energy applications.
Figure 5.12 shows that increasing Tgeo decreases the energetic and exergetic

35 35
Qcon = 150 kW
Qcon = 175 kW
31 Qcon = 200 kW 32
Tgeo = 573 K
Qcooling,House (kW)
PowerHouse (kW)

27 29

23 26

19 23

15 20
1.5 1.75 2 2.25 2.5 2.75 3
mgeo (kg/s)

FIGURE 5.9 Effect of mass flow rate of the geothermal source fluid on electric power and cooling
supplied to a building. Adapted from Dincer and Rosen (2013).
5.3 Trigeneration 191

5.5 1.3
COPen at Qcon = 150 kW
5 COPen at Qcon = 175 kW 1.2
COPen at Qcon = 200 kW
4.5 1.1
COPex at Qcon = 150 kW
4 COPex at Qcon = 175 kW
1
COPex at Qcon = 200 kW
COPen

3.5 0.9

COPex
Tgeo = 573 K
3 0.8

2.5 0.7

2 0.6

1.5 0.5

1 0.4
1.5 1.75 2 2.25 2.5 2.75 3
mgeo (kg/s)

FIGURE 5.10 Effect of mass flow rate of the geothermal source fluid on energetic and exergetic
coefficient of performances (COPs). Adapted from Dincer and Rosen (2013).

0.58 0.108
mH2, precooled at Qcon = 150 kW
0.56 mH2, precooled at Qcon = 175 kW 0.104
0.54 mH2, precooled at Qcon = 200 kW 0.1
0.52 mH2, liquefied at Qcon = 150 kW 0.096
mH2 precooled (kg/s)

mH2, liquefied at Qcon = 175 kW mH2, liquefied (kg/s)


0.5 0.092
mH2, liquefied at Qcon = 200 k
0.48 0.088
Tgeo = 573 K
0.46 0.084
0.44 0.08
0.42 0.076
0.4 0.072
0.38 0.068
0.36 0.064
0.34 0.06
1.5 1.75 2 2.25 2.5 2.75 3
mgeo (kg/s)

FIGURE 5.11 Effect of mass flow rate of the geothermal source fluid on rate hydrogen gas precooled
and liquefied. Adapted from Dincer and Rosen (2013).

COPs of the QEAS. For a fixed ṁgeo of 2.0 kg/s, the energetic and exergetic COPs
vary from 3.7 to 2.9 and from 1.1 to 0.7, respectively, as Tgeo increases from 473
to 573 K for condenser loads of 150, 175, and 200 kW. An increase in the tem-
perature of the geothermal source water results in a higher energy transfer to the
VHTG, due to the higher potential of losing heat to the low-temperature stream
192 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

4 1.3
COPen at Qcon = 150 kW COPex at Qcon = 150 kW
3.9 1.25
COPen at Qcon = 175 kW COPex at Qcon = 175 kW
3.8 COPen at Qcon = 200 kW COPex at Qcon = 200 kW 1.2
3.7 mgeo = 2 kg/s 1.15
3.6 1.1
3.5 1.05

COPen

COPex
3.4 1
3.3 0.95
3.2 0.9
3.1 0.85
3 0.8
2.9 0.75
2.8 0.7
470 492 514 536 558 580
Tgeo (K)

FIGURE 5.12 Effect of geothermal source temperature on energetic and exergetic COPs.
Adapted from Dincer and Rosen (2013).

entering the VHTG from the VHHX. This increase in energy input to the QEAS
results in a lower QEAS output in terms of cooling load, for a fixed condenser
load. The decrease in cooling load reduces the COPs of the QEAS. As the per-
formance of the QEAS degrades, its capability of precooling hydrogen also
decreases. In addition, an increase in Tgeo also affects the mass of precooled
hydrogen gas and the mass of hydrogen gas liquefied, as can be seen in
Fig. 5.13. The quantities of hydrogen gas precooled and liquefied vary from
0.5 to 0.41 kg/s and from 0.005 to 0.09 kg/s, respectively (Fig. 5.14).

0.504 0.11
0.495 0.1
0.486 mH2, precooled at Qcon = 150 kW 0.09
mH2, precooled at Qcon = 175 kW
mH2 precooled (kg/s)

0.477 mH2, precooled at Qcon = 200 kW 0.08


mH2, liquefied (kg/s)

0.468 0.07
0.459 0.06
0.45 0.05
0.441 Tgeo = 573K 0.04
0.432 mH2, liquefied at Qcon = 150 kW 0.03
0.423 mH2, liquefied at Qcon = 175 kW 0.02
mH2, liquefied at Qcon = 200 kW
0.414 0.01
0.405 0
470 492 514 536 558 580
Tgeo (K)

FIGURE 5.13 Effect of geothermal source temperature on the rate hydrogen gas precooled and liquefied.
Adapted from Dincer and Rosen (2013).
5.4 Integrated Systems 193

1.1 0.22
mH2, precooled at Qcon = 150 kW
1 mH2, precooled at Qcon = 175 kW 0.2

0.9 mH2, precooled at Qcon = 200 kW 0.18


mH2, liquefied at Qcon = 150 kW
mH2 precooled (kg/s)

mH2, liquefied (kg/s)


0.8 mH2, liquefied at Qcon = 175 kW 0.16
mH2, liquefied at Qcon = 200 kW
0.7 0.14
Tgeo = 573 K
0.6 0.12
mgeo = 2 kg/s
0.5 0.1

0.4 0.08

0.3 0.06

0.2 0.04
270 275 280 285 290 295 300 305
Ambient temperature (K)
FIGURE 5.14 Effect of ambient temperature on the rate hydrogen gas precooled and liquefied. Adapted
from Dincer and Rosen (2013).

5.4 INTEGRATED SYSTEMS


As pointed out earlier, integrated energy systems integrate multiple processes in
beneficial ways. Integrated energy systems often are taken to include cogenera-
tion, trigeneration, multigeneration, and district energy systems, but they are
often broader. Integration of energy systems can involve locating industrial
facilities in close proximity so wastes (heat or material) from one can poten-
tially be provided to another that needs them as inputs. Integration can also
be used to permit energy systems to allow cascading of waste heat at specific
temperatures to users progressively needing heat at those temperatures.
Multigeneration, the simultaneous generation of more than one product (usu-
ally more than three) from a process, is a particularly interesting type of inte-
grated energy system. It is usually achieved by combining several processes and
often uses a single-energy input. Integrated systems often involve the produc-
tion with multiple product outputs, sometimes using multiple systems, from
one or multiple inputs. When producing cooling and heating using multige-
neration systems, absorption cooling and heating systems are often incorpo-
rated as such systems can utilize waste heat from thermal power plants to
produce cooling and heating simultaneously. Furthermore, multigeneration
systems can be integrated with drying processes or with an electrolyzer or other
process for hydrogen production.

5.4.1 Case Study: Integrated System for HVACR Applications


The integrated system proposed by Ghosh and Dincer (2015) consists of two
open cycles, namely, a heating and cooling cycle and a closed refrigerant cycle.
194 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

The heating and cooling cycle uses an atmospheric air-water mixture as the
working fluid while the refrigerant cycle uses R-134a. The following psychro-
metric processes are utilized in the integrated system:

• Cooling with dehumidification


• Space cooling
• Evaporative cooling (cooling tower)
• Heating with humidification
• Space heating

The integrated system produces multiple outputs depending on the loading con-
ditions. For instance, on a hot day, atmospheric air is first cooled in the evapo-
rator of the vapor compression cycle. The heat rejected by the atmospheric air is
utilized by the evaporator to increase the temperature of R134a. The cooled air is
then used for space cooling and further utilized to cool hot water for residential
purposes. But on a cold day, the atmospheric air is first heated and humidified
using the heat rejected in the condenser of the vapor compression cycle. The air is
then used for space heating and followed by ventilation.
Figure 5.15 shows a schematic diagram of the integrated system for heating,
ventilation, air conditioning, and refrigeration (HVACR) applications. On a
hot day, atmospheric air at 32 °C is first cooled and dehumidified to 20 °C
at state point 2. This air is used for space cooling for a residential application
and exits at 28 °C at state point 4. This air is further used to cool water from 40
to 32 °C in the cooling tower for residential purposes. The air is released to the
atmosphere from the cooling tower at 33 °C.
On a cold day, atmospheric air at 10 °C is first heated and humidified to 48 °C
from state point 12 to 15. The heated air is then used for space heating for res-
idential or industrial purposes, leaving the building at 32 °C.

5.4.1.1 Analysis
Comprehensive energy and exergy analyses are performed for the integrated
system, providing substantial information on its performance, efficiency,
and emissions. The following input data and assumptions are utilized in the
analysis of the system:

• Atmospheric air can be treated as an air-water mixture.


• The ambient temperature and pressure, respectively, are 297 K and
101.3 kPa on an average day, 305 K and 101.3 kPa on a hot summer day,
and 283 K and 101.3 kPa on a cold winter day.
• The ambient relative humidity of atmospheric air is 0.95.
• No heat losses occur in the expansion valve in the vapor
compression cycle.
• The compression of refrigerant in the compressor is isentropic.
Saturated vapor E
F
14 D

12 13 15 16 17
V
Refrigerant cycle (R-134a)

Atmospheric air II
10 9 Equipment list
Water I Compressor
II Condenser
III Expansion valve
III I IV Evaporator
V Cooling tower
5 VI Fan

11 8

V B Processes
A Cooling with dehumidification
IV
Hot water B Space cooling

6 C Evaporative cooling
D Heating with humidification
Cold water C 4 2 3 1
A

5.4
E Space heating
7 F Ventilation
Condensate

Integrated Systems
Makeup water
FIGURE 5.15 Integrated system process flow diagram. Adapted from Ghosh and Dincer (2015).

195
196 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

The efficiency for different psychrometric subprocesses is determined following


Ratlamwala and Dincer (2012). The efficiency of the cooling with dehumidi-
fication process (A in Fig. 5.15) is based on the concept that the efficiency of
the system is expressible as the ratio of the energy carried by the exiting stream
to the heat rejected by the system. This definition is founded on the purpose of
the system being to cool the incoming stream; therefore, heat rejected by the
system is the actual input to the system. Consequently, we can write
m_ a1 h1
ηcd ¼ (5.8)
Q_ A
m_ a1 ex1
ψ cd ¼   (5.9)
T0
Q_ A 1 
T

For the space cooling process (B in Fig. 5.15), we define the efficiency as the
ratio of the change in energy/exergy by the energy/exergy input to the process,
where the change in energy/exergy is the difference between the energy/exergy
of the stream entering the system and energy/exergy of the stream exiting.
That is,

m_ a4 h4 + Q_ B
ηsc ¼ (5.10)
m_ a2 h2
 
T0
m_ a4 ex4 + Q_ B 1 
T
ψ sc ¼ (5.11)
m_ a2 ex2

With this definition, energy/exergy output is heat rejected from the process.
For the evaporative cooling process in the cooling tower (C in Fig. 5.15), the
efficiency is the ratio of the energy gained by the process to the energy provided.
Then, we can write energy and exergy efficiencies for the process as
m_ a5 h5 + m_ w7 h7
ηct ¼ (5.12)
m_ a4 h4 + m_ w6 h6
m_ a5 ex5 + m_ w7 ex7
ψ ct ¼ (5.13)
m_ a4 ex4 + m_ w6 ex6

In these definitions, the energy gained by the process is the difference between
the energy carried by the exiting stream and the energy carried by the entering
stream (Dincer, 2012). However, energy provided to the process is the energy of
the input stream and energy of the water.
For the heating with humidification process (D in Fig. 5.15), the efficiency is
based on the desired output of the process being the amount of energy gained
by the process and the required input to the process being the energies added to
5.4 Integrated Systems 197

the process via heat and hot water. The desired output is found by subtracting
the energy of the stream leaving the process to the energy of the stream entering.
The required input to the process is found by adding the heat added to the pro-
cess and energy carried by the input hot water. Consequently, the energy and
exergy efficiencies for the process can be expressed as follows:

m_ a15 h15  m_ a12 h12


ηhh ¼ (5.14)
Q_ in, 1215 + m_ w14 h14

m_ a15 ex15  m_ a12 ex12


ψ hh ¼   (5.15)
T0
Q_ in, 1215 1  + m_ w14 h14
T

For the space heating process (E in Fig. 5.15), the efficiency is defined as the ratio of
change in energy/exergy by the energy/exergy input to the process. With this def-
inition, the change in energy/exergy is the difference between the energy/exergy of
the stream entering the process and the energy/exergy of the stream leaving. More-
over, the energy/exergy input to this process is heat provided. As a result,
m_ a16 h16  m_ a15 h15
ηsh ¼ (5.16)
Q_ E
m_ a16 ex16  m_ a15 ex15
ψ sh ¼   (5.17)
T0
Q_ E 1 
T

The efficiencies of the integrated system can be defined differently under differ-
ent loading conditions. For instance, on a hot summer day with an ambient tem-
perature of 32 °C, the requirement could be of 100% cooling load and 0%
heating load. The energy efficiency for this loading condition would be the total
useful energy output (Q_ C for space cooling, the change in energy between hot
and cold waters in the cooling tower and Q_ cond , the useful heat from the con-
denser) divided by the total input energy (input air and compressor work input):

Q_ C + m_ w6 ðh7  h6 Þ + Q_ cond
ηc ¼ (5.18)
m_ a1 h1 + W_ comp

Correspondingly, the exergy efficiency for a 100% cooling load can be


defined as
   
T0 T0
Q_ C 1  + m_ w6 ðex7  ex6 Þ + Q_ cond 1 
T2 T9
ψc ¼ (5.19)
m_ a1 ex1 + W_ comp

On a cold winter day with an ambient temperature of 10 °C, the requirement


could be of 0% cooling load and 100% heating load. The energy efficiency for
198 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

this loading condition would be the total useful energy output (Q_ h for space
heating) divided by the total input energy (input air, compressor work input,
and the evaporator input energy). That is,

Q_ h
ηh ¼ (5.20)
m_ a12 h12 + W_ comp + Q_ evap

Correspondingly, the exergy efficiency for a 100% heating load can be defined as
 
T12
Q_ h 1 
T15
ψh ¼   (5.21)
T12
m_ a12 ex12 + W_ comp + Q_ evap 1 
T15

The overall energy and exergy efficiencies of the integrated system on an average
day with an ambient temperature of 24 °C can be defined according to a
requirement of 50% cooling load and 50% heating load, as follows:

Q_ c + m_ w6 ðh7  h6 Þ + Q_ h
ηsys ¼ (5.22)
m_ a1 h1 + m_ a12 h12 + W_ comp
   
T0 T12
Q_ c 1  + m_ w6 ðex7  ex6 Þ + Q_ h 1 
T2 T15
ψ sys ¼ (5.23)
m_ a1 ex1 + m_ a12 ex12 + W_ comp

5.4.1.2 Results and Discussion


Table 5.6 lists the thermodynamic properties at each state point in the inte-
grated system. Based on the balance equations, assumptions, and values con-
sidered, the following efficiency values for the overall system and different
loading conditions are calculated using the software package, Engineering
Equation Solver (EES):
ηc ¼ 18:5% ψ c ¼ 18:1%
ηh ¼ 77:4% ψ h ¼ 75:4%
ηsys ¼ 18:6% ψ sys ¼ 33:3%

Both energy and exergy efficiency values for each subprocess are also calculated.
The following values are obtained:
ηcd ¼ 68:0% ψ cd ¼ 16:2%
ηsc ¼ 100% ψ sc ¼ 94:6%
ηct ¼ 100% ψ ct ¼ 90:6%
ηhh ¼ 100% ψ hh ¼ 77:3%
ηsh ¼ 100% ψ Sh ¼ 97:4%
W_ comp ¼ 10, 642kW
5.4 Integrated Systems 199

Table 5.6 Thermodynamic Properties at Each State Point for the Integrated System
State Point Fluid Type P (kPa) T (K) h (kJ/kg) s (kJ/kg-K) ex (kJ/kg) ṁ (kg/s)

1 Air 101.3 305.2 106.6 5.977 0 1105


2 Air 101.3 293.2 42.3 5.76 1.833 1105
3 Water 101.3 301.2 117.3 0.4088 0.1097 19.89
4 Air 101.3 296 45.22 5.77 5.77 1105
5 Air 101.3 306 98.45 5.95 5.95 1105
6 Water 101.3 313.2 167.5 0.5723 0.4337 1725
7 Water 101.3 305 133.4 0.462 0.0001139 1725
8 R-134a 1000 277.2 252.8 0.9293 30.08 472.3
9 R-134a 1000 316.3 275.3 0.9293 52.61 472.3
10 R-134a 101.3 312.5 107.3 0.3919 48.61 472.3
11 R-134a 101.3 277.2 107.3 0.4045 44.76 472.3
12 Air 101.3 283 27.98 5.71 0 1105
13 Air 101.3 321.9 147.9 6.111 2.188 1105
14 Saturated vapor 101.3 373.1 2676 7.355 439 19.89
15 Air 101.3 321.9 147.9 6.111 2.188 1105
16 Air 101.3 305 70.02 5.855 1.229 1105

Note that Ẇcomp is the work rate required for the compressor to reach a pressure
of 1 MPa. The energy efficiency of space heating is calculated to be 100%
because no heat loss to the surroundings is assumed. Any heat rejected/given
to the system is utilized as useful output (heating or cooling). Figure 5.16 pre-
sents the ratio of exergy destruction of each component in the integrated system
to the total exergy destruction of the system. It is seen that the condenser con-
tributes the most to the overall exergy destruction and the compressor the least.

0.40

0.35
Exergy destruction ratio

0.30

0.25

0.20

0.15

0.10

0.05

0.00
Cooling Compressor Condenser Space Evaporator Space Throttling
tower cooling heating valve

FIGURE 5.16 Ratio of exergy destruction of component to the total exergy destruction. Adapted from
Ghosh and Dincer (2015).
200 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.7 Breakdown by Component of Exergy Destruction Rates


for the Integrated System
Component Exergy Destruction Rate (kW)

Cooling tower 77.22


Compressor 0
Condenser 6425
Space cooling 108.8
Evaporator 4600
Space heating 5929
Throttling valve 1819
Integrated system 18,959
Source: Ghosh and Dincer (2015).

This is somewhat artificial since the compressor is assumed to be isentropic,


which leads to no exergy destruction. Space heating also contributes notably
to the exergy destruction, mainly due to the large change in exergy between
the inlet and outlet streams. Table 5.7 presents the exergy destruction rates
for the overall integrated system and its components.

5.5 DISTRICT HEATING AND COOLING


As world demand for fossil fuels increases and energy supplies become harder
to access, the price of energy is expected to rise. Governments and businesses are
paying increasingly close attention to the role of energy in the design and oper-
ation of buildings and communities. District energy or DHC can help commu-
nities advance sustainable growth goals and manage the changing nature of risk
in the generation and delivery of energy while contributing to broader commu-
nity economic development. Part of the interest of businesses, industries, and
municipalities in DHC systems is due to concerns about the reliability of
municipal infrastructure and the effects of GHGs, coupled with opportunities
presented by the emergence of alternative fuels and improvements in technol-
ogy that allows for local-scale energy production.
District energy is a recognized approach for meeting the heating, cooling, and
domestic hot water needs of buildings and communities, which can also sup-
port the process heating requirements of local industry. District energy serves to
manage the thermal needs of energy consumers at a building or community
level. As a management system, district energy can help accommodate and meet
the energy demands of buildings and industries, which vary in terms of usage
amount, rate, and pattern. By linking buildings and industrial activities
together through a thermal network, district energy aggregates the varying
5.5 District Heating And Cooling 201

energy demands into a steady thermal load that can be efficiently managed
(Dincer and Rosen, 2013).
District energy should not be confused with energy generation technologies,
since no energy is produced. District energy provides a medium that allows
for the transfer of energy. The medium used to transport energy from an energy
supplier to an energy consumer is usually steam, hot water, or chilled water, but
other media can also be used.
A DHC system may be designed with a central energy plant, a series of smaller
plants, or multiple plants connected by pipes that provide space heating, air
conditioning, hot water, steam, and chilled water to any group of buildings.
Some district energy systems provide electric power using cogeneration or
CHP (Dincer and Rosen, 2013). DHC generally involves of three subprocesses:

i. Collection and/or generation of thermal energy


ii. Distribution of the thermal energy from plant sites to a network of
energy consumers
iii. Transfer of the thermal energy to the energy consumer

As a result, individual buildings served by a district energy system do not need


their own boilers or furnaces, chillers, or air conditioners. The district energy
system replaces those devices, mainly with a simple heat exchanger. District
energy can provide numerous benefits:

• Improved energy efficiency


In some instances, district energy can improve efficiency. In part, this is
because when steam, hot water, or chilled water is transferred to a
customer’s building, they are ready to use. Other energy supplies like
natural gas or fuel oil may be only 80% or less efficient at a building as
they must be combusted. In addition, district energy systems can use the
waste heat rejected from burning fuel to produce electricity at a power
plant, increasing the overall efficiency with which useful energy is
extracted from the fuel.
In some instances, the rejected waste heat can be used to drive turbines
and generate electricity. This cogeneration arrangement can produce both
heating and cooling plus electricity for customers. A CHP system may
have double the fuel efficiency of an electricity generation plant and can
also lower the emissions typically associated with conventional fossil
fuel-powered electrical production. The less fuel used, the less sulfur
dioxide, carbon dioxide, and other combustion products are emitted to
the environment.
• Reduced environmental impact
District energy enables building owners and managers to use less energy,
improve operating efficiency, and protect the environment. With district
202 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

energy, building managers no longer need to burn fuels or store or use


refrigerants on site, so the site is safer and more environmentally sound
and does not require smokestacks. Instead, fuel and refrigerants are used
at district energy plants. These systems normally employ stringent
emission controls—more so than individual buildings—providing air
quality benefits.
• Fuel flexibility
An important advantage of district energy is that since it serves many
customers with one or more energy conversion devices, performance
flexibility may exist that is not found in individual buildings. For instance,
many district energy systems can use a variety of conventional fuels
such as coal, oil, and natural gas, depending on which is most competitive
at the time. Further, a large district energy system can also transition to
use nuclear or renewable thermal energy. The latter can include various
forms of biomass (e.g., wood and food processing waste), geothermal
heat, and cogenerated heat.
• Ease of operation and maintenance
District energy systems provide heating and/or cooling to user buildings.
Building owners do not need boilers or chillers, so there is less
maintenance, monitoring, and equipment permitting. District energy
customers also avoid the need for fuel delivery, handling, and storage so
there are fewer safety and liability concerns for employees and building
occupants. The use of district energy service frees up valuable building
space by eliminating the need for mechanical rooms.
• Reliability
District energy systems normally provide heating and/or cooling with
high reliability, often operating at a reliability of 99.999%. District energy
owners usually provide backup systems.
• Comfort and convenience for customers
District energy service allows building operators to manage and control
their indoor environments, providing comfort at all outdoor
temperatures. In addition, district energy reduces vibrations and noise
problems associated with central HVAC equipment that can be annoying
to building occupants.
• Reduced life-cycle costs
Since buildings using district energy do not need boilers or chillers,
building owners and managers reduce their up-front capital requirements
and their ongoing, operating, maintenance, and labor costs considerably.
That reduces financial risk and can improve return on investment.
District energy can lead to reduced property taxes associated with new
boiler and chiller installations, insurance and annual maintenance
contracts, and costs associated with operating boilers and chillers.
In addition, district energy systems have the flexibility to use a variety
5.5 District Heating And Cooling 203

of fuel sources in larger, more economical volumes—from oil, to natural


gas, to coal, to biomass—reducing the impact of supply and price
variations.
• Decreased building capital costs
Buildings connected to district energy systems often have lower capital
costs for their energy equipment because they do not require conventional
boilers and chillers, and they save building space, allowing it to be used
for revenue generation.
• Improved architectural design flexibility
District energy provides greater building design flexibility, since boilers or
furnaces are avoided, as are smokestacks and cooling towers on roofs.
Architects can design or renovate buildings to be more versatile and
esthetically pleasing for both occupants and the community when district
energy is employed.
Figure 5.17 shows the typical layout of a DHC system that provides thermal
energy to several end users, including offices, industries, institutions, and res-
idential buildings.
Energy can be input to DHC systems using various devices that include conven-
tional fossil fuel boilers, biomass boilers, CHP, heat pumps, and technologies
that recover waste heat from municipal, commercial, and industrial activities.
The types of energy generation technologies and components used in a DHC
system are selected and/or designed to exploit available local fuel sources
and waste heat sources. Table 5.8 identifies some of the energy generation tech-
nologies and fuel sources used in DHC systems. Figure 5.18 illustrates the oper-
ation of a DHC system (which is CHP-based).

FIGURE 5.17 Layout of a district heating and cooling system.


204 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.8 Common Technologies and Fuels Used for District Energy
Boiler/Generator Product
Plant Type Technology Possible Fuel Sources Output

Heat only (hot water and steam) Combustion turbine Natural gas, liquid fuels Steam/electricity
and CHP 5–25 MW
Reciprocating engine Natural gas, diesel, biogas, Hot water
liquid fuels 0.5–7 MW
Microturbine Natural gas, hydrogen, Hot water
propane, diesel 25–500 kW
Fuel cell Natural gas, hydrogen Hot water/steam
1 kW–10 MW
Stirling engine Natural gas, biogas Hot water
1–25 kW
Chilled water Lake/ocean water Water Cold water
Chiller Steam/hot water/electricity Chilled water
Thermal storage Incorporates one or more large storage tanks of hot or chilled water or ice linked to
a district energy system. Storage types include steel storage tanks, aquifers, and
boreholes
Source: Canadian District Energy Association (2008).

Fuel source Central production plant Electricity utility


• Waste heat • Steam
Fuel Electricity Distributes
• Biomass • Electricity
electricity to local
• Fuel oil • Hot water
• Natural gas • Chilled water customer base

Government Business
applications applications
Steam
Electricity
Hot water
Chilled water
Commercial/
Residential
industrial
applications
applications

FIGURE 5.18 Typical district energy configuration. Modified from Canadian District Energy Association,
2008.
5.5 District Heating And Cooling 205

5.5.1 Case Study: Cogeneration-Based District Energy


A major cogeneration-based DHC project in Downtown Edmonton, Alberta,
Canada, is considered. The system (Edmonton Power, 1991; MacRae, 1992)
has (i) an initial supply capacity of 230 MW (thermal) for heating and 100 MW
(thermal) for cooling, (ii) the capacity to displace about 15 MW of electric power
used for electric chillers through district cooling, and (iii) the potential to increase
the efficiency of the Rossdale power plant that would cogenerate to provide the
steam for DHC from about 30% to 70%, respectively. The design includes the
potential to expand the supply capacity for heating to about 400 MW (thermal).
The design incorporated central chillers and a district cooling network. Screw
chillers were to be used originally, and absorption chillers in the future.
Central chillers are often favored because (i) the seasonal efficiency of the
chillers can increase due to the ability to operate at peak efficiency more often
in a central large plant and (ii) lower chiller condenser temperatures (e.g.,
20 °C) can be used if cooling water from the environment is available to the
central plant, relative to the condenser temperatures of approximately 35 °C
needed for air-cooled building chillers. These two effects can lead to central
large chillers having almost double the efficiencies of distributed small chillers.
There are two main stages in this analysis. First, the design for cogeneration-
based DHC (Edmonton Power, 1991; MacRae, 1992) is evaluated thermody-
namically. Then, the design is modified by replacing the electric centrifugal
chillers with heat-driven absorption chillers (first single- and then double-effect
types) and reevaluated by Dincer and Rosen (2013).

5.5.1.1 Original System


The cogeneration-based district energy system considered (see Fig. 5.19)
includes a cogeneration plant for heat and electricity and a central electric
chiller that produces a chilled fluid. Hot water is produced, to satisfy all heating
requirements of the users, at a temperature and pressure of 120 °C and 2 bar,
respectively. The heat is distributed to the users via heat exchangers, district
heating grids, and user heat-exchanger substations. A portion of the cogener-
ated electricity is used to drive a central centrifugal chiller and the remaining
electricity is used for other purposes (e.g., export, driving, and other electrical
devices). The central chiller produces cold water at 7 °C, which is distributed to
users via district cooling grids.

5.5.1.2 Modified System


For the cogeneration-based district energy system using absorption chillers, the
design is modified by replacing the electric chiller with single-effect absorption
chillers (see Fig. 5.20). Hot water is produced at 120 °C and 2 bar to satisfy all
206 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

·
Wnet
Waste
A
· C E
Ef
Heating 50 ⬚C Domestic
Cogeneration Hot water supply (80 ⬚C) substation • u,w hot water
60 ⬚C QH
· •u
Air QH Hot water supply (60 ⬚C) QH Space heating 22 ⬚C
• u,s
40 ⬚C QH

· Cold water supply (7 ⬚C) 15 ⬚C Space cooling


Wch Central chiller · •u • u,r
(centrifugal chiller) QC Warm water return (15 ⬚C) QC 22 ⬚C QC
Cooling substation 22 ⬚C

B D F

· ·
Waste (=Wch + QC)
Production of electricity, Transport of heat and cool End use
heat and cool via district energy heating and cooling

FIGURE 5.19 Simplified diagram of the cogeneration-based district energy system of Rossdale Power
Plant. The system, which uses electric chillers, is divided into six subsections within three categories.
On the left are production processes, including cogeneration of electricity and heat (A) and chilling
(B). In the middle are district energy transport processes, including district heating (C) and district cooling
(D). On the right are end-user processes, including user heating (E) and user cooling (F). Adapted
from Dincer and Rosen, 2013.

heating requirements of the users and to drive the central absorption chiller.
A small portion of the cogenerated electricity is used to drive the absorption
solution and refrigeration pumps, and the remaining electricity is used for pur-
poses other than space cooling.
This cogeneration-based district energy system is then further modified by
replacing the electric centrifugal chillers with double-effect absorption chillers
(see Fig. 5.20). The system is similar to the cogeneration-based district energy
system using single-effect absorption chillers, except that higher-quality heat
(170 °C and 8 bar) is produced to drive the double-effect absorption chillers.

5.5.1.3 Approach and Data


The plant is divided into subsections for analysis purposes. Efficiencies of the
individual subsystems and the overall system are examined. Also, several
selected combinations of the subsystems are evaluated to pinpoint better the
locations and causes of inefficiencies.
For the analysis, the year is divided into two seasonal periods (see Table 5.9).
Period 1 (October to April) has an environmental temperature of 0 °C and is
considered to be a winter period with only a heating demand. Period 2
5.5 District Heating And Cooling 207

Wnet
Waste

A
Ef

Cogeneration

Air QH

Wch
Central chiller
(absorption chiller) Qc
Qgen
B

Waste (=Wch + Qc + Qgen)

FIGURE 5.20 Modified version of production processes (units A and B) for the simplified diagram in
Fig. 5.19. In the modified system, the electric chillers are replaced with absorption chillers (single- or
double-effect), driven mainly by heat from the cogeneration plant. The rest of the system in Fig. 5.19
(units C–F) remains unchanged in the modified system. The temperature of the heating medium supplied
to the absorption chillers is higher for the double-effect chiller relative to the single-effect machine.
Adapted from Dincer and Rosen, 2013.

Table 5.9 Monthly Heating and Cooling Load Breakdown (in %) in the Design Area of Edmonton,
Alberta
Period 1 (Winter) Period 2 (Summer)

Load Oct. Nov. Dec. Jan. Feb. Mar. Apr. Total May Jun. Jul. Aug. Sep. Total

Heating 6.90 12.73 16.83 18.67 14.05 12.95 7.34 89.46 2.39 1.56 1.34 1.92 3.33 10.54
Cooling 0 0 0 0 0 0 0 0 10.62 22.06 32.00 26.80 8.52 100
Source: Edmonton Power (1991).

(May to September) has an environmental temperature of 30 °C and is consid-


ered to be a summer period with a cooling demand and a small heating
demand for hot water heating. The small variations in plant efficiency that
occur with changes in environmental temperature are neglected here.
Rossdale power plant has an annual free cooling of 33 GWh/year; the cooling
requirement of the chilling plant is 169 GWh/year. The COP of the centrifugal
chiller in the design is 4.5 (Edmonton Power, 1991). Thus, the annual electricity
208 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

supply rate to the chiller is Ẇch ¼ 169/4.5 ¼ 38 GWh/year. For the chilling oper-
ation, including free cooling and electrical cooling, COP ¼ (169 + 33)/38 ¼ 202/
38 ¼ 5.32. The net electricity output Ẇnet of the combined cogeneration/chiller
portion of the system is 433  38 ¼ 395 GWh/year, where the electricity genera-
tion rate of the cogeneration plant is 433 GWh/year.
The overall energy efficiency of the proposed cogeneration plant is 85%, the elec-
tric efficiency (i.e., the efficiency of producing electricity via cogeneration) is
25%, and the heat production efficiency is 60%. Also, the total heating require-
ment of the buildings in the design region is Q_ H ¼ 1040GWh=year for space and
hot water heating, and the cooling requirement is Q_ C ¼ 202GWh=year for space
cooling (DOE, 2000). The total fuel energy input rate can be evaluated for the
cogeneration plant using electric chillers as E_f ¼ 1064=0:6 ¼ 1733GWh=year.
Since 33 GWh/year of this cooling is provided through free cooling, the cooling
requirement of the chilling plant is 169 GWh/year (Edmonton Power, 1991).
The COP of the single-effect absorption chiller used here is taken to be 0.67,
a typical value (Colen, 1990). Therefore, the annual heat required to drive the
single-effect absorption machine is Q_ gen ¼ 169=0:67 ¼ 252GWh=year. The total
fuel energy input rate to the cogeneration plant can thus be evaluated as
E_f ¼ ð1040 + 252Þ=0:6 ¼ 2153GWh=year.
As mentioned above, steam is required at higher temperatures and pressures to
drive the double-effect absorption chillers, and more electricity is curtailed as
higher-quality heat or more heat is produced. It is assumed that the overall energy
efficiency of the proposed cogeneration plant is unchanged (85%) in period 2.
Only the electric and heat efficiencies are changed due to more heat being pro-
duced in this period, when the absorption chiller is in operation. Thus, in periods
1 and 2, respectively, the electric efficiency (i.e., the efficiency of producing elec-
tricity via cogeneration) is 25% and 21% (Rosen and Le, 1998), and the heat pro-
duction efficiency is 60% and 64% (Rosen and Le, 1995). The COP of the
double-effect absorption chiller used here is taken to be 1.2, a typical value
(Colen, 1990). Therefore, the annual heat required to drive the double-effect
absorption machine is Q_ gen ¼ 169=1:2 ¼ 141GWh=year. The total fuel energy
input rate to the cogeneration plant can be evaluated as the sum of the fuel energy
input rate to the plant in the two periods. Thus, E_f ¼ 1942GWh=year.
The COP for the chilling operation, including free cooling, using single-effect
absorption cooling is COP ¼ 202/252 ¼ 0.80 and, using double-effect absorp-
tion cooling, is COP ¼ 202/141 ¼ 1.43. It is noted for the absorption chiller
cases that, since the work required to drive the solution and refrigeration
pumps is very small relative to the heat input (often less than 0.1%), this work
is neglected here.
For simplicity, economics and part-load operation are not considered here, so
the results and findings are thus correspondingly limited. Also, several
5.5 District Heating And Cooling 209

simplifying energy-related assumptions are used to make the section concise


and direct while still permitting the differences between the energy and exergy
results to be highlighted.

5.5.1.4 Preliminary Analysis


In exergy analysis, the temperatures at different points in the system (see
Figs. 5.19 and 5.20) are important. It is assumed that the average supply and
return temperatures, respectively, are 80 and 60 °C for district heating and
7 and 15 °C for district cooling (Edmonton Power, 1991). It is also assumed
that the supply and return temperatures, respectively, are 60 and 40 °C for
the user-heating substation and 15 and 22 °C for the user-cooling substation.
Furthermore, it is assumed that the user room temperature is constant
throughout the year at 22 °C.
An equivalent heat-transfer temperature Tequiv between the supply (subscript 1)
and return (subscript 2) temperatures can be written as
h1  h2
Tequiv ¼ (5.24)
s1  s2

where h and s denote specific enthalpy and specific entropy, respectively. For
district heating, the equivalent temperature is 70 °C for the supply system
and 50 °C for the user substation, while for district cooling, the equivalent tem-
perature is 11 °C for the supply system and 19 °C for the user substation (Rosen
and Le, 1998).
In Figs. 5.19 and 5.20, the system boundaries are for simplicity assumed to be
located sufficiently far from the sources of losses that the temperature associ-
ated with such losses is equal to the temperature of the environment. The ther-
mal exergy losses are then reduced to zero, but accounted for in the system
irreversibilities.
Table 5.9 shows that 89.46% and 10.54% of the total annual heat loads occur
in periods 1 and 2, respectively. Since there is assumed to be no space heating
demand in period 2, the 10.54% quantity is taken to be the heat needed for
water heating (which is assumed constant throughout the year). Table 5.9 also
presents the space cooling breakdown in period 2. Annual energy transfer rates
for the cogeneration-based district energy system are shown in Table 5.10, with
details distinguished where appropriate for the three chiller options consid-
ered. The data in Table 5.10 are used to calculate exergy efficiencies for the sys-
tems for each period and for the year.

5.5.1.5 Results
For the cogeneration-based district energy system using electric chillers,
single-effect absorption chillers, and double-effect absorption chillers,
210 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.10 Energy Transfer Rates (in GWh/year) for the Cogeneration-Based
District Energy System in Edmonton, Alberta
Type of Energy Period 1, To 5 0° C Period 2, To 5 30° C

District heating, Q_ H 0.8946  1040 ¼ 930 0.1054  1040 ¼ 110


Water heating, Q_
W
H (22 GWh/year/month)  0.1054  1040 ¼ 110 (or
7 month ¼ 154 22 GWh/year/month)
Space heating, Q_ H
S
930  154 ¼ 776 0
Space cooling, Q_ C 0 1.00  202 ¼ 202

Electric chiller case


Total electricity, Ẇ 0.8946  433 ¼ 388 0.1054  433 ¼ 45.6
Input energy, Ėf 0.8946  1733 ¼ 1551 0.1054  1733 ¼ 183

Single-effect absorption chiller case


Heat to drive absorption 0 1.00  252 ¼ 252
chiller, Q_ gen
Total electricity, Ẇ 0.8946  433 ¼ 388 25/60  (110 + 252) ¼ 151
Input energy, Ėf 0.8946  1733 ¼ 1551 (110 + 252)/0.6 ¼ 603

Double-effect absorption chiller case


Heat to drive absorption 0 1.00  141 ¼ 141
chiller, Q_ gen
Total electricity, Ẇ 0.8946  433 ¼ 388 21/64  (110 + 141) ¼ 82
Input energy, Ėf 0.8946  1733 ¼ 1551 (110 + 141)/0.64 ¼ 391

Tables 5.10–5.13 list the energy and exergy efficiencies evaluated for the main
system components, for several subsystems composed of selected combina-
tions of the components, and for the overall system.
The efficiencies calculated through this study are presented in Table 5.11 for
each of the six main components of the cogeneration-based district energy sys-
tem (Figs. 5.19 and 5.20), for the three chiller cases considered. Also listed in
Table 5.11 are efficiencies, broken down by function category (production,
transport, and use), for the three main subsystems identified in Figs. 5.19
and 5.20, namely,

• production of electricity, heat, and cool (including cogeneration and chilling);


• transport of heat and cool (consisting of DHC, also known in
combination as district energy); and
• end-use heating and cooling (for space and hot water heating and space
cooling).

The efficiencies of the overall cogeneration-based district energy system, for the
three chiller cases considered, are presented in Table 5.12. The efficiencies for the
heating and cooling sides of the overall system are also presented in Table 5.12.
5.5 District Heating And Cooling 211

Table 5.11 Efficiencies for the six Components and Three Main Function-Based Subsystems
of the Cogeneration-Based District Energy System, Considering Three Types of Chillers
Energy Efficiency, η (%) Exergy Efficiency, ψ (%)

1-Stage 2-Stage 1-Stage 2-Stage


Centrifugal Absorption Absorption Centrifugal Absorption Absorption
Subsystem
Chiller Chiller Chiller Chiller Chiller Chiller

Production of electricity, heat, and cool


Cogeneration 85 85 85 37 37 37
Chilling 450a 67a 120a 36 23 30
Combined 94 83 88 35 35 35
cogeneration
and chilling

Transport of heat and cool


District 100 100 100 74 74 74
heating
District 100 100 100 58 58 58
cooling
Combined 100 100 100 73 73 73
DHC

End-use heating and cooling


End-use 100 100 100 54 54 54
heating
End-use 100 100 100 69 69 69
cooling
Combined 100 100 100 55 55 55
end-use
heating and
cooling
a
These are COP values when divided by 100.

The efficiencies determined here are presented in Table 5.13 for the portion of
the cogeneration-based district energy system involving the distribution of ther-
mal energy (heat or cool). This subsystem comprises the district and end-use
heating and cooling components. Efficiencies for the heating and cooling por-
tions of this thermal energy distribution subsystem are also given in Table 5.13.
The principal process occurring in all of the components of this subsystem is
heat transfer.
Utilities that provide electricity, heating, and cooling services by operating dis-
trict energy systems are normally mainly concerned with the processes involved
in producing these energy forms and transporting them to users. The end-use of
the energy commodities is left to the users. Hence, from the perspective of dis-
trict energy utilities, it is useful to know the efficiencies of the combined
212 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

Table 5.12 Efficiencies for the Overall System and the Subsystems Representing the Heating and
Cooling Sides of the Cogeneration-Based District Energy System
Energy Efficiency, η (%) Exergy Efficiency, ψ (%)

1-Stage 2-Stage 1-Stage 2-Stage


Centrifugal Absorption Absorption Centrifugal Absorption Absorption
Chiller Chiller Chiller Chiller Chiller Chiller

Heating side 85 85 85 30 31 31
(cogeneration,
district heating,
and end-use
heating)
Cooling side 532a 80a 143a 14 9 12
(chilling, district
cooling, and end-
use cooling)
Overall system 94 83 88 28 29 29
a
These are COP values when divided by 100.

Table 5.13 Efficiencies for the Thermal-Energy Distribution Portion of the System (i.e., the Combined
Transport and End-Use Subsystems)
Energy Efficiency, η (%) Exergy Efficiency, ψ (%)

1-Stage 2-Stage 1-Stage 2-Stage


Centrifugal Absorption Absorption Centrifugal Absorption Absorption
Chiller Chiller Chiller Chiller Chiller Chiller

District and end- 100 100 100 40 40 40


use heating
District and end- 100 100 100 41 41 41
use cooling
Combined district 100 100 100 40 40 40
energya and end-
use heating and
cooling
a
District energy is combined DHC.
5.5 District Heating And Cooling 213

Table 5.14 Efficiencies for Subsystems of the Cogeneration-Based District Energy System,
Selected to Reflect the Perspective of a Production Utilitya
Energy Efficiency, η (%) Exergy Efficiency, ψ (%)

1-Stage 2-Stage 1-Stage 2-Stage


Centrifugal Absorption Absorption Centrifugal Absorption Absorption
Chiller Chiller Chiller Chiller Chiller Chiller

Cogeneration and 85 85 85 34 35 34
district heating
Chilling and 532b 80b 143b 21 14 18
district cooling
Combined 94 83 88 32 32 32
cogeneration,
chilling, and
district energyc
a
District energy production utilities are usually responsible for producing thermal energy and transporting it to users, but not for the
end-use processes.
b
These are COP values when divided by 100.
c
District energy is combined DHC.

production and transport subsystem. Energy and exergy efficiencies for this
subsystem, for the three chiller cases considered, are given in Table 5.14. The
efficiencies for the heating and cooling portions of this subsystem are also
included in Table 5.14.

5.5.1.6 Discussion
Overall energy efficiencies (Table 5.12) are seen to vary, for the three system
alternatives considered, from 83% to 94%, and exergy efficiencies from 28%
to 30%. Tables 5.10–5.13 demonstrate that energy efficiencies do not provide
meaningful and comparable results relative to exergy efficiencies when the
energy products are in different forms. For example, the energy efficiency of
the overall process using electric chillers is 94%, which could lead one to
believe that the system is very efficient. The exergy efficiency of the overall pro-
cess, however, is 28%, indicating that the process is far from ideal thermody-
namically. The exergy efficiency is much lower than the energy efficiency in
part because heat is being produced at a temperature (120 °C) higher than
the temperatures actually needed (22 °C for space heating and 40 °C for hot-
water heating). The low exergy efficiency of the chillers (see Table 5.11) is
largely responsible for the low exergy efficiency for the overall process.
The exergy efficiencies of the chilling, district cooling, and end-use cooling sub-
systems, respectively, are 36%, 58%, and 69% (see Table 5.11). For the combi-
nation that includes all three subsystems mentioned above, the exergy
214 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

efficiency takes on a relatively low value of 14% (Table 5.12). This low effi-
ciency value can be explained by noting that the cool water supply temperature
(11 °C) needed for space cooling (to 22 °C) is relatively near to the environ-
mental temperatures (in summer). The exergy of the cool is small compared
with the work input to drive the electric centrifugal chiller. The excess exergy
input via work is destroyed due to irreversibilities.
The exergy-based efficiencies in Tables 5.10–5.13 are generally different than
the energy-based ones because the energy efficiencies utilize energy quantities
that are in different forms, while the exergy efficiencies provide more meaning-
ful and useful results by evaluating the performance and behavior of the sys-
tems using work equivalents for all energy forms. The exergy and energy for
electricity are the same while the exergy for the thermal energy forms encoun-
tered here is less than the corresponding energy.
The results for cogeneration-based district energy systems using absorption
chillers (single-effect and double-effect types) and using electric chillers are,
in general, similar.
Generally, the results appear to indicate that the three integrated cogeneration
and district energy systems considered have similar efficiencies. It is likely,
therefore, that the choice of one option over another will be strongly dependent
on economics and other factors (e.g., environmental impact, space availability,
and noise limitations).
Finally, integrated district energy systems may involve thermal energy storage.
For example, a ground-coupled heat pump system can extract low-grade heat,
which may be deposited in the ground during summer using the waste heat
from a central chiller and/or by natural means, for space heating in winter. This
low-grade heat can also be extracted using a heat pump for domestic hot water
during both winter and summer. Utilizing thermal energy storage may increase
energy and exergy efficiencies of building energy systems.

5.6 CLOSING REMARKS


The importance of cogeneration, trigeneration, integrated, and district energy
systems is highlighted in this chapter, as is the usefulness of exergy analysis
in assessing and improving such systems. The energy and exergy analyses of case
studies for these systems suggest that all can yield benefits and that multigenera-
tion (i.e., cogeneration and trigeneration) systems can enhance efficiency nota-
bly in some situations.
The efficiencies and losses presented of the many complex components and
subsystems that comprise cogeneration and district energy systems highlight
the important insights provided by exergy analysis. This is particularly true
5.6 Closing Remarks 215

when these systems are integrated since different energy forms are simulta-
neously produced in multigeneration-based district energy systems. Although
energy and exergy values in general differ, exergy analysis provides more mean-
ingful efficiencies than energy analysis and pinpoints the locations and causes
of inefficiencies more accurately.
The results further indicate that the complex array of energy forms involved in
cogeneration-based district energy systems, as well as other systems examined
in this chapter, make them difficult to assess and compare thermodynamically
without exergy analysis. This difficulty is primarily attributable to the different
nature and quality of the three product energy forms: electricity, heat, and cool.
This understanding is important for designers of such systems in development
and optimization activities and in selecting the proper type of system for differ-
ent applications and situations.

Nomenclature
COP coefficient of performance
E energy
ex specific exergy
Ex exergy
h specific enthalpy
H enthalpy
m mass
P pressure
Q heat
R universal gas constant
s specific entropy
S entropy
T temperature
w specific work
W work

Greek symbols
η energy efficiency
ρ density
τ exergetic temperature factor
ψ exergy efficiency

Subscripts
0 environmental state
a air
c cooling
216 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

cd cooling with dehumidification


comp compressor
con condenser
ct cooling tower
dest destruction
en energy
equiv. equivalent
eva evaporator
ex exergy
f fuel
g flue gases
gen generation
geo geothermal
h heating
hh heating with humidification
H heating
in inlet
p pump
Q heat
sc space cooling
sh space heating
sys system
w water
W work

Superscripts
˙ rate with respect to time
s space
w water

Acronyms
AHU air handling unit
CCHP combined cooling, heat, and power
CCS carbon dioxide (CO2) capture and storage
CHP combined heat and power
DHC district heating and cooling
GHG greenhouse gas
HP high pressure
HTG high-temperature generator
HVAC heating, ventilation, and air conditioning
HVACR heating, ventilation, air conditioning, and refrigeration
IPCC Intergovernmental Panel on Climate Change
LH Linde-Hampson
LHV lower heating value
LP low pressure
References 217

LTG low-temperature generator


MTG medium-temperature generator
ORC organic Rankine cycle
QEAS quadruple effect absorption system
SOFC solid oxide fuel cell
VHTG very-high-temperature generator
VHHX very-high-temperature heat exchanger

References
Ahmadi, P., Dincer, I., Rosen, M.A., 2012. Exergo-environmental analysis of an integrated organic
Rankine cycle for trigeneration. Energy Convers. Manag. 64, 447–453.
Alanne, K., Saari, A., 2006. Distributed energy generation and sustainable development. Renew.
Sust. Energ. Rev. 10, 539–558.
Al-Sulaiman, F.A., Hamdullahpur, F., Dincer, I., 2011. Trigeneration: a comprehensive review based
on prime movers. Int. J. Energy Res. 35, 233–258.
Al-Sulaiman, F.A., Hamdullahpur, F., Dincer, I., 2012. Performance assessment of a novel system
using parabolic trough solar collectors for combined cooling, heating, and power production.
Renew. Energy 48, 161–172.
Blarke, M.B., 2012. Towards an intermittency-friendly energy system: comparing electric boilers and
heat pumps in distributed cogeneration. Appl. Energy 91, 349–365.
Canadian District Energy Association, 2008. The New District Energy: Building Blocks for Sustain-
able Community Development. http://www.ontario-sea.org/Storage/32/2406_The_New_
District_Energy_-_Building_Blocks_for_Sustainable_Community_Development.pdf, On-line
handbook.
CH2MHILL, 2007. City of Brockville water pollution control centre cogeneration feasibility assess-
ment. http://city.brockville.on.ca/index.cfm?ID¼314&Download¼%23%24N%27*%0A&
f¼383.
Christidis, A., Koch, C., Pottel, L., Tsatsaronis, G., 2012. The contribution of heat storage to the prof-
itable operation of combined heat and power plants in liberalized electricity markets. Energy
41, 75–82.
Colen, H.R., 1990. HVAC Systems Evaluation. R.S. Means Company, Kingston, MA.
Cook, B., Gazzano, J., Gunay, Z., Hiller, L., Mahajan, S., Taskan, A., Vilogorac, S., 2012. The smart
meter and a smarter consumer: quantifying the benefits of smart meter implementation in the
United States. Chem. Cent. J. 6, 1–16.
Cutter, E., Woo, C.W., Kahrl, F., Taylor, A., 2012. Maximizing the value of responsive load. Electr. J.
25, 6–16.
Depuru, S.S.S.R., Wang, L., Devabhaktuni, V., 2011. Smart meters for power grid: challenges, issues,
advantages and status. Renew. Sust. Energ. Rev. 15, 2736–2742.
Dincer, I., 2012. Thermal energy storage systems as a key technology in energy conservation. Int. J.
Energy Res. 26, 567–588.
Dincer, I., Acar, C., 2015. A review on clean energy solutions for better sustainability. Int. J. Energy
Res. 39 (5), 585–606.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, 2nd ed.,
Elsevier Science, Oxford.
Dincer, I., Zamfirescu, C., 2012. Potential options to greenize energy systems. Energy 46, 5–15.
218 C HA PT E R 5 : Cogeneration, Multigeneration, and Integrated Energy Systems

DOE, 2000. Combined heat and power: a federal Manager’s resource guide. Final report, U.S.
Department of Energy, Federal Energy Management Program, Washington, DC.
Edmonton Power, 1991. City of Edmonton district energy development phase: section 2. Engineer-
ing report.
Ghosh, S., Dincer, I., 2015. Development and performance assessment of a new integrated system
for HVAC&R applications. Energy 80, 159–167.
Hogerwaard, J., Dincer, I., Zamfirescu, C., 2013. Analysis and assessment of a new organic Rankine
based heat engine system with/without cogeneration. Energy 62, 300–310.
International Energy Agency, 2008. Combined heat and power. Technical report, http://www.iea.
org/publications/freepublications/publication/chp_report.pdf.
International Energy Agency, 2014. Key world energy statistics. Technical report, http://www.iea.
org/publications/freepublications/publication/KeyWorld2014.pdf.
Joung, M., Kim, J., 2013. Assessing demand response and smart metering impacts on long-term elec-
tricity market prices and system reliability. Appl. Energy 101, 441–448.
Liu, M., Shi, Y., Fang, F., 2014. Combined cooling, heating and power systems: a survey. Renew.
Sust. Energ. Rev. 35, 1–22.
Ludig, S., Haller, M., Bauer, N., 2011. Tackling long-term climate change together: the case of flex-
ible CCS and fluctuating renewable energy. Energy Procedia 4, 2580–2587.
MacRae, K.M., 1992. Realizing the Benefits of Community Integrated Energy Systems. Canadian
Energy Research Institute, Calgary, Alberta.
Meibom, P., Kiviluoma, J., Barth, R., Brand, H., Weber, C., Larsen, H.V., 2007. Value of electric heat
boilers and heat pumps for wind power integration. Wind Energy 10, 321–337.
Nuytten, T., Claessens, B., Paredis, K., van Bael, J., Six, D., 2013. Flexibility of a combined heat and
power system with thermal energy storage for district heating. Appl. Energy 104, 583–591.
Ozcan, H., Dincer, I., 2013. Thermodynamic analysis of an integrated SOFC, solar ORC and absorp-
tion chiller for tri-generation applications. Fuel Cells 13 (5), 781–793.
Pehnt, M., 2008. Environmental impacts of distributed energy systems—the case of micro cogen-
eration. Environ. Sci. Pol. 11, 25–37.
Procter, R., 2013. Integrating time-differentiated rates, demand response, and smart grid to manage
power system costs. Electr. J. 26, 50–60.
Ratlamwala, T.A.H., Dincer, I., 2012. Efficiency assessment of key psychometric processes. Int. J.
Refrig. 36, 1142–1153.
Ratlamwala, T.A.H., Dincer, I., Gadalla, M., 2012. Thermodynamic analysis of a novel integrated
geothermal based power generation-quadruple effect absorption cooling-hydrogen liquefac-
tion system. Int. J. Hydrog. Energy 37, 5840–5849.
Rosen, M.A., Le, M.N., 1995. Efficiency measures for processes integrating combined heat and
power and district cooling. In: Thermodynamics and the Design, Analysis and Improvement
of Energy Systems, AES-vol. 35. American Society of Mechanical Engineers, New York,
pp. 423–434.
Rosen, M.A., Le, M.N., 1998. Thermodynamic assessment of the components comprising an inte-
grated system for cogeneration and district heating and cooling. In: Proceedings of the ASME
Advanced Energy Systems Division, vol. 38, pp. 3–11.
Rosen, M.A., Le, M.N., Dincer, I., 2005. Efficiency analysis of a cogeneration and district energy sys-
tem. Appl. Therm. Eng. 25, 147–159.
Srikhirin, P., Aphornratana, S., Chungpaibulpatana, S., 2001. A review of absorption refrigeration
technologies. Renew. Sust. Energ. Rev. 5, 343–372.
References 219

Suleman, F., Dincer, I., Agelin-Chaab, M., 2014. Development of an integrated renewable energy
system for multigeneration. Energy 78, 196–204.
U.S. Department of Energy Environmental Protection Agency, 2012. Combined heat and power—a
clean energy solution. http://www.epa.gov/chp/documents/clean_energy_solution.pdf.
Wu, D.W., Wang, R.Z., 2006. Combined cooling, heating and power: a review. Prog. Energy Com-
bust. Sci. 32, 459–495.
CHAPTER 6

Heat Storage Systems

6.1 INTRODUCTION
Heat storage, also known as thermal energy storage (TES), generally involves
the temporary storage of high- or low-temperature thermal energy for later
use. Dincer and Rosen (2011) described TES as “an advanced energy technol-
ogy that is attracting increasing interest for thermal applications such as space
and water heating, cooling, and air conditioning.”
Examples of heat storage applications include storage of solar energy for over-
night heating, of summer heat for winter use, of winter ice for space cooling in
summer, and of heat or cool generated electrically during off-peak hours for use
during subsequent peak demand hours. In this regard, a heat storage system is
in many instances an useful device for offsetting temporal mismatches between
thermal energy availability and demand.
All heat storage systems have three functions:

• Charge: a heat source is used to provide heat to the storage medium.


• Storage: a medium is used to store the heat for later use. The storage
medium may be located at the heat source, the discharge, or
somewhere else.
• Discharge: heat is extracted from the storage medium in a controlled
fashion for use.

Additionally, all heat storage systems consist of three basic parts:

• Storage material and, if applicable, a container for the storage material


• A heat exchanger to facilitate heat transfer to and from the storage material
• A control system that facilitates the charging and discharging of the
thermal storage

Despite the commonalities in functions and basic parts, there are notable var-
iations in the way heat storage systems are configured. Some are designed spe-
cifically for a particular residence or application and are correspondingly 221

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00006-5


© 2015 Elsevier Inc. All rights reserved.
222 C HA PT E R 6 : Heat Storage Systems

tailored. Others follow a more general design approach, with small modifica-
tions for the particular installation or application.
Heat storage systems for heating or cooling capacity are often utilized in appli-
cations where the occurrence of a demand for energy and that of the most favor-
able supply of energy, based on economic or other factors, are not coincident.
Thermal storages are used in energy conservation, industry, commercial build-
ing, and solar energy systems. The storage medium can be located in storages of
various types, including tanks, ponds, caverns, and underground aquifers.
For electrical heating systems, heat storage can be used to enable the purchase
of off-peak power, which costs less than power during peak usage times for util-
ities that provide off-peak power pricing. This type of system is used on a daily
or twice-daily charging schedule. During off-peak periods (typically early after-
noon and between late evening and early morning), an electric heater or
ground source heat pump (GSHP) is used to heat a tank of water or another
storage system. The heat storage then provides heat to a distribution system
during the remainder of the day. Heat storage can also be used as a thermal
dump for excess electricity, such as when a renewable electrical source, like solar
photovoltaic panels, produces more electricity than the grid needs. Excess elec-
tricity can be converted to heat, stored, and used for space heating or other ther-
mal needs. In combined heat and power systems, heat storage allows for more
continuous operation. Many cogeneration systems operate to meet thermal
demand, resulting in excess electricity at times when it is not needed. It also
causes systems to cycle on and off to meet partial load heat demands. Heat stor-
age allows systems to operate for longer periods of time, by having it charge the
storage instead of cycling on and off (Haeseldonckx et al., 2007).
Many other heat storage applications exist, including providing energy with high
reliability for buildings such as hospitals that experience large consequences if
there are interruptions in service. Heat storage can also be used in space cooling
applications, for instance, by using the storage of winter ice to provide space cool-
ing during the summer (Dincer and Rosen, 2013; Dincer and Dost, 1996). Heat
storage can usually be used for many applications in most types of buildings.
From a thermodynamic point of view, there are three types of heat storage: chem-
ical, latent, and sensible. Latent heat storages use phase change materials (PCMs).
In sensible heat storage, the storage medium remains in a single phase, while in
latent heat storage, the storage medium undergoes a phase change. Sensible heat
storage systems (e.g., liquid water systems) exhibit changes in temperature in the
store as heat is added or removed. In latent heat storage systems (e.g., liquid
water/ice systems and eutectic salt systems), the storage temperature remains
fixed during the phase change portion of the storage cycle. In chemical energy
storage, heat is stored in chemical reactions that can readily be reversed. At pre-
sent, this type of system is not commonly used for residential applications and is
6.1 Introduction 223

Table 6.1 Typical Performance Parameters for Sensible, Latent, and Chemical Heat Storage
Technologies
Storage
Heat Storage Type Capacity (kWh/t) Power (MW) Efficiency (%) Period

Sensible (e.g., water) 10-50 0.001-10 50-90 Daily-monthly


Latent (e.g., phase change materials) 50-150 0.001-1 75-90 Hourly-monthly
Chemical 120-250 0.01-1 75-100 Hourly-daily
Source: IEA (2011).

mainly undergoing development, so it is not discussed further in this chapter.


Typical performance data, including capacity, power, efficiency, and storage
periods, are shown in Table 6.1 for the main heat storage types.
Heat storage systems are used in a wide variety of applications and are designed
to operate on a cyclical basis (e.g., daily, weekly, and seasonally). Heat storage
systems achieve benefits by fulfilling one or more of the following purposes:

• Generation capacity increase: Demands for heating, cooling, and electricity


are seldom constant over time, and the excess generation capacity
available during low-demand periods can be used to charge a heat storage
in order to increase the effective generation capacity during high-demand
periods. This process allows a smaller production unit to be installed
(or to add capacity without purchasing additional units) and results in
a higher load factor for the units:
• Enhanced operation of cogeneration plants: Combined heat and power, or
cogeneration, plants are generally operated to meet the demands of the
connected thermal load, which often results in excess electric generation
during periods of low electricity use. By incorporating a heat storage
system, the plant need not follow a load and instead can be dispatched in
more advantageous ways (within some constraints).
• Energy purchase shift to low-cost periods: This use is the demand-side
application of the first purpose listed and allows energy consumers subject
to time-of-day pricing to shift energy purchases from high- to low-cost
periods.
• Increased system reliability: Any form of energy storage, from the
uninterruptible power supply of a small personal computer to a large
pumped storage, normally increases system reliability.
• Functional integration: In applications where on-site water storage is needed
for fire protection, it may be feasible to incorporate thermal storage into a
common storage tank. Likewise, equipment designed to solve power-
quality problems may be adaptable to energy storage purposes.
224 C HA PT E R 6 : Heat Storage Systems

The most significant benefit of a heat storage system is often cited as its ability to
reduce electric costs by using off-peak electricity to produce and store energy for
daytime cooling. Indeed, heat storage systems successfully operate in offices,
hospitals, schools, universities, airports, etc., in many countries, shifting energy
consumption from periods of peak electricity rates to periods of lower rates.
That benefit is accompanied by the additional benefit of lower demand charges.
Having investigated methods for evaluating and comparing heat storage sys-
tems for many years, the present authors observe that, while many technically
and economically successful thermal storages are in operation, no generally
valid basis for comparing the achieved performance of one storage with that
of another operating under different conditions has found broad acceptance.
The energy efficiency, the ratio of the energy recovered from storage to that
originally inputted, is conventionally used to measure heat storage system
performance. The energy efficiency, however, is inadequate because it does
not take into account important factors like how nearly the performance
approaches ideality, storage duration, and temperatures of the supplied and
recovered thermal energy and of the surroundings.
Exergy analysis provides an illuminating, rational, and meaningful alternative
for assessing and comparing heat storage systems. In particular, exergy analysis
yields efficiencies that provide a true measure of how nearly actual performance
approaches the ideal and identifies more clearly than energy analysis the mag-
nitudes, causes, and locations of thermodynamic losses. Consequently, exergy
analysis can assist in improving and optimizing heat storage system designs.
Using information in the authors’ recent book on heat storage systems (Dincer
and Rosen, 2011), this chapter describes the application of exergy analysis to
heat storage systems and demonstrates the usefulness of such analyses in pro-
viding insights into heat storage system behavior and performance, for heating,
ventilation, and air conditioning (HVAC) and other applications. Key thermo-
dynamic considerations in heat storage system evaluation are discussed, and
the use of exergy in evaluating a heat storage system is detailed.

6.2 PERFORMANCE CONSIDERATIONS IN HEAT


STORAGE SYSTEMS
This section provides aspects of thermodynamics most relevant to energy and
exergy analyses. We define thermodynamics as the science of energy (which comes
from first law of thermodynamics (FLT)) and exergy (which comes from the sec-
ond law of thermodynamics (SLT)). Fundamental principles and such related
issues as reference-environment selection, efficiency definition, and material-
properties acquisition are discussed. General implications of exergy analysis results
are discussed, and a step-by-step procedure for energy and exergy analyses is given.
6.2 Performance Considerations In Heat Storage Systems 225

6.2.1 Principal Thermodynamic Factors in Heat


Storage Systems
Several of the principal thermodynamic considerations in the evaluation and
comparison of heat storage systems are described in the next several paragraphs.
Energy and exergy: Energy and exergy are significant quantities in evaluating heat
storage systems. Exergy analysis complements energy analysis and circumvents
many of the difficulties associated with conventional energy-based heat storage
system methods by providing a more rational evaluation and comparison
basis.
Temperature: Exergy reflects the temperature of a heat transfer and the degrada-
tion of heat quality through temperature loss. Exergy analysis applies equally
well to systems for storing thermal energy at temperatures above and below
the temperature of the environment because the exergy associated with such
energy is always greater than or equal to zero. Energy analysis is more difficult
to apply to such storage systems because efficiency definitions have to be care-
fully modified when cooling capacity, instead of heating capacity, is stored or
when both warm and cool reservoirs are included. Thus, exergy analysis pro-
vides for more rational evaluation of heat storage systems for cooling or heating
capacity.
Efficiencies: The evaluation of a heat storage system requires a measure of per-
formance that is rational, meaningful, and practical. A more perceptive basis
than energy efficiency is needed if the true usefulness of thermal storages is
to be assessed and so permits maximization of their economic benefit. Exergy
efficiencies provide rational measures since they assess the approach to ideal
heat storage system performance.
Losses: With energy analysis, all losses are attributable to energy releases across
system boundaries. With exergy analysis, losses are divided into two types:
exergy releases from the system and internal exergy consumptions. The latter
include reductions in availability of the stored heat through mixing of warm
and cool fluids. The division of exergy losses allows the causes of inefficiencies
to be accurately identified and improvement effort to be effectively allocated.
Stratification: Thermal stratification within a heat storage system reduces tem-
perature degradation. In many practical cases, a vertical cylindrical tank with
a hot water inlet (outlet) at the top and a cold water inlet (outlet) at the bottom
is used. The hot water and cold water in the tank usually are stratified initially
into two layers, with a mixing layer in between. The degree of stratification is
affected by the volume and configuration of the tank, the design of the inlets
and outlets, the flow rates of the entering and exiting streams, and the durations
of the charging, storing, and discharging periods. Increasing stratification
improves heat storage system efficiency relative to a thermally mixed storage
tank. Four primary factors degrade stored energy by reducing stratification:
226 C HA PT E R 6 : Heat Storage Systems

• Heat leakages to or from the environment


• Heat conduction and convection from the hot portions of the storage
fluid to the colder portions
• Vertical conduction in the tank wall
• Mixing during charging and discharging periods (often the main cause of
loss of stratification)

The effects of stratification are more clearly assessed with exergy than energy.
Through carefully managing the injection, recovery, and holding of heat (or
cold) so that temperature degradation is minimized, better storage-cycle perfor-
mance can be achieved (as measured by better thermal energy recovery and
temperature retention, that is, increased exergy efficiency).
Storage duration: Rational evaluation and comparison of heat storage systems
must account for storage duration. The length of time thermal energy is
retained in a heat storage system does not enter into expressions for efficiency,
although it is clearly a dominant consideration in overall heat storage system
effectiveness. The relation between storage duration and effectiveness permits
an approach for comparing heat storage systems using a time parameter.
Reference-environment temperature: Since heat storage evaluations based on
energy and exergy are affected by the value of the reference-environment tem-
perature T0, temporal and spatial variations of T0 must be considered (espe-
cially for heat storage systems with storage periods of several months).
There is a growing interest in the use of diurnal or daily heat storage systems for elec-
trical load management in both new and existing buildings. Heat storage technol-
ogies allow electricity consumption costs to be reduced by shifting electrical heating
and cooling demands to periods when electricity prices are lower, for instance, dur-
ing the night. Load shifting can also reduce demand charges, which can represent a
significant proportion of total electricity costs for commercial buildings.
Some important technical requirements for heat storage systems are as follows:

• High energy density (per unit mass and/or per unit volume) in the storage
material
• Good heat transfer between heat transfer fluid (HTF) and the storage
medium
• Mechanical and chemical stability of the storage material
• Chemical compatibility between HTF, heat exchanger, and storage medium
• Complete reversibility for a large number of charging/discharging cycles
• Low thermal losses
• Ease of control

Also, when designing a heat storage system, cost of the (i) storage material,
(ii) heat exchanger, and (iii) space and the enclosure for the heat storage system
should be considered.
6.2 Performance Considerations In Heat Storage Systems 227

6.2.2 Energy and Exergy Analyses of Heat


Storage Systems
Environmental problems associated with energy utilization span a growing
spectrum of environmental quality and natural ecosystem issues. The SLT is
useful in providing insights into environmental impact. In conjunction with
this, exergy appears to be an effective measure of the potential of a substance
to impact the environment.
Exergy is defined as the maximum amount of work that can be produced by a
stream of matter, heat, or work as it comes to equilibrium with a reference envi-
ronment. It is a measure of the potential of a stream to cause change, as a con-
sequence of not being completely stable relative to the reference environment.
Exergy is not subject to a conservation law; rather, exergy is consumed or
destroyed, due to irreversibilities in any process. For exergy analysis, the state
of the reference environment, or the reference state, must be specified. This
is commonly done by specifying the temperature, pressure, and chemical com-
position of the reference environment.
Exergy analysis is a method that uses the conservation of mass and conservation
of energy principles together with the SLT for the design and analysis of energy
systems. The exergy method can be suitable for furthering the goal of more effi-
cient energy-resource use, for it enables the locations, types, and true magni-
tudes of wastes and losses to be determined. Therefore, exergy analysis can
reveal whether or not and by how much it is possible to design more efficient
energy systems by reducing the inefficiencies in existing systems.
Recently, exergy analysis has been recognized as the more powerful method for
performance evaluations and design of heat storage systems than energy anal-
ysis (Rosen et al., 1988; Taylor et al., 1991; Krane and Krane, 1992; Rosen,
1992; Dincer et al., 1997). For an overall heat storage process or any subpro-
cesses (i.e., charging, storing, and discharging), overall energy and exergy bal-
ances; individual energy and exergy balances for charging, storing, and
discharging periods; overall energy and exergy efficiencies; and individual
energy and exergy efficiencies for charging, storing, and discharging periods
are summarized in Table 6.2.
The term efficiency can be used in at least two ways: first law efficiency (i.e.,
energy efficiency) and second law efficiency (i.e., exergy efficiency). The energy
efficiency merely reflects the standard task of energy storage, as the ratio of use-
ful energy output to total energy input. However, the exergy efficiency incorpo-
rates the notion of increasing thermodynamic unavailability, as reflected by
increasing entropy, in a process or subprocess. That is why storage exergy effi-
ciencies are always lower that the energy efficiencies; process irreversibilities
destroy some of the input energy (Rosen, 1992; Rosen and Hooper, 1996;
Dincer et al., 1997).
228
C HA PT E R 6 :
Table 6.2 Energy and Exergy Balance and Efficiency Equations for a Heat Storage System

Heat Storage Systems


Description Energy Equationa Exergy Equationa
 
Balances for overall storage process ΔE ¼ Qc  ðQl + Qd Þ ΔEx ¼ ExQ, c  ExQ, l + ExQ, d  Exdest
or any subprocesses
Individual balances Ql ¼ Ql, c + Ql, s + Ql, d Exdest ¼ Exdest, c + Exdest, s + Exdest, d
ΔE ¼ ΔEc + ΔEs + ΔEd Exl ¼ Exl, c + Exl, s + Exl, d
ΔEx ¼ ΔExc + ΔExs + ΔExd
Charging period ΔEc ¼ Qc  Ql, c ΔExc ¼ ExQ, c  Exl, c  Exdest, c
Storing period ΔEs ¼ Ql, s ΔExc ¼ Exl, s  Exdest, s
Discharging period ΔEd ¼ Qd  Ql, d ΔExd ¼ ExQ, d  Exl, d  Exdest, d
Balances for initial (i) and final (f) states ΔE ¼ Ef  Ei   ΔEx ¼ Exf  Exi  
ΔEc ¼ Ef, c  Ei, c Ei, c ¼ Ei ΔExc ¼ Exf, c  Exi, c Exi, c ¼ Exi
   
ΔEs ¼ Ef, s  Ei, s Ei, s ¼ Ef, c ΔExs ¼ Exf, s  Exi, s Exi, s ¼ Exf, c
       
ΔEd ¼ Ef, d  Ei, d Ei, d ¼ Ef, s and Ef, d ¼ Ef ΔExd ¼ Exf, d  Exi, d Exi, d ¼ Exf, s and Exf, d ¼ Exd
Energy content for a particular state E R ¼ E  E0
with respect to a reference ΔE ¼ EfR  EiR (applies to charging, storage, and discharging periods)
environment
Overall efficiencies Qd Ql + ΔE ExQ, d ExQ, l + ΔEx + Exdest
η¼ ¼1 ψ¼ ¼1
Qc Qc ExQ, c ExQ, c
Efficiencies for charging period ΔEc Q l, c ΔExc Exl, c + Exdest, c
ηc ¼ ¼1 ψc ¼ ¼1
Qc Qc ExQ, c ExQ, c
Efficiencies for storing period ΔEc + ΔEs Q l, s ΔExc + ΔExs Exl, s + Exdest, s
ηs ¼ ¼1 ψs ¼ ¼1
ΔEc ΔEc ΔExc ΔExc
Efficiencies for discharging period Qd Ql, d + ΔE ExQ, d Exl, d + ΔEx + Exdest, d
ηd ¼ ¼1 ψd ¼ ¼1
ΔEc + ΔEs ΔEc + ΔEs ΔExc + ΔExs ΔExc + ΔExs
a
c, charging; d, discharging; dest, destruction; f, final; i, initial; l, loss (leakage); R, reference; s, storing.
Source: Dincer (2002).
6.2 Performance Considerations In Heat Storage Systems 229

As seen in Table 6.2, in a heat storage system, there are three steps (subpro-
cesses) and these must be taken into consideration for comprehensive energy
and exergy analyses. In the analysis, it is usually assumed that there are no heat
losses to the surroundings from the charging or discharging fluids (all heat
removed from the charging fluid is added to the storage medium, and for
the discharging period, all heat added to the discharging fluid originates in
the storage medium). This assumption is most valid for small rates of heat loss
from the relevant fluids. Rosen (1992) pointed out that in practical systems, the
flows of the charging and discharging fluids are often steady and have time-
dependent thermal properties.
Note that all four energy efficiencies in Table 6.2 become identical if
ERf ¼ ERi ¼ 0, and then, these equations cannot provide rational measures of
the performance of the system. In order to provide rational measures, △E must
be less than zero. Similar expression can be written for the exergy efficiencies.

6.2.3 Environmental Impacts of Heat Storage Systems


Heat storage systems can contribute significantly to meeting society’s desire for
more efficient, environmentally benign energy use, particularly in the areas of
building heating and cooling and electric power generation. By reducing energy
consumption, the utilization of heat storage results in two significant environ-
mental benefits: (i) conservation of fossil fuels through efficiency increases
and/or fuel substitution and (ii) reductions in emissions of such pollutants
as CO2, SO2, NOx, and chlorofluorocarbons (CFCs).
Heat storage systems can impact air quality at building sites by reducing emis-
sions of (i) the amount of ozone-depleting CFC and hydrochlorofluorocarbons
(HCFC) refrigerants in chillers and (ii) combustion gases from fuel-fired heat-
ing and cooling equipment. Heat storage systems help reduce CFC use in two
main ways. First, since cooling systems with heat storage require less chiller
capacity than conventional systems, they use fewer or smaller chillers with less
refrigerant. Second, using heat storage systems can offset the lost cooling capac-
ity that sometimes occurs when existing chillers are converted to more benign
refrigerants, making it easier for building operators to switch refrigerants.
The potential aggregate air-emission reductions at power plants due to heat
storage systems can be significant. For example, heat storage systems have been
shown to reduce CO2 emissions in the United Kingdom by 14-46% by shifting
electric load to off-peak periods (Beggs, 1994), while an EPRI-cosponsored
analysis found that TES could reduce CO2 emissions by 7% compared to
conventional electric cooling technologies (Reindl, 1994). Also, using the
California Energy Commission data indicating that existing gas plants produce
about 0.06 kg of NOx and 15 kg of CO2 per 293,100 kWh of fuel burned, and
230 C HA PT E R 6 : Heat Storage Systems

assuming that heat storage system installations save an average of 6% of the


total cooling electricity needs, heat storage systems could possibly eliminate
annual emissions of about 560 tons of NOx and 260,000 tons of CO2 statewide
(California Energy Commission, 1996).

6.3 CLASSIFICATION OF HEAT STORAGE SYSTEMS


The suitability of a particular technology for an individual application can be
broadly evaluated in terms of technical potential. Heat storage systems can be
classified based on storage duration, desired output temperature and capacity
(or size/load), and storage medium. These criteria can be used as a starting
point in determining suitability for particular applications (Hauer et al., 2013).
Technical, sizing, economic, energy saving, and environmental criteria are used
to evaluate heat storage systems and applications. Each of these should be con-
sidered carefully for successful heat storage system implementation. Indepen-
dent technical criteria for storage systems are difficult to establish since they
usually are case-specific and are closely related to and generally affected by
the economics of the resultant systems. Although appropriate trade-offs must
be made among criteria, certain technical factors are generally desirable,
including

• storage capacity,
• storage duration,
• technical availability,
• integrability with other thermal systems,
• reliability,
• applicability,
• lifetime,
• size,
• cost,
• efficiency,
• safety,
• installation, and
• environmental standards.

A heat storage system designer requires technical information on heat storage


systems, such as the types of storage available, the amount of storage required,
and the effect of storage on system performance, reliability, and cost. Heat stor-
age systems can be difficult to employ at sites with space restrictions. Also, heat
storages often have significant first capital costs. Financial analyses of heat stor-
age system-based projects can be complex, although most consulting energy
engineers are capable of evaluating relevant financial parameters and benefits.
6.3 Classification Of Heat Storage Systems 231

6.3.1 Storage Duration


Heat storages can be designed to store heat for various time periods, from hours
to months. In each case, storage systems are charged with heat or cold, which is
then stored for a length of time before discharging occurs. Different heat storage
systems classified by their storage duration along with examples are the
following:

• Less than 1 day (hourly loads): Electrical thermal storage (ETS) devices and
off-peak utility power systems, which might charge thermal mass each
afternoon and night
• One day (diurnal storage and daily loads): Solar or biomass thermal space
heating systems in mild climates
• Several days (daily loads for a few days due to intermittent supply): Solar
thermal space heating systems and domestic hot water in mild climates
that typically experience cloudy days
• Several months (seasonal storage): Space heating systems that use seasonally
available solar resource to charge storage, for example, charge during
the summer for use in the winter or vice versa

The length of time and the amount of usable heat stored in the medium depend
on the user’s needs, the heat source, and the system design. However, there are
advantages and disadvantages to short- and long-time frame systems that
are important to consider when deciding on a heating system. Systems
that are recharged daily have a smaller size, and thus a lower capital cost, than
seasonal systems. Smaller systems are typically manufactured off-site. Seasonal
storage, while more expensive and voluminous, reduces the reliance on any sin-
gle day of thermal charging, which increases system reliability. For instance, a
few cloudy days do not affect significantly solar thermal seasonal storage over a
year. Also, seasonal storage can be used for district heating systems, heating sys-
tems that provide heat to multiple buildings, where capital and maintenance
costs can be shared.

6.3.2 Temperature Range


Heat storage systems can store heat at a range of temperatures. In fact, they can
also be used in space cooling applications, in which some systems store ice to
be used for cooling during the summer. The storage temperature, or tempera-
ture range in the case of sensible storage, is determined by the type of space
heating (or cooling) application, the temperature of the heat source, and the
temperature of the demand-side distribution system. Practical limitations are
also imposed by the storage container, particularly since a greater temperature
difference between the storage material and the ambient temperature raises the
rate of heat loss through the container walls, increasing the need for insulation.
232 C HA PT E R 6 : Heat Storage Systems

Heating systems are often classified as high-temperature systems or low-


temperature systems. Low-temperature systems typically store heat below the
boiling point of water (100 °C) and are used for space heating applications with
hydronic distribution systems. For instance, baseboards require heat in the
range of 65-85 °C, and radiant floors require temperatures of 30-40 °C.
Higher-temperature storage, or systems that store heat above 100 °C, can be
used for forced-air systems: in these systems, air is often blown over rocks or
ceramic bricks that store heat. High-temperature heat storage can also be
stepped down for other uses. Temperature ranges and applications of cold-,
low-, medium-, and high-temperature heat storage systems are listed in
Table 6.3. In some instance, these heat storage systems can be made more effec-
tive by designing them to maintain high-temperature heat in the inner regions
of the storage and lower temperatures at the outer regions.

6.3.3 Storage Capacity


The capacity (or size or scale) of a heat storage depends on several factors, such
as (i) heating/cooling requirement (heat load), (ii) heat source capacity,
(iii) storage material heat capacity, (iv) storage duration, and (v) standby loss.
Optimal sizing of the thermal storage system is important to achieve appropri-
ate performance. Without a supplemental heat source, undersized thermal stor-
age systems are insufficient to meet heating demands. Oversized systems have a
higher capital cost, can cost more to maintain, and can waste energy through
standby losses. Sizing is of increased significance for seasonal storage systems as
even optimally sized systems have large space requirements and installation

Table 6.3 Temperature Ranges and Applications of Cold-, Low-, Medium-, and
High-Temperature Heat Storage Systems
Temperature
Type Range (°C) Application Additional Information

Cold <10 Building cooling Antifreeze mixture is needed to avoid freezing in


heat exchanger fluid
No auxiliary conventional heat source is needed
Low 10-30 Low-temperature residential Artificial charging requirement (with solar
heating collectors)
Medium 30-50 Residential and commercial Artificial charging requirement (with solar
heating (e.g., schools and offices) collectors or waste heat)
High >50 District heating No heat pump requirements
Need for auxiliary burners
Need for high-performance storage media

Sources: Hahne (2000) and Chwieduk (2012).


6.3 Classification Of Heat Storage Systems 233

Table 6.4 Near-Term Suitability Criteria for Determining Prime Heat Storage Technologies
Heat Storage Type Examples Near-Term Beneficial Areas

Small scale Ice storage, hot and cold water Higher demand variability (i.e., more peak-like demands,
tanks with much hot or cold needed at one time or another)
Large scale Underground thermal energy Significant waste heat resources, concentrated heating
storage (UTES), molten salts or cooling demand, or large amounts of concentrating solar
power (CSP)

Source: IEA (2015).

costs. Examples of small- and large-scale heat storage technologies are pre-
sented in Table 6.4, along with comments on their near-term suitability.
A need exists for improved heat storage-sizing techniques, as analyses of appli-
cations reveal both undersized and oversized systems. Undersizing can result in
poor levels of indoor comfort, while oversizing results not only in higher than
necessary initial costs but also in the potential wasting of electricity if more
energy is stored than required. Another requirement for successful heat storage
is proper installation and control. State-of-the-art and properly designed and
controlled storage systems often do not use more energy than conventional
heating and cooling equipment.
Performance data describing the use of heat storage systems for heating and
cooling by shifting peak loads to off-peak periods have been reported and show
the potential for such technologies to be substantial. The initial costs of such
systems can be lower than those for other systems. To yield the benefits, new
construction techniques are required together with the use of more sophisti-
cated thermal-design calculations that are, at present, not well known to many
builders and designers.

6.3.4 Underground Thermal Energy Storage


Underground heat storage, or underground thermal energy storage (UTES), has
storing temperature range from around 0 °C to up to 40-50 °C. This operating
temperature range is suitable for heating and cooling applications in HVAC.
UTES systems have been demonstrated in many cases to be advantageous for
energy management to provide good economic returns. For example, UTES sys-
tems have been successfully integrated with heating, cooling, and air
conditioning applications in many projects (Kizilkan and Dincer, 2012).
As an increasingly used storage technology, UTES makes use of the under-
ground as a storage medium for both heat storage and cold storage. UTES tech-
nologies include borehole storage, aquifer storage, cavern storage, and pit
234 C HA PT E R 6 : Heat Storage Systems

storage. The choice of technology depends in large part on local geologic con-
ditions, as discussed below:
• Borehole storage is based on vertical heat exchangers installed in the
ground, which ensure the transfer of thermal energy to and from the
ground layers (e.g., clay, sand, and rock). Many projects aim for seasonal
storage of solar heat in summer to heat buildings in winter. Ground heat
exchangers are also frequently used in combination with heat pumps
where the ground heat exchanger extracts low-temperature heat from
the soil.
• Aquifer storage uses a natural underground water-permeable layer as a
storage medium. The transfer of thermal energy is achieved by mass
transfer (i.e., extracting/reinjecting water from/into the underground
layer). Most applications deal with the storage of winter cold to be used
for the cooling of large office buildings and industrial processes in the
summer. A major prerequisite for this technology is the availability of
suitable geologic formations.
• Cavern storage and pit storage are based on large underground water
reservoirs created in the subsoil to serve as heat storages. These storage
options are technically feasible, but applications are limited because of the
high investment costs.
The two main types of UTES considered in more detail in this chapter are,
namely, borehole TES (BTES) and aquifer TES (ATES). Heat pumps are com-
monly incorporated with these systems to utilize the underground storage as
a heating or a cooling source. Additionally, for the same purpose, these systems
can be connected to solar collectors, surface water, waste heat sources, etc. The
main distinction between BTES and ATES is that the former is often (but not
always) closed with no direct contact between the heat pump fluid and the
underground water. In an ATES, the underground water is extracted and
pumped through heat exchangers to recover the available heat.
For high-temperature (i.e., above 100 °C)-sensible heat storage, the storage
medium of choice is usually a liquid (e.g., oil or molten salts, the latter for tem-
peratures up to 550 °C). For very high temperatures, solid materials (e.g.,
ceramics and concrete) can be used, but most high-temperature-sensible heat
storage options are undergoing development or demonstration.
A common technology for heat storage today is the domestic hot water
tanks. Other technologies, such as ice and chilled-water storage, are starting
to play important roles in several countries, including Australia, the United
States, China, and Japan, as utilities seek to reduce peak loads and con-
sumers seek to lower their electricity bills. UTES systems are frequently
found in Canada, Germany, and many European countries (IEA, 2011).
Table 6.5 depicts a range of heat storage technologies and their characteris-
tics and efficiencies.
6.3 Classification Of Heat Storage Systems 235

Table 6.5 Current Status and Applications of Selected Heat Storage Technologies
Initial
Heat Investment
Storage Cost (USD/ Primary
Technology Locationa Efficiency (%) kW) Application Sample Applications

UTES Supply 50-90 3400-4500 Long-term storage Drake Landing Solar


Community (Canada)
Molten salts Supply 40-93 400-700 High-temperature Gemasolar CSP plant
(Spain)
Thermo Supply, 80-99 1000-3000 Low-, medium-, and Thermal Control Systems
chemical demand high-temperature (TCS) for CSP plants (R&D)
Solid media Demand 50-90 500-3000 Medium-temperature Residential electric
storage thermal storage (the
United States)
Ice storage Demand 75-90 6000-15,000 Low-temperature Tokyo Denki University
(Japan), China Pavilion
project (China)
Hot water Demand 50-90 N/Ab Medium-temperature Peak demand reduction
storage (France)
(residential)
Cold water Demand 50-90 300-600 Low-temperature Shanghai Pudong
storage International Airport
(China)
a
Typical locations in current heat storage systems. These locations may change as the technologies evolve.
b
Energy storage capabilities present in hot water storage tanks can be utilized for negligible additional cost.
Sources: IEA (2011, 2015).

Heat storage systems can take many forms to suit a variety of applications, such as
off-peak heating and air conditioning and industrial/process heating and cool-
ing. Selecting a storage and its characteristics usually requires a detailed feasibility
study. The analysis is involved and best accomplished following an established
procedure. Data needed for a feasibility analysis can include (i) an hour-by-hour
load profile for the design day and (ii) a description of a baseline nonstorage sys-
tem, including chiller capacity, operating conditions, and efficiency. The descrip-
tion of a heat storage system often stipulates the following:

• Sizing basis (full storage, load leveling, or demand limiting)


• Sizing calculations showing chiller capacity and storage capacity and
considering required supply temperature
• Design operating profile, showing load, chiller output, and the amount of
heat added to or taken from storage for each hour of the design day
• Chiller operating conditions while charging the storage and if applicable
when meeting the load directly
• Chiller efficiency under each operating condition
• Description of the system control strategy, for the design day and
part-load operation
236 C HA PT E R 6 : Heat Storage Systems

Table 6.6 Summary of Classification of Heat Storage Systems


Heat Storage
Classification Criteria Type Notes and Applications

Temperature High temperature Solar power-based, operating temperature is higher than 350 °C
Low temperature Operating temperature is lower than 100 °C
Below zero (CTES) Water with glycol, snow, ice storage, ice slurry storage, etc.
Duration Seasonal Long term, in areas with wide seasonal temperature variation
Weekly Mostly residential
Daily Commercial buildings with high daytime and low nighttime loads
Hourly Buildings with highly variable hourly loads
Storage material (solid, Water Most common medium especially for HVAC applications
liquid, gas, and latent) Ice High storage capacities in limited spaces
Oil Mainly in solar thermal power plants where oil can be used both
for storage medium and as HTF
Sand and rock Cost-effective storage mediums
Concrete Low cost, widely available, and can be easily shaped
PCM Materials (e.g., salt compounds) undergo phase change
Molten salt Used for high-temperature (usually above 300 °C) applications
UTES ATES Storage medium is underground water where hot and cold
thermal fluids are in direct contact (open system)
BTES Storage without direct contact of the hot and cold thermal fluids
(closed system)
Cavern TES Rock caverns utilized for TES

An operating cost analysis essentially includes the following:


• An evaluation of demand savings
• A determination of changes in energy consumption and cost
• A description and justification of the assumptions used for annual energy
demand and use estimates
As mentioned earlier, heat storage systems are classified according to various
criteria (e.g., storing period, storing medium, temperature level, and applica-
tion). Classifications of heat storage systems are summarized in Table 6.6. Sev-
eral factors are desirable for specific heat storage applications, including high
specific heat capacity, long-term stability under cyclic operation, and low cost.

6.4 HEAT STORAGE SYSTEMS FOR HEATING


APPLICATIONS
Heat storage systems for heating applications mainly use low-grade heat such as
solar energy or waste heat from power generation plants and industrial thermal
6.4 Heat Storage Systems For Heating Applications 237

Table 6.7 Thermal Capacities of Some Common Heat Storage Materials at


20 °C
Specific Heat Volumetric Heat
Material Density (kg/m3) (kJ/kg K) Capacity (kJ/m3 K)

Clay 1458 0.879 1280


Brick 1800 0.837 1510
Sandstone 2200 0.712 1570
Wood 700 2.39 1670
Concrete 2000 0.88 1760
Glass 2710 0.837 2270
Aluminum 2710 0.896 2430
Iron 7900 0.452 3570
Steel 7840 0.465 3680
Gravely earth 2050 1.84 3770
Magnetite 5177 0.752 3890
Water 988 4.182 4170
Sources: Norton (1992) and Dincer (1998).

processes for short- and long-term storage purposes. For short-term heat stor-
age, there are a number of storage media considered in practice (e.g., water, oil,
molten salts, molten metals, bricks, sand, and soil). Large aquifers, rock beds,
solar ponds, and large tanks are used for long-term (e.g., annual) storage. Some
commonly used heat storage materials and their properties are summarized in
Table 6.7.
It is desirable that a storage material be inexpensive and has a good thermal
capacity factor. Another important parameter in heat storage is the rate at which
heat can be released and extracted. This leads to the ability of a material to store
heat, which is a function of thermal diffusivity. For this reason, iron shot is
an excellent thermal storage medium, having both high heat capacity and high
thermal conductance. For high-temperature heat storage (up to several hundred
degrees centigrade), iron or iron oxide is as good as water per unit volume of
storage. The cost is moderate for either pellets of the oxide or metal balls. Since
iron and its oxide have equal performance, the slow oxidization of the metal in
a high-temperature liquid or air system does not degrade performance.
Water as a heat storage medium has an excellent specific heat and is both inex-
pensive and chemically stable. If it is employed above 100 °C, the system has to
be pressurized, which adds significantly to costs. For such a case, the limitation
of water is its critical point (374 °C). In addition, a number of heat-resistant oils
(e.g., Therminol) are available in the market, which can be readily used without
238 C HA PT E R 6 : Heat Storage Systems

pressurization at temperatures over a broad temperature range (10 to


320 °C). However, while the average specific heat of water is 4.19 kJ/kg K, most
oils have specific heats only about 2.3 kJ/kg K (Diamant, 1984). Another dis-
advantage of these oils is the liability of high-temperature cracking, polymeri-
zation, and the formation of volatile products. Other storage substances
include molten salts and molten metals. Molten salts can also be employed
for high-temperature storage with an average specific heat of 1.5 kJ/kg K
(Diamant, 1984), but they have some disadvantages, such as solidification tem-
peratures of a minimum of 150 °C and corrosion effects. Molten metals (e.g.,
liquid sodium) can be used in the unpressurized state at temperatures up to
760 °C with an average specific heat of 1.3 kJ/kg K (Diamant, 1984), implying
some disadvantages, for example, handling problems.
Rock is another good heat storage material from the standpoint of cost, but its
thermal capacity is only half that of water. Past studies have demonstrated that
the rock storage bin is practical and its main advantage is that it can easily be
used for heat storage above 100 °C. Therefore, rocks are sometimes preferred
over water for solar systems. However, air/rock solar systems can make provi-
sion for partial heat storage in water for domestic hot water use. The amount of
heat stored in rocks, compared with an equivalent amount of heat stored in
water, occupies approximately three times as much volume. This apparent dis-
advantage is partly offset by costs associated with water containment. Adding
the higher cost and maintenance of a liquid collector and the economics often
favor the use of air collectors with rock storage for domestic heating applica-
tions. Combining water with air/rock systems has become a standard form
of solar system application. The objective is to provide a portion of energy
needs for domestic hot water without a significant reduction in the solar energy
supply for space heating. Essentially, this hybrid system is composed of a stan-
dard air/rock solar system, a counterflow heat exchanger, and a small water
store. Conventional air/rock solar systems for space heating are only used dur-
ing the heating season, lengthening the payback period, because the solar sys-
tem is inoperative for several months of the year. Adding hot water capability to
the basic air/rock system increases the capital cost slightly. So, for a relatively
small additional investment, a solar thermal system can provide 100% of
the domestic hot water during summer and proportionally less as the seasons
change to and from winter.

6.4.1 Thermodynamic Analysis of Heat Storage Systems


for Heating Applications
The subprocesses of a storage for heating capacity are shown in Fig. 6.1. Ther-
modynamic analyses of the heat storage system are conducted based on three
subprocesses: charging, storing, and discharging.
6.4 Heat Storage Systems For Heating Applications 239

Charging Storing Discharging


b c

a d

Ql

Time
FIGURE 6.1 Three processes in a general heat storage system for heating capacity: charging (left),
storing (middle), and discharging (right). The heat leakage from the system Ql is illustrated for the
storing process but can occur in all three processes. Modified from Dincer (2002).

6.4.1.1 Charging Process


During the charging period, the total heat input (charge), Qc, provided to the
heat storage system can be expressed as
Qc ¼ mc cp ðTc, i  Tc, o Þ (6.1)

where cp, Tc,i, and Tc,o denote specific heat capacity, charging inlet temperature,
and outlet temperature of the HTF, respectively. Also, mc represents the total
mass accumulated and/or transported over the charging period, tc (s), and
can be written as
ð tc
mc ¼ m_ ðtc Þ dtc (6.2)
0

For a constant mass flow rate, ṁ(tc) can be denoted as ṁ (kg/s) and the above
equation can be rewritten as
_ c
mc ¼ mt (6.3)

The total exergy input to the heat storage system during the charging process is
given by
  
Tc, i
ExQ, c ¼ mc cp ðTc, i  Tc, o Þ  T0 ln (6.4)
Tc, o

6.4.1.2 Storing Process


During the storing process, the heat storage system (for heating applications)
loses heat due to the energy interaction between the system and its
240 C HA PT E R 6 : Heat Storage Systems

surrounding. This interaction is mainly through heat transfer along the system
boundaries since the storage temperature (Ts) is higher than the surrounding
temperature (T0).This heat loss rate (Q_ l ) can be assessed as
Q_ l ¼ UAðTs  T0 Þ (6.5)

where U and A represent overall heat transfer coefficient and storage surface
area, respectively.
The overall heat loss over the entire storing period is given by
Ql ¼ ms cp ðΔTs Þ (6.6)

where mc, cp, and ΔTs denote mass, specific heat capacity, and the difference
between initial (Ts,i) and final (Ts,f) temperatures of the heat storage medium,
respectively. Note that ΔTs is calculated as
ΔTs ¼ Ts, i  Ts, f (6.7)

The total exergy loss from the heat storage system during the storing process is
given by
 
T0
ExQ, l ¼ Ql 1  (6.8)
Ts

6.4.1.3 Discharging Process


During the discharging period, the total heat output (discharge), Qd, provided
by the heat storage system is given as
Qd ¼ md cp ðTd, o  Td, i Þ (6.9)

where cp, Td,i, and Td,o denote specific heat capacity, discharging inlet temper-
ature, and discharging outlet temperature of the HTF, respectively. The term md
represents the total mass accumulated and/or transported over the discharging
period, td (s), and can be written as
ð td
md ¼ m_ ðtd Þ dtd (6.10)
0

For a constant mass flow rate, ṁ(td) can be denoted as ṁ (kg/s) and the above
equation can be rewritten as
_ d
md ¼ mt (6.11)

The total exergy discharged from the heat storage system during the discharging
process is expressible as
  
Td, o
ExQ, d ¼ md cp ðTd, o  Td, i Þ  T0 ln (6.12)
Td, i
6.4 Heat Storage Systems For Heating Applications 241

6.4.1.4 Energy and Exergy Efficiencies


The energy efficiency of the heat storage system can be evaluated over the entire
operating cycle (charging, storing, and discharging) as the ratio of the total
energy recovered from the system during discharging (Qd) to the total energy
charged into the system (Qc). That is,
Qd
η¼ (6.13)
Qc

Similarly, the exergy efficiency of the system can be evaluated for the overall
storage period as
ExQ, d
ψ¼ (6.14)
ExQ, c

6.4.2 Illustrative Example


Consider two heat storage systems (Fig. 6.2), X and Y, in surroundings at 25 °C.
Each storage receives 104,650 kJ of heat from a stream of 500 kg of water, which
is cooled from 80 to 30 °C. Therefore, the heat input to the storage during the
charging period for each storage unit is
Qc ¼ mc cp ΔT ¼ 500kg  4:186kJ=kg°C  ð80  30Þ°C ¼ 104, 650kJ

Charging
80 °C

500 kg

T0 = 25 °C 30 °C T0 = 25 °C

35 °C 75 °C

h = 90% h = 90%
4500 kg 500 kg
y = 25% y = 83%

30 °C 30 °C
Discharging after 1 day Discharging after 100 days
(case X) (case Y)
FIGURE 6.2 Storage process for two systems, X and Y, showing charging process (top), which is identical
for both systems, and discharging processes for system X (bottom left) and system Y (bottom right).
242 C HA PT E R 6 : Heat Storage Systems

6.4.2.1 Energy Efficiency Comparison


6.4.2.1.1 System X
After 1 day, the heat of 94,185 kJ is recovered during the discharging period
from storage system X by a stream of 4500 kg of water being heated from 30
to 35 °C, that is,
Qd ¼ md cp ΔT ¼ 4500kg  4:186kJ=kg°C  ð35  30Þ°C ¼ 94,185kJ

Therefore, the energy efficiency for the heat storage system X is


Qd 94,185kJ
ηX ¼ ¼ ¼ 0:90
Qc 104, 650kJ

The heat loss to the surroundings during storage is


Ql ¼ Qc  Qd ¼ 104, 650kJ  94, 185kJ ¼ 10, 465kJ

6.4.2.1.2 System Y
Storage system Y stores the heat for 100 days. For this system, the heat recovered
during discharging, the energy efficiency, and the heat rejection to the sur-
rounding can be evaluated in a similar way as for system X. After the storing
period, an energy quantity of 94,185 kJ is recovered during the discharging
period by heating a stream of 500 kg of water from 30 to 75 °C, determined
as follows:
Qd ¼ md cp ΔT ¼ 500kg  4:186kJ=kg°C  ð75  30Þ°C ¼ 94, 185kJ

The energy efficiency for system Y can be found following the method for sys-
tem X:
Qd 94, 185kJ
ηY ¼ ¼ ¼ 0:90
Qc 104,650kJ

The heat loss to the surroundings during storage can be determined as follows:
Ql ¼ Qc  Qd ¼ 104, 650kJ  94, 185kJ ¼ 10, 465kJ

Note that the ability of storing sensible heat in a given tank (or container) is
dependent on the value for the material of ρcp.
Both storage systems have the same efficiency based on energy or the FLT. But
system Y, which stores heat for 100 days rather than 1 day and returns the heat
at the much more useful temperature of 75 °C rather than 35 °C, exhibits con-
siderably better performance. It is clear that a more perceptive measure of com-
parison than that provided by the energy efficiency of the storage cycle is
needed if the true usefulness of a heat storage system is to be assessed and a
6.4 Heat Storage Systems For Heating Applications 243

rational basis for the optimization of its economic value established. This can
be achieved via the exergy efficiency, which is a better measure of the thermo-
dynamic effectiveness of the heat storage system. An efficiency defined simply
as the percentage of the total energy stored in a system that can be recovered
ignores the quality of the recovered energy and so cannot provide a measure
of ideal performance as noted earlier. An efficiency based on exergy provides
a better measure.

6.4.2.2 Exergy Efficiency Comparison


Consider the aforementioned example of heat storage systems X and Y. For the
corresponding cases, an exergy analysis is conducted below. The exergy efficien-
cies can be straightforwardly obtained using the equations in Table 6.2. The
exergy change during the charging period can be obtained as
  
T1
ExQ, c ¼ mc cp ðT1  T2 Þ  T0 ln
T2

or
  
353
ExQ, c ¼ 500kg  4:186kJ=kgK  ð353  303Þ  298  ln K ¼ 9386:88kJ
303

6.4.2.2.1 System X
The exergy change during the discharging period for system X can be calculated
as follows:
  
T1
ExQ, d ¼ md cp ðT1  T2 Þ  T0 ln
T2

or
  
308
ExQ, d ¼ 4500kg  4:186kJ=kgK  ð308  303Þ  298  ln K ¼ 2310:18kJ
303

Therefore, the exergy efficiency for storage X is


ExQ, d 2310:18kJ
ψX ¼ ¼ ¼ 0:25
ExQ, c 9316:88kJ

6.4.2.2.2 System Y
As pointed out earlier, heat is recovered from the storage after 100 days by a
stream of 500 kg of water entering at 30 °C and leaving at 75 °C. The exergy
change of storage system Y can be obtained as follows:
  
T1
ExQ, d ¼ md cp ðT1  T2 Þ  T0 ln
T2
244 C HA PT E R 6 : Heat Storage Systems

or
  
348
ExQ, d ¼ 500kg  4:186kJ=kgK  ð348  303Þ  298  ln K ¼ 7819:52kJ
303

The exergy efficiency for storage Y can then be determined:


ExQ, d 7819:52kJ
ψY ¼ ¼ ¼ 0:83
ExQ, c 9316:88kJ

The performance of heat storage systems X and Y is compared in Table 6.8. Both
storage systems exhibit the same energy efficiencies (90%) despite having
markedly different storage periods. This observation indicates that the FLT is
not sufficient to distinguish the performance of these two heat storage systems.
This requirement suggests exergy methods are needed. The distinction between
the two storage systems (X and Y) is easily and clearly made using exergy anal-
ysis. The exergy efficiencies differ at 25% for system X and 83% for system Y.
This advantage of heat storage system Y over system X is due to its higher heat
recovery temperature. This example illustrates in a practical manner the con-
cepts discussed in this chapter and highlights the importance of exergy, rather
than energy, methods in evaluating the performance of heat storage systems.
The results are consistent with those obtained previously by Rosen et al.
(1988) and Rosen (1992) for energy storage systems.

Table 6.8 Performance Comparison of Heat Storage Systems X and Y


System X System Y

General parameters
Storing period (days) 1 100
Charging-fluid inlet temperature (°C) 80 80
Charging-fluid outlet temperature (°C) 30 30
Discharging-fluid inlet temperature (°C) 30 30
Discharging-fluid outlet temperature (°C) 35 75
Energy parameters
Energy input (kJ) 104,650 104,650
Energy recovered (kJ) 94,185 94,185
Energy loss (kJ) 10,465 10,465
Energy efficiency (%) 90 90
Exergy parameters
Exergy input (kJ) 9387 9387
Exergy recovered (kJ) 2310 7820
Exergy loss (kJ) 7077 1567
Exergy efficiency (%) 25 83
6.4 Heat Storage Systems For Heating Applications 245

More broadly, important advantages of exergy analysis can be listed as follows:


• It provides more proper accounting of the loss of availability of heat in
storage systems using the conservation of mass and energy principles
together with the SLT for the goals of design and analysis.
• It gives more meaningful and useful information than energy analysis
regarding the efficiency, losses, and performance for heat storage systems.
• It more correctly reflects the thermodynamic and economic values of the
operation of heat storage systems.
• It is a useful technique for revealing by how much it is possible to design
more efficient heat storage systems by reducing the inefficiencies.

6.4.3 Case Study of a Macroscale Application: Borehole


Storage at UOIT (Canada)
The University of Ontario Institute of Technology (UOIT) in Oshawa, Ontario,
has a unique borehole heat storage system with HTF of a 15% glycol/water mix,
circulating through over 150 km of 400 polypropylene piping, which was pressure
tested to 300 psi. The system has two mechanical rooms, one of them is for a
boiler system and geothermal heat pumps and the other for chillers. The system
injects heat from the geothermal loop into the ground during summer and uses
this heat in the winter to provide low-temperature hydronic heating at 52 °C.
A challenge with the borehole heat storage system was its high capital cost,
which was mainly due to the use of high-efficiency HVAC equipment and
the construction of the geothermal field. This was offset by significant operating
cost savings each year. The system attained annual energy savings of 40% in the
heating and 16% in the cooling modes. The payback period was 3-5 years for
the high-efficiency HVAC equipment and 7.5 years for the well field.

6.4.4 Thermodynamic Analysis of ATES


Underground aquifers are sometimes used for heat storage systems (Jenne,
1992). Table 6.9 lists examples of buildings using ATES-based systems for

Table 6.9 Selected ATES Systems for Heating and Cooling Applications
Building Location Initial Operating Date Capacity (kW)

IBM office Zoetermeer (NL) 1992 700


Nike office Hilversum (NL) 1999 2000
Maira hospital Overpelt (BE) 2005 1500
IKEA store Duiven (NL) 1999 750
Westway housing London (the United 2006 250
project Kingdom)
Source: Hendricks et al. (2008).
246 C HA PT E R 6 : Heat Storage Systems

Cold Heat
demand demand

Aquifer Aquifer

FIGURE 6.3 ATES working principle for both cooling (left) and heating (right) modes.

providing a range of cooling/heating loads. The storage medium in many ATES


systems remains in a single phase during the storing cycle, so that temperature
changes are exhibited in the store as heat is added or removed. The working
principle of an ATES system is presented in Fig. 6.3 for both heating and
cooling modes.
ATES systems have been demonstrated to be feasible commercially and envi-
ronmentally, and some of these systems have been assessed and compared ther-
modynamically while operating at the same or varied environmental
conditions. An early study of ATES, including both energy and exergy analyses
and efficiency evaluations, was made by Rosen (1999). His approach is fol-
lowed here to assess energetically and exegetically the performance of an ATES
in different operating modes (heating/cooling).
The application of exergy analysis to ATES systems is described in this section.
For an elementary ATES model, expressions are presented for the injected and
recovered quantities of energy and exergy and for efficiencies. The impact of
introducing a threshold temperature below which residual heat remaining in
the aquifer water is not considered worth recovering is examined. ATES exergy
efficiencies are demonstrated to be more useful and meaningful than energy
efficiencies because the former account for the temperatures associated with
thermal energy transfers and consequently assess how nearly ATES systems
approach ideal thermodynamic performance. ATES energy efficiencies do
not provide a measure of approach to ideal performance and, in fact, are often
misleadingly high because some of the thermal energy can be recovered at tem-
peratures too low for useful purposes.
6.4 Heat Storage Systems For Heating Applications 247

6.4.4.1 ATES Model


The charging of the ATES occurs over a finite time period tc and after a holding
interval discharging occurs over a period td. The working fluid is water, having a
constant specific heat cp, and assumed incompressible. The temperature of the
aquifer and its surroundings prior to heat injection is T0, the reference-
environment temperature. Only heat stored at temperatures above T0 is consid-
ered, and pump work is neglected.
During charging, heated water at a constant temperature Tc is injected at
a constant mass flow rate ṁc into the ATES. After a storing period, dischar-
ging occurs, during which water is extracted from the ATES at a constant
mass flow rate ṁd.The fluid discharge temperature is taken to be a function
of time, that is, Td ¼ Td(t). The discharge temperature after an infinite
time is taken to be the temperature of the reference environment, that
is, Td(1) ¼ T0, and the initial discharge temperature is taken to be
between the charging and reference-environment temperatures, that is,
T0  Td(0)  Tc.
Many discharge temperature-time profiles are possible. Here, the discharge
temperature is taken to decrease linearly with time from an initial value
Td(0) to a final value T0. The final temperature is reached at a time tf and
remains fixed at T0 for all subsequent times, that is,
8 9
< ðTd ð0Þ  T0 Þt =
Td ð0Þ  ð0  t  tf Þ
Td ðt Þ ¼ tf (6.15)
: ;
T0 ðtf  t  t1 Þ

The simple linear discharge temperature-time profile is sufficiently realistic yet


simple. The temperature-time profiles considered in the present model for the
fluid flows during the charging and discharging periods are summarized in
Fig. 6.4.

Charging Discharging
Temperature, T

Tc
Td(0)

To

0 tc 0 tf
Time, t Time, t
FIGURE 6.4 Temperature-time profiles assumed for the charging and discharging periods in the ATES
model considered. Modified from Dincer and Rosen (2013).
248 C HA PT E R 6 : Heat Storage Systems

The two main types of thermodynamic losses that occur in ATES systems are
accounted for in the model:

• Energy losses. Energy injected into an ATES that is not recovered is


considered lost. Thus, energy losses include energy remaining in the
ATES and energy injected into the ATES that is convected in a water
flow or is transferred by conduction far enough from the discharge
point that it is unrecoverable.
• Mixing losses. As heated water is pumped into an ATES, it mixes with the water
already present (which is usually cooler), resulting in the recovered water
being at a lower temperature than the injected water. In the present model,
this loss results in the discharge temperature Td being at all times less than
or equal to the charging temperature Tc, but not below the reference-
environment temperature T0 (i.e., T0  Td(t)  Tc for 0  t  1).

6.4.4.2 Energy and Exergy Analyses


The energy and exergy injected into the ATES during charging and recovered
during discharging are evaluated. The energy flow associated with a flow of liq-
uid at a constant mass flow rate ṁ, for an arbitrary period of time with T a func-
tion of t, is
ðt
E¼ E_ðt Þdt (6.16)
0

where the integration is performed over the time period, and the energy flow
rate at time t is

E_ðt Þ ¼ mc
_ p ðT ðt Þ  T0 Þ (6.17)

Here, c denotes the specific heat of the liquid. Combining Eqs. (6.16) and
(6.17) for constant ṁ, cp, and T0,
ðt
_ p
E ¼ mc ðT ðt Þ  T0 Þdt (6.18)
0

The corresponding exergy flow is


ðt
Ex ¼ _ ðt Þdt
Ex (6.19)
0

where the exergy flow rate at time t is


 
_ ðt Þ ¼ mc T ðt Þ
Ex _ p ðT ðt Þ  T0 Þ  T0 ln (6.20)
T0

Combining Eqs. (6.19) and (6.20), and utilizing Eq. (6.18),


6.4 Heat Storage Systems For Heating Applications 249

ðt   ðt  
T ðt Þ T ðt Þ
_ p
Ex ¼ mc ðT ðt Þ  T0 Þ  T0 ln _ p T0 ln
dt ¼ E  mc dt (6.21)
0 T0 0 T0

6.4.4.2.1 Charging and Discharging


The energy input to the ATES during charging, for a constant water injection
rate ṁc and over a time period beginning at zero and ending at tc, is expressed
by Eq. (6.18) with T(t) ¼ Tc. That is,
ð tc
Tc
Ec ¼ m_ c cp ðTc ðt Þ  T0 Þdt ¼ m_ c cp tc (6.22)
0 T0

The corresponding exergy input is expressed by Eq. (6.21), with the same con-
ditions as for Ec. Thus, after integration,
 
Tc Tc
_
Exc ¼ mc cp tc ðTc  T0 Þ  T0 ln ¼ Ec  m_ c cp tc T0 ln (6.23)
T0 T0

The energy recovered from the ATES during discharging, for a constant water
recovery rate ṁd and for a time period starting at zero and ending at td, is
expressed by Eq. (6.18) with T(t) as in Eq. (6.15). Thus,
ð td
θð2tf  θÞ
Ed ¼ m_ d cp ðTd ðt Þ  T0 Þdt ¼ m_ d cp ½Td ð0Þ  T0  (6.24)
0 2tf

where

td ð0  td  tf Þ
θ¼ (6.25)
tf ðtf  td  t1 Þ

The corresponding exergy recovered is expressed by Eq. (6.21), with the same
conditions as for Ed. Thus,
ð td   ð td  
Td ðt Þ Td ðt Þ
Exd ¼ m_ d cp ðTd ðt Þ  T0 Þ  T0 ln dt ¼ Ed  m_ d cp T0 ln dt (6.26)
0 T0 0 T0

Here,
ð td   ð td  
Td ðt Þ aθ + b b
ln dt ¼ ln ðat + bÞdt ¼ ln ðaθ + bÞ  θ  ln ðbÞ (6.27)
0 T0 0 a a

where
T0  Td ð0Þ
a¼ (6.28)
T0 Tf
Td ð0Þ
b¼ (6.29)
T0
250 C HA PT E R 6 : Heat Storage Systems

When td  tf, the expression for the integral in Eq. (6.27) reduces to
ð td    
Td ðt Þ Td ð0Þ Td ð0Þ
ln dt ¼ tf ln 1 (6.30)
0 T0 Td ð0Þ  T0 T0

6.4.4.2.2 Balances and Efficiencies


An ATES energy balance taken over a complete charging-discharging cycle states
that the energy injected is either recovered or lost. A corresponding exergy bal-
ance states that the exergy injected is either recovered or lost, where lost exergy is
associated with both waste exergy emissions and internal exergy consumptions
due to irreversibilities.
If f is defined as the fraction of injected energy Ec that can be recovered if the
length of the discharge period approaches infinity (i.e., water is extracted until
all recoverable energy has been recovered), then
Ed ðtd ! 1Þ ¼ fEc (6.31)

It follows from the energy balance that (1  f) Ec is the energy irretrievably lost
from the ATES. Clearly, f varies between zero for a thermodynamically worth-
less ATES to unity for an ATES having no energy losses during an infinite dis-
charge period. But mixing in the ATES can still cause exergy losses even if f ¼ 1.
Since Ec is given by Eq. (6.22) and Ed(td ! 1) by Eq. (6.24) with θ ¼ tf,
Eq. (6.31) may be rewritten as
m_ d cp ðTd ð0Þ  T0 Þtf
¼ f m_ c cp ðTc  T0 Þtc (6.32)
2

or
tf m_ d ðTd ð0Þ  T0 Þ
f¼ (6.33)
2tc m_ c ðTc  T0 Þ

For either energy or exergy, the efficiency is the fraction, taken over a complete
cycle, of the quantity input during charging that is recovered during dischar-
ging, while loss is the difference between input and recovered amounts
of the quantity. Hence, the energy loss as a function of the discharge time
period is given by [Ec  Ed(td)], while the corresponding exergy loss is given
by [Exc Exd(td)]. Energy losses do not reflect the temperature degradation
associated with mixing, while exergy losses do.
The energy efficiency η for an ATES, as a function of the discharge time period, is
given by
Ed ðtd Þ m_ d ðTd ð0Þ  T0 Þ θð2tf  θÞ
ηðtd Þ ¼ ¼ (6.34)
Ec m_ c ðTc  T0 Þ 2tf tc
6.4 Heat Storage Systems For Heating Applications 251

and the corresponding exergy efficiency ψ by


Exd ðtd Þ
ψ ðtd Þ ¼ (6.35)
Exc

The energy efficiency in Eq. (6.34) simplifies when the discharge period td
exceeds tf, that is, η(td  tf) ¼ f.
In practice, it is not economically feasible to continue the discharge period until
as much recoverable heat as possible is recovered. As the discharge period
increases, water is recovered from an ATES at ever-decreasing temperatures
(ultimately approaching the reference-environment temperature To), and the
energy in the recovered water is of decreasing usefulness. Exergy analysis reflects
this phenomenon, as the magnitude of the recovered exergy decreases as the
recovery temperature decreases. To determine the appropriate discharge period,
a threshold temperature Tt is often introduced, below which the residual energy
in the aquifer water is not considered worth recovering from an ATES. For the
linear temperature-time relation used here (see Eq. 6.15), it is clear that no ther-
mal energy could be recovered over a cycle if the threshold temperature exceeds
the initial discharge temperature, while the appropriate discharge period can be
evaluated using Eq. (6.15) with Tt replacing Td(t) for the case where
To  Tt  Td(0). Thus,
8 9
< Td ð0Þ  Tt =
ðT0  Tt  Td ð0ÞÞ
td ¼ Td ð0Þ  T0 (6.36)
: ;
0 ðTd ð0Þ  Tt Þ

In practice, a threshold temperature places an upper limit on the allowable dis-


charge time period. Utilizing a threshold temperature usually has the effect of
decreasing the difference between the corresponding energy and exergy
efficiencies.
Nonetheless, ATES performance measures based on exergy are more useful and
meaningful than those based on energy. Exergy efficiencies account for the tem-
peratures associated with heat transfers to and from an ATES, as well as the
quantities of heat transferred, and consequently provide a measure of how
nearly ATES systems approach ideal performance. Energy efficiencies account
only for quantities of energy transferred and can often be misleadingly high,
for example, in cases where heat is recovered at temperatures too low to be use-
ful. The use of an appropriate threshold recovery temperature can partially
avoid the most misleading characteristics of ATES energy efficiencies.
As a case study, which is based on Eqs. (6.15)–(6.36), energy and exergy ana-
lyses of an ATES are performed using experimental data from the first of four
short-term ATES test cycles for the Upper Cambrian Franconia-Ironton-
Galesville confined aquifer. The test cycles were performed at the University
252 C HA PT E R 6 : Heat Storage Systems

Charging

Volumetric
Discharging

rate (L/s)
20

flow
20
18 18
16 16
150 150

Temperature,
100 100

T (⬚C)
50 50

0 0
0 2 8 14 16 0 2 4 6 8
Time, t (days) Time, t (days)

FIGURE 6.5 Observed values for the temperature and volumetric flow rate of water, as a function of time
during the charging and discharging periods, for the experimental test cycles used in the ATES
case study. Modified from Dincer and Rosen (2013).

of Minnesota’s St. Paul campus from November 1982 to December 1983


(Hoyer et al., 1985). During the test, water was pumped from the source well,
heated in a heat exchanger, and returned to the aquifer through the storage well.
After storage, energy was recovered by pumping the stored water through a heat
exchanger and returning it to the supply well. The storage and supply wells are
located 255 m apart.
For the test cycle considered here, the water temperature and volumetric flow
rate vary with time during the injection and recovery processes as shown in
Fig. 6.5. The storage period duration (13 days) is also shown. Charging
occurred during 5.24 days over a 17-day period. The water temperature and vol-
umetric flow rate were approximately constant during charging and had mean
values of 89.4 °C and 18.4 L/s, respectively. Discharging also occurred over
5.24 days, approximately with a constant volumetric flow rate of water and lin-
early decreasing temperature with time. The mean volumetric flow rate during
discharging was 18.1 L/s, and the initial discharge temperature was 77 °C,
while the temperature after 5.24 days was 38 °C. The ambient temperature
was reported to be 11 °C.

6.4.4.2.3 Simplifications, Analysis, and Results


In subsequent calculations, mean values for volumetric flow rates and charging
temperature are used. Also, the specific heat and density of water are both taken to
be fixed, at 4.2 kJ/kg K and 1000 kg/m3, respectively. Since the volumetric flow
rate (in L/s) is equal to the mass flow rate (in kg/s) when the density is
1000 kg/m3, m_ c ¼ 18:4kg=s and m_ d ¼ 18:1kg=s. Also, the reference-environment
temperature is fixed at the ambient temperature, that is, T0 ¼ 11 °C ¼ 284 K.
During charging, it can be shown using Eqs. (6.22) and (6.23), with
tc ¼ 5.24 day ¼ 453,000 s and Tc ¼ 89.4 °C ¼ 362.4 K, that

Ec ¼ 18:4kg=s  4:2kJ=kgK  453, 000s  ð362:4  284ÞK ¼ 2:74  109 kJ


6.4 Heat Storage Systems For Heating Applications 253

and
 
362:4K
Exc ¼ 2:74  109 kJ  18:4kg=s  4:2kJ=kgK  453, 000s  284K  ln
284K
¼ 0:32  109 kJ

During discharging, the value of the time tf is evaluated using the linear
temperature-time relation of the present model and the observations that
Td(t ¼ 5.24 day) ¼ 38 °C ¼ 311 K and Td(0) ¼ 77 °C ¼ 350 K. Then, using
Eq. (6.15) with t ¼ 5.24 day,
5:24day
38 °C ¼ 77°C  ð77 °C  11 °CÞ 
tf

which can be solved to show that tf ¼ 8.87 day. Thus, with the present linear
model, the discharge water temperature would reach T0 if the discharge period
was lengthened to almost 9 days. In reality, the rate of temperature decline
would likely decrease, and the discharge temperature would asymptotically
approach T0. The value of the fraction f can be evaluated with Eq. (6.33) as
8:87day  18:1kg=s  ð77  11Þ°C
f¼ ¼ 0:701
2  5:24day  18:4kg=s  ð89:4  11Þ°C

Thus, the maximum energy efficiency achievable is approximately 70%. With


these values and Eqs. (6.28) and (6.29), it can be shown that
ð11  77Þ°C
a¼ ¼ 0:0262day 1
284K  8:87day
350K
b¼ ¼ 1:232
284K

Consequently, expressions dependent on discharge time period td can be writ-


ten and plotted (see Fig. 6.6) for Ed, Exd, η, and ψ using Eqs. (6.24)–(6.26),
(6.33), and (6.34) and for the energy loss (Ec  Ed) and exergy loss (Exc  Exd).

6.4.4.2.4 Discussion
Both energy and exergy efficiencies in Fig. 6.6 increase from zero to maximum
values as td increases. Further, the difference between the two efficiencies
increases with increasing td. This latter point demonstrates that the exergy effi-
ciency gives less weight than the energy efficiency to the energy recovered at
higher td values since it is recovered at temperatures nearer to the reference-
environment temperature.
Several other points in Fig. 6.6 are worth noting. First, for the conditions spec-
ified, all parameters level off as td approaches tf and remain constant for td  tf.
Second, as td increases towards tf, the energy recovered increases from zero to a
254 C HA PT E R 6 : Heat Storage Systems

Energy and exergy quantities (109 kJ)


td = tf

2 Energy loss

1
Energy recovery, Ed
Exergy loss
Exergy recovery, Ed
0
100

Energy efficiency, h
Efficiencies (%)

50

Exergy efficiency, y
0
0 5 10
Discharge time period, td (days)

FIGURE 6.6 Variation of several calculated energy and exergy quantities and efficiencies, as a function
of discharge time period, for the ATES case study. Modified from Dincer and Rosen (2013).

maximum value, while the energy loss decreases from a maximum of all the
input energy to a minimum (but nonzero) value. The exergy recovery and
exergy loss functions behave similarly qualitatively. but exhibit much lower
magnitudes.
The difference between energy and exergy efficiencies is due to temperature dif-
ferences between the charging-fluid and discharging-fluid flows. As the dischar-
ging time increases, the deviation between these two efficiencies increases
(Fig. 6.6) because the temperature of recovered heat decreases (Fig. 6.5). In this
case, the energy efficiency reaches approximately 70% and the exergy efficiency
40% by the completion of the discharge period, even though the efficiencies are
both 0% when discharging commences.
To further illustrate the importance of temperature, a hypothetical modifica-
tion of the present case study is considered. In the modified case, all details
are as in the original case except that the temperature of the injection flow dur-
ing the charging period is increased from 89.4 to 200 °C (473 K), while the
duration of the charging period is decreased from its initial value of 5.24 days
6.4 Heat Storage Systems For Heating Applications 255

(453,000 s) so that the energy injected does not change. By equating the energy
injected during charging for the original and modified cases, the modified
charging-period duration tc0 can be evaluated as a function of the new injection
flow temperature Tc0 as follows:
Tc  T0 ð89:4  11Þ°C
tc0 ¼ tc ¼ 453,000s  ¼ 188, 000s
Tc0  T0 ð200  11Þ°C

The modified exergy input during charging can then be evaluated as


 
473K
Ex0c ¼ 2:74  109 kJ  18:4kg=s  4:2kJ=kgK  188, 000s  284K  ln
284K
¼ 0:64  109 kJ

This value is double the exergy input during charging for the original case, so,
since the discharging process remains unchanged in the modified case, the
exergy efficiency (for any discharging time period) is half that for the original
case. The altered value of exergy efficiency is entirely attributable to the new
injection temperature and occurs despite the fact that the energy efficiency
remains unchanged.
If a threshold temperature is introduced and arbitrarily set at 38 °C (the actual
temperature at the end of the experimental discharge period of 5.24 day), then
the data in Fig. 6.6 for td ¼ 5.24 day apply and one can see that

• the exergy recovered (0.127  109 kJ) is almost all (91%) of the exergy
recoverable in infinite time (0.139  109 kJ), while the energy recovered
(1.60  109 kJ) is not as great a portion (83%) of the ultimate energy
recoverable (1.92  109 kJ);
• the exergy loss (0.19  109 kJ) exceeds the exergy loss in infinite time
(0.18  109 kJ) slightly (by 5.5%), while the energy loss (1.14  109 kJ)
exceeds the energy loss in infinite time (0.82  109 kJ) substantially
(by 39%); and
• the exergy efficiency (40%) has almost attained the exergy efficiency
attainable in infinite time (43.5%), while the energy efficiency
(58%) is still substantially below the ultimate energy efficiency
attainable (70%).

To gain confidence in the model and the results, some of the quantities calcu-
lated using the linear model can be compared with the same quantities as
reported in the experimental paper (Hoyer et al., 1985):

• The previously calculated value for the energy injection during charging of
2.74  109 kJ is 1.1% less than the reported value of 2.77  109 kJ.
• The energy recovered at the end of the experimental discharge period of
td ¼ 5.24 days can be evaluated with Eq. (6.24) as
256 C HA PT E R 6 : Heat Storage Systems

Ed ð5:24day Þ ¼ 18:1kg=s  4:2kJ=kg°C  ð77  11Þ°C


5:24  ð2  8:87  5:24Þ
 86,400s=day
2  8:87
¼ 1:60  10 kJ
9

which is 1.8% less than the reported value of 1.63  109 kJ.
• The energy efficiency at td ¼ 5.24 day can be evaluated with Eq. (6.25) as

1:60  109 kJ
ηð5:24day Þ ¼ ¼ 0:584
2:74  109 kJ

which is 1.0% less than the reported value of 0.59 (referred to as the
“energy recovery factor”).

6.5 HEAT STORAGE SYSTEMS FOR COOLING


APPLICATIONS
In heat storage systems for cooling applications, or cold TES (CTES), “cold” is
stored in a thermal storage mass. In most conventional cooling systems, there
are two major components (Dincer and Rosen, 2011):

• Chiller: to cool a fluid such as water


• Distribution system: to transport the cold fluid from the chiller to where
cooling is needed (e.g., to cool air in a building)

In conventional building systems, the chiller operates only when the building
occupants require cold air. In a cooling system incorporating heat storage sys-
tems, the chiller also operates at times other than when the cooling is needed.
During the past two decades, heat storage technologies, especially cold storage,
have matured and are now accepted as a proved energy technologies. However,
the predicted payback period of a potential cold storage installation is often not
sufficiently attractive to give it priority over other energy-efficient technologies.
This determination often is made because full advantage is not made of the
many potential benefits of cold storage or because the cold storage sizing is
not optimized. Some recommendations for optimizing the payback period
of CTES systems are discussed below.
For new facilities, cold storage should be integrated carefully into the overall
building and its energy systems so as to exploit the potential benefits of CTES,
including

• reduced pipe and pump sizes for chilled-water distribution,


• reduced duct and fan sizes for low-temperature air distribution, and
• correspondingly reduced operating costs.
6.5 Heat Storage Systems For Cooling Applications 257

Smaller chiller and electrical systems lead to initial cost advantages. The sizing
of the cold storage system should be optimized, as opposed to the typical pro-
cess of considering full storage and one or two levels of partial-storage versus a
conventional system. A practical method to assist in determining the optimum
system size should be developed. Also, the value should be accounted for of the
gain in usable building space due to less space being required for mechanical
system components when CTES is used.
For existing facilities, the potential advantages of CTES that should be evaluated
include

• modifying the existing chillers to make ice versus the purchase of a new
machine,
• using spare chiller capacity by adding a CTES system,
• using cold storage to increase cooling capacity in situations where chiller
and electrical service capacity are fully utilized,
• sizing the cold storage system optimally as opposed to taking the best of
only a few options, and
• using available low-temperature air and water to advantage through “free
cooling” where practical.

6.5.1 CTES Storage Media Selection and Characteristics


The storage medium largely determines the storage volume and the size and
configuration of the HVAC system and components. The main options include
chilled water, ice, and eutectic salts. Ice systems offer the densest storage capac-
ity but have the most complex charge and discharge equipment. Water systems
offer the lowest storage density and are the least complex. Eutectic salts have
intermediate characteristics. Details on each storage medium are now provided:
Chilled water: Chilled-water systems require the largest storage tanks but can
easily interface with existing chiller systems. Chilled-water CTESs use the sen-
sible heat capacity of water to store cooling capacity. They operate at temper-
ature ranges (3.3-5.5 °C) compatible with standard chiller systems and are
most economic for systems greater than 2000 ton/h in capacity.
Ice: Ice systems use smaller tanks and offer the potential for the use of low-
temperature air systems but require more complex chiller systems. Ice CTES sys-
tems use the latent heat of fusion of water (335 kJ/kg) to store cooling capacity.
To store energy at the temperature of ice requires refrigeration equipment that
provides charging fluids at temperatures below the normal operating range of
conventional air conditioning equipment. Special ice-making equipment or
standard chillers modified for low-temperature service are used. The low
chilled-water supply temperatures available from ice storage allow the use of
cool-air distribution, the benefits of which include the ability to use smaller
258 C HA PT E R 6 : Heat Storage Systems

fans and ducts and the introduction of less humid air into occupied spaces.
With ice as the storage medium, there are several technologies available for
charging and discharging the storage. Ice-harvesting systems feature an evapo-
rator surface on which ice is formed and periodically released into a storage
tank that is partially filled with water. External melt ice-on-coil systems use sub-
merged pipes through which a refrigerant or secondary coolant is circulated. Ice
accumulates on the outside of the pipes. Storage is discharged by circulating the
warm return water over the pipes, melting the ice from the outside. Internal
melt ice-on-coil systems also feature submerged pipes on which ice is formed.
Storage is discharged by circulating warm coolant through the pipes, melting
the ice from the inside. The cold coolant is then pumped through the building
cooling system or used to cool a secondary coolant that circulates through the
building’s cooling system. Encapsulated ice systems use water inside sub-
merged plastic containers that freeze and thaw as cold or warm coolant is cir-
culated through the storage tank holding the containers. Ice slurry systems store
water or water/glycol solutions in a slurry state (a partially frozen mixture of
liquid and ice crystals that looks like slush). To meet a cooling demand, the
slurry may be pumped directly to the load or to a heat exchanger cooling a sec-
ondary fluid that circulates through the building chilled-water system. Internal
melt ice-on-coil systems are the most commonly used type of ice storage tech-
nology in commercial applications. External melt and ice-harvesting systems
are more common in industrial applications, although they can also be applied
in commercial buildings and district cooling systems. Encapsulated ice systems
are also suitable for many commercial applications. Ice slurry systems have not
been widely used in commercial applications.
Eutectic salts: Eutectic salts can use existing chillers but usually operate at
warmer temperatures than ice or chilled-water systems. Eutectic salts use a com-
bination of inorganic salts, water, and other elements to create a mixture that
freezes at a desired temperature. The material is encapsulated in plastic con-
tainers that are stacked in a storage tank through which water is circulated.
The most commonly used mixture for thermal storage freezes at 8.3 °C, which
allows the use of standard chilling equipment to charge the storage, but leads to
higher discharge temperatures. These temperatures, in turn, limit the operating
strategies that may be applied. For example, eutectic salts may only be used in
full storage operation if dehumidification requirements are low.
In summary, a CTES system can benefit users in three ways:

• Lower electricity rates: With CTES, chillers can operate at night to meet the
daytime cooling needs, taking advantage of lower off-peak electricity
consumption rates.
• Lower demand charges: Many commercial customers pay a monthly
electrical demand charge based on the largest amount of electricity used
6.5 Heat Storage Systems For Cooling Applications 259

during any 30 min period of the month. CTES reduces peak demands, by
shifting those demands to off-peak periods. Furthermore, some utilities
provide a rebate for shifting electrical demand to nighttime or other off-peak
periods.
• Lower air conditioning system and compressor costs: Without CTES, large
compressors capable of meeting peak cooling demands are needed,
whereas smaller and less expensive units are sufficient when CTES is
used. Also, since water from a CTES may be colder than conventional
chilled water, smaller pipes, pumps, and air handlers may be integrated
into the building design to reduce costs further.
Another example of a heat storage system for cooling applications is the use of
thermal storage to take advantage of off-peak electricity tariffs. Chiller units can
be run at night when the cost of electricity is relatively low. These units are used to
cool down a thermal storage, which then provides cooling for air conditioning
throughout the day. Not only electricity costs are reduced, but also the efficiency
of the chiller is increased because of the lower nighttime ambient temperatures,
and the peak electricity demand is reduced for electrical supply utilities.

6.5.2 Thermodynamic Analysis of Heat Storage Systems


for Cooling Applications
Exergy analysis in some ways complements energy analysis for heat storage sys-
tem performance evaluation. For assessing cold TES, it is more evident that
exergy methods are superior, whether they are applied to the overall heat stor-
age process or its charging, storing, and discharging subprocesses. This may or
may not be clear from the energy and exergy balances as well as efficiencies pre-
sented in Table 6.2 and in the literature (Rosen, 1992; Dincer et al., 1997). The
exergy efficiency incorporates the notion of increasing thermodynamic unavail-
ability, as reflected by increasing entropy, in a system or process. This key dif-
ference between exergy and energy efficiencies occurs because the former
accounts for the losses due to irreversibilities, in addition to losses associated
with waste emissions. Many have recognized that this difference is particularly
important when assessing CTES and applied exergy analysis to it (Krane, 1987;
Rosen, 1992; Bejan, 1994; Dincer et al., 1997; Domanski and Fellah, 1998).
The three main processes in a heat storage for cooling capacity are illustrated in
Fig. 6.7. A thermodynamic analysis is described for the charging, storing, and
discharging processes of the heat storage system and the overall cycle.

6.5.2.1 Charging Process


During the charging period, the total “cold” (charge), Qc, provided to the heat
storage system can be expressed as
Qc ¼ mc cp ðTc, o  Tc, i Þ (6.37)
260 C HA PT E R 6 : Heat Storage Systems

Charging Storing Discharging


b c

a d

Ql

Time

FIGURE 6.7 Three processes in a general heat storage system for cooling capacity: charging (left),
storing (middle), and discharging (right). The heat leakage into the system Ql is illustrated for the
storing process but can occur in all three processes. Modified from Dincer and Rosen (2011).

where cp, Tc,i, and Tc,o denote specific heat capacity, and charging inlet temper-
ature, and charging outlet temperature of the HTF, respectively. The term mc
represents the total mass accumulated and/or transported over the charging
period, tc (s), and can be written as
ð tc
mc ¼ m_ ðtc Þ dtc (6.38)
0

For a constant mass flow rate, ṁ(tc) can be denoted as ṁ (kg/s) and the above
equation can be rewritten as
_ c
mc ¼ mt (6.39)

Total exergy charged into the heat storage system during charging is given by
  
Tc, o
ExQ, c ¼ mc cp ðTc, o  Tc, i Þ  T0 ln (6.40)
Tc, i

6.5.2.2 Storing Process


During the storing process, the heat storage system (for cooling applications)
gains heat (or loses “cold”) due to energy interactions between the system and
its surrounding. This interaction is mainly through heat transfer along the sys-
tem boundaries since the storage temperature (Ts) is lower than the tempera-
ture of the surroundings (T0).This heat gain (or cold loss) rate (Q_ l ) can be
assessed as

Q_ l ¼ UAðT0  Ts Þ (6.41)

where U and A represent overall heat transfer coefficient and area, respectively.
6.5 Heat Storage Systems For Cooling Applications 261

The overall heat loss over the entire storing time is given by
Ql ¼ ms cp ðΔTs Þ (6.42)

where mc, cp, and ΔTs denote mass, specific heat capacity, and the difference
between initial (Ts,i) and final (Ts,f) temperatures of the heat storage medium,
respectively. The term ΔTs is calculated as
ΔTs ¼ Ts, f  Ts, i (6.43)

The total exergy loss from the heat storage system during the storing process is
given by
 
Ts
ExQ, l ¼ Ql 1  (6.44)
T0

6.5.2.3 Discharging Process


During the discharging period, the total cooling output (discharge), Qd, pro-
vided by the heat storage system can be written as
Qd ¼ md cp ðTd, i  Td, o Þ (6.45)

where cp, Td,i, and Td,o denote specific heat capacity, the discharging inlet tem-
perature, and discharging outlet temperature of the HTF, respectively. The term
md represents the total mass accumulated and/or transported over the dischar-
ging period, td (s), and can be written as
ð td
md ¼ m_ ðtd Þ dtd (6.46)
0

For a constant mass flow rate, ṁ(td) can be denoted as ṁ (kg/s) and the above
equation can be rewritten as
_ d
md ¼ mt (6.47)

The total exergy discharged from the heat storage system during the discharging
process is given by
  
Td, i
ExQ, d ¼ md cp ðTd, i  Td, o Þ  T0 ln (6.48)
Td, o

6.5.2.4 Energy and Exergy Efficiencies


The energy efficiency of heat storage system can be evaluated over the entire
operating cycle (charging, storing, and discharging) as the ratio of the total
262 C HA PT E R 6 : Heat Storage Systems

cooling recovered from the system during the discharging process (Qd) to the
total cooling charged into the system (Qc). This can be expressed as
Qd
η¼ (6.49)
Qc

Similarly, the exergy efficiency of the system can be evaluated for the overall
storage period as
ExQ, d
ψ¼ (6.50)
ExQ, c

6.5.3 Illustrative Example


Energy and exergy analyses of four CTESs are performed utilizing Eqs. (6.37)–
(6.50). In each case, the CTES has identical initial and final states, so that the
CTES operates in a cyclic manner, continuously charging, storing, and dischar-
ging. The cold storage cases, differentiated by their main characteristics, are as
follows:

• Sensible heat storage, with a fully mixed storage fluid


• Sensible heat storage, with a linearly stratified storage fluid
• Latent heat storage, with a fully mixed storage fluid
• Combined latent and sensible heat storage, with a fully mixed
storage fluid

6.5.3.1 Assumptions and Specified Data


The following assumptions are made for each of the cases:

• Storage boundaries are nonadiabatic.


• Heat gain from the environment during charging and discharging is
negligibly small relative to heat gain during the storing period.
• The external surface of the storage tank wall is at a temperature 2 °C greater
than the mean storage-fluid temperature.
• The mass flow rate of the HTF is controlled so as to produce constant inlet
and outlet temperatures.
• Work interactions, and changes in kinetic and potential energy terms, are
negligibly small.

Specified data for the four cases are presented in Table 6.10 and relate to the
diagram in Fig. 6.10. In Table 6.10, Tb and Td are the charging and discharging
outlet temperatures of the HTF, respectively. The subscripts 1, 2, and 3 indicate
the temperature of the storage fluid at the beginning of charging, storing, or dis-
charging, respectively. Also l indicates the liquid state and s indicates the solid
state for the storage fluid at the phase change temperature.
6.5 Heat Storage Systems For Cooling Applications 263

Table 6.10 Specified Temperature Data for the Cases in the CTES Example
System

Temperature (°C) I II III IV

Tb 4 15 1 1
Td 11 11 10 10
T1 10.5 19/2a 0 (l) 8
T2 5 17/7a 0 (s) 8
T3 6 18/6a 0 (l&s) 0 (l&s)
a
When two values are given, the storage fluid is vertically linearly stratified and the first and second
values are the temperatures at the top and bottom of the storage fluid, respectively.

In addition, for all cases, the inlet temperatures are fixed for the charging-fluid
flow (Ta) as 10 °C and for the discharging-fluid flow (Tc) as 20 °C. For cases
involving latent heat changes (i.e., solidification), F ¼ 10%. The specific heat cp
is 4.18 kJ/kg K for both the storage and HTFs. The phase change temperature of
the storage fluid is 0 °C. The configuration of the storage tank is cylindrical with
an internal diameter of 2 m and internal height of 5 m. Environmental condi-
tions (T0 and P0) are 20 °C and 1 atm.

6.5.3.2 Results and Discussion


The results for the four systems are listed in Table 6.11 and include overall and
subprocess efficiencies, input and recovered cold quantities, and energy and
exergy losses. The overall and subprocess energy efficiencies are identical for sys-
tems I and II and for systems III and IV. In all systems, the energy efficiency values
are high. The different and lower exergy efficiencies for all cases indicate that

Table 6.11 Energy and Exergy Quantities for the Cases in the CTES Example
Energy Quantities Exergy Quantities
Period or Quantity I II III IV I II III IV

Efficiencies (%)
Charging (1) 100 100 100 100 51 98 76 77
Storing (2) 82 82 90 90 78 85 90 85
Discharging (3) 100 100 100 100 38 24 41 25
Overall 82 82 90 90 15 20 28 17

Input recovered and lost quantities (MJ)


Input 361.0 361.1 5237.5 6025.9 31 23 500 575
Recovered 296.0 295.5 4713.8 5423.3 4.6 4.6 142 94.7
Loss (external) 65.7 65.7 523.8 602.6 2.9 2.9 36.3 48.9
Loss (internal) – – – – 23 16 321 431
264 C HA PT E R 6 : Heat Storage Systems

energy analysis does not account for the quality of the “cold” energy, as related to
temperature, and considers only the quantity of “cold” energy recovered.
The input and recovered quantities in Table 6.11 indicate the quantity of “cold”
energy and exergy input to and recovered from the storage. The energy values
are much greater than the exergy values because, although the energy quantities
involved are large, the energy is transferred at temperatures only slightly below
the reference-environment temperature and therefore is of limited usefulness.
The cold losses during storage, on an energy basis, are entirely due to cold losses
across the storage boundary (i.e., heat infiltration). The exergy-based cold
losses during storage are due to both cold losses and internal exergy losses
(i.e., exergy consumptions due to irreversibilities within the storage). For the
present systems, in which the exterior surface of the storage tank is assumed
to be 2 °C warmer than the mean storage-fluid temperature, the exergy losses
include both external and internal components. Alternatively, if the heat trans-
fer temperature at the storage tank external surface is at the environment tem-
perature, the external exergy losses would be zero and the total exergy losses
would be entirely due to internal consumptions. If heat transfer occurs at
the storage-fluid temperature, on the other hand, more of the exergy losses
would be due to external losses. In all cases, the total exergy losses, which
are the sum of the internal and external exergy losses, remain fixed.
The four systems demonstrate that energy and exergy analyses give different
insights for CTES systems. Both energy and exergy analyses account for the
quantity of energy transferred in storage processes. Exergy analyses take into
account the loss in quality of “cold” energy and thus more correctly reflect
the actual value of the CTES.
In addition, exergy analysis is conceptually more direct when applied to CTES
systems because cold is treated as a useful commodity. With energy analysis,
flows of heat rather than cold are normally considered. Thus, energy analyses
become convoluted and confusing as one must deal with heat flows while
accounting for the fact that cold is the useful input and product recovered
for CTES systems. Exergy analysis inherently treats any quantity that is out of
equilibrium with the environment (be it colder or hotter) as a valuable com-
modity and thus avoids the intuitive conflict in the expressions associated with
CTES energy analysis. The concept that cold is a valuable commodity is both
logical and in line with intuition when applied to CTES systems.

6.6 CASE STUDIES


Thousands of heat storage systems have been operating for years in the world,
particularly in developed countries; in hospitals, schools, universities, airports,
6.6 Case Studies 265

government facilities, and private office buildings; and in industrial process


cooling applications. Described below are several reported case studies by
the IEA (1994) and ARI (1997) that demonstrate how heat storage systems pro-
vide energy savings and reduce the environmental impact and that illustrate
some clever applications of heat storage system equipment in new buildings
to reduce initial costs.

6.6.1 Microscale Application: Advanced Heat Storage


System through PCMs
Most of the laptop computers are cooled by a fan operating at high temperatures
of the central process unit (CPU). If the CPU of the computer becomes too warm,
its capacity declines. A sheet of PCM placed under the computer can help in cool-
ing the computer. The “storage” is regenerated on shutdown of the computer
since the room temperature is lower than the melting point of the PCM. The
IEA (2015) reports the following details for this system when the PCM is applied:
• The time before the fan starts is increased from 20 min from start of the
computer to 4 h.
• Energy consumption is reduced by 25%, allowing longer battery
operation of the computer.
• The lifetime of the computer (the CPU) is believed to increase due to
lower operating temperatures.

6.6.2 Anova Verzekering Co. Building (The Netherlands)


In this application, a groundwater aquifer heat storage system was installed as
part of a space-conditioning unit in a renovated office building of Anova
Verzekering Co. This building application of heat storage reduced energy use
and pollutant emissions (Table 6.12). An electric heat pump supplies hydronic
heating and cooling. Accounting for the subsidy of USD 212,000 received from
the Dutch government, which is equivalent to 20% of the total initial system
costs, the reduced energy costs due to heat storage system were predicted to lead
to a payback period within of 6.5 years for the additional investment costs due
to the heat storage system. In the case study, primary energy consumption
decreases due to the heat storage system by over 40%, even though electricity
use increases, and emissions decrease by about 40%.

6.6.3 Kraft General Foods Headquarters Building


(Northfield, Illinois)
All daytime air conditioning loads are being met at this facility by melting ice
that is made and stored overnight. It was anticipated that additional loads from
future expansion could be met by operating some of the chillers during the day
as well as at night. The building was designed to pump 2.2 °C water to the air
266 C HA PT E R 6 : Heat Storage Systems

Table 6.12 Annual Energy Savings and Emission Reductions for the Case Studya
Reduction for Heat Storage
System

Commodity Conventional System Heat Storage System Amount %

Consumptions
Natural gas (m3) 215,800 95,500 120,300 56
Electricity (kWh) 395,550 511,500 84,000 21
Primary energyb 322,000 179,000 143,000 44

Emissions
CO2 (kg) 608,000 346,000 262,000 43
NOx – – – 40
SO2 – – – 40
a
For further details: IEA (1994).
b
Primary energy is calculated as the equivalent amount of natural gas based on the assumption that 0.25 m3 gas is used in the
generation of 1 kWh electricity.
Source: Dincer and Rosen (2001).

handling units, which in turn provide 7.2 °C air to the building. These temper-
atures, which are lower than for nonheat storage-based systems, permit the use
of smaller pipes, pumps, air handling units, and ductwork, resulting in lower
initial capital costs for the system. Annual electric bills for this building are
nearly USD 200,000 lower than for an almost identical building, just 5 km
away, which does not use a heat storage system.

6.6.4 Chrysler’s New Technology Development Center


(Auburn Hills, Michigan)
Since opening in 1990, Chrysler’s technology development center has achieved
both equipment and operating cost savings by using a 68,000 ton/h chilled-
water heat storage system. The heat storage capacity allowed the center’s chiller
plant to be downsized from 17,710 tons, which would have been needed to
meet peak cooling loads, to 11,385 tons. Chilled water is stored in the heat stor-
age system at night and supplements chiller operation during peak cooling con-
ditions the following day. Reduced chiller costs more than offset the cost of the
heat storage system installation, resulting in initial savings of USD 3.6 million.
In addition, the heat storage system shifts over 5000 kW of peak electrical
demand to off-peak periods, saving over USD 1 million/year.

6.6.5 San Francisco Marriott Hotel (San Francisco,


California)
Using a heat storage system in tandem with a real-time pricing strategy from the
local utility, this hotel predicted savings of USD 135,000 in annual cooling
6.7 Detailed Illustrative Example: Exergy Analysis Of Uoit BTES 267

costs. Only 1800 ton/h of ice storage is needed, enough to satisfy the 450 ton
cooling load during the daily peak-rate time period, which lasts only 2-3 h
under the real-time pricing schedule. Over one-third of the installed cost of
the heat storage system will be covered by a rebate from the utility; the remain-
ing amount is expected to be recouped in less than two years of operation.

6.6.6 Texas A&M University (Corpus Christi, Texas)


In August 1997, the Central Power and Light (CPL) Co. presented Texas A&M
University Corpus Christi with a USD 431,800 incentive award for the univer-
sity’s participation in a heat storage system program designed for substantial
energy savings. The university system uses the bulk of its electricity during off-
peak evening hours, allowing CPL to shift some of its electric load from peak
usage times and to share the annual cost savings of up to USD 150,000 with
the university. Texas A&M University Corpus Christi invested approximately
USD 900,000 to purchase and install the heat storage system equipment, which
became operational in January 1995, and predicted recovery of that cost through
energy savings within 5 years. The USD 431,800 incentive includes USD 20,400
for installing high-efficiency equipment for cooling and heating the campus. The
remaining incentive was provided through CPL’s energy efficiency program,
wherein the university was offered USD 200 for every kW of electricity load shift
from CPL’s peak daytime load to off-peak evening hours. The university reduced
electricity peak demand by approximately 2057 kW, earning a USD 411,400
incentive. CPL worked with the university to install an 11,800 ton/h thermal
storage system with a water storage capacity of 5300 m3. The university also
installed a 500 ton industrial heat pump for heating, which CPL estimates will
save the university approximately USD 90,000 a year in energy costs. The heat
pump captures waste heat from the university’s 3000 ton chiller plant and recir-
culates it into areas of the campus needing heating. The heat storage system tank
stores chilled water, which is produced by the conventional air conditioning sys-
tem during the night and is then used to cool buildings during the day, when the
highest demand is placed on the air conditioning system.
These case studies and many others demonstrate that heat storage technology
offers compelling energy, environmental, diversity, and economic benefits to
owners and regions. As heat storage systems is increasingly commercialized,
policies should be considered that can help increase the market penetration
of TES beneficially.

6.7 DETAILED ILLUSTRATIVE EXAMPLE: EXERGY


ANALYSIS OF UOIT BTES
The campus considered in this exergy assessment is the UOIT, in Oshawa,
Ontario, Canada. The campus includes seven buildings, most of which are
268 C HA PT E R 6 : Heat Storage Systems

designed to be heated and cooled using GSHPs in conjunction with a BTES,


with the aim of reducing energy-resource use, environmental emissions, and
financial costs. The analyses in this case study consider 10 buildings since
the whole system was designed for 10 buildings. This university BTES field
is the largest and deepest in Canada, and the geothermal well field is one of
the largest in North America (Dincer and Rosen, 2011). The system in cooling
mode is illustrated in Fig. 6.8.
The total cooling load of the campus buildings is about 7000 kW. Test drilling
programs were carried out to determine the feasibility of thermal storage in the
overburden and bedrock formations at the site. Using the thermal conductivity
test results, it was determined that a field of 370 boreholes, each 200 m in
depth, would be required to meet the energy demand. The Swedish practice
of water-filled borehole heat exchangers (BHEs) was utilized instead of the
North American practice of grouted BHEs.

A1 A2 A3 A4 A5 A6 A7 A8 A9 A10

21 24 27 30 33 36 39 42 45 48

20 22 23 25 26 28 29 31 32 34 35 37 38 40 41 43 44 46 47 49

5
1

13 14 15 16 17 18

1 4 5 8 9 12
Chillers Heat pumps 1 Heat pumps 2
7×240 kW 7×176 kW 7×176 kW
2 3 6 7 10 11

52 51 54 53 56 55

62 68 74
60 61 66 67 72 73

57 Cooling 63 Cooling 69 Cooling


tower 1 tower 1 tower 1

58 59 64 65 70 71

Borehole heat exchangers


370 ×200 m

FIGURE 6.8 Flow diagram of the GSHP/BTES system at the UOIT, in the cooling mode. Modified from
Kizilkan and Dincer (2012).
6.7 Detailed Illustrative Example: Exergy Analysis Of Uoit BTES 269

The university’s central plant includes a cooling and heating system for the cam-
pus, utilizing the BTES. Chillers are used to provide energy and pumps convey
the working fluid between the buildings and the BTES. Additional heat pump
modules assist in cooling. Chilled water is supplied from two multistack
chillers, each having seven modules, and two sets of heat pumps each with
seven modules. Chillers are variable displacement centrifugal units with mag-
netic bearings that allow for excellent part-load performance. The condenser
water enters the borehole field, which retains the heat from the condensers
for use in the winter (when the heat pumps reverse) and provides low-
temperature hot water for the campus (Dincer and Rosen, 2011).
A glycol solution, encased in polyethylene tubing, circulates through an inter-
connected, underground network. A 15% glycol solution is the fluid that is cir-
culated through the BHE mounted in the ground. Inlet and outlet temperatures
of the solution to and from the ground are 29.4 and 35 °C, respectively. The
glycol solution concentration is 30% and is circulated between the system
and buildings to transfer heat. Inlet and outlet temperatures of the solution
to and from the fan coils are 5.5 and 14.4 °C, respectively. During the winter,
the fluid, circulating through tubing extended into the wells, collects heat from
the earth and transports it into the buildings. In summer, the system reverses to
extracts heat from the buildings and transmits it to the ground.

6.7.1 Analysis Assumptions


In the analysis, the refrigerant R507A is taken to be used in the chiller system
and R407C in the heat pump system. Heat transfer to the system and work trans-
fer from the system are taken to be positive, and the reference-environment
temperature and pressure are taken to be 24 °C and 98.825 kPa, respectively.
It is assumed that processes are at steady-state and steady-flow conditions; the
isentropic efficiencies of the chiller and heat pumps are 85% and 80%, respec-
tively; the compressor is adiabatic and has mechanical and electrical efficiencies
of 80% and 84%, respectively; the circulating pumps have mechanical and elec-
trical efficiencies of 85% and 88%, respectively; power inputs to the fan coils and
potential and kinetic energy effects are negligible; and there are no chemical or
nuclear reactions.

6.7.2 Exergy Balances and Efficiencies


Exergy balances, which allow exergy destruction rates and exergy efficiencies to
be determined, for the main components of the GSHP/BTES system in Fig. 6.8,
that is, the compressor (comp), condenser (cond), expansion valve (ev), evap-
orator (e), fan coil (fc), borehole heat exchanger (bhe), and circulating pump
(p), respectively, are as follows:
Ex _ comp, in  Ex
_ dest, comp ¼ Ex _ comp, out + W_ comp (6.51)
270 C HA PT E R 6 : Heat Storage Systems

   
Ex _ cond, in  Ex
_ dest, cond ¼ Ex _ cond, out + Ex _ hw, out
_ hw, in  Ex (6.52)

Ex _ ev, in  Ex
_ dest, ev ¼ Ex _ ev, out (6.53)
   
Ex _ e, in  Ex
_ dest, e ¼ Ex _ e, out + Ex _ cw, out
_ cw, in  Ex (6.54)
 
 
_ dest,fc ¼ Ex
Ex _ cw, out + Q_ fc 1  T0
_ cw, in  Ex (6.55)
Tfc
 
 
_ dest, bhe ¼ Ex
Ex _ hw, out + Q_ bhe 1  T0
_ hw, in  Ex (6.56)
Tbhe

Ex _ p, in  Ex
_ dest, p ¼ Ex _ p, out + W_ p (6.57)

Corresponding exergy efficiencies for the BTES system components are as


follows:

_ comp, out  Ex
Ex _ comp, in
ψ comp ¼ (6.58)
W_ comp
 
_ hw, out  Ex
Ex _ hw, in
ψ cond ¼   (6.59)
_ cond, in  Ex
Ex _ cond, out

_ ev, out
Ex
ψ ev ¼ (6.60)
Ex_ ev, in

 
_ cw, in  Ex
Ex _ cw, out
ψe ¼   (6.61)
_ e, out  Ex
Ex _ e, in

 
_ cw, in  Ex
Ex _ cw, out
ψ fc ¼   (6.62)
_ T0
Qfc 1 
Tfc
 
_ hw, in  Ex
Ex _ hw, out
ψ bhe ¼   (6.63)
T0
Q_ bhe 1 
Tbhe

_ p, out  Ex
Ex _ p, in
ψ dest, p ¼ (6.64)
W_ p

For the overall system, the exergy efficiency can be written as


X _
_ out
Ex _ Q
Ex output
ψs ¼ ¼X X (6.65)
Ex_ in W_ comp + W_ p
6.7 Detailed Illustrative Example: Exergy Analysis Of Uoit BTES 271

6.7.3 Performance Assessment


The performance of the BTES system is described here, based on an earlier anal-
ysis (Kizilkan and Dincer, 2012). In that work, the university data, temperature,
mass flow rate, and exergy data for R507A, R407C, and the glycol solution are
given, for the state points in Fig. 6.8.
The exergy destruction rate, the relative irreversibility (RI), and the exergy effi-
ciency are listed in Table 6.13 for the overall system and each of its components.
The exergy efficiency for the BTES system on a product/fuel basis is 62% and the
overall exergy destruction rate is 1346 kW. The greatest exergy destruction rates
in the BTES system occur in the compressors of the chiller and heat pumps, fol-
lowed by the condenser, the expansion valve, and the evaporator. The data in
Table 6.13 are visually presented in terms of RI and exergy efficiency (ψ) in
Figs. 6.9 and 6.10, for the compressors, condensers, expansion valves, evapo-
rators, fan coils, BHEs, cooling towers, and pumps.

Table 6.13 Exergy-Based Performance Parameters for the GSHP/BTES System and Its Components
Ėxdest Ėxdest
Component (kW) RI (%) ψ (%) Component (kW) RI (%) ψ (%)

Compressor 1 78.28 5.816 86.62 BHE 1 27.31 2.029 41.86


Compressor 2 56.64 4.208 87.33 BHE 2 20.24 1.504 41.86
Compressor 3 56.64 4.208 87.33 BHE 3 20.24 1.504 41.86
Condenser 1 126.6 9.402 29.68 Cooling tower 1 28.15 2.091 54.58
Condenser 2 103.4 7.684 27.68 Cooling tower 2 20.86 1.55 54.57
Condenser 3 103.4 7.684 27.68 Cooling tower 3 20.86 1.55 54.57
Expansion valve 1 146.8 10.91 86.06 Fan coil pump A1 0.747 0.055 6.97
Expansion valve 2 81.96 6.089 86.05 Fan coil pump A2 1.301 0.096 6.97
Expansion valve 3 81.96 6.089 86.05 Fan coil pump A3 1.827 0.135 6.97
Evaporator 1 90.70 6.738 49.66 Fan coil pump A4 1.218 0.09 6.97
Evaporator 2 99.75 7.41 39.68 Fan coil pump A5 1.467 0.109 6.97
Evaporator 3 99.75 7.41 39.68 Fan coil pump A6 0.883 0.065 6.97
Fan coil A1 7.209 0.535 69.28 Fan coil pump A7 0.457 0.033 6.97
Fan coil A2 7.213 0.535 69.28 Fan coil pump A8 0.866 0.064 6.97
Fan coil A3 6.931 0.514 69.28 Fan coil pump A9 0.468 0.034 6.97
Fan coil A4 7.209 0.535 69.28 Fan coil pump A10 0.627 0.046 6.97
Fan coil A5 7.209 0.535 69.28 BHE pump 1 0.034 0.002 3.25
Fan coil A6 6.119 0.454 69.28 BHE pump 2 0.03 0.002 3.25
Fan coil A7 5.005 0.371 69.28 BHE pump 3 0.03 0.002 3.25
Fan coil A8 13.98 1.039 69.28 Cooling tower pump 1 2.537 0.188 3.25
Fan coil A9 3.389 0.251 69.28 Cooling tower pump 2 1.092 0.081 3.25
Fan coil A10 3.532 0.262 69.28 Cooling tower pump 3 1.092 0.081 3.25
Overall 1346.15 – 62.07
272 C HA PT E R 6 : Heat Storage Systems

Cooling towers Fan coil pumps


5% 1%
BHEs
5%
Compressors
Fan coils 14%
5%

Evaporators Condensers
22% 25%

Expansion
valves
23%

FIGURE 6.9 Relative irreversibilities of overall sets of condensers, expansion valves, evaporators,
compressors, fan coils, BHEs, cooling towers, and pumps for the system in Fig. 6.8.

100
90
80
Exergy efficiency (%)

70
60
50
40
30
20
10
0

FIGURE 6.10 Average exergy efficiencies of condensers, expansion valves, evaporators, compressors,
fan coils, BHEs, cooling towers, and pumps and the exergy efficiency for the overall system in Fig. 6.8.

The inlet glycol solution temperature is an important and representative param-


eter of the BTES system. The glycol solution temperature entering the refriger-
ation system (i.e., the condenser) is higher than the ground temperature in
summer conditions, because of heat rejection from the circulating glycol solu-
tion to the ground. It is shown in Fig. 6.11 that exergy destruction rate increases
as the entering glycol solution temperature decreases.
The overall exergy destruction rate of the system is shown in Fig. 6.12 as evap-
orator temperature varies and in Fig. 6.13 as condenser temperature varies. The
overall exergy destruction rate decreases and the overall exergy efficiency
increases with increasing of evaporator temperature and decreasing of con-
denser temperature. The variations are almost linear.
2200
0.76
2000
0.72
1800
Exdest,tot (kW)
0.68
1600 esys

esys
0.64
1400
Exdest,tot 0.6
1200

1000 0.56

800 0.52
28 29 30 31 32 33 34
TBW,in (°C)

FIGURE 6.11 Variation of exergy destruction rate and exergy efficiency with inlet glycol solution
temperature (Dincer and Rosen, 2013).

1400 0.67

Exdest,tot
1300 0.66
Exdest,tot (kW)

esys
1200 0.65

1100 0.64
esys

1000 0.63
−6 −4 −2 0 2 4 6
TE (°C)

FIGURE 6.12 Variation of exergy destruction rate and exergy efficiency with evaporator temperature
(Dincer and Rosen, 2013).

3000 0.66

2600
0.64
Exdest,tot (kW)

2200
esys

0.62
1800

Exdest,tot esys
0.6
1400

1000 0.58
40 45 50 55 60 65 70 75
TC (°C)

FIGURE 6.13 Variation of exergy destruction rate and exergy efficiency with condenser temperature
(Dincer and Rosen, 2013).
274 C HA PT E R 6 : Heat Storage Systems

The results of the GSHP/BTES exergy analysis are somewhat sensitive to varia-
tions in reference-environment properties and in glycol solution concentration,
as shown in previous analyses (Kizilkan and Dincer, 2012).

6.8 CLOSING REMARKS


Heat storage is an advanced energy technology. Heat storage has been attracting
increasing interest in several applications, for example, active and passive solar
heating, water heating, cooling, and air conditioning. Heat storage systems
often are the most economical energy storage technology for building heating,
cooling, and air conditioning applications.
Some key points raised in this chapter can be summarized as follows:

• The selection of the heat storage systems mainly depends on the storage
period required (e.g., diurnal, weekly, or seasonal), economic viability,
operating conditions, etc. Several parameters influence the viability of any
heat storage system, for example, facility thermal loads, thermal and
electrical load profiles, availability of waste or excess thermal energy,
electrical costs and rate structures, type of thermal generating equipment,
and end use type and demand.
• Heat storage systems can play a significant role in meeting society’s needs
for more efficient, environmentally benign energy use in various sectors
and are an important technology for addressing mismatches between
times of energy supply and demand.
• Using heat storage systems can lead to substantial energy savings (up to
50% when implemented with appropriate demand-side management
strategies) and emission reductions of greenhouse gases like CO2, SO2,
and NOx (about 40%).
• Substantial energy savings can be realized by heat storage systems when
implementing the techniques such as using waste energy and surplus heat,
reducing electrical demand charges, and avoiding heating, cooling, or air
conditioning equipment purchases.
• For design, performance evaluation, and optimization of heat storage
systems, both energy and exergy analyses should be considered. But exergy
analysis should be preferred as it provides an effective method that
integrates the conservation of mass and conservation of energy principles
together with the SLT.
• Heat storage systems are being increasingly applied and continue to attract
new interest, for a range of applications, for example, active and passive
solar heating, water heating, cooling, and air conditioning. Also, heat
storage systems provide an advanced energy technology for building
heating and cooling applications, which sometimes constitutes the most
economical storage technology.
6.8 Closing Remarks 275

Nomenclature
A surface area
cp specific heat capacity
e specific energy
E energy
ex specific exergy
Ex exergy
f energy fraction recovery
F fraction of storage-fluid mass in liquid phase
h specific enthalpy
H enthalpy
m mass
P pressure
Q heat
t time
T temperature
U overall heat transfer coefficient
W work

Greek symbols
η energy efficiency
θ time
ρ density
ψ exergy efficiency

Subscripts
0 environmental state
bhe borehole heat exchanger
c charging
comp compressor
cond condenser
cw cooling water
d discharging
dest destruction
e evaporator
ev expansion valve
f final
fc fan coil
gen generation
hw heating water
i initial, inlet, input
l loss (leakage)
o outlet, output
p pump
Q heat
R reference
s storing
W work
276 C HA PT E R 6 : Heat Storage Systems

Superscripts
˙ rate with respect to time
0
modified
R reference

Acronyms
ARI Air Conditioning and Refrigeration Institute
ATES aquifer thermal energy storage
BHE borehole heat exchanger
BTES borehole thermal energy storage
CTES cold (cool) thermal energy storage
CFC chlorofluorocarbon
CPU central process unit
CSP concentrated solar power
EPRI Electric Power Research Institute
ETS electrical thermal storage
FLT first law of thermodynamics
GSHP ground source heat pump
HCFC hydrochlorofluorocarbon
HTF heat transfer fluid
HVAC heating, ventilation, and air conditioning
IEA International Energy Agency
PCM phase change material
RI relative irreversibility
SLT second law of thermodynamics
TES thermal energy storage
UOIT University of Ontario Institute of Technology
UTES underground thermal energy storage

References
ARI, 1997. Thermal Energy Storage: A Solution for Our Energy, Environmental and Economic Chal-
lenges. The Air-Conditioning and Refrigeration Institute, Arlington, VA.
Beggs, C.B., 1994. Ice thermal storage: impact on United Kingdom carbon dioxide emissions. Build.
Serv. Eng. Res. Technol. 15, 756–763.
Bejan, A., 1994. Entropy Generation through Heat and Fluid Flow. Wiley and Sons, New York, NY.
California Energy Commission, 1996. Source energy and environmental impacts of thermal energy
storage. Technical report no. P500-95-005. California Energy Commission, California.
Chwieduk, D.A., 2012. Solar-Assisted Heat Pumps. Comprehensive Renewable Energy. Elsevier,
Poland.
Diamant, R.M.E., 1984. Energy-Conservation Equipment. The Architectural Press, London.
Dincer, I., 1998. Evaluation and selection of energy storage systems for solar thermal applications.
In: Dincer, I., Ayhan, T. (Eds.), Proceedings of the TIEES-98, Second Trabzon International
Energy and Environment Symposium. Begell House, pp. 168–172.
References 277

Dincer, I., 2002. On thermal energy storage systems and applications in buildings. Energy Build.
34, 377–388.
Dincer, I., Dost, S., 1996. A perspective on thermal energy storage systems for solar energy appli-
cations. Int. J. Energy Res. 20, 547–557.
Dincer, I., Rosen, M.A., 2001. Energetic, environmental, and economic aspects of thermal energy
storage systems for cooling capacity. Appl. Therm. Eng. 21, 1105–1117.
Dincer, I., Rosen, M.A., 2011. Thermal Energy Storage: Systems and Applications, second ed. John
Wiley and Sons, New York.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, 2nd ed.
Elsevier, Oxford.
Dincer, I., Dost, S., Li, X., 1997. Performance analyses of sensible heat storage systems for thermal
applications. Int. J. Energy Res. 21, 1157–1171.
Domanski, R., Fellah, G., 1998. Thermoeconomic analysis of sensible heat, thermal energy storage
systems. Appl. Therm. Eng. 18, 693–704.
Haeseldonckx, D., Peeters, L., Helsen, L., D’haeseleer, W., 2007. The impact of thermal storage on
the operational behavior of residential CHP facilities and the overall CO2 emissions. Renew.
Sust. Energ. Rev. 11, 1227–1243.
Hahne, E., 2000. The ITW solar heating system: an old timer fully in action. Sol. Energy
69, 469–493.
Hauer, A., Quinnell, J., Lavemann, E., 2013. Energy Storage Technologies—Characteristics, Com-
parison, and Synergies. Wiley-VCH Verlag GmbH & Co., Weinheim, Germany.
Hendricks, M., Snijders, M., Boid, N., 2008. Underground thermal energy storage for efficient heat-
ing and cooling of buildings. In: 1st International Conference on Industrialised, Integrated,
Intelligent Construction, Loughborough.
Hoyer, M.C., Walton, M., Kanivetsky, R., Holm, T.R., 1985. Short-term aquifer thermal energy
storage (ATES) test cycles, St. Paul, Minnesota, USA. In: Proceedings of 3rd International
Conference on Energy Storage for Building Heating and Cooling, Toronto, Canada,
pp. 75–79.
International Energy Agency, 1994. Energy storage. Heat Pump Cen. Newsl. 12, 8.
International Energy Agency, 2011. Technology roadmap: energy efficient buildings: heating and
cooling equipment. http://www.iea.org/publications/freepublications/publication/buildings_
roadmap.pdf.
International Energy Agency, 2015. Technology Roadmap Energy Storage. http://www.iea.org/
publications/freepublications/publication/TechnologyRoadmapEnergystorage.pdf.
Jenne, E.A., 1992. Aquifer Thermal Energy (Heat and Chill) Storage. Pacific Northwest Lab,
Richland, WA.
Kizilkan, O., Dincer, I., 2012. Exergy analysis of borehole thermal energy storage system for build-
ing cooling applications. Energy Build. 49, 568–574.
Krane, R.J., 1987. A second law analysis of the optimum design and operation of thermal energy
storage systems. Int. J. Heat Mass Transf. 30, 43–57.
Krane, R.J., Krane, M.J.M., 1992. The optimum design of stratified thermal energy storage systems.
Part I. Development of the basic analytical model. ASME J. Energy Resour. Technol.
114, 197–203.
Norton, B., 1992. Solar Energy Thermal Technology. Springer, London.
Reindl, D.T., 1994. Characterizing the marginal basis source energy emissions associated with com-
fort cooling systems. Report no. TSARC 94-1, Thermal Storage Applications Research Center,
USA.
278 C HA PT E R 6 : Heat Storage Systems

Rosen, M.A., 1992. Appropriate thermodynamic performance measures for closed systems for ther-
mal energy storage. ASME J. Energy Resour. Technol. 114, 100–105.
Rosen, M.A., 1999. Second-law analysis of aquifer thermal energy systems. Energy 24, 167–182.
Rosen, M.A., Hooper, F.C., 1996. Second law analysis of thermal energy storage systems.
In: Proceedings of the TIEES-96, First Trabzon International Energy and Environment Sympo-
sium. Karadeniz Technical University, Trabzon, Turkey, pp. 373–379.
Rosen, M.A., Hooper, F.C., Barbaris, L.N., 1988. Exergy analysis for the evaluation of the perfor-
mance of closed thermal energy storage systems. ASME J. Energy Resour. Technol.
110, 255–261.
Taylor, M.J., Krane, R.J., Parsons, J.R., 1991. Second law optimization of a sensible heat thermal
energy storage system with a distributed storage element. Part I. Development of the analytical
model. ASME J. Energy Resour. Technol. 113, 20–26.
CHAPTER 7

Renewable Energy-Based Building


HVAC Systems

7.1 INTRODUCTION
The use of renewable energy in building heating, ventilation, and air condition-
ing (HVAC) systems is examined in this chapter. Three case studies are consid-
ered, each with a different renewable energy-based HVAC system. The main
HVAC processes considered are heating and cooling. Energy and exergy ana-
lyses of the systems in the case studies are utilized to assess the performances
of the overall systems and their heating and cooling systems. Parametric ana-
lyses are applied to assess the effects of varying important design and operating
parameters on the energy and exergy efficiencies of heating, cooling, and overall
systems.
The gradual depletion of fossil fuel reserves and other concerns with their use
have fostered interest in alternative energy sources that are sustainable over
time and environmentally benign. As a consequence, renewable energy
resources have received much attention. The main renewable energy resources
used at present are hydraulic, solar, geothermal, wind, and biomass (Ahmadi
et al., 2013). Biomass is mainly obtained from living or dead matter (Cohce
et al., 2011). Solar energy can be harvested in various ways, for example, solar
photovoltaics and solar thermal energy (Khalid et al., 2015a). Solar thermal
collectors are considered in the case studies in this chapter.
A significant portion of the total energy use in the world is attributable to the
residential sector. Depending on the country and its climatic conditions, energy
use by residential buildings ranges from 16% to 50% of the total energy con-
sumption (Caliskan et al., 2011). In Canada, for example, residential energy
use accounted for 16% of total energy use in 2010, as shown in Fig. 7.1. Of this,
80% was used for space heating and hot water heating, as shown in Fig. 7.2. For
Saudi Arabia, electricity consumption in the residential sector accounts for 50%
of the total electricity consumption, as can be seen in Fig. 7.3 for the year 2011.
Given that a large part of the energy in many countries is used in the residential
sector, it is advantageous to have efficient and clean systems to satisfy the 279

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00007-7


© 2015 Elsevier Inc. All rights reserved.
280 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

12% 3%
38%
16% Industrial
Transportation
Residential
Commercial
31% Agriculture

FIGURE 7.1 Breakdown of energy use by the sector in Canada, based on data reported by Natural
Resources Canada (2011).

4% 2%
14%
Space heating

17% Water heating


63% Appliances
Lighting
Space cooling

FIGURE 7.2 Distribution of residential energy use by end use in Canada, based on data reported
by Natural Resources Canada (2011).

2% 3%
11% Residential
Commercial
19% 50%
Industrial
Government
15% Agricultural
Others

FIGURE 7.3 Breakdown of electricity use by the sector in Saudi Arabia, based on data provided
by SEC (2011).

demands. In most developing and industrialized countries, most of the electric-


ity is generated by combusting fossil fuels, which leads to emissions of green-
house gases and other environmental pollutants and depletes supplies of
finite energy resources. Much of the electricity use in the residential sector is
for HVAC applications, so renewable energy systems for HVAC applications
7.2 Case Study 1 281

can be beneficial. Designing building energy systems to utilize renewable


energy helps reduce energy utilization and the corresponding emissions. The
types of renewable energy resources most commonly used in buildings include
solar, biomass, wind, and geothermal (Khalid et al., 2015b), and these are uti-
lized in the HVAC system case studies examined in this chapter. In the remain-
der of this chapter, each of the three case studies is described and assessed using
energy and exergy analyses. Energy and exergy efficiencies of heating, cooling,
and overall systems are determined, and the effects are examined of varying
selected design and operating parameters.

7.2 CASE STUDY 1


In this case study, a biomass-based system for building cooling and heating is
considered. The main products are hot steam, which is used for heating, and
chilled water, which is used for cooling.

7.2.1 System Description


Figure 7.4 presents a diagram of the biomass-based system for cooling and heat-
ing buildings. Atmospheric air enters the combustion chamber at state 1, and
the fuel (biomass) enters the combustion chamber at state 2. The fuel and air
mix and burn inside the combustion chamber, and the hot combustion gases
exit the combustion chamber at state 3. Cold water at state 5 enters heat
exchanger 1 and is heated by the hot combustion gases, which exit heat
exchanger 1 at state 4. Steam exiting heat exchanger 1 at state 6 is utilized
for space heating and hot water heating. The hot combustion gases exiting heat
exchanger 1 drive the generator of the vapor absorption chiller and then are
emitted to the environment at state 7.

7.2.2 Assumptions in Thermodynamic Analyses


The proposed systems considered in case studies 1-3 are assessed with energy
and exergy analyses. The following assumptions are invoked during the ana-
lyses in all case studies:

• The reference-environment state has a temperature T0 ¼ 298 K and a


pressure P0 ¼ 100 kPa.
• The changes in kinetic and potential energy and exergy terms are
negligible.
• The pressure losses in all heat exchangers and pipelines are negligible.
• Air is treated as an ideal gas.
• The combustion gases exiting the combustion chamber behave like air.
• The chemical exergy is not considered in the vapor absorption chiller, as it
is not relevant.
282 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

Biomass
2 Combustion
1 gases
Air I 3
5 Water

II
6 Steam

7 4

IV 8 III Building

14 15 Chilled
9 water
Water
IX
V
13 16
10 VIII X

12 17

VI 11 VII

I Combustion chamber VI Evaporator


II Heat exchanger VII Absorber
III Generator VIII Solution pump
IV Condenser IX Solution heat exchanger
V Throttling valve 1 X Throttling valve 2

FIGURE 7.4 Biomass-based system for building cooling and heating considered in case study 1.

7.2.3 Balance Equations


Mass, energy, entropy, and exergy rate balances for the overall system and its
components are written in this subsection for case study 1.

7.2.3.1 Combustion Chamber


Applying the principle of conservation of mass, mass rate balances can be writ-
ten for the combustion chamber as follows:

m_ 1 + m_ f ¼ m_ 3 (7.1)

m_ f ¼ m_ 2 (7.2)
7.2 Case Study 1 283

An energy rate balance for the combustion chamber can be expressed as


follows:

m_ 1 h1 + m_ f hf ¼ m_ 3 h3 + Q_ cc (7.3)

where hf is taken to be the lower heating value (LHV) of the fuel.


Applying an entropy rate balance yields

Q_
m_ 1 s1 + m_ f sf + S_gen, CC ¼ m_ 3 s3 + cc (7.4)
Ts

An exergy rate balance, which can be used to determine the exergy destruction
rate in the combustion chamber, follows
 
_ d, cc + Q_ cc 1  T0
m_ 1 ex1 + m_ f exf ¼ m_ 3 ex3 + Ex (7.5)
Ts

where Ts denotes combustion temperature and exf exergy of the fuel, which can
be determined using exergy-to-energy ratio ϕ as follows:
exf
ϕ¼ (7.6)
LHV

The exergy-to-energy ratio for a fuel CαHβNδOΥ can be expressed as follows


(Szargut et al., 1988):
 
β δ Υ β
Φ ¼ 1:0401 + 0:1728 + 0:0432 + 0:2169 1  0:2062 (7.7)
α α δ α

Also, the specific exergy of the gases at state 3 can be expressed as the sum of the
physical exergy exph3 and the chemical exergy exch3:
ex3 ¼ exch3 + exph3 (7.8)

The chemical exergy of the combustion gases can be determined as


X X
exch3 ¼ xk exkch + RT0 xk ln xk (7.9)

and the physical exergy as


exph3 ¼ h2  h0  T0 ðs3  s0 Þ (7.10)

7.2.3.2 Heat Exchanger 1


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
heat exchanger 1 as follows:
m_ 5 ¼ m_ 6 (7.11)
284 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

m_ 3 h3 + m_ 5 h5 ¼ m_ 4 h4 + Q_ loss1 + m_ 6 h6 (7.12)

Q_
m_ 3 s3 + m_ 5 s5 + S_gen, he ¼ m_ 4 s4 + loss1 + m_ 6 s6 (7.13)
Ts
 
m_ 3 ex3 + m_ 5 ex5 ¼ m_ 4 ex4 + Ex _ d, he + Q_ loss1 1  T0 + m_ 6 ex6 (7.14)
TS

7.2.3.3 Generator
Mass rate balances for the generator of the vapor absorption chiller can be writ-
ten as follows, for the combustion gases and the LiBr, respectively:
m_ 4 ¼ m_ 7 (7.15)

m_ 14 ¼ m_ 15 + m_ 8 (7.16)

The energy rate balance can be applied to the generator as follows:


m_ 14 h14 + m_ 4 h4 ¼ m_ 7 h7 + m_ 15 h15 + m_ 8 h8 (7.17)

The heat input rate to the generator Q_ d can be expressed as

Q_ d ¼ m_ 4 ðh4  h7 Þ (7.18)

The entropy generation rate in the generator can be calculated as follows:

m_ 14 s14 + m_ 4 s4 + S_gen, g ¼ m_ 7 s7 + m_ 15 s15 + m_ 8 s8 (7.19)

The exergy destruction rate in the generator Ėxd,g can be determined using an
exergy rate balance:
_ d, g
m_ 14 ex14 + m_ 4 ex4 ¼ m_ 7 ex7 + m_ 15 ex15 + m_ 8 ex8 + Ex (7.20)

7.2.3.4 Condenser
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the condenser as follows:
m_ 8 ¼ m_ 9 (7.21)

m_ 8 h8 ¼ m_ 9 h9 + Q_ c (7.22)

Q_ c
m_ 8 s8 + S_gen, c ¼ m_ 9 s9 + (7.23)
T0
 
T0
m_ 8 ex8 ¼ m_ 9 ex9 + Q_ c 1  _ d, c
+ Ex (7.24)
TS
7.2 Case Study 1 285

where Q_ c denotes the heat rejection rate from the condenser.

7.2.3.5 Throttling Valve 1


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
throttling valve 1 as follows:
m_ 9 ¼ m_ 10 (7.25)

m_ 9 h9 ¼ m_ 10 h10 (7.26)

m_ 9 s9 + S_gen, tv1 ¼ m_ 10 s10 (7.27)

_ d, tv1
m_ 9 ex9 ¼ m_ 10 ex10 + Ex (7.28)

7.2.3.6 Evaporator
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the evaporator as follows:
m_ 10 ¼ m_ 11 (7.29)

m_ 10 h10 + Q_ e ¼ m_ 11 h11 (7.30)

Q_ e _
m_ 10 s10 + + Sgen, e ¼ m_ 11 s11 (7.31)
T0
 
T0
m_ 10 ex10 + Q_ e _ d, e
 1 ¼ m_ 11 ex11 + Ex (7.32)
Te

where Q_ e denotes the heat absorption rate by the evaporator.

7.2.3.7 Absorber
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the absorber as follows:

m_ 12 ¼ m_ 11 + m_ 17 (7.33)

m_ 11 h11 + m_ 17 h17 ¼ m_ 12 h12 + Q_ a (7.34)

Q_ a
m_ 11 s11 + m_ 17 s17 + S_gen, a ¼ m_ 12 s12 + (7.35)
T0
 
T0
m_ 11 ex11 + m_ 17 ex17 ¼ m_ 12 ex12 + Q_ a 1  _ d, a
+ Ex (7.36)
Ts

where Q_ a denotes the heat loss rate from the absorber.


286 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

7.2.3.8 Solution Pump


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the solution pump as follows:
m_ 12 ¼ m_ 13 (7.37)

m_ 12 h12 + W_ sp ¼ m_ 13 h13 (7.38)

m_ 12 s12 + S_gen, sp ¼ m_ 13 s13 (7.39)

m_ 12 ex12 + W_ sp ¼ m_ 13 ex13 + Ex
_ d, sp (7.40)

7.2.3.9 Solution Heat Exchanger


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the solution heat exchanger as follows:

m_ 13 + m_ 15 ¼ m_ 14 + m_ 16 (7.41)

m_ 13 h13 + m_ 15 h15 ¼ m_ 14 h14 + m_ 16 h16 (7.42)

m_ 13 s13 + m_ 15 s15 + S_gen, she ¼ m_ 14 s14 + m_ 16 s16 (7.43)

_ d, she
m_ 13 ex13 + m_ 15 ex15 ¼ m_ 14 ex14 + m_ 16 ex16 + Ex (7.44)

7.2.3.10 Throttling Valve 2


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
throttling valve 2 as follows:
m_ 16 ¼ m_ 17 (7.45)

m_ 16 h16 ¼ m_ 17 h17 (7.46)

m_ 16 s16 + S_gen, tv2 ¼ m_ 17 s17 (7.47)

_ d, tv2
m_ 16 ex16 ¼ m_ 17 ex17 + Ex (7.48)

7.2.4 Energy Efficiencies of the System and Its Main


Sections for Case Study 1
For the system considered in case study 1, expressions for the energy efficiencies
of the heating and cooling sections and the overall system are developed in this
section.

7.2.4.1 Heating Cycle


The energy efficiency of the heating cycle can be expressed as
7.2 Case Study 1 287

m_ 6 h6  m_ 5 h5
ηen, he ¼ (7.49)
m_ 3 h3  m_ 4 h4

7.2.4.2 Vapor Absorption Chiller


For the vapor absorption chiller, a coefficient of performance (COP) can be
used to express the energetic performance. The energetic COP of the vapor
absorption chiller is written here as

Q_ e
COPen, ac ¼ (7.50)
Q_ d

7.2.4.3 Overall System


The energy efficiency of the overall system can be expressed as follows:

Q_ e + m_ 4 h4  m_ 3 h3
ηen, ov1 ¼ (7.51)
m_ f LHV

7.2.5 Exergy Efficiencies of the System and Its Main


Sections for Case Study 1
For the system considered in case study 1, expressions are developed for the
exergy efficiencies of the heating and cooling sections and the overall system,
corresponding to the energy efficiencies in the previous subsection.

7.2.5.1 Heating Cycle


The exergy efficiency of the heating cycle can be expressed as
m_ 6 ex6  m_ 5 ex5
ηex, he ¼ (7.52)
m_ 3 ex3  m_ 4 ex4

7.2.5.2 Vapor Absorption Chiller


The exergetic COP of the vapor absorption chiller, which expresses that device’s
exergetic performance (Dincer and Rosen, 2013), can be written as
 
T0
Q_ e 1
T
COPex, ac ¼  e  (7.53)
T0
Q_ d 1 
Ts

7.2.5.3 Overall System


The exergy efficiency of the overall system can be written as follows:
288 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

 
T0
Q_ e  1 + m_ 4 ex4  m_ 3 ex3
Te
ηex, ov1 ¼ (7.54)
m_ f exf

7.2.6 Results and Discussion


To perform energy and exergy analyses of the system in case study 1, values of
mass flow rate, temperature, pressure, specific enthalpy, and specific exergy are
calculated for each state of the developed system (see Table 7.1), using Engi-
neering Equation Solver (EES). For the considered system, the overall energy
and exergy efficiencies are found to be 25.1% and 2.4%, respectively. The exergy
efficiency of the overall system is much lower than the energy efficiency. This is
primarily because in this case, the system products are heating and cooling,
which are low-quality energy forms. The energy efficiency does not differentiate
between the qualities of products, while the exergy efficiency does. The exergetic
and energetic COPs of the vapor absorption chiller are found to be 0.31 and
0.61, respectively.
Figure 7.5 shows the exergy destruction rates for the major components of the
system in case study 1. It is clear from Fig. 7.5 that the combustion chamber has
the greatest exergy destruction rate of all system components, followed by heat

Table 7.1 Input and Calculated Process Data for the Biomass-Based System with Vapor
Absorption Chiller (Case Study 1)
State No. Fluid Type T (°C) P (kPa) ṁ (kg/s) h (kJ/kg) ex (kJ/kg)

1 Air 25.0 100 0.49 298.6 0.0


2 Biomass 25.0 100 0.07 18,588.0 24,757.0
3 Combustion gases 800.0 100 0.56 1130.0 428.6
4 Combustion gases 150.0 100 0.56 424.8 20.9
5 Water 25.0 100 0.75 104.8 0.0
6 Steam 120.0 200 0.75 503.8 53.0
7 Flue gases 70.0 100 0.56 343.8 3.0
8 Water 78.1 7.4 0.01 2646.0 125.6
9 Water 40.0 7.4 0.01 167.7 1.4
10 Water 1.4 0.68 0.01 167.7 10.6
11 Water 1.4 0.68 0.01 2503.0 208.6
12 Water/LiBr 34.0 0.68 0.8 90.5 59.6
13 Water/LiBr 34.0 7.4 0.8 90.5 59.6
14 Water/LiBr 69.8 7.4 0.8 162.4 66.2
15 Water/LiBr 80.0 7.4 0.79 185.7 74.1
16 Water/LiBr 43.2 7.4 0.79 112.8 65.9
17 Water/LiBr 36.2 0.68 0.79 112.8 78.7
7.2 Case Study 1 289

1000
Exergy destruction rate (kW)

100

10

1
Combustion Heat Throttling Evaporator Generator
chamber exchanger valve 2
components
FIGURE 7.5 Exergy destruction rates for the major components of the system in case study 1.

0.65 0.55

0.6
0.5
0.55

0.45
0.5
COPen

0.45
0.4 COPex

0.4
0.35
0.35 COPen,ac
COPex,ac
0.3 0.3
25 27 29 31 33 35
T0 (°C)

FIGURE 7.6 Effects of varying ambient temperatures on the energetic and exergetic COPs of the vapor
absorption chiller in case study 1.

exchanger and throttling valve 2. Efforts thus appear merited to attempt to


reduce the exergy destruction rates in these devices, provided it can be done
cost-effectively.
Figure 7.6 shows the effect of varying ambient temperatures on the energetic
and exergetic COPs of the vapor absorption chiller of the system. As the ambi-
ent temperature increases from 25 to 35 °C, the energetic COP does not change,
but the exergetic COP increases from 0.35 to 0.54. In Fig. 7.7, the effects of
varying ambient temperatures on the energy and exergy efficiencies of the heat-
ing cycle for system 1 are shown. As the ambient temperature increases, the
energy efficiency remains fixed, but the exergy efficiency decreases. Figure 7.8
290 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

0.8 0.18

0.7
0.17
0.6

Energy efficiency

Exergy efficiency
0.16
0.5

0.4
0.15

0.3
0.14
0.2 hen,he
hex,he
0.1 0.13
25 27 29 31 33 35
T0 (°C)

FIGURE 7.7 Effects of varying ambient temperatures on the energy and exergy efficiencies
of the heating cycle in case study 1.

0.3 0.025

0.25 0.024
Energy efficiency

Exergy efficiency
0.2 0.023

0.15 0.022

0.1 0.021

0.05 hen,ov1 0.02


hex,ov1
0 0.019
25 27 29 31 33 35
T0 (°C)

FIGURE 7.8 Effects of varying ambient temperatures on the overall energy and exergy efficiencies
of the system in case study 1.

shows the variation of the energy and exergy efficiencies of the overall system
with the ambient temperature. The overall exergy efficiency decreases from
17.4% to 13.9% as the ambient temperature increases from 25 to 35 °C, while
energy efficiency remains unchanged.
The effects of varying steam mass flow rates on the energetic COP of the vapor
absorption chiller and energy efficiencies of the overall system and its heating
cycle are shown in Fig. 7.9. As the mass flow rate rises from 0.7 to 0.8 kg/s, the
7.2 Case Study 1 291

0.9 0.9

0.8 0.8

0.7 0.7

Exergy efficiency
0.6 0.6
COPen

0.5 0.5
COPen,ac
hen,ov1
hen,he
0.4 0.4

0.3 0.3

0.2 0.2
0.7 0.72 0.74 0.76 0.78 0.8
m6 (kg/s)

FIGURE 7.9 Effects of varying mass flow rates on the energetic COP of the vapor absorption chiller and
the energy efficiencies of the overall system and its heating cycle, for case study 1.

energy efficiency of the heating cycle increases from 70.6% to 80.7%, while
the energetic COP does not change. As a consequence of the change in
the energy efficiency of the heating cycle, the energy efficiency of the overall
system also rises, from 23.5% to 26.6%.
Figure 7.10 shows the effect of varying steam mass flow rates on the exergetic
COP of the vapor absorption chiller and the exergy efficiencies of the overall

0.35 0.2

0.3
0.16
0.25
Exergy efficiency

0.12
0.2
COPex

0.15
0.08
COPex,ac
hex,ov1
0.1
hex,he
0.04
0.05

0 0
0.7 0.72 0.74 0.76 0.78 0.8
m6 (kg/s)

FIGURE 7.10 Effect of varying mass flow rates on the exergetic COP of the vapor absorption chiller and
the exergy efficiencies of the overall system and its heating cycle.
292 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

system and the heating cycle. With increasing mass flow rate, the exergy output
of the heating cycle increases, resulting in an increase in the exergy efficiencies
of the overall system and the heating cycle.

7.3 CASE STUDY 2


The system presented in case study 2 uses a combination of two renewable
energy sources (solar and wind) to generate three main product outputs: cool-
ing, hot water, and heating.

7.3.1 System Description


Figure 7.11 shows the system in case study 2 for heating and cooling using a
solar thermal collector and wind turbine. The solar radiation is incident on
the flat plate collector, and Duratherm oil at state 1 enters the collector and
is heated. The oil leaves the collector at state 2 and enters heat exchanger 2,
which also acts as a storage tank and is utilized in heating the water that enters
the storage tank at state 4. After heating the water in the storage tank, the oil is
pumped back to the collector. The temperature of the hot water taken from the
storage tank at state 5 is not sufficiently high to drive the generator of the vapor
absorption chiller. So, the water at state 5 is further heated using an electric
heater to attain the required temperature and exits the heater at state 7. Part
of the electricity generated by the wind turbine is supplied to the electric heater.
Heat is transferred from the hot water at state 6 to the generator of the vapor
absorption chiller, and the water exiting the generator at state 7 is utilized to
satisfy the hot water requirement of the building. The electricity generated by
the wind turbine system leaves at state 20 and is divided into two parts, one
directed to the heat pump and the other to the electric heater.

7.3.2 Balance Equations


Mass, energy, entropy, and exergy rate balances for the overall system and its
components are written in this subsection for case study 2.

7.3.2.1 Flat Plate Collector


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the flat plate collector as follows:
m_ 1 ¼ m_ 2 (7.55)

m_ 1 h1 + Q_ sol ¼ m_ 2 h2 + Q_ loss, f pc (7.56)


7.3 Case Study 2 293

I Flat plate collector X Solution pump


II Storage tank XI Solution heat exchanger
III Pump XII Throttling valve 2
IV Heater XIII Wind system
Hot water
V Generator XIV Compressor
VI Condenser XV Condenser 1
VII Throttling valve 1 XVI Expansion valve Water
VIII Evaporator XVII Evaporator 1
IX Absorber
23 24
XV
21 XVI

XIV 22 XVII 25
20 7
Hot water
6 8
XIII V VI

14 15 9
IV
2 5
XI VII

13 16
I
XII 10
1 II X
12 17
11
4 IX VIII
Chilled
Water
water
III

Water
Building
FIGURE 7.11 Wind and solar thermal energy-based systems for cooling and space heating and hot
water heating considered in case study 2.

Q_ sol _ Q_ loss, f pc
m_ 1 s1 + + Sgen, f pc ¼ m_ 2 s2 + (7.57)
T0 T0
   
T0 T0
m_ 1 ex1 + Q_ sol 1  ¼ m_ 2 ex2 + Q_ loss, f pc 1  _ d, f pc
+ Ex (7.58)
TS Ts

where

Q_ sol ¼ q  Area (7.59)


294 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

7.3.2.2 Heater
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the heater as follows:

m_ 5 ¼ m_ 6 (7.60)

m_ 5 h5 + Q_ heater ¼ m_ 6 h6 (7.61)

Q_ heater ¼ W_ elec, heater (7.62)

m_ 5 s5 + S_gen, heater ¼ m_ 6 s6 (7.63)

m_ 5 ex5 + W_ elec, heater ¼ m_ 6 ex6 + Ex


_ d, heater (7.64)

7.3.2.3 Compressor
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the compressor as follows:
m_ 22 ¼ m_ 23 (7.65)

m_ 22 h22 + W_ comp ¼ m_ 23 h23 (7.66)

m_ 22 s22 + S_gen, comp ¼ m_ 23 s23 (7.67)

m_ 22 ex22 + W_ comp ¼ m_ 23 ex23 + Ex


_ d, comp (7.68)

7.3.2.4 Condenser 1
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
condenser 1 as follows:
m_ 23 ¼ m_ 24 (7.69)

m_ 23 h23 ¼ Q_ c1 + m_ 24 h24 (7.70)

Q_ c1
m_ 23 s23 + S_gen, c1 ¼ m_ 24 s24 + (7.71)
T0
 
T0
m_ 23 ex23 ¼ m_ 24 ex24 + Q_ c1 1  _ d, c1
+ Ex (7.72)
Ts

7.3.2.5 Expansion Valve


Mass, energy, entropy, and exergy rate balances, respectively, can be written for
the expansion valve as follows:
m_ 24 ¼ m_ 25 (7.73)
7.3 Case Study 2 295

m_ 24 h24 ¼ m_ 25 h25 (7.74)

m_ 24 s24 + S_gen, ev ¼ m_ 25 s25 (7.75)

_ d, ev
m_ 24 ex24 ¼ m_ 25 ex25 + Ex (7.76)

7.3.2.6 Evaporator 1
Mass, energy, entropy, and exergy rate balances, respectively, can be written for
evaporator 1 as follows:
m_ 25 ¼ m_ 22 (7.77)

m_ 25 h25 + Q_ e1 ¼ m_ 22 h22 (7.78)

Q_ e1 _
m_ 25 s25 + + Sgen, e1 ¼ m_ 22 s22 (7.79)
T0
 
T0
m_ 25 ex25 + Q_ e1 _ d, e1
 1 ¼ m_ 22 ex22 + Ex (7.80)
Te1

where Q_ e1 denotes the heat absorbed by evaporator 1.

7.3.3 Energy Efficiencies of the System and its Main


Sections for Case Study 2
For the system considered in case study 2, expressions for the energy efficiencies
of the heating and cooling sections and the overall system are developed in this
section.

7.3.3.1 Heat Pump


For the heat pump, a COP can be used to express energetic performance as
m_ 21 h21  m_ 22 h22
COPen, hp ¼ (7.81)
W_ comp

7.3.3.2 Vapor Absorption Chiller


For the vapor absorption chiller, a COP can be used to express energetic perfor-
mance as
m_ 10 h10  m_ 11 h11
COPen, ac ¼ (7.82)
m_ 6 h6  m_ 7 h7

7.3.3.3 Overall System


The energy efficiency of the overall system can be expressed as follows:
296 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

ðm_ 21 h21  m_ 22 h22 + m_ 10 h10  m_ 11 h11 + m_ 7 h7  m_ 4 h4 Þ


ηen, ov2 ¼   (7.83)
Q_ sol, f pc + En
_ wind

7.3.4 Exergy Efficiencies of the System and Its Main


Sections for Case Study 2
For the system considered in case study 2, expressions are developed for the
exergy efficiencies of the heating and cooling sections and the overall system,
corresponding to the energy efficiencies in the previous subsection.

7.3.4.1 Heat Pump


The exergetic COP of the heat pump, which expresses that device’s exergetic per-
formance, can be expressed as
m_ 21 ex21  m_ 22 ex22
COPex, hp ¼ (7.84)
W_ comp

7.3.4.2 Vapor Absorption Chiller


The exergetic COP of the vapor absorption chiller, which expresses that device’s
exergetic performance, can be expressed as
m_ 10 ex10  m_ 11 ex11
COPex, ac ¼ (7.85)
m_ 6 ex6  m_ 7 ex7

7.3.4.3 Overall System


The exergy efficiency of the overall system can be written as follows:
ðm_ 21 ex21  m_ 22 ex22 + m_ 10 ex10  m_ 11 ex11 + m_ 7 ex7  m_ 4 ex4 Þ
ηex, ov2 ¼   (7.86)
_ sol, f pc + Ex
Ex _ wind

7.3.5 Results and Discussion


In order to perform the energetic and exergetic analyses of the system presented
in case study 2, the values of temperature, mass flow rate, pressure, specific
enthalpy, and specific exergy are calculated for each state of the developed
system (see Table 7.2). The developed system used the thermodynamic values
provided in the EES software. The energy and exergy efficiencies of the overall
system are found to be 73.49% and 4.75%, respectively. The energetic and exer-
getic COPs of the heat pump are found to be 4.54 and 0.22, respectively.
The system present in the case study is also analyzed component-wise, that is,
the exergy destruction of the main components of the system is presented in
Fig. 7.12.
7.3 Case Study 2 297

Table 7.2 Input and Calculated Process Data for Wind and Solar Thermal System with Heat
Pump (Case Study 2)
State No. Fluid Type T (°C) P (kPa) ṁ (kg/s) h (kJ/kg) ex (kJ/kg)

1 Oil 40.0 100 3.50 40.8 1.1


2 Oil 80.0 100 3.50 107.9 7.3
3 Oil 40.0 100 3.50 40.8 1.1
4 Water 25.0 100 1.12 104.8 0.0
5 Water 75.0 100 1.12 314.0 15.9
6 Water 85.0 100 1.12 355.9 22.4
7 Water 75.0 100 1.12 314.0 15.9
8 Water 78.1 7.4 0.014 2646.0 125.6
9 Water 40.0 7.4 0.014 167.7 1.4
10 Water 1.4 0.68 0.014 167.7 10.6
11 Water 1.4 0.68 0.014 2503.0 208.6
12 Water/LiBr 34.0 0.68 1.00 90.5 59.6
13 Water/LiBr 34.0 7.4 1.00 90.5 59.6
14 Water/LiBr 69.8 7.4 1.00 162.4 66.2
15 Water/LiBr 80.0 7.4 0.98 185.7 74.1
16 Water/LiBr 43.2 7.4 0.98 112.8 78.7
17 Water/LiBr 36.2 0.68 0.98 112.8 65.9
18 – – – – – –
19 – – – – – –
20 R-134a 5.0 200 1.57 248.7 17.4
21 R-134a 57.4 1000 1.57 290.7 52.4
22 R-134a 34.3 1000 1.57 99.2 43.1
23 R-134a 10.0 200 1.57 99.9 37.1

35
Exergy destruction rate (kW)

30

25

20

15

10

0
Condenser 1 Compressor Throttling Evaporator 1 Generator Evaporator Expansion
valve 2 valve
Unit

FIGURE 7.12 Exergy destruction rates of the major components of the system presented in case
study 2.
298 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

0.8 0.7

0.65
0.7
0.6

0.6
0.55

COPen

COPex
0.5
0.5

0.45
0.4
COPen,ac 0.4
COPex,ac
0.3 0.35
25 27 29 31 33 35
T0 (°C)

FIGURE 7.13 Effects of varying ambient temperatures on the energetic and exergetic COPs of the vapor
absorption chiller of the system in case study 2.

Figure 7.13 shows the effect of varying ambient temperatures on the energetic
and exergetic COPs of the vapor absorption chiller of the system considered in
case study 2. As the ambient temperature rises from 25 to 35 °C, the energetic
COP remains the same, but the exergetic COP increases from 0.37 to 0.65.
In Fig. 7.14, the effect of ambient temperature on the energetic and exergetic
COPs of the heat pump is shown. As the ambient temperature increases, the

5 0.25

4
0.2

3
COPen

COPex

0.15
2

0.1
1
COPen,hp
COPex,hp
0 0.05
25 27 29 31 33 35
T0 (°C)

FIGURE 7.14 Effects of varying ambient temperatures on the energetic and exergetic COPs of the heat
pump of the system in case study 2.
7.3 Case Study 2 299

0.8 0.05

0.045
0.6
Energy efficiency

Exergy efficiency
0.04
0.4
0.035

0.2
hen,ov2 0.03
hex,ov2

0 0.025
25 27 29 31 33 35
T0 (°C)

FIGURE 7.15 Effects of varying ambient temperatures on the overall energy and exergy efficiencies
of the system in case study 2.

energetic COP remains unchanged, while the exergetic COP decreases.


Figure 7.15 shows the change of energy and exergy efficiencies of the overall
system with the ambient temperature. The overall energy efficiency is
unchanged, while exergy efficiency decreases from 4.7% to 2.6%, respectively.
In Fig. 7.16, as the generator outlet temperature for the system in case study 2
rises from 70 to 80 °C, the energetic COP of the vapor absorption chiller

5 0.77

0.76
4
0.75
Energy efficiency

3
0.74
COPen

0.73
2

0.72
1
COPen,ac 0.71
COPen,hp
hen,ov2
0 0.7
70 72 74 76 78 80
T7 (°C)

FIGURE 7.16 Effects of varying generator outlet temperatures on the energy efficiency of the overall
system and the energetic COPs of the vapor absorption chiller and heat pump, for case study 2.
300 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

0.8 0.054

0.052
0.6

Exergy efficiency
0.05

COPex
0.4 0.048

0.046
0.2

COPex,ac 0.044
COPex,hp
hex,ov2
0 0.042
70 72 74 76 78 80
T7 (°C)

FIGURE 7.17 Effects of varying generator output temperatures on the exergy efficiency of the overall
system and the exergetic COPs of heat pump and vapor absorption chiller, for case study 2.

increases from 0.49 to 1.47, while the energy efficiency of the overall system
increases from 70.4% to 76.5%. But the energetic COP of the heat pump does
not change, primarily because it requires electricity to operate.
Figure 7.17 shows the effect of generator output temperature on the exergy effi-
ciency of the overall system and the exergetic COPs of the heat pump and vapor
absorption chiller. As the generator output temperature increases from 70 to
80 °C, the exergetic COP of the vapor absorption chiller increases from 0.25
to 0.75, while the exergy efficiency of the overall system increases from 4.3%
to 5.2%. Since the output of the vapor absorption cycle is fixed thus on increas-
ing the generator output temperature, the energetic and exergetic COPs of the
vapor absorption chiller increase.

7.4 CASE STUDY 3


The system considered in case study 3 utilizes a combination of two renewable
energy sources (solar and wind) to provide two main product outputs: cooling
and heating. Although the main inputs and products for this case study and
case study 2 are similar qualitatively, the system arrangement and components
for the two cases differ.

7.4.1 System Description


Figure 7.18 shows the wind and solar thermal energy-based systems for heating
and cooling. The solar radiation striking the flat plate collector heats the Dur-
atherm oil entering at state 1. The oil exits the collector at state 2 and passes
7.4 Case Study 3 301

I Flat plate collector VI Compressor


II Storage tank VII Condenser
III Pump VIII Expansion valve
IV Heater IX Evaporator 10
V Wind system 11
VII

VIII
8
9
VI IX 12

V 6

Chilled water
IV
2
5

Water
Hot water

1 II

Water
III
Building
3

FIGURE 7.18 Wind and solar thermal energy-based systems for cooling and heating considered
in case study 3.

through heat exchanger 2, which also acts as a storage tank and is utilized in
heating the water that enters the storage tank at state 4. After heating the water
in the storage tank, the oil is pumped back to the collector. The hot water out
from the storage tank at state 5 is further heated and converted to steam using
an electric heater, which is driven by a portion of the electricity generated by the
wind turbine system. The steam at state 6 is utilized for space heating and hot
water heating in the buildings. The evaporator of the vapor compression chiller
absorbs the heat from the surroundings to provide space cooling. The compres-
sor of the vapor compression chiller is driven by part of the electricity generated
by the wind turbine system.
302 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

7.4.2 Balance Equations


Mass, energy, entropy, and exergy rate balances can be written for the system
presented in case study 3 in a similar way as done for case studies 1 and 2.

7.4.3 Energy Efficiency of System in Case Study 3


For the system considered in case study 3, expressions for the energy efficiencies
of the heating and cooling sections and the overall system are developed in this
section.

7.4.3.1 Vapor Compression Chiller


For the vapor compression chiller, a COP can be used to express energetic per-
formance as
m_ 9 h9  m_ 12 h12
COPen, vc ¼ (7.87)
W_ comp

7.4.3.2 Heating Cycle


The energy efficiency of the heating cycle can be expressed as
m_ 6 h6  m_ 5 h5
ηen, he ¼ (7.88)
Q_ heater

7.4.3.3 Overall System


The energy efficiency of the overall system is written as follows:
ðm_ 9 h9  m_ 12 h12 + m_ 6 h6  m_ 5 h5 Þ
ηen, ov3 ¼   (7.89)
Q_ sol, f pc + En
_ wind

7.4.4 Exergy Efficiencies of the System and Its Main


Sections for Case Study 3
For the system considered in case study 3, expressions are developed for the
exergy efficiencies of the heating and cooling sections and the overall system,
corresponding to the energy efficiencies in the previous subsection.

7.4.4.1 Vapor Compression Chiller


The exergetic COP of the vapor compression chiller can be expressed as
m_ 9 ex9  m_ 12 ex12
COPex, vc ¼ (7.90)
W_ comp
7.4 Case Study 3 303

7.4.4.2 Heating Process


The exergy efficiency of the heating cycle can be expressed as
m_ 6 ex6  m_ 5 ex5
ηex, he ¼ (7.91)
Q_ heater

7.4.4.3 Overall System


The exergy efficiency of the overall system can be written as follows:
ðm_ 9 ex9  m_ 12 ex12 + m_ 6 ex6  m_ 5 ex5 Þ
ηex, ov3 ¼   (7.92)
_ sol, f pc + Ex
Ex _ wind

7.4.5 Results and Discussion


The system in case study 3 is analyzed through energy and exergy analyses. In
order to perform the analysis, the values of mass flow rate, temperature, pres-
sure, specific enthalpy, and specific exergy are calculated for each state of the
developed system (see Table 7.3). The energy and exergy efficiencies of the
overall system are found to be 25.1% and 3.6%, respectively, while the ener-
getic and exergetic COPs of the vapor compression chiller are found to be
2.95 and 0.39, respectively. Figure 7.19 shows the exergy destruction rates of
the major components of the system in case study 3.
Figure 7.20 shows the effect of ambient temperature on the energetic and exer-
getic COPs of the vapor compression chiller of the system considered in case

Table 7.3 Input and Calculated Process Data for Wind and Solar Thermal System with Vapor
Compression Chiller (Case Study 3)
State No. Fluid Type T (°C) P (kPa) ṁ (kg/s) h (kJ/kg) ex (kJ/kg)

1 Oil 40.0 100 3.0 40.8 1.11


2 Oil 85.0 100 3.0 116.7 8.86
3 Oil 40.0 100 3.0 40.8 1.11
4 Water 25.0 100 1.21 104.8 0.0
5 Water 70.0 100 1.21 293.0 13.01
6 Water 90.0 100 1.21 377.0 26.09
7 – – – – – –
8 – – – – – –
9 R-134a 5.0 200 1.61 248.7 17.4
10 R-134a 65.3 1200 1.61 295.5 56.7
11 R-134a 41.2 1200 1.61 110.2 43.7
12 R-134a 10.0 200 1.61 110.2 35.7
304 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

35

Exergy destruction rate (kW)


30

25

20

15

10

0
Condenser 1 Compressor Compressor Condenser Evaporator Heater Expansion
valve
Unit

FIGURE 7.19 Exergy destruction rates of the major components of the system in case study 3.

3.2 0.52

0.5
2.4
0.48

0.46
COPen

COPex
1.6
0.44

0.42
0.8
COPen,ac 0.4
COPex,ac
0 0.38
25 27 29 31 33 35
T0 (°C)

FIGURE 7.20 Effects of varying ambient temperatures on the energetic and exergetic COPs of the vapor
compression chiller of the system in case study 3.

study 3. As the ambient temperature increases from 25 to 35 °C, the energetic


COP does not change, but the exergetic COP increases from the 0.39 to 0.50.
In Fig. 7.21, the effects of varying ambient temperatures on the energy and
exergy efficiencies of the heating cycle of the system in case study 3 are shown.
As the ambient temperature increases, the energy efficiency does not change,
but the exergy efficiency decreases from 15.6% to 12.7%.
Figure 7.22 shows the change of energy and exergy efficiencies of the overall sys-
tem with the ambient temperature. The overall exergy efficiency changes from
3.6% to 4.1%, while the energy efficiency remains unchanged. There is increase
in exergy efficiency of the system because the rate at which the output of the vapor
compression chiller is more than the rate at which the output of the heating cycle
decreases, resulting in the overall increase in the exergy efficiency of the system.
7.5 Comparison Of Case Studies 305

0.16
1
0.155

0.8 0.15
Energy efficiency

Exergy efficiency
0.6 0.145

0.14
0.4
0.135
0.2
hen,he 0.13
hex,he
0 0.125
25 27 29 31 33 35
T0 (°C)

FIGURE 7.21 Effects of varying ambient temperatures on the energy and exergy efficiencies of the
heating cycle in case study 3.

0.3 0.042

0.24
Energy efficiency

Exergy efficiency
0.04
0.18

0.12
0.038

0.06
hen,ov3
hex,ov3
0 0.036
25 27 29 31 33 35
T0 (°C)

FIGURE 7.22 Effects of varying ambient temperatures on the energy and exergy efficiencies of the
overall system in case study 3.

7.5 COMPARISON OF CASE STUDIES


A general trend observed for all case studies is that they have high energy effi-
ciencies but low exergy efficiencies. This trend is due to the fact that energy anal-
ysis does not distinguish between the qualities of the energy output, while
exergy does, and in all the cases, the outputs are cooling and heating, which
306 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

are low-quality energy at near-ambient conditions. Another common trend is


that in all the three case studies, the overall exergy efficiencies decrease as the
ambient temperature increases.
The case studies also demonstrate the outcome if an absorption chiller is uti-
lized, instead of a vapor compression chiller, for cooling (see case studies 1
and 2). Of the systems presented in the three case studies, case study 2 is dem-
onstrated for the design conditions to be the most advantageous from the exer-
getic point of view.
More generally, the results for all case studies suggest that renewable energy
should be used judiciously in buildings, accounting for the demand and other
factors.

7.6 CLOSING REMARKS


The use of renewable energy in building HVAC systems is described, focusing on
heating and cooling. This topic is illustrated in this chapter by developing three
different systems (in three case studies) for sustainable HVAC applications in
buildings and investigated them with energy and exergy analyses. The energy
efficiencies for the systems presented in case studies 1, 2, and 3 are found to
be 25.1%, 73.4%, and 25.1%, respectively. So, the system in case study 2 is
observed to be superior on the basis of overall energy efficiency. But the exergy
efficiencies for the systems in case studies 1, 2, and 3 are found to be 2.4%, 4.7%,
and 3.6%, respectively, suggesting that the system in case study 2 is the most
advantageous on the basis of overall exergy efficiency. For all systems, the exergy
efficiencies are quite low compared with the energy efficiencies, a trend mainly
attributable to the outputs of the systems. All systems produce heating and
cooling, which are of low quality and low-energy quantities. Energy analysis
cannot distinguish between low- and high-quality outputs, while the exergy effi-
ciency can. All systems considered are driven by a combination of renewable
energy sources, suggesting that renewable energy has significant potential for
beneficial HVAC applications in buildings.

Nomenclature
A area (m2)
COP coefficient of performance
ex specific exergy (kJ/kg)
Ėx exergy rate (kW)
h specific enthalpy (kJ/kg)
LHV lower heating value (kJ/kg)
ṁ mass flow rate (kg/s)
P pressure (kPa)
References 307

q solar radiation received per unit area (W/m2)


Q_ heat transfer rate (kW)
s specific entropy (kJ/kg K)
T temperature (°C)
Ẇ work rate or electric power (kW)

Greek Letters
η efficiency
Φ exergy-to-energy ratio

Subscript
a absorber
ac absorption chiller
c condenser
comp compressor
d destruction
e evaporator
elec electric
en energy
ex exergy
ev expansion valve
g generator
gen generation
he heat exchanger
hp heat pump
p pump
s source
she solution heat exchanger
sol solar thermal
sp solution pump
tv throttle valve
vc vapor compression
1, 2,… state numbers

References
Ahmadi, P., Dincer, I., Rosen, M.A., 2013. Development and assessment of an integrated biomass-
based multi-generation energy system. Energy 56, 155–166.
Caliskan, H., Dincer, I., Hepsali, A., 2011. Exergetic and sustainability performance comparison of
novel and conventional air cooling systems for building applications. Energy Build.
43, 1461–1472.
Cohce, M.K., Dincer, I., Rosen, M.A., 2011. Energy and exergy analyses of a biomass-based hydrogen
production system. Bioresour. Technol. 102, 8466–8474.
308 C HA PT E R 7 : Renewable Energy-Based Building HVAC Systems

Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, second
ed. Elsevier, Oxford, UK.
Khalid, F., Dincer, I., Rosen, M.A., 2015a. Energy and exergy analyses of a solar-biomass integrated
cycle for multigeneration. Sol. Energy 112, 290–299.
Khalid, F., Dincer, I., Rosen, M.A., 2015b. Development and analysis of sustainable energy systems
for building HVAC applications. Appl. Therm. Eng. 87, 389–401.
Natural Resources Canada, 2011. https://www.nrcan.gc.ca/energy/resources (accessed 24.01.15).
Saudi Electricity Company, 2011. Annual report on energy consumption, Riyadh, Saudi Arabia.
Szargut, J., Morris, D.R., Steward, F.R., 1988. Exergy Analysis of Thermal, Chemical, and Metallur-
gical Processes. Hemisphere Publishing Corporation, New York.
CHAPTER 8

Exergy-Related Methods

8.1 INTRODUCTION
Energy use is essentially governed by thermodynamic principles, and therefore,
an understanding of thermodynamic aspects of energy can help us understand
pathways to sustainable development. In this regard, it is necessary to define ther-
modynamics as the science of energy that comes from the first law of thermody-
namics and exergy that comes from the second law of thermodynamics. Exergy
analysis is useful for improving the efficiency of energy-resource use, for it quan-
tifies the locations, types, and magnitudes of wastes and losses. In general, more
meaningful efficiencies are evaluated with exergy analysis rather than energy
analysis, since exergy efficiencies are always a measure of how nearly the effi-
ciency of a process approaches the ideal. Therefore, exergy analysis identifies
accurately the margin available to design more efficient energy systems by reduc-
ing inefficiencies. It is crystal clear that thermodynamic performance is best eval-
uated using exergy analysis because it provides more insights and is more useful
in efficiency-improvement efforts than energy analysis.
Exergy analysis represents a relatively recent and exciting innovation in thermo-
dynamics and energy systems, although it is an old concept historically as it
comes from the second law of thermodynamics. Exergy methods have received
notable attention only over the past few decades. Although that attention has
grown during that period, it has remained limited and applications of exergy
analysis in the heating, refrigeration, and air conditioning industries have
been few and far between.
Various exergy-related methods exist that can be used to better understand
and improve energy and engineering processes and systems. Many of these
are applicable to heating, refrigerating, and air conditioning, and these are
described in this chapter. This section focuses on design for energy and environ-
ment for buildings, with special attention on heating, ventilation, and air
conditioning (HVAC) systems. Methods for determining appropriate energy
sources and efficiencies are examined as are methods for enhancing systems
by reducing energy losses to facilitate better use of resources. 309

Exergy Analysis of Heating, Refrigerating, and Air Conditioning. http://dx.doi.org/10.1016/B978-0-12-417203-6.00008-9


© 2015 Elsevier Inc. All rights reserved.
310 C HA PT E R 8 : Exergy-Related Methods

8.2 DESIGN FOR ENERGY AND ENVIRONMENT


The design of a product or process must be well done in order to ensure its objec-
tive is achieved. In a comprehensive design process, a number of factors are con-
sidered (e.g., customer satisfaction, material availability, economics, ability to
meet the demand, safety, life cycle assessment (LCA), efficiency, reliability, legal
and regulatory compliance, manufacturability, construction requirements, life-
time, and environmental impact) (Graedel and Allenby, 1995). Among these fac-
tors, “environmental impact” is becoming increasingly important in product and
process design. Including environmental impact as a design factor is often
referred to as design for environment (DFE) and accounts for the environmental
impacts of resource use, emissions of wastes and pollutants, and other activities.
Environmental impact should be considered throughout the design process and is
better addressed at an early stage in the design, as doing so makes it easy to enhance
the environmental aspects of designs and incorporate improvements. The life-
times of products and processes vary significantly, often from years to decades.
So it is important to make appropriate DFE decisions, as the results of these
decisions are often in effect throughout the lifetime of the product or process.
The energy sector is important in many economies, making the design of energy
systems significant in terms of such factors as efficiency and environmental
impact. DFE activities related to energy use often have two foci (Rosen,
2004): (i) the use of clean energy sources and (ii) increased efficiency.
There are various stages in designing a product or process where DFE can be
implemented. The various stages where DFE can be utilized for HVAC systems
are shown in Fig. 8.1. In a process, DFE can be implemented by enhancing
operating practices, using up-to-date and environmentally friendly technology,
incorporating low-toxicity materials, etc. In a product, DFE can be taken into
account in numerous ways, for example, redesigning the part and making
assembly and transport more efficient and environmentally benign.
The many levels in a design for which DFE can be implemented are shown in
Fig. 8.2. Every level of design is sensitive to environmental impact and involves
numerous possible measures. For example, design for energy efficiency can
involve reducing energy use, utilizing flexible and renewable energy sources,
and mitigating possible CO2 emissions. Figure 8.2 shows, for many design fac-
tors, a key objective of the numerous designs for environment measures.
A number of important factors should be considered when designing a product
or process for a HVAC system. For instance, a HVAC system should exhibit the
following:
• High efficiency
• Cost effectiveness
• Effective energy utilization
• Sustainability
8.2 Design for Energy and Environment 311

Design for environment

Product changes Process changes

Change of Technology Improved operating


Redesign of parts materials changes practices

Redesign of assembly
Less material variety New technology Training
Material purification Increased automation Better material handling
Redesign of product Avoidance of toxics Layout changes Maintenance
Improved equipment Efficient management
Stream segregation
Redesign of transport Inventory control

FIGURE 8.1 Various stages where DFE can be utilized for HVAC systems. Modified from Dartmouth (2015).

Design for environment

Level Objective

Design for manufacturability Pollution prevention during manufacturing, reduce and safer material and process

Design for energy efficiency Reduced, flexible and renewable energy use, zero emission, carbon neutrality

Design for zero toxics Zero damage to organism (animal, human, plant)
Design for dematerialization Reduction in quantity of material without reducing quality and required strength

Design for packaging Minimize packaging, change in packaging method


Design for logistics Local material use, reduce transportation
Design for longevity Long life of product, better serviceability
Design for modularity Easy upgrade , deg. of separation and recombination
Design for serviceability Ease of repairs, recapture of used/broken parts
Design for use of recycled materials Reducing waste
Design for reduced material variety Less use of different types of materials
Design for healthy materials Use of environmental friendly materials
Design for disassembly Reuse of components, quick and cheap disassembly
Design for recycling Greater materials recovery, safe disposal of nonrecyclables material
Design for economic recycling Promote recycling
Design for compostability Biodegradation of nonrecyclable materials
Design for energy recovery Safe incineration and composting of residues
Design for compliance Satisfy regulations and prepare for future

FIGURE 8.2 Various levels of DFE and their objectives. Modified from Dartmouth (2015).
312 C HA PT E R 8 : Exergy-Related Methods

Much research has been undertaken to improve the efficiency of HVAC systems
and to reduce their environmental impacts. These studies use various methods
and approaches to reduce the negative impacts of HVAC systems on the envi-
ronment. Several of these methods and approaches are given as follows:

• Development of low-cost and high-efficiency technology and focusing


research and development on new technology
• Deployment of new technologies to retrofit old buildings
• Use of highly efficient heating, ventilation, and cooling systems
• Strictly following standard for HVAC systems and components
• Designing buildings such that they exploit solar energy gains to the
building in winter and reduce those gains in the summer
• Use of highly reflective surfaces, good insulation and well-sealed
high-performance windows, and light-colored roofs in buildings
• Use of high-efficiency single or integrated water heating systems, where
the integration can be with space heating or cooling systems
• Use of low-carbon fuels in HVAC
• Use of renewable energy for heating, cooling, and ventilation,
including the utilization of solar energy for water and space heating
and absorption chiller cycles driven by solar energy for space cooling
• Use of natural ventilation through modified building designs
• Multigeneration or combined heat and power (CHP) for heating, cooling,
and power generation
• Proper sizing of HVAC systems during design

It is very important to connect all the stages of a product or process by noting


the interconnectivity of material, energy, and waste flows. One approach for
representing the connections for a product or process and for identifying the
harmful impacts of a product/process on the environment is a flow diagram.
An environmentally responsible product (ERP) matrix is helpful for analyzing
the life cycle environmental impacts of HVAC systems (Grote and Antonsson,
2009). With this approach, a raking system is applied to each process and stage
of the life cycle of the HVAC system. A number is assigned to each process and
stage between 0 and 4 based on the degree of environmental impact, where the-
following value assignments are permitted in the matrix for a comparative
ranking:

4 No environmental impact
3 Minimal environmental impact, that is, less than the expected average
2 Moderate environmental impact, that is, about the expected average
1 Substantial environmental impact, that is, more than the expected average
0 Very high environmental impact

An example of an MRP matrix for HVAC systems is presented in Table 8.1.


8.3 Life Cycle Assessment 313

Table 8.1 Environmentally Responsible Product (ERP) Matrix for HVAC


Material Energy Solid Liquid Gas
Process Choice Use Residues Residues Residues

Raw material 3 3 2 3 3
extraction
Manufacturing 2 2 2 2 2
Packaging 3 3 2 3 4
Use 4 2 3 3 1
Recycling/ 3 3 2 1 3
waste

An overall MRP rating can be computed by summing the matrix values:


XX
RERP ¼ Mi, j
i j

Based on this rating, comparisons can be made of the environmental impact of


two similar systems, namely, an old HVAC system and a new HVAC system,
using modern technology and also based on their energy sources.

8.3 LIFE CYCLE ASSESSMENT


LCA is a tool for determining the environmental impacts during all the stages of
the life cycle of a product or process (Graedel and Allenby, 1995, 2010). The life
cycle includes raw material and resource extraction; material and resource pro-
cessing; manufacturing, distribution, use, repair, and maintenance; and dis-
posal or recycling.

8.3.1 Stages of LCA


A LCA of a product or process includes four stages: (1) goal and scope defini-
tion, (2) inventory analysis, (3) impact assessment, and (4) improvement anal-
ysis. These four stages, which are shown and further detailed in Fig. 8.3, are
interrelated and provide a comprehensive understanding of the inputs and out-
puts or energy and materials for any process.

8.3.1.1 Goal and Scope Definition


The first stage of LCA, goal and scope definition, involves specifying the inten-
tion or objective of the LCA and the scope or boundaries of the assessment, as
well as the application and audience.
In this section, the specific application considered is HVAC. Several studies
have been reported on LCAs of HVAC systems. These have yielded good under-
standing of the environmental issues associated with HVAC. The studies have
314 C HA PT E R 8 : Exergy-Related Methods

Goal and scope definition


Process system boundary is defined to include processes from raw materials
extraction, energy generation, and disposal

Inventory analysis
Energy and material inputs, wastes, and emissions data are collected and the
environmental load (EL) quantified

Interpretation
Impact assessment
Examines potential and actual Environmental and human health effects from
the use of resources and environmental releases

Improvement analysis
Need and opportunities to reduce the impact of the product and service on the
environment

FIGURE 8.3 Framework of LCA.

also pointed out some critical issues and problems related to environmental
impact and potential measures to reduce it.

8.3.1.2 Inventory Analysis


The second stage of LCA is inventory analysis, which is often the most time-
and effort-consuming phase of the entire assessment. It involves data acquisi-
tion to develop an inventory of all energy and material flows, over the entire life
cycle, for the product or process considered. This LCA stage is often extensive
and challenging because of the breadth of work involved to collect all data
at each stage of the life cycle. The data often differ by LCA stage, for example,
material and resource extraction and processing and transportation.
For HVAC system, it is often difficult to collect all the inventory data regarding
material and energy flow, as HVAC is a complex combination of components
(e.g., furnace, boiler, air conditioner, heat pump, fan coil, ducting, pipes, and
radiators). These components are themselves made of numerous subcompo-
nents and materials (e.g., steel, iron, aluminum, and copper). For example,
based on an analysis by Viral et al. (2008), the type and quantity of material,
the embodied energy and exergy, and the carbon dioxide emissions, for various
HVAC components, over their estimated lifetimes, are given in Table 8.2. The
8.3 Life Cycle Assessment 315

Table 8.2 Material, Embodied Energy and Exergy and Carbon Dioxide Emissions for Various Central
Devices and Distribution Components in HVAC Systems, Over Their Estimated Lifetimes
Total Total Total CO2 Estimated
Mass Embodied Embodied Emission Lifetime
Component Material (kg) Energy (MJ) Exergy (MJ) (kg) (year)

Furnace Steel 46 924.6 896.3 63.0 20


Galvanized steel 18 702.0 680.5 97.2
Aluminum 9 1395.0 1332.3 74.2
Copper 3 146.1 140.9 16.5
Boiler Cast iron 145 3625.0 3516.9 277.0 35
Galvanized steel 18 702.0 680.5 97.2
Copper 1 48.7 47.0 5.5
Air conditioner Steel 78 1567.8 1519.9 106.9 20
Galvanized steel 35 1365.0 1323.3 189.0
Copper 17 827.9 798.3 93.5
Aluminum 17 2635.0 2516.6 140.1
R-407C 6 282.0 259.1 18.0
(refrigerant)
Heat pump Steel 101 2030.1 1968.1 138.4 20
Galvanized steel 32 1248.0 1209.9 172.8
Copper 17 827.9 798.3 93.5
R-407C 7 329.0 302.2 21.0
(refrigerant)
Fan coil unit Steel 48 964.8 935.3 65.8 25
Galvanized steel 26 1014.0 983.0 140.4
Copper 2 97.4 93.9 11.0
Duct Galvanized steel 265 10335.0 10019.2 1431.0 35
Fiberglass 140 3430.0 3309.3 210.0
Pipes and Steel 415 8341.5 8086.6 568.6 35
radiators Copper 82 3993.4 3850.5 451.0
Fiberglass 202 4949.0 4774.8 303.0
Modified from Viral et al. (2008).

embodied energy is defined as the cumulative amount of energy needed to


extract, process, manufacture, and transport a product or service to its point
of use, while embodied exergy is defined as the amount of work the system
can perform when it is brought into thermodynamic equilibrium with the envi-
ronment (Jørgensen et al., 2004).
The HVAC system is divided in Table 8.2 into seven subsystems and compo-
nents and the type of material used in their manufacture is specified. The total
embodied energy for each system and component is calculated based on the
quantity and type of material used in its manufacture and is illustrated in
316 C HA PT E R 8 : Exergy-Related Methods

Figs. 8.4 and 8.5. In Fig. 8.4, it is seen that, in terms of components, the greatest
total embodied energy is used in manufacturing materials for pipes and radi-
ator. In Fig. 8.5, it is seen that, in terms of materials, the maximum total embod-
ied energy is used in manufacturing galvanized steel and normal steel for HVAC
components.
From Figs. 8.4 and 8.5, it can be observed that for HVAC devices, the greatest
total embodied energy is used in the piping and radiator and that for HVAC
materials, the greatest total embodied energy is used in galvanized steel, while
the lowest total embodied energy is used in the fan coil unit and refrigerant
(R-407C).

18,000
Total embodied energy (MJ)

16,000
14,000
12,000
10,000
8000
6000
4000
2000
0
p

ct

s
r
e

ni
ile

ne

or
m
ac

Du
il u
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi
FIGURE 8.4 Total embodied energy in various HVAC components.

18,000 Total embodied energy 2500

Total CO2 emission (kg)


16,000 Total embodied exergy
energy/exergy (MJ)

14,000 Total CO2 emission 2000


Total embodied

12,000
1500
10,000
8000 1000
6000
4000 500
2000
0 0
el

s
r
ee

nt
pe

s
ro
e

la
ra
St

st

in

ti
op

rg
ge
um

as
d

be
ze

fri
C
Al

Fi
ni

(re
va

7C
al
G

0
-4
R

FIGURE 8.5 Total CO2 emissions and total embodied energy and exergy for materials used in HVAC
systems.
8.3 Life Cycle Assessment 317

If we consider the embodied energy per unit mass of material used in different
devices, the picture for total embodied energy use changes, as can be seen
in Figs. 8.6 and 8.7. Figure 8.6 shows that the greatest embodied energy per
unit mass of material is used in the air conditioner, while the lowest embodied
energy per unit mass of material is used in piping and the radiator. Figure 8.7
demonstrates that the greatest embodied energy per unit mass of type of
material is used in aluminum and the lowest embodied energy per unit mass
of type of material is used in normal steel.

45.0
Embodied energy (MJ/kg of material)

40.0
35.0
30.0
25.0
20.0
15.0
10.0
5.0
0.0
p

ct

s
r
e

ni
ile

ne

or
m
ac

Du
il u
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi

FIGURE 8.6 Embodied energy per unit mass of material for various HVAC components.
Embodied energy /exergy (MJ/kg)

CO2 emission (kg/kg of material)

180 Embodied energy 9


160 8
Embodied exergy
140 7
120 CO2 emission 6
100 5
80 4
60 3
40 2
20 1
0 0
el

um

ss
r
ee

nt
pe

ro
e

la
ra
St

st

in

ti
op

rg
ge
um

as
d

be
ze

fri
C
Al

Fi
ni

(re
va

C
al

07
G

-4
R

FIGURE 8.7 CO2 emissions and total embodied energy and exergy per unit mass of material for various
materials used in HVAC components.
318 C HA PT E R 8 : Exergy-Related Methods

8.3.1.3 Impact Assessment


The third stage of LCA is impact assessment, which focuses on the environmental
impacts associated with the quantities identified in the inventory analysis stage.
In this stage, various types of flows (e.g., emissions of CO2 equivalent) and their
environmental impacts (e.g., climate change) are classified as shown in Fig. 8.8.
The information at each stage is evaluated in the interpretation phase of LCA.
In the assessment of HVAC systems considered here, the quantity of emissions
is calculated in terms of CO2 released during the processes of material produc-
tion using the total embodied energy. Figure 8.9 shows the total CO2 emission
mass attributable to the total embodied energy use for various HVAC devices. It
is observed that the greatest total CO2 emission is associated with the duct and
the lowest with the fan coil unit. The quantity of CO2 emissions is dependent
on the quantity of embodied energy, as well as the type and quantity of mate-
rial. As can be seen in Fig. 8.5, the greatest CO2 emission occurs in manufactur-
ing galvanized steel and the lowest in refrigerant production.

Natural Type of Impact over


resources emission environment

Natural Raw Waste Global


gas/coal/fuel material material warming

Stratospheric
Energy Different CO, CO2
ozone depl.
production material
production
Water Air
HVAC vapour pollution
component
manufacturing Spread of
Waste water
diseases
Transportation CH4, NOx, Ecosystem
and installation SOx, VOCs change

Waste heat, Water


Uses (heating or
persistent pollution
cooling)
free radical
Soil
Dust, heavy pollution
Waste metal
Recycle/reuse

FIGURE 8.8 Simplified life cycle diagram of HVAC system, showing life cycle stages, emission types,
and the corresponding environmental impacts.
8.3 Life Cycle Assessment 319

1800
1600
Total CO2 emission (kg)

1400
1200
1000
800
600
400
200
0

ct

s
r
e

ni
ile

ne

or
m
ac

Du
lu
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi
FIGURE 8.9 Total carbon dioxide emissions for various HVAC devices.

4.5
Mass of CO2/kg of material

4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
p

ct

s
r
e

ni
ile

ne

or
m
ac

Du
il u
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi

FIGURE 8.10 CO2 emission in different HVAC appliances per kg of material.

As observed earlier, the graph in Fig. 8.7 as presented for total CO2 emission for
HVAC devices changes notably when the CO2 emission per unit mass of mate-
rial is examined instead (see Fig. 8.10). Figure 8.10 shows that the CO2 emis-
sion mass per unit mass of material used is greatest for the duct and lowest for
the pipe and radiator, whereas Fig. 8.7 shows that the CO2 mass emission per
mass of material is greatest for aluminum and lowest for steel.

8.3.1.4 Improvement Analysis


The fourth and final stage of LCA is improvement analysis, in which the results
of a life cycle improvement (LCI) analysis and a life cycle inventory analysis
320 C HA PT E R 8 : Exergy-Related Methods

(LCIA) are combined and conclusions are drawn and recommendations made.
Life cycle interpretation combines and corelates all the four stages of LCA and
can help decision makers in identifying and ranking multiple environmental
improvement options.

8.3.2 Exergetic Life Cycle Assessment


Exergy can be defined as the maximum useful work that can be obtained from a
process or system in a reference environment and can be used via exergy anal-
ysis to provide a clear understanding of thermodynamic efficiencies and losses
for the process or system. Exergetic life cycle assessment (ExLCA) is a useful tool
for evaluating the impact on the environment of a process or product and is an
extension of LCA that considers flows of exergy. ExLCA can beneficially be
applied HVAC systems to determine the exergy flows throughout the lifetime
of HVAC components and processes and useful to reduce exergy destructions
and increase process efficiency and effectiveness.
Note that exergy losses are not incurred on a onetime basis, but rather occur
throughout the lifetime of a product or process. This is reflected in ExLCA.
Exergy, sustainability, and environmental impact are interrelated. For instance,

• if the exergy efficiency of a process approaches 100%, then environmental


impact approaches zero (in such a case, exergy is converted from one
source to another without any consumptions and waste emissions);
• but when exergy efficiency approaches 0%, then sustainability also
approaches zero because no useful output is obtained and all the input
exergy is wasted (Dincer and Rosen, 2013).

Figure 8.11 shows a flowchart for the stages of ExLCA. Applying the ExLCA
flowchart in Fig. 8.11, the embodied exergy for HVAC systems is calculated
using the embodied energy for each device. The reference environment condi-
tions are taken to be the ambient conditions, and here, the ambient tempera-
ture and pressure are taken to be 25 °C and 101.325 kPa, respectively. For the
HVAC system considered earlier, Fig. 8.12 shows that the total embodied exergy
is greatest in piping and the radiator and lowest in the fan coil unit. In terms of
the type of material, the total embodied exergy is greatest for galvanized steel
and lowest for the refrigerant.
The embodied exergy per unit mass of material used in various HVAC devices
provides a different viewpoint than that for the total embodied exergy.
Figure 8.13 shows that the embodied exergy per mass of material is greatest
for the air conditioner and lowest for the piping and radiator. Yet Fig. 8.7 shows
that embodied exergy per unit mass of type of material is greatest in aluminum
and lowest in normal steel.
8.3 Life Cycle Assessment 321

FIGURE 8.11 Flowchart for ExLCA.

18,000.0
Total embodied exergy (MJ)

16,000.0
14,000.0
12,000.0
10,000.0
8000.0
6000.0
4000.0
2000.0
0.0
p

ct

s
r
ce

ni
ile

ne

or
m

Du
a

lu
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi

FIGURE 8.12 Total embodied exergy for various HVAC devices.


322 C HA PT E R 8 : Exergy-Related Methods

45.0

Embodied exergy (MJ/kg of material)


40.0
35.0
30.0
25.0
20.0
15.0
10.0
5.0
0.0

ct

s
r
e

ni
ile

ne

or
m
ac

Du
lu
Bo

pu

at
tio
rn

di
co
Fu

di

at

ra
on

He

d
Fa
rc

an
Ai

s
pe
Pi
FIGURE 8.13 Embodied exergy per unit mass of material for various HVAC devices.

8.4 ENERGY RETROFITS


Retrofitting generally refers to adding new features to an older system or repla-
cing old technology with more advanced versions. Retrofits can be undertaken
for various reasons, including reducing operating expenses by exploiting the ben-
efits of new technologies, optimizing existing systems, and increasing lifetime.
Proper operation of an HVAC system is important for ensuring occupant com-
fort and health, as well as good economics and efficiency. Improper HVAC
operation can lead to a poor air quality, high power consumption, environ-
mental damage, and other problems. Consequently, retrofits are often used
to improve HVAC operation. Change in codes and standard are also factors that
necessitate HVAC retrofits. The emergence of new technologies like variable
speed drives and features like enhanced control systems often provide addi-
tional incentives for system retrofits. Retrofitting of an HVAC system often
reduces electrical power consumption, improves reliability, and helps meet
environmental standards.
Retrofitting for an HVAC system can be complicated. The approach can vary
from retrofit measures a homeowner can undertake to complete retrofit by
an HVAC expert. Some of the main steps that homeowners can undertake
are as follows:

• Replacing old thermostats with advanced programmable models and


utilizing thermostats in all the rooms so as to increase HVAC system
effectiveness and decrease energy consumption
8.4 Energy Retrofits 323

• Replacing old windows with new ones that are thermally superior
• Replacing old heating and cooling equipment with new modern
versions, which can reduce energy use by 30-80% and CO2 emissions
by 30-100% (Biden, 2009; Wikipedia, 2015)
• Adding or improving insulation of walls, roofs, and floors
• Switching to high-efficiency lighting like compact fluorescent or
LED bulbs
• Adding renewable energy (solar, wind, biogas, etc.) in place of
conventional forms
• Substituting materials that have low volatile organic compounds (VOCs)
to improve indoor air quality
• Substituting appliances with low energy consumption models
• Undertaking adjustments (e.g., tune HVAC systems and control air leaks)

A complete retrofitting requires homeowners and property managers to seek


external expertise in order to design a retrofit program and estimate its costs
and benefits. One external source of expertise is the local utility company, which
often can evaluate individual homes and their HVAC systems and provide rec-
ommendations for measures. A building HVAC contractor or a consulting engi-
neer familiar with HVAC systems can also suggest appropriate measures.
The process of retrofitting should start with a proper evaluation of the existing
HVAC system and building. A comprehensive plan can help avoid problems
that may detract from achieving such objectives as increasing HVAC system effi-
ciency and reliability. The main motives of an HVAC retrofit usually include
increasing system efficiency and effectiveness, improving economics, and
addressing environmental concerns.
HVAC system retrofits generally lead to reduced operating and maintenance
costs and increased efficiency while improving tenant comfort and satisfaction
and often saving money. Possible efficiency upgrades for a building’s HVAC
system include upgrading motors, enhancing controls, incorporating occu-
pancy sensors, adding energy recovery devices, and improving air flow and dis-
tribution. Most of these measures will decrease the energy consumption of a
building. Performing major retrofits of occupied buildings for significant
energy efficiency improvements can be challenging, as work typically needs
to be done over an extended period and at odd hours of the day and without
significant disturbances. Other site-specific complexities may exist for building
energy-related systems when the HVAC system is located and controlled on
multiple floors of a building and when inadequate knowledge and documen-
tation about the system are available. Proper planning, expert engineering sup-
port, and effective communications among project team members are required
for effective retrofits in general and especially when site-specific challenges are
present.
324 C HA PT E R 8 : Exergy-Related Methods

Retrofitting can involve not only changing system technologies but also more
complex measures such as the following:

• Reducing peak electrical demands by advantageously distributing the load


over the entire day or part of it
• Replacing on-site (decentralized) heating with central cogeneration
heating using natural gas
• Replacing on-site (decentralized) cooling with central absorption chillers,
driven by waste heat or cogenerated heat
• Substituting utilized energy sources with more economic alternatives

Substituting the energy source and system within an HVAC system can signif-
icantly affect the levelized cost of operation and efficiency of the system. A case
study is now described in which the effect of a change in energy source on the
levelized cost of power generation is investigated.

8.4.1 Case Study


The cost associated with using energy resources, especially renewable energy
sources, is an important criteria for selecting them. The capital cost of an energy
system based on renewable energy is usually higher than an energy system
based on conventional fossil fuels. But the operating cost is usually much smal-
ler in a renewable energy-based system than in a fossil fuel-based energy system.
Here, a case study is described for a typical house in a particular location
(Oshawa, Ontario, Canada), in which the effect is analyzed using various
energy sources in terms of the overall cost of power generation. The character-
istics of the house considered are listed in Table 8.3. It is assumed that there are
six people living in the house in total and that the heating load remains the
same throughout the day. The heating load calculations are done based
on the characteristics of the house and a reference environment temperature.
A heat pump is considered for supplying the required heat at a constant coef-
ficient of performance (COP) of 3.5. Two renewable energy sources (solar and
wind) and two nonrenewable sources (diesel and gasoline) are considered as
energy sources in the cost analyses. HOMER Pro microgrid, an online software
analysis tool, is used to calculate the overall cost of power generation and to
determine the required size of the system in terms of power output.
The house considered is assumed to be old and to have an average filtration
quality and a basement. The lower monthly average temperature given in
Table 8.3 is taken to be the outside design temperature and the inside design
temperature is taken to be 21 °C. Based on these design temperatures and
the house characteristics, monthly average heating load has been calculated
(see Table 8.4).
8.4 Energy Retrofits 325

Table 8.3 Details of the House in the Case Study


Characteristics Details

Total wall area (m2) 428


Total ceiling area (m2) 89
Gross wall type Wood frame, R-11 cavity insulation
Ceiling construction 3-1/2 in., R-11 insulation
Floor type Basement
Filtration quality Average
Number of residents 6

Table 8.4 Design Temperature, Wind Speed, Solar Radiation for a Location and Load Detail
Average Average
High Low Design Wind Average Solar Average Electric
Temp. Temp. Temp. Speed Radiation Heating Load
Month (°C) (°C) (°C) (m/s) (kWh/m2/day) Load (kW) (kW)

January 1 8 8 6.46 1.38 11 3.14


February 0 7 7 6.04 2.23 10.5 3.00
March 4 4 4 6.12 3.3 9.1 2.60
April 11 2 2 6.05 4.55 7.3 2.09
May 17 8 8 5.23 5.5 5.5 1.57
June 22 13 13 4.88 6.04 3.7 1.06
July 25 16 16 4.55 6.02 1.82 0.52
August 24 16 16 4.48 5.22 1.81 0.52
September 20 12 12 5 4.07 3.9 1.11
October 13 6 6 5.61 2.67 5.9 1.69
November 7 1 1 6.28 1.6 7.6 2.17
December 2 4 4 6.41 1.18 9.5 2.71
Year 12 4 4 5.59 3.65 6.47 1.85
average

Table 8.4 also shows the average wind speed and the average solar radiation in
each month for the selected location. As noted earlier, four energy sources are
considered.

8.4.1.1 Wind
First, wind is considered as the primary energy source, with a wind turbine pro-
viding the energy input to the house. Figure 8.14 shows the wind-based renew-
able energy system analyzed with the HOMER Pro software.
Figure 8.15 shows the total net present cost of the wind-based renewable energy
system, broken down by type of cost (e.g., capital, replacement, operating, and
326 C HA PT E R 8 : Exergy-Related Methods

Primary load 1
Generic 10 kW
44 kWh/day
5.6 kW peak

Converter S6CS25P
AC DC

FIGURE 8.14 Wind-based renewable energy system analyzed with HOMER Pro.

Cash flow summary


50,000 Capital
Replacement
Net present cost ($)

Operating
40,000 Fuel
Salvage
30,000

20,000

10,000

–10,000
Capital Replacement Operating Fuel Salvage

FIGURE 8.15 Cash flow summary of wind-based renewable energy system according to the type
of cost.

Cash flow summary


50,000 Generic 10 kW
Surrette 6CS25F
Converter
40,000
Net present cost ($)

30,000

20,000

10,000

−10,000
Capital Replacement Operating Fuel Salvage

FIGURE 8.16 Component breakdown of cash flow summary for wind-based renewable energy system.

salvage), while Fig. 8.16 shows the total net present cost and a breakdown by
component. The capital and replacement costs are much higher for the battery
than for the remaining system components. Wind turbines are cheaper in terms
of initial capital cost and in terms of operating cost. There is no fuel cost associ-
ated with the system, as all the required power is supplied by the wind turbine.
8.4 Energy Retrofits 327

Figure 8.17 shows the monthly average electricity production from wind tur-
bine. The highest electricity production occurs in January (5.45 kW) and lowest
in August (1.9 kW). The results identify that two wind turbines, each with a
capacity of 10 kW, should be used to produce the required power.
Note that lead acid batteries are used to store any excess energy produced by the
wind turbine. Figure 8.18 shows the state of charge (SOC) of the battery bank
throughout the year. The batteries are observed to remain almost fully charged
in the months of July and August due to low heating loads in the summer. The
capacity of battery bank is 40 kW.
An inverter is used to convert the direct current (DC) power produced by wind
turbine to alternating current (AC) power. The power output from the inverter
throughout the year is shown in Fig. 8.19. The maximum capacity of the
inverter is 6 kW. The levelized cost of power generation from the wind-based
renewable energy system is found to be $0.443 kWh1. (Note that all monetary
values in this chapter are in 2014 US $.)

8.4.1.2 Solar
In order to compare one renewable energy source with another, a solar
photovoltaic (PV) system is used to generate electricity from solar radiation.

Monthly average electric production


6 Wind
5
Power (kW)

4
3
2
1
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
FIGURE 8.17 Monthly average electricity production rate from wind turbine.

Battery state of charge monthly averages


100
Average value (%)

90 Max
80 Daily high
70 Mean
60 Daily low
Min
50
40
30
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month
FIGURE 8.18 SOC of battery bank throughout the year.
328 C HA PT E R 8 : Exergy-Related Methods

Inverter output power monthly averages


6

Average value (kW)


5 Max
Daily high
4 Mean
3 Daily low
2 Min
1
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month
FIGURE 8.19 Power output from the inverter throughout the year.

Figure 8.20 shows the solar-based renewable energy system analyzed with the
HOMER Pro software. Monthly averages of the solar radiation for the selected
location are given Table 8.4.
Figure 8.21 shows the total net present cost of the solar-based renewable energy
system, broken down by type of cost (e.g., capital, replacement, operating, and

Primary load 1 PV
44 kWh/day
5.6 kW peak

Converter S6CS25P
AC DC

FIGURE 8.20 Solar-based renewable energy system analyzed with HOMER Pro.

Cash flow summary


250,000 Capital
Replacement
200,000
Net present cost ($)

Operating
Fuel
150,000 Salvage

100,000

50,000

−50,000
Capital Replacement Operating Fuel Salvage

FIGURE 8.21 Cash flow summary of solar-based renewable energy system according to the type
of cost.
8.4 Energy Retrofits 329

salvage), while Fig. 8.22 shows the total net present cost and a breakdown by
component. From Figs. 8.21 and 8.22, it can be clearly seen that capital cost is
much higher for the PV collectors than that for the remaining components.
Figure 8.23 shows the monthly average electricity production from the PV
array. It is observed that the electricity production is greatest in June
(13 kW) and lowest in December (3.75 kW), while the intensity of solar radi-
ation is greatest in June (6.04 kWh/m2/day) and lowest in December
(1.18 kWh/m2/day).
Figure 8.24 shows the power output of the PV array during the daytime. As solar
radiation is only available during the daytime, there is no power output during
the night. Batteries are used to supply electrical power in the absence of solar
radiation. The capacity of the PV array is 65 kW.
Figure 8.25 shows the SOC of the lead acid battery bank used to store any excess
energy produced by the PV array, throughout the year. The batteries remain
almost fully charged from May to September, mainly due to the low heating
load and high levels of solar radiation. The capacity of the battery bank is also
65 kW.

Cash flow summary


250,000 PV
Surrette 6CS25P
200,000
Net present cost ($)

Converter
150,000

100,000

50,000

−50,000
Capital Replacement Operating Fuel Salvage

FIGURE 8.22 Breakdown by component of the cash flow summary for the solar-based renewable
energy system.

Monthly average electric production


14 PV
12
10
Power (kW)

8
6
4
2
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
FIGURE 8.23 Monthly average electricity production from the PV array.
330 C HA PT E R 8 : Exergy-Related Methods

Pv array power output monthly averages


60

Average value (kW)


50 Max
Daily high
40 Mean
30 Daily low
20 Min
10
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month
FIGURE 8.24 Power output from the PV array during the daytime throughout the year.

Battery state of charge monthly averages


100
Average value (%)

90 Max
80 Daily high
70 Mean
60 Daily low
Min
50
40
30
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month
FIGURE 8.25 SOC of battery bank throughout the year.

An inverter converts the DC power produced by PV array to AC power. The


design capacity of the inverter is 6 kW, and its power output throughout the
year is shown in Fig. 8.26. The levelized cost of power generation from this
solar-based renewable energy system is found to be $1.53 kWh1.

8.4.1.3 Diesel
To understand the effect of change in energy source from renewable to nonre-
newable on the levelized cost of power generation, two nonrenewable energy

Inverter output power monthly averages


6
Average value (kW)

5 Max
Daily high
4
Mean
3 Daily low
2 Min
1
0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month
FIGURE 8.26 Power output from the inverter throughout the year.
8.4 Energy Retrofits 331

Primary load 1 S6CS25P


Generator 1
44 kWh/day
5.6 kW peak
AC DC

FIGURE 8.27 Fossil fuel-based energy system analyzed with HOMER Pro.

sources (diesel and gasoline) are considered. Figure 8.27 shows the fossil fuel-
based energy system, which considers both diesel and gasoline fuels.
Figure 8.28 shows the total net present cost for the nonrenewable-based energy
system, where the initial capital cost is only for the generator. Batteries were
considered in the initial design, but the results from HOMER Pro demonstrated
that batteries were not needed. From Fig. 8.28, it can be seen that the capital
cost is very small compared to the fuel cost, which, for diesel fuel, is taken
to be $1.22 L1.
Figures 8.29 and 8.30 show the electricity production rate from the generator,
where diesel fuel is used. The greatest electricity production rate is observed to

Cash flow summary


120,000 Capital
Replacement
100,000
Net present cost ($)

Operating
80,000 Fuel
Salvage
60,000
40,000
20,000
0
–20,000
Capital Replacement Operating Fuel Salvage

FIGURE 8.28 Cash flow summary of diesel-based energy system broken down by type of cost.

Monthly average electric production


3.5 Generator 1
3.0
2.5
Power (kW)

2.0
1.5
1.0
0.5
0.0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec
FIGURE 8.29 Monthly average power generation from diesel generator.
332 C HA PT E R 8 : Exergy-Related Methods

Generator 1 electrical output monthly averages


4

Average value (kW)


Max
3 Daily high
Mean
2 Daily low
Min
1

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month

FIGURE 8.30 Power output from diesel generator throughout the year.

occur in January (3.1 kW) and the lowest in July (1.42 kW). The levelized cost
of power generation from the generator using diesel fuel is found to be
$0.74 kWh1.

8.4.1.4 Gasoline
In order to compare the nonrenewable energy sources, diesel fuel is replaced
with gasoline and the system is reanalyzed. The initial cost for both systems
is the same, but the cost of gasoline is taken to be $1.01 L1. Figure 8.31 shows
the total net present cost of this nonrenewable energy system, in which gasoline
is the energy source.
Figures 8.32 and 8.33 show the electricity production rate from the generator
using gasoline. The greatest and lowest electricity production rates from the
generator are found to be the same as in the previous case where diesel fuel
is used, that is, the greatest and lowest electricity production rates occur in
January (3.1 kW) and July (1.42 kW), respectively. The levelized cost of power
generation from the generator using gasoline is found to be $0.65 kWh1.

Cash flow summary


Capital
Replacement
100,000 Operating
Net present cost ($)

Fuel
Salvage

50,000

Capital Replacement Operating Fuel Salvage

FIGURE 8.31 Cash flow summary of gasoline-based energy system broken down by type of cost.
8.4 Energy Retrofits 333

Monthly average electric production


3.5 Generator 1
3.0
2.5
Power (kW)

2.0
1.5
1.0
0.5
0.0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec

FIGURE 8.32 Monthly average electricity production rate from gasoline generator.

Generator 1 electrical output monthly averages


4
Average value (kW)

Max
3 Daily high
Mean
2 Daily low
Min
1

0
Jan Feb Mar Apr May Jun Jul Aug Sep Oct Nov Dec Ann
Month

FIGURE 8.33 Power output from gasoline generator throughout the year.

8.4.2 Comparison
Four energy sources are considered to generate electricity for heating purposes
and cost analyses are performed using HOMER Pro software. Two of the energy
sources are renewable (wind and solar) and two are fossil fuels (diesel fuel and
gasoline). The life span for all of the energy systems is considered to be 25 years.
The capacities of the various components (PV, generator, battery, converter,
and wind turbine) obtained with HOMER Pro are given in Table 8.5. The results
indicate that there is no need for battery storage for the diesel and gasoline-
based cases, where a generator is used to produce the electricity. For a similar
demand, the required installed capacity of a diesel generator is found to be
much smaller than the installed capacities of the PV, the array, and the wind
turbine, which reduces its initial capital cost. The initial capital cost is much less
for the diesel- or gasoline-based energy systems ($4000) than the solar-based
energy system ($242,300). But the fuel cost in the case of wind- and solar-based
energy systems is zero, while it is $4451 year1 and $3684 year1 for the diesel-
and gasoline-based energy systems, respectively. In terms of emissions, there
are few emissions from the solar- and wind-based energy systems, whereas
there are for the fossil fuel-based systems. The CO2 emissions are higher for
the energy system based on diesel fuel (18,788 kg/year) than for the system
334 C HA PT E R 8 : Exergy-Related Methods

Table 8.5 Comparison of Performance for Energy Systems Using Four Energy Sources
Energy Source

Parameter Wind Solar Diesel Gasoline

Life (year) 25 25 25 25
Installed capacity of wind turbine (kW) 20 0 0 0
Installed capacity of PV array (kW) 0 65 0 0
Installed capacity of generator (kW) 0 0 4 4
Installed capacity of battery (kW) 40 65 0 0
Installed capacity of converter (kW) 6 6 0 0

Costs
Initial capital cost ($) 49,800 242,300 4000 4000
Replacement cost ($) 29,927 95,403 28,691 28,691
Fuel cost ($/year) 0 0 4451 3684.6
Operating cost ($/year) 3141 5326 11,620 10,122
Total net present value ($) 89,950 310,387 152,542 133,389
Cost of energy ($/kWh) 0.443 1.531 0.74 0.651

Emissions
Carbon dioxide (kg/year) 0 0 18,788 16,559
Carbon monoxide (kg/year) 0 0 46.4 46.4
Unburned hydrocarbons (kg/year) 0 0 5.14 5.14
Particulate matter (kg/year) 0 0 3.5 3.5
Sulfur dioxide (kg/year) 0 0 37.7 34
Nitrogen oxides (kg/year) 0 0 414 414
Note: All cost and values are in 2014 US $.

based on gasoline (16,559 kg/year). Other types of emissions are similar for
both diesel- and gasoline-based energy systems (see Table 8.5).
Figure 8.34 compares the levelized cost of power generation using the four dif-
ferent energy sources considered. It can be seen that the levelized cost of power
generation is lowest for the wind-based system ($0.44 kWh1) and highest for
the solar-based system ($1.53 kWh1). In the case of fossil fuel-based energy
systems, it is more economical and environmentally benign to use gasoline
instead of diesel fuel, as the levelized cost of power generation and CO2 emis-
sions are both lower for gasoline-based energy system. The results suggest that it
is advantageous to use the wind-based energy system for this particular location
to reduce the levelized cost of power generation and environmental impact in
terms of CO2 emissions. More generally, of course, the levelized cost of power
generation varies spatially and temporally on the basis of available renewable
resources and other factors.
8.5 Energy Substitution 335

Levelized cost of electricity ($/kWh) 1.8


1.6
1.4
1.2
1
0.8
0.6
0.4
0.2
0
Wind Solar Diesel Gasoline
Energy source
FIGURE 8.34 Comparison of levelized cost of power generation for energy systems using different
energy sources.

8.5 ENERGY SUBSTITUTION


HVAC systems can use conventional and/or renewable energy resources. Each
has advantages and disadvantages, in terms of technical performance, econom-
ics, environmental impact, and other factors, and many investigations have
been reported on these. For instance, exergy and energy analyses have been
reported using four heating options (heat pump, condensing boiler, conven-
tional boiler, and solar collector) for a building using fossil fuels and renewable
energy resources (Balta et al., 2010). Through a case study, the environmental
impacts of HVAC systems were shown to depend mainly on the type of energy
resource used, type of equipment, distribution method, and material used in
the HVAC system (Prek, 2004). Similar results were reported by Heikkila
(2004), who found environmental impact to depend greatly on the form
and source of energy and the type of heating and cooling systems.

8.5.1 Energy Sources for HVAC


Energy consumption patterns in heating, cooling, and ventilation mainly
depend on the building location and time of year, as well as the type of HVAC
devices used. Numerous sources of energy are available for HVAC systems, and
substitution of one energy resource with another can increase or decrease the
associated environmental impact and affect other factors. The energy resources
available for HVAC systems can be grouped generally into two main types:

1. Nonrenewable energy sources (including most conventional energy


sources)
2. Renewable energy sources
336 C HA PT E R 8 : Exergy-Related Methods

These are illustrated Fig. 8.35, with details for each type (only considering
major sources). Some energy resources can be classified as both renewable
and nonrenewable energy sources, like energy from waste, as shown in
Fig. 8.35. Each energy resource has advantages and disadvantages in terms of
utilization efficiency, economics, environmental impact, and other factors.
Usually, the conversion of these resources to thermal energy or electricity for
heating, cooling, and ventilation causes impacts on the environment, in part
due to the low conversion efficiencies of some processes.

8.5.1.1 Nonrenewable Energy Sources


Nonrenewable energy resources (e.g., fossil fuels) are those that are not replen-
ished at the rate at which they are used. As previously discussed, HVAC systems
can be driven by several major nonrenewable energy sources, namely, petro-
leum, coal, natural gas, and nuclear energy. A breakdown of these sources is
shown in Fig. 8.36.
Each energy resource in Fig. 8.36 has its own characteristics, for example, com-
position and energy content, and all the fossil fuels produce CO2 when com-
busted. It is useful to compare nonrenewable energy resources based on their
energy (or heat) content and CO2 emissions, as one input for determining the
most suitable energy source for a given HVAC system. The heat content, carbon
content, and total CO2 emissions for selected nonrenewable energy resources
are listed in Table 8.6 (TCR, 2013). The CO2 emissions are dependent on the
carbon content of the fuel.
Figure 8.37 compares the specific heat contents of several fuels and shows that
the specific heat content is highest for kerosene (14.40 kWh/kg) and lowest for
lignite (4.59 kWh/kg). Note that the heat content of a fuel is based here on its
higher heating value (HHV). Globally, the most widely used energy sources for
HVAC systems are oil and natural gas. Around the globe, natural gas is mostly
used for direct heating while oil is usually indirectly used for heating and cool-
ing by converting it into electricity (ETP, 2010; IEA, 2013).

Sources of energy for HVAC

NonRenewable Renewable

Petroleum Coal Natural gas Nuclear Waste Others Solar Wind Hydro Geothermal Biomass Wave

FIGURE 8.35 Classification of major renewable and nonrenewable energy sources for HVAC.
8.5 Energy Substitution 337

Non-renewable
sources of energy for HVAC

Petroleum Coal Natuaral gas Nuclear

Kerosene Anthracite

Diesel Bituminous

Gasoline Lignite

LPG
Subbituminous
Ethane

Propane

Butane

Naptha

Heavy oils

Light oils

FIGURE 8.36 Classification of major nonrenewable energy sources for HVAC.

The carbon content per unit energy is shown in Fig. 8.38 for various fuels. The
carbon content is highest for anthracite (96.36 g C/kWh) and lowest for natural
gas (49.34 g C/kWh). Much CO2 is produced when burning a nonrenewable
fuel in the presence of air for space heating, electricity generation, and other
purposes, as shown in Fig. 8.39 on an energy basis and in Fig. 8.40 on a mass
basis. The CO2 emission per unit energy production is observed in Fig. 8.39 to
be highest for petroleum coke (0.349 kg CO2/kWh) and lowest for natural gas
(0.181 kg CO2/kWh), while the CO2 emission per mass of fuel is seen in
Fig. 8.40 to be highest for still gas (3.47 kg CO2/kg) and lowest for lignite
and ethylene (1.51 kg CO2/kg).
Figure 8.41 compares the heat content and CO2 emissions from combustion for
various fuels, on a unit mass basis. The ratio of CO2 emission per unit energy is
observed to be highest for natural gas and lowest for anthracite.

8.5.1.2 Case Study for Nonrenewable Energy Sources


To demonstrate the effects of fuel substitution in an HVAC system on CO2
emissions, a case study involving a hypothetical house is analyzed considering
338 C HA PT E R 8 : Exergy-Related Methods

Table 8.6 Characteristics of Several Fuel Types


Heat Carbon Content CO2 Emission per CO2 Emission
Energy Content per Unit Energy Unit Energy per Unit Mass
Source Fuel Type (kWh/kg) (g C/kWh) (kg CO2/kWh) (kg CO2/kg)

Coal and Anthracite 8.11 96.36 0.353 2.86


coke Bituminous 8.05 86.91 0.319 2.57
Subbituminous 5.57 90.29 0.331 1.84
Lignite 4.59 89.67 0.329 1.51
Coke 8.01 94.96 0.348 2.79
Natural gas Natural gas 12.65 49.34 0.181 2.29
Petroleum Still gas 13.50 62.10 0.228 3.47
products Kerosene 14.40 69.98 0.257 3.27
LPG 13.26 58.62 0.215 2.85
Propane (liquid) 14.26 57.19 0.210 2.99
Ethane 9.37 58.28 0.214 2.00
Ethylene 6.56 62.75 0.230 1.51
Isobutane 12.65 60.39 0.221 2.80
Butane 13.01 60.63 0.222 2.89
Natural gasoline 11.98 62.20 0.228 2.73
Special naphtha 12.62 67.32 0.247 3.11
Petroleum coke 4.96 95.30 0.349 1.73
Heavy gas oils 12.32 69.71 0.256 3.15
Light oil 11.70 59.83 0.260 2.60
Abbreviation: LPG, liquid petroleum gas.
Modified from TCR (2013).

16.0
Specific heat content (kWh/kg)

14.0
12.0
10.0
8.0
6.0
4.0
2.0
0.0
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil

Coal and coke Natural gas Petroleum products

FIGURE 8.37 Specific heat contents for selected nonrenewable energy sources.
8.5 Energy Substitution 339
Specific carbon content (g C/kWh)

100
90
80
70
60
50
40
30
20
10
0
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil
Coal and coke Natural gas Petroleum products

FIGURE 8.38 Specific carbon contents for selected nonrenewable energy sources.
Specific CO2 emission (kg CO2/kWh)

0.40
0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil

Coal and coke Natural gas Petroleum products

FIGURE 8.39 Carbon dioxide emission per unit energy content associated with combustion for selected
nonrenewable energy sources.

two energy inputs. The annual heating requirement of the house is taken to be
30,000 kWh (see Table 8.7).
The following assumptions are invoked to simplify the analyses (Rosen et al.,
1996):
• The overall fuel-to-electricity conversion efficiency for all fuel types of is
38%. This efficiency actually depends on the fuel type and the system
used, for example, a conventional coal-fired power plant typically has a
much lower efficiency than a gas turbine plant. A common conversion
340 C HA PT E R 8 : Exergy-Related Methods

Specific CO2 emission (kg CO2/kg of fuel)


4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0

Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil
Coal and coke Natural gas Petroleum products

FIGURE 8.40 Carbon dioxide emission per unit fuel mass associated with combustion for selected
nonrenewable energy sources.

4.0 Specific CO2 emission Specific heat content 16


Specific CO2 emission (kg CO2/kg)

Specific heat content (kWh/kg)


3.5 14
3.0 12
2.5 10
2.0 8
1.5 6
1.0 4
0.5 2
0.0 0
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil
Coal and coke Natural gas Petroleum products

FIGURE 8.41 Comparison of specific heat content and specific CO2 emission from combustion, on a
mass basis, for selected nonrenewable energy sources.

efficiency, roughly based on the average of efficiencies for different


systems, is used for all fuels here to simplify the analyses.
• The overall fuel-to-heat efficiency for the direct heating process for all
types of fuel is 90%. This efficiency also depends on the type of fuel and
conversion system in practice, but a constant value is used, again to
simplify the analyses.
• The efficiency of an electric resistance heating is 100%, which is
reasonable since the conversion of electricity to heat incurs only very
minor losses.

Three options are considered for the energy input to the house:
8.5 Energy Substitution 341

Table 8.7 Annual Fuel Requirement, Rate of CO2 Emission and CO2 Emission Reduction Through
Different Options in a Typical House
Annual CO2 Emissions to Heat a Typical
House Having Heat Load 30,000 kWh/year
Option 1: Option 2: Option 3: CO2
CO2 CO2 Emission
Emission Emission Reduced by
CO2 While Reduced Using Natural
Emission Fuel Producing in Direct Gas in Direct
(kg CO2/ Required Electricity Heating Heating
Source Fuel Type kWh) (kg/year) (kg/year) (kg/year) (kg/year)

Coal and Anthracite 0.353 3700 27,900 16,100 5700


coke Bituminous 0.319 3730 25,200 14,500 4600
Subbituminous 0.331 5400 26,100 15,100 5000
Lignite 0.329 6540 25,960 15,000 4900
Coke 0.348 3740 27,500 15,900 5600
Natural gas Natural gas 0.181 2370 14,300 8300 0
Petroleum Still gas 0.228 2220 18,000 10,400 1600
products Kerosene 0.257 2080 20,300 11,700 2500
LPG 0.215 2260 17,000 9800 1100
Propane (liquid) 0.21 2100 16,600 9600 960
Ethane 0.214 3200 16,900 9800 1100
Ethylene 0.23 4570 18,100 10,500 1600
Isobutane 0.221 2370 17,500 10,100 1300
Butane 0.222 2300 17,600 10,100 1400
Natural gasoline 0.228 2500 18,000 10,400 1600
Special naphtha 0.247 2380 19,500 11,300 2200
Petroleum coke 0.349 6040 27,600 15,900 5600
Heavy gas oils 0.256 2430 20,200 11,700 2500
Light oil 0.26 2560 20,500 11,900 2600

1. The annual rate of CO2 emissions if fuel is used for electricity generation
with an overall fuel-to-electricity conversion efficiency of 38%, and
then, the electricity is supplied to an electric heater for space heating
with an efficiency of 100%
2. Reduction in annual rate of CO2 emissions, relative to option 1, if fuel is
used directly for space heating with a 90% efficiency
3. Reduction in annual rate of CO2 emissions, relative to option 1, if the
various conventional fuels that are currently used for HVAC are
substituted with natural gas and natural gas is used directly for space
heating with an efficiency of 90%
342 C HA PT E R 8 : Exergy-Related Methods

In option 1, fuel is converted to electricity, which is used for space heating. In


option 2, fuel is used directly for space heating and the reduction in annual CO2
emissions relative to option 1 is calculated. In option 3, all fuels are substituted
with natural gas, which is used directly for space heating, and the reductions in
annual CO2 emissions relative to option 2 are calculated. The annual CO2 emis-
sions are compared in Table 8.7 for the three options.
The quantity of each type of fuel required, the CO2 emissions while producing
electricity for HVAC systems, and the CO2 emissions reduced while using fuel
directly for heating are determined for a one-year period (see Table 8.7). It can
be seen in Table 8.7 that the quantity of fuel required is greatest for lignite
(6540 kg/year) due to its low heat content and lowest for kerosene
(2100 kg/year) due to its high heat content. The reduction in annual CO2 emis-
sions when using fuel directly for heating is also calculated for various fuels (see
Fig. 8.42). The results show that CO2 emissions per year can be greatly reduced
by using fuel for direct heating. The largest reduction in CO2 emissions per year
occurs for anthracite (16,100 kg/year) when it is used directly for heating in
place of electric heating.
The reductions in CO2 emissions per year are also determined from substituting
various fuels with natural gas in direct heating (see Fig. 8.43). The results show
that annual CO2 emissions can be greatly reduced by replacing coal, coke, and
petroleum coke with natural gas and somewhat reduced by replacing other fos-
sil fuels with natural gas.

CO2 emission reduced in direct heating


CO2 emission reduction rate (kg/year)

18,000 CO2 emission reduced while using natural gas in direct heating
16,000
14,000
12,000
10,000
8000
6000
4000
2000
0
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil

Coal and coke Natural gas Petroleum products

FIGURE 8.42 Annual CO2 emission reduction with direct heating and reduction by replacing with natural
gas in direct heating, for various fuels.
8.6 Integration of Energy Systems 343
Reduction in CO2 emission rate by
natural gas substitution (kg/year)

6000
5000
4000
3000
2000
1000
0
Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil
Coal and coke Natural gas Petroleum products

FIGURE 8.43 Annual reduction in CO2 emissions from substituting natural gas with other nonrenewable
fuels, in direct heating.

8.6 INTEGRATION OF ENERGY SYSTEMS


Enhancing the performance and efficiency of energy systems is an important
challenge for the energy sector. Integration of energy systems is one way to
address this challenge and especially to enhance overall efficiency. There are var-
ious ways of integrating energy systems and they can be categorized in many
ways, such as based on the output from the integrated system. Figure 8.40 shows
the various ways of integrating energy systems for HVAC, based on the number of
outputs, namely, cogeneration, trigeneration, and ultimately multigeneration.
Cogeneration is a technique that is used to produce two useful outputs with one
process, often with increased efficiency and reduced environmental impacts
than would be the case using a separate energy system for each useful output.
Cogeneration often is taken to be the same as combined power and heat (CHP),
but more generally, it can apply to a process for any two coproducts. Here, we
focus on the CHP interpretation. Cogeneration applications are numerous and
vary from large-scale industrial uses to small onsite residential or commercial
building systems. Cogeneration is a proved and reliable technology, which is
often more cost-effective and efficient than separate systems for heat and elec-
tricity. Many industries reject a significant amount of waste heat, which can be
used to produce steam or hot water for space and hot water heating and/or
space cooling using absorption chiller. Since thermal energy is one product
of cogeneration, it is possible to integrate the system with HVAC to produce
space heating or cooling. In addition, the other product, electricity, can be used
to drive electric heaters, air conditioners, or heat pumps.
Similarly, trigeneration is a technique for producing three useful outputs
(electricity, heat, and cooling) simultaneously from one or more energy inputs.
344 C HA PT E R 8 : Exergy-Related Methods

Energy systems
integration for HVAC
Ways of integration

Cogeneration Trigeneration Multigeneration

Outputs

Heating and Heating and Heating and cooling Heating and cooling
Heating and cooling Heating and cooling
cooling power and power and drying and ventilation and
and power and ventilation
etc. power etc.

Heating and Cooling and Heating and power Cooling and power
ventilation ventilation Heating and power Cooling and power
and ventilation and ventilation
and ventilation and ventilation
and drying etc. and drying etc.
Cooling and
power

Primary energy sources

Solar
Energy sources
Coal
Biomass

Oil Fossil fuel Renewable


Geothermal

Natural gas
Wind

FIGURE 8.44 Possible means of energy systems integration for HVAC.

The possible useful trigeneration outputs from a HVAC perspective are shown
in Fig. 8.44. When more than three outputs are produced in an integrated sys-
tem, it is usually referred to as a multigeneration system (see Fig. 8.44). The pri-
mary energy source for such systems can be fossil fuels or renewable energy or a
combination of both. Figure 8.44 shows the major primary energy sources for
cogeneration, trigeneration, and multigeneration systems as applicable
to HVAC.

8.7 ALLOCATION OF ENVIRONMENTAL EMISSIONS


Allocating emissions for energy systems having multiple inputs and multiple
products can be challenging. This activity is important as many governing bod-
ies are defining guidelines for emission trading, which allocate emissions based
on product outputs or on some other basis. Possible methods for allocating
emissions for the process of CHP have been developed (Rosen, 2008). There
is no single emission allocation method that is universally accepted, so work
is ongoing in this field. These emission allocation methods can be complex
and difficult to apply and challenging to translate into policy. For instance,
for CHP systems, output-based allocation methods can be based on efficiency,
work potential, and heat content, and the results obtained with such methods
8.7 Allocation of Environmental Emissions 345

are often not consistent. As multigeneration systems are more complex, the
application of emission allocation methods for them is a greater challenge.
A further use of these methods may be the allocation of carbon credits, since air
pollution and greenhouse gas (GHG) emission avoidance schemes may result
in carbon credits, which will need to be distributed appropriately. To provide
an understanding of this field and its state of development, emission allocation
methods for CHP systems are described in this section.

8.7.1 CHP Systems


CHP systems are characterized by the simultaneous production of electricity
and useful heat from one input, often a fossil fuel. These systems are usually
very efficient, sometimes over 80% on an energy basis, mainly because the
waste heat is delivered as a product. Although electrical output is somewhat cur-
tailed in shifting from electricity generation to cogeneration, the net output of
the plant increases greatly. The increased efficiency usually permits CHP sys-
tems to exhibit reduced overall energy inputs and emissions. The heat from
CHP can be used for direct domestic hot water or space heating, district heating,
and industrial process heating. The heat can also be used for cooling by using it
to drive absorption chillers. The capacity and other characteristics of a CHP
plant can be decided in various ways and depend on the nature of the electrical
and thermal loads being served. CHP can be used to supply a base load (elec-
trical or thermal) or it may follow the temporal variation of one of these loads.
CHP plants are sometimes coupled with energy storage systems to overcome
problems due to fluctuating loads.

8.7.2 Exergy Analysis


Exergy is loosely the work potential of a system or, more rigorously, the max-
imum amount of work that can be produced from a system or flow as it comes
to equilibrium with a reference environment. The reference environment is
specified in terms of its temperature, pressure, and chemical composition.
The magnitude of exergy depends on the state of a system or flow and that
of the reference environment, so the results of an exergy analysis are dependent
on the selection of the reference environment. Exergy analysis applies the sec-
ond law of thermodynamics along with the conservation of mass and energy
principles for design and improvement efforts. Unlike energy, exergy is not con-
served. Exergy analysis leads to more meaningful understanding of thermody-
namic performance than energy analysis. In particular, exergy analysis
identifies and quantifies the margin available to improve the efficiency of pro-
cesses and systems by reducing inefficiencies. Exergy analysis is thought by
many to evaluate thermodynamic performance better than energy analysis
346 C HA PT E R 8 : Exergy-Related Methods

because it provides more insights and is more useful in efficiency-improvement


efforts. One important application area of exergy analysis is CHP.

8.7.3 Carbon Dioxide Allocation Methods for CHP


Numerous methods have been developed for allocating carbon dioxide emis-
sions from CHP systems to the electrical and thermal energy products, which
can be utilized for HVAC processes like space heating or other tasks. Such
methods seek to allocate carbon dioxide emissions so as to ensure equitable
credits for different stakeholders like owners and users of thermal and electrical
energy. Similarly, methods have been developed for allocating fuel use, and
these methods are somewhat related to allocation methods for emissions.
For carbon dioxide emissions, the relation is relatively straightforward, as the
fuel allocation can be multiplied by a carbon dioxide emission factor.
This section describes six technically and/or economically based allocation
methods for carbon dioxide from CHP and is drawn from an earlier report
by Rosen (2008). The allocation methods for carbon dioxide from CHP are pre-
sented here in terms of fractions of carbon dioxide emissions allocated to each
product. Of course, emissions can also be allocated to the products of a CHP
plant based on the agreement between the stakeholders involved in such pro-
jects or based on other factors than those considered in this section.
Allocation by product energy. Allocations are made in proportion to the energy of
the products. That is, the fractional allocation of carbon dioxide to the product
electricity fE is determined as
E
fE ¼ (8.1)
ðE + H Þ

and the fractional allocation of carbon dioxide to the product thermal energy fH
is determined as
Q
fH ¼ (8.2)
ðE + H Þ

Here, E and H are the electrical and thermal energy product outputs, respec-
tively. The term H is either the direct heat energy transferred or the net energy
transferred in a heat exchanger. Although simple, this method neglects energy
quality. It also ignores the fact that the input fuel quantities required to produce
the same quantity (units of energy) of electrical and thermal energy differ, so
appropriate shares of emissions allocated to the products are not obtained.
Allocation by product exergy. Allocations are made in proportion to the exergy of
the products:
ExE
fE ¼ (8.3)
ðExE + ExH Þ
8.7 Allocation of Environmental Emissions 347

ExH
fH ¼ (8.4)
ðExE + ExH Þ

where ExE and ExH are the respective net exergy outputs of the CHP electrical
and thermal products. The electrical exergy and energy are the same, but the
thermal exergy corresponding to the thermal energy output is evaluated by mul-
tiplying the exergetic temperature factor Ht:
To
Ht ¼ 1  (8.5)
T

Here, T and To denote the of thermal product and reference environment tem-
peratures, respectively. The variation of Ht with T is shown in Table 8.8. When
the thermal energy is delivered via heat exchangers, the exergy delivered ExH is
evaluated by exergy difference of incoming and outgoing flows. Note that the
exergy method results are sensitive to the choice of reference environment. This
is often characterized by ambient conditions, but not always. In one study
(Strickland and Nyober, 2002), for example, the reference environment tem-
perature was set to the process temperature of 100 °C. For consistency of anal-
ysis, a uniform reference environment temperature is useful.
The exergy-based method accounts for both the quantity and quality of energy.
This method is more logical in approach than the energy allocation method
discussed previously.
Note that there can be significant differences in the energy and exergy values of
thermal energy products. These are compared for thermal energy in Figs. 8.45
and 8.46, which show the variation in energy, exergy, and exergy-to-energy
ratio with change in operating temperature for space heating and space cooling,
respectively. The exergy values correspond to 100 J of thermal energy and are
calculated across a typical range of operating temperatures for HVAC systems.
Note that the ratio of the exergy and energy varies from zero to greater than
unity as the temperature of heat transfer varies from reference environment
temperature to negative temperatures. It is observed for heating that the exergy
values vary from zero to unity when the temperature of heat transfer varies from
reference environment temperature to infinity and that when temperatures are

Table 8.8 Variation in Exergetic Temperature Factor With Temperature


for a Reference Environment Temperature of T0 ¼ 300 K
Thermal Product Temperature, T (K) Exergetic Temperature Factor, Ht

400 0.25
600 0.50
1200 0.75
348 C HA PT E R 8 : Exergy-Related Methods

120 0.35

100 0.3

Energy and exergy (kJ)


0.25

Exergy to energy ratio


80 Energy (kJ)
Exergy (kJ)
0.2
Ratio of exergy to energy
60
0.15

40
0.1

20
0.05

0 0
288 293 313 333 353 373 393 413
Temperature (K)
FIGURE 8.45 Variation in energy, exergy, and exergy-to-energy ratio with temperature for space
heating.

120 0.00

100 −0.05
Energy and exergy (kJ)

80

Exergy to energy ratio


−0.10
60
Energy (kJ) −0.15
40 Exergy (kJ)
Ratio of exergy to energy −0.20
20
−0.25
0

−20 −0.30

−40 −0.35
283 273 263 253 243 233 223
Temperature (K)
FIGURE 8.46 Variation in energy, exergy, and exergy-to-energy ratio with temperature for space
cooling.

below the reference environment temperature, the exergy values are negative,
meaning the exergy associated with thermal energy transfer is input to the sys-
tem to facilitate cooling. That is, the energy and exergy transfers are in the oppo-
site direction.
8.7 Allocation of Environmental Emissions 349

Allocation by product economic value. Allocations are made in proportion to the


economic values of the products:
cE E
fE ¼ (8.6)
ðcE E + cQ HÞ

cH H
fH ¼ (8.7)
ð cE E + cQ H Þ

Here, cE and cH denote, respectively, the unit economic values of the CHP elec-
trical and thermal energy products. The economic value is a general economic
measure, which may reflect price, production cost, or some other value. Note
with this method that when the economic values are based on the energy, the
exergy-based values can be used to allocate the respective fractions for carbon
credits.
This method provides a simple way for effective decision making for owners of
CHP plants. It does not require economic values of both products and can
instead use the ratio of values to allocate the carbon credits to the respective
products as follows:

E
fE ¼   (8.8)
E + HðcE =cH Þ1

E
fH ¼ (8.9)
½E ðcE =cH Þ + H

Here, cE/cH denotes the ratio of the unit economic value for electricity to that for
thermal energy. The economic value is generally lower for the thermal product,
so usually, cE =cH > 1.
Allocation based on incremental fuel consumption to electrical production. Emissions
are allocated based on the fuel consumed by electrical and thermal products,
with the electrical product considered the by-product of thermal energy pro-
duction. The fuel consumed in thermal energy production is evaluated inde-
pendently (e.g., the boiler efficiency is taken into account if the thermal
energy is in the form of steam), as follows:
H
FH ¼ (8.10)
ηb

where FH is the fuel consumed to produce thermal energy and ηb is the effi-
ciency of a device that produces the same thermal energy as the CHP system.
Then, the fuel consumed for electricity generation is evaluated, by subtracting
the amount of fuel required for the thermal energy:
FE ¼ F  FH (8.11)
350 C HA PT E R 8 : Exergy-Related Methods

Consequently, the emission allocation fractions are evaluated as follows:

FH H
fH ¼ ¼ (8.12)
F Fηb

FE
fE ¼ ¼ 1  fH (8.13)
F

The aforementioned method is consistent with the “fuel charge to power”


(FCP) method, which is commonly applied by consulting firms.
Allocation based on incremental fuel consumption to thermal energy production.
Emissions are evaluated by calculating the fuel consumed in the CHP system
for electrical and thermal products. The main difference with the previous
method is that the thermal energy is taken as by-product of electricity genera-
tion. The share of fuel consumption used to produce electricity is calculated
using efficiency values of an independent system (e.g., a reference power plant):

E
FE ¼ (8.14)
ηpp

where ηpp represents the efficiency of a device that produces the same thermal
energy as the CHP system. The fuel consumed in thermal energy production is
calculated as

FH ¼ F  FE (8.15)

Then, the emission allocations are determined as the respective ratios of fuel
consumed in the thermal and electrical energy production:
FE E
fE ¼ ¼ (8.16)
F Fηpp

FH
fH ¼ ¼ 1  fE (8.17)
F

Allocation based on shared emission savings. Allocations are based on the indepen-
dent calculation of fuel used in producing electrical and thermal products. The
fuel amounts are calculated for both products using reference efficiencies (e.g.,
a steam boiler and a power plant):

H
FH ¼ (8.18)
ηb

E
FE ¼ (8.19)
ηpp
8.8 Utilization of Renewable and Sustainable Energy 351

Then, emission allocations are calculated as follows:


! !
FE E E H
fE ¼ ¼ + (8.20)
F ηpp ηpp ηb
  !
FH H E H
fH ¼ ¼ + (8.21)
F ηb ηpp ηb

This method allocates emissions similarly to methods mentioned in the previ-


ous section and does not treat the products as primary or secondary.

8.8 UTILIZATION OF RENEWABLE AND


SUSTAINABLE ENERGY
Renewable energy sources are those that are replenished at the same rate at
which they are used. The main renewable energy sources that can be used to
drive HVAC systems are solar, wind, hydro, geothermal, wave, and biomass.
Further details of these energy sources are shown in Fig. 8.47.

8.8.1 Major Renewable Energy Sources


Renewable energy can be used for HVAC, as shown in Fig. 8.47. Renewable
energy can also be exploited to offset HVAC needs through building design,
for example, making use of natural heating, cooling, and ventilation where

Renewable sources
of energy for HVAC

Solar Wind Geothermal Hydro Wave Biomass

PV Micro
Solid Liquid Gaseous
Solar thermal Mini

MCSW Kraft black Biogas


Small liquor

Wood & WR Ethanol Land fill gas


Large

Agricultural Biodiesel
by-products

Solid
by-products

PV, Photovoltaic; MCSW, Municiple Solid Waste; WR, Wood Residuals

FIGURE 8.47 Breakdown of nonrenewable energy sources utilizable for HVAC.


352 C HA PT E R 8 : Exergy-Related Methods

possible. Electricity generation efficiencies and life cycle CO2 emissions are
given in Table 8.9 for various types of renewable energy. The variation in the
overall energy efficiencies shown in Table 8.9 is due to changes in the environ-
mental conditions and the type of conversion system used. For example, in the
case of hydropower, the overall efficiency of the power plant depends on the
type of turbine used and mass flow rate of water, while in the case of geothermal
energy, the overall efficiency depends on the temperature of water.
The specific CO2 emissions in Table 8.9 are seen to be much smaller for
renewable rather than no-renewable resources. The minimum, maximum,
and 50th percentile life cycle-specific CO2 emissions from several renewable
energy sources are adapted from a study by the IPCC (2011). There is a large
difference between the minimum and maximum values of the life cycle-
specific CO2 emission for renewable sources because in the study, a number
of references are considered (based on a literature review of LCAs of GHG
emissions from electricity generation), and these often involve different
parameter values. The 50th percentile life cycle-specific CO2 emission is high-
est for solar PVs (46 g CO2/kWh) and lowest for hydroelectric power
(4 g CO2/kWh). Thus, from the perspective of environmental impact due
to carbon dioxide emissions, it may be advantageous to utilize hydroelectric
power to drive HVAC systems.
One type of renewable energy for HVAC, biomass, is described in detail here, as
an illustration of the amount of detailed information that can be obtained and
is required in assessments.

Table 8.9 Energy Efficiencies and Life Cycle Carbon Dioxide Emissions for
Selected Renewable Energy Sources
Life Cycle CO2 Emission per Unit of
Electricity Produced
Overall
Energy 50th
Energy Efficiency Minimum Percentile Maximum
Source (%) (g CO2/kWh) (g CO2/kWh) (g CO2/kWh)

Wind 30-45 2 12 81
Solar thermal 15-25 5 22 217
Solar PV 12-22 7 46 89
Hydroelectric 85-95 0 4 43
Geothermal 20-35 6 45 79
Ocean energy 2-4 2 8 23
Adapted from IPCC (2011).
8.8 Utilization of Renewable and Sustainable Energy 353

8.8.2 Biomass
Biomass is a biological material derived from living or recently living organ-
isms. It can be used directly as energy source or converted to other fuels like
biogas or biodiesel. Biomass can be separated into three categories: solid, liq-
uid, and gaseous (see Fig. 8.47). The specific heat content, the carbon content,
and the unit CO2 emissions for various biomass types are listed in Table 8.10.
Biomass can be used for applications like heating, cooling, and electricity gen-
eration (Filho and Badr, 2004) and is often beneficial, especially due to its
availability and low operating cost. Biogas and biodiesel are often the best
suited types of biomass for driving HVAC systems due their high energy con-
tents and low environmental impacts.

Table 8.10 Biomass Fuels Types, Specific Heat Contents and Unit Carbon Emissions
Heat Carbon Content CO2 Emission per CO2 Emission
Form of Biomass Fuel Content per Unit Energy Unit Energy per Unit Mass
Biomass Type (kWh/kg) (g C/kWh) (kg CO2/kWh) (kg CO2/kg)

Solid Municipal solid 3.21 84.42 0.309 0.99


waste
Tires 8.68 80.01 0.293 2.55
Plastics 12.28 69.78 0.256 3.14
Petroleum coke 9.69 95.30 0.349 3.39
(solid)
Wood and wood 4.97 87.28 0.320 1.59
residuals (12%
MC)
Agricultural 2.67 109.97 0.403 1.07
by-products
Peat 2.58 104.07 0.382 0.99
Solid 8.34 98.20 0.360 3.00
by-products
Liquid Kraft black liquor 3.87 87.86 0.322 1.25
(hardwood)
Kraft black liquor 3.95 88.55 0.325 1.28
(softwood)
Ethanol (100%) 8.24 63.70 0.234 1.93
Biodiesel (100%) 11.26 68.72 0.252 2.84
Gas Biogas (captured 12.80 48.45 0.178 2.27
methane)
Landfill gas (50% 7.65 48.45 0.178 1.36
CH4/50%CO2)
Abbreviation: MC, moisture content.
354 C HA PT E R 8 : Exergy-Related Methods

Figure 8.48 compares the specific heat contents of various biomass fuels, in
terms of HHV. The specific heat content is seen to be greatest for biogas
(12.80 kWh/kg) and lowest for peat (2.58 kWh/kg).
Carbon dioxide is produced when biomass is combusted, and Fig. 8.49
compares the carbon content per unit energy for various types of biomass.
The carbon content per unit energy is seen to be greatest for agricultural by-
products (109.97 g C/kWh) and lowest for biogas (captured methane)
(48.45 g C/kWh). The quantities of specific CO2 emissions from combustion
of various types of biomass are shown in Fig. 8.50 on an energy basis and
Fig. 8.51 on a mass basis.

14
Specific heat content (kWh/kg)

12
10
8
6
4
2
0
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

(12% MC)

Agricultural by-products

Peat

Solid by-products
Wood and wood residuals

(hardwood)

(softwood)

Ethanol (100%)

Biodiesel (100%)

Biogas (captured
methane)
Landfill gas (50%
CH4/50%CO2)
Kraft black liquor

Kraft black liquor


Solid Liquid Gas

FIGURE 8.48 Specific heat contents of various biomass fuels.

120
Specific carbon content (g C/kWh)

100
80
60
40
20
0
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

(12% MC)

Agricultural by-products
Wood and wood residuals

Peat

Solid by-products

Kraft black liquor


(hardwood)
Kraft black liquor
(softwood)

Ethanol (100%)

Biodiesel (100%)

Biogas (captured
methane)
Landfill gas (50%
CH4/50%CO2)

Solid Liquid Gas

FIGURE 8.49 Carbon content per unit energy for various types of biomass.
8.8 Utilization of Renewable and Sustainable Energy 355
Specific CO2 emission (kg CO2 /kWh)

0.45
0.40
0.35
0.30
0.25
0.20
0.15
0.10
0.05
0.00
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

(12% MC)

Agricultural by-products
Wood and wood residuals

Peat

Solid by-products

Kraft black liquor


(hardwood)
Kraft black liquor
(softwood)

Ethanol (100%)

Biodiesel (100%)

Landfill gas (50%


CH4/50%CO2)
Biogas (captured methane)
Solid Liquid Gas

FIGURE 8.50 CO2 emission per unit energy from combustion for various types of biomass.
Specific CO2 emission (kg CO2/kg)

4.0
3.5
3.0
2.5
2.0
1.5
1.0
0.5
0.0
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

(12% MC)

Agricultural by-products

Solid by-products
Wood and wood residuals

Peat

Kraft black liquor


(hardwood)
Kraft black liquor
(softwood)

Ethanol (100%)

Biodiesel (100%)

Biogas (captured
methane)
Landfill gas (50%
CH4/50%CO2)

Solid Liquid Gas

FIGURE 8.51 CO2 emission per unit mass for various types of biomass.

The CO2 emission from combustion per unit energy production is observed in
Fig. 8.50 to be greatest for agricultural by-products (0.403 kg CO2/kWh) and
lowest for biogas (0.176 kg CO2/kWh). The CO2 emission from combustion
per unit mass is observed in Fig. 8.51 to be greatest for petroleum coke (solid)
(3.39 kg CO2/kg) and lowest for municipal solid waste (0.99 kg CO2/kg).
Figure 8.52 compares unit-specific heat contents and specific CO2 emissions for
biomass combustion, on mass bases. The ratio of CO2 emission per unit energy
is seen to be greatest for biogas and lowest for solid by-products.
356 C HA PT E R 8 : Exergy-Related Methods

4.0 Specific CO2 emission Specific heat content 16

Specific CO2 emission (kg CO2/kg)

Specific heat content (kWh/kg)


3.5 14
3.0 12
2.5 10
2.0 8
1.5 6
1.0 4
0.5 2
0.0 0

Anthracite

Bituminous

Subbituminous

Lignite

Coke

Natural gas

Still gas

Kerosene

LPG

Propane (liquid)

Ethane

Ethylene

Isobutane

Butane

Natural gasoline

Special naptha

Petroleum coke

Heavy gas oils

Light oil
Coal and coke Natural gas Petroleum products

FIGURE 8.52 Comparison of specific carbon contents and specific CO2 emissions for various types of
biomass, on mass bases.

8.8.3 Case Study for Renewable Energy Sources


The effect of substituting various fuels for HVAC systems on environmental
impact in terms of CO2 emissions per year is investigated. A similar approach
is used as in the case study of nonrenewable sources. A hypothetical house with
an assumed annual heating requirement of 30,000 kWh is considered (see
Table 8.11), and the same assumptions regarding device efficiencies are
invoked as in the case of case study for nonrenewable sources.
Four options, all involving biomass fuels, are considered for the energy input
to the house and comparisons are made in terms of CO2 emission per year in
each case:

1. Biomass is used for electricity generation, with an overall fuel-to-


electricity energy conversion efficiency of 38%, and the electricity is
supplied to an electric resistance space heater having an energy efficiency
of 100%.
2. Biomass is used directly for space heating, with a 90% overall fuel-to-heat
energy efficiency.
3. Biomass is replaced with natural gas, which is directly used for space
heating with a 90% overall fuel-to-heat energy efficiency.
4. Biomass is replaced with biogas prepared from biomass, which is directly
used for space heating with a 90% overall fuel-to-heat energy efficiency.

In option 1, biomass is converted to electricity, which is used for heating, and


the total CO2 emission per year is calculated. In option 2, biomass is directly
used for heating and the CO2 emission reduction per year compared to option
1 is calculated. In option 3, all the biomass is replaced with natural gas, which is
used for direct heating, in order to observe the effect on the CO2 emission
8.8 Utilization of Renewable and Sustainable Energy 357

Table 8.11 Annual Biomass Requirement and CO2 Emission and Emission Reduction for Various
Biomass Options in a Typical House
Annual CO2 Emissions to Heat a Typical House
Having a Heat Load of 30,000 kWh/year
Option 3:
CO2 Option 4:
Emission CO2
Option 1: Option 2: Reduced Emission
CO2 CO2 by Using Reduced
Emission Emission Natural by Using
While Reduced Gas in Biogas in
CO2 Biomass Producing in Direct Direct Direct
Form of Biomass Emission Required Electricity Heating Heating Heating
Biomass Type (kg CO2/kWh) (kg/year) (kg/year) (kg/year) (kg/year) (kg/year)

Solid Municipal 3.21 9330 24,400 14,100 4280 4390


solid waste
Tires 8.68 3460 23,200 13,400 3750 3850
Plastics 12.28 2440 20,200 11,700 2500 2600
Petroleum 9.69 3100 27,600 15,900 5620 5730
coke (solid)
Wood and 4.97 6040 25,300 14,600 4640 4750
wood
residuals
(12% MC)
Agricultural 2.67 11,300 31,800 18,400 7410 7520
by-products
Peat 2.58 11,600 30,100 17,400 6690 6800
Solid by- 8.34 3600 28,400 16,400 5970 6100
products
Liquid Kraft black 3.87 7750 25,400 14,700 4700 4820
liquor
(hardwood)
Kraft black 3.95 7600 25,600 14,800 4790 4900
liquor
(softwood)
Ethanol 8.24 3640 18,400 10,700 1750 1860
(100%)
Biodiesel 11.26 2660 19,900 11,500 2370 2500
(100%)
Gas Biogas 12.80 2340 14,000 8100 110 0.0
(captured
methane)
Landfill gas 7.65 3920 14026.6 8100 110 0.0
(50%CH4/
50%CO2)
358 C HA PT E R 8 : Exergy-Related Methods

reduction per year. The annual CO2 emission reduction compared to option 2
by substituting natural gas for biomass is calculated. In option 4, biomass is
replaced with biogas (prepared from biomass) and the biogas is used for direct
heating to assess the effect on CO2 emission reduction per year. The four
options are compared in terms of the quantity of CO2 emission per year (see
Table 8.11). We determine the quantity of each type of biomass required per
year, the CO2 emissions per year from producing electricity for HVAC systems,
the CO2 emissions per year while using biomass directly for heating, and the
reduction in CO2 emissions per year from using biomass directly for heating.
It can be seen in Table 8.11 that the greatest quantity of biomass required is for
agricultural by-product (11,300 kg/year) due to its low specific heat content
and the lowest quantity for biogas (2340 kg/year) due to its high specific heat
content. The reduction in CO2 emissions per year when using biomass directly
for heating is also calculated for various biomass types (see Fig. 8.53). The
results show that CO2 emissions per year can be greatly reduced by using bio-
mass for direct heating. The largest reduction in annual CO2 emissions occurs
for agricultural by-product (18,400 kg/year) when it is used directly for heating
in place of electric heating.
The reductions in CO2 emissions per year are also determined from substituting
various biomass types with natural gas and biogas in direct heating (see
Fig. 8.54). The results show that CO2 emissions per year can be greatly reduced
by replacing agricultural by-product, peat, and solid by-product with natural

CO2 emission while producing electricity


Annual CO2 emission and emission reduction

CO2 emission reduced in direct heating


35,000
30,000
25,000
rate (kg/year)

20,000
15,000
10,000
5000
0
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

Wood and wood residuals


(12% MC)

Agricultural by-products

Peat

Solid by-products

Kraft black liquor (NA


hardwood)

Kraft black liquor (softwood)

Ethanol (100%)

Biodiesel (100%)

Biogas (captured methane)

Landfill gas (50%


CH4/50%CO2)

Solid Liquid Gas

FIGURE 8.53 Annual CO2 emission from electricity production and CO2 emission reduction in direct
heating.
8.8 Utilization of Renewable and Sustainable Energy 359

CO2 emission reduced by using natural gas in direct heating


CO2 emission reduced by using biogas in direct heating
Annual CO2 emission reduction rate (kg/year)

8000
7000
6000
5000
4000
3000
2000
1000
0
Municipal solid waste

Tires

Plastics

Petroleum coke (solid)

(12% MC)
Wood and wood residuals

Agricultural by-products

Kraft black liquor

Biogas (captured methane)


Peat

Solid by-products

Kraft black liquor (NA


hardwood)

(softwood)

Ethanol (100%)

Biodiesel (100%)

Landfill gas (50%


CH4/50%CO2)
Solid Liquid Gas

FIGURE 8.54 Annual CO2 emission reduction by substituting natural gas and biogas in place of different
biomass sources, in direct heating.

gas and biogas and somewhat reduced by replacing other biomass types with
natural gas and biogas. Natural gas is a nonrenewable energy source but the
results nonetheless suggest that natural gas is more suitable than most biomass
types for direct heating in terms of the quantity of CO2 emission. But in terms of
cost and renewability, biomass is more suitable.

8.8.4 Comparative Study of Energy Consumption


in Three Residential Buildings
The effects of changing the energy source for HVAC systems, and of building
design, are investigated through a case study. In the case study, three different
hypothetical buildings and several energy sources for them are considered. The
effects of various combinations of building design and energy source on envi-
ronmental impact measures are analyzed.
For accurate and meaningful analyses, it is necessary to quantify for the build-
ings the energy requirements and their breakdown by application. Although
the energy needs of a building vary spatially and with building characteristics,
three typical buildings are selected for the current study.
The electricity demand of each building is assumed to be constant (20 kW). A
breakdown of the energy use in the building is shown in Fig. 8.55. For the pre-
sent case study, only HVAC loads of the buildings, namely, space heating, space
cooling, and ventilation loads, are considered. These are 56%, 5%, and 9% of
the total load, respectively.
360 C HA PT E R 8 : Exergy-Related Methods

60

Portion of energy use in building (%)


50

40

30

20

10

0
Common Common House Space Lighting Space Water
auxilary auxilary appliances cooling heating heating
equipment motors

FIGURE 8.55 Breakdown of energy use in an average residential building (U.S. EIA, 2001).

8.8.4.1 Description of Residential Buildings


Building 1 is taken to be an old building located where weak environmental
standards are applied during design and construction. Poor insulation is uti-
lized and the HVAC technology is old and inefficient. A coal-fired thermal
power plant is taken to be the electricity source for building 1 and the building
design allows for no natural space heating and cooling. A heat pump with a low
COP is used for space heating and cooling.
Building 2 is taken to be a new building for which modern standards are fol-
lowed during design and construction. Reflective materials are used with stan-
dard insulation levels to reduce radiative heat transfer losses. The building has a
high-efficiency HVAC system using modern technology. As for building 1, a
coal-fired thermal power plant is taken to be the electricity source and the
building design allows for no natural heating and cooling. A heat pump with
a high COP is used for heating and cooling.
Building 3 is taken to be a new building, for which modern standards are fol-
lowed during design and construction. Reflective materials are used with stan-
dard insulation levels to reduce radiative heat transfer losses. The building has a
high-efficiency HVAC system using modern technology. In this case, the build-
ing design allows some natural heating and cooling (20% heating and 15%
cooling) and full natural ventilation. The remaining heating and cooling load
is supplied through a solar PV system, as shown in Fig. 8.56. A heat pump with a
high COP is used for the remaining heating and cooling.
The parameter values used in the case studies for the three buildings are given
in Table 8.12. A heat pump is used for both heating and cooling in all buildings.
High environmental impact

– No solar heat input in


Building 1 (HVACsystem) winter
Heating (0%)
Heating (100%)
Electricity • Weak environmental standards applied Natural – No reduction in cooling
Thermal Cooling (0%) load in summer due to
power plant
Cooling (100%) • Low efficiency HVAC system building design
• Old technology Ventilation (0%)
Ventilation (100%)
• No use of reflective material and poor vinsulation – No natural ventilation
due to building design

Low environmental impact

8.8
– No solar heat input in
Building 2 (HVACsystem) winter

Utilization of Renewable and Sustainable Energy


Heating (0%)
Heating (100%)
Electricity
• Modern standard followed Natural – No reduction in cooling
Thermal Cooling (0%) load in summer due to
Cooling (100%) • High efficiency HVAC system building design
power plant
• Modern technology Ventilation (0%)
Ventilation (100%)
• Use of reflective material and proper insulation – No natural ventilation
due to building design

Zero environmental impact

– Solar heat input in


Building 3 (HVACsystem) winter
Heating (80%) Heating (20%)
Solar Electricity
• Modern standard followed Natural – Reduction in cooling
Cooling (15%) load in summer due to
photovoltaic Cooling (85%) • High efficiency HVAC system building design
system • Modern technology Ventilation (100%)
Ventilation (0%)
• Use of reflective material and proper insulation – Natural ventilation due
to building design

FIGURE 8.56 Energy sources for HVAC systems for three residential buildings and their environmental impacts.

361
362 C HA PT E R 8 : Exergy-Related Methods

Table 8.12 Total Load Requirements and Other Design Parameters of Three Building Case Studies
Parameter Building 1 Building 2 Building 3

Input parameters
Electrical load of building (kW) 20 20 20
Space heating load (kW) 11.20 11.20 11.20
Space cooling load (kW) 1.0 1.0 1.0
Ventilation load (kW) 1.80 1.80 1.80
COP of heat pump in heating mode 2.50 5.0 5.0
COP of heat pump in cooling mode 1.5 4.0 4.0
Energy efficiency of mechanical ventilation fan system (%) 85 95 N/A
Overall energy efficiency of thermal power plant (%) 34 34 N/A
Energy efficiency of solar photovoltaic power plant (%) N/Aa N/A 22

Output parameters
Space heating load reduced through building design (kW) 0.0 0.0 2.20
Space cooling load reduced through building design (kW) 0.0 0.0 0.20
Ventilation load reduced through building design (kW) 0.0 0.0 1.80
Electricity demand of building HVAC (kW) 7.30 4.40 2.0
Actual energy required to produce electricity for HVAC (kW) 18.20 13.0 9.10
CO2 emission rate (g/s) 1.90 1.10 0.0
SO2 emission rate (g/s) 1.60 0.9 0.0
NOx emission rate (g/s) 0.008 0.006 0.0

Note that only operational CO2, SO2, and NOx emission rates are considered here, not full life cycle emissions.
a
N/A, not applicable.

The COP of the heat pump is considered for building 1 to be 2.5 in heating mode
and 1.5 in cooling mode and, for buildings 2 and 3, to be 5.0 in heating mode and
4.0 in cooling mode. The overall energy efficiency is taken to be 34% for the coal-
fired thermal power plant and 22% for the solar PV system. The energy efficiency of
the mechanical ventilation fan system is assumed to be 85% in building 1 and 95%
in building 2. There is no mechanical ventilation fan system in building 3, as there
is 100% natural ventilation in the building. For CO2, SO2, and NOx emissions,
unit values per kW of electricity production reported in the literature for a coal-
fired power plant are used (Mittal et al., 2012).
The examination of energy demand in the residential buildings (see Fig. 8.57)
shows that building 1 has the highest rate of electricity consumption
(7.30 kW), whereas building 3 has the lowest (2.0 kW). The high rate of elec-
tricity use exhibited by building 1 mainly occurs because its heat pump has a
low COP, while the rate of electricity consumption for building 3 is low because
a high-efficiency heat pump is used and natural heating, cooling, and ventilation
are incorporated. The differences in the overall energy efficiencies of the systems
8.8 Utilization of Renewable and Sustainable Energy 363

Rate of electricity consumption of building HVAC


20
Rate of electricity consumption and
Rate of energy required to produce required electricity
18
atcual energy required (kW)

16
14
12
10
8
6
4
2
0
Building 1 Building 2 Building 3
FIGURE 8.57 Rate of electricity consumption and actual energy requirement in three buildings.

cause their actual energy demands to vary. Figure 8.57 shows that the energy rate
needed to produce the required electricity is highest for building 1 (18.20 kW),
based on the 38% energy efficiency of the coal-fired power steam power plant,
and lowest for building 3 (9.10 kW). Note that this value is still a high proportion
of the rate of electricity consumption in building 3. This is mainly due to the low
overall energy efficiency of the solar PV system, which is taken to be 22%.
Figure 8.58 shows for the three buildings considered the rate of electricity con-
sumption and the CO2, SO2, and NOx emission rates. The rate of CO2 emission

1 8.0
Rate of electricity consumption of building HVAC
Rate of electricity consumption of building

0.9 CO2 emission rate


7.0
SO2 emission rate
0.8 NOx emission rate
6.0
0.7
Emission rate (g/s)

5.0
HVAC (kW)

0.6

0.5 4.0

0.4 3.0
0.3
2.0
0.2
1.0
0.1

0 0.0
Building 1 Building 2 Building 3
FIGURE 8.58 Rate of electricity consumption of HVAC system of buildings and respective CO2, SO2, and
NOx emission rates for the three buildings.
364 C HA PT E R 8 : Exergy-Related Methods

is observed in Fig. 8.58 to be 1.90 g/s for building 1, 1.10 g/s for building 2, and
0.0 g/s for building 3. Similarly, the SO2 emission rate is observed to be 1.6 g/s
for building 1, 0.9 g/s for building 2, and 0.0 g/s for building 3. The emission
rates are much lower for NOx than for CO2 and SO2 for buildings 1 and 2 and
zero for building 3. Figure 8.58 shows that building 3 has a zero emission rate
for CO2, SO2, and NOx, because all the electricity demand in building 3 is ful-
filled through the solar PV system and the building design.

8.9 CLOSING REMARKS


Various exergy-related methods that can be used to better understand and
improve HVAC processes and systems are described. Environmental impacts
are also considered, as these should be considered while designing HVAC sys-
tems. Emissions from HVAC in terms of CO2, SO2, and NOx emissions can be
significantly reduced by using high-efficiency HVAC systems and can be
reduced to zero by using natural heating, cooling, ventilation, and renewable
sources of energy. For comprehensive assessments, economic analyses of sys-
tem, and use of modern technology and building design having natural heating
and cooling, are advantageous. LCA, although challenging to analyze, gives a
clear picture of the environmental impact during the lifetime of a product or
process and where changes can be made to reduce the impact. Proper HVAC
retrofitting can lead to more efficient, reliable, and cost-effective systems. Retro-
fitting can be applied to the input energy source in order to reduce the levelized
cost of power generation for HVAC and environmental impact, and it is site-
specific. Emission allocation methods are discussed based on the outputs,
and of these methods, the exergy-based approach is found to be most consistent
and logical. Various nonrenewable and renewable energy sources available for
HVAC systems are described, including heat content, carbon content, and CO2
emissions for selected fuels. CO2 emissions are determined, considering both
conventional and renewable energy sources, and compared with different
options like electricity conversion and direct heating.

Nomenclature
cE unit economic value of the electrical product of CHP
cH unit economic value of the thermal product of CHP
C total CO2 emissions from CHP
CE CO2 emissions associated with electrical energy product produced by CHP
CH CO2 emissions associated with thermal energy product produced by CHP
E net output of electrical energy from CHP
ExE net output of electrical exergy from CHP
ExFE fuel exergy consumption associated with generating electricity by CHP
ExFH fuel exergy consumption associated with producing thermal exergy product by CHP
References 365

ExH net output of thermal exergy from CHP


fE fraction of CHP emissions allocated to electrical product
fH fraction of CHP emissions allocated to thermal product
F total primary fuel energy consumed by CHP system
FE fuel consumption attributed to electricity generation
FH fuel consumption attributed to the production of thermal energy
H net output of thermal energy from CHP
T temperature
To temperature of reference environment

Greek symbols
φ CO2 emission coefficient for a fuel
ηb energy efficiency of independent device (e.g., boiler) for thermal energy
ηE energy efficiency of generating electrical energy by CHP
ηH energy efficiency of producing thermal energy by CHP
ηPP energy efficiency of independent device (e.g., power plant) for electrical energy
ψE exergy efficiency of generating electricity by CHP
ψE exergy efficiency of producing thermal energy product by CHP
τ exergetic temperature factor

Acronyms
CHP combined heat and power
COP coefficient of performance
DFE design for environment
ExLCA exergetic life cycle assessment
HHV higher heating value
HVAC heating, ventilation, and air conditioning
LCA life cycle assessment
LCIA life cycle inventory analysis
LPG liquid petroleum gas
MCSW municipal solid waste
PV photovoltaic
SOC state of charge
VOC volatile organic compound
WR wood residuals

References
Balta, M.T., Dincer, I., Hepbasli, A., 2010. Performance and sustainability assessment of energy
options for building HVAC applications. Energy Build. 42 (8), 1320–1328.
Biden, J., 2009. Recovery through Retrofit, Middle Class Task Force Council on Environmental
Quality. EERE (Energy efficiency and renewable energy), U.S.. https://www.whitehouse.gov/
assets/documents/Recovery_Through_Retrofit_Final_Report.pdf (accessed 20.03.15).
Dartmouth, 2015. Design for the environment (DfE). http://engineering.dartmouth.edu/
d30345d/courses/engs171/DfE.pdf (accessed 22.03.15).
366 C HA PT E R 8 : Exergy-Related Methods

Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, second
ed. Elsevier, Oxford, UK.
EIA (Energy Information Administration), 2001. Residential energy consumption survey, U.S.
http://www.eia.gov/consumption/residential/data/2001/ (accessed 11.06.14).
ETP (Energy Technology Perspectives), 2010. Scenarios & Strategies to 2050. International Energy
Agency, France.
Filho, P.A., Badr, O., 2004. Biomass resources for energy in North-Eastern Brazil. Appl. Energy
77 (1), 51–67.
Graedel, T.E., Allenby, B.R., 1995. Industrial Ecology. Prentice Hall, Englewood Cliffs, NJ.
Graedel, T.E., Allenby, B.R., 2010. Industrial Ecology and Sustainable Engineering. Prentice Hall,
Upper Saddle River, NJ.
Grote, K.H., Antonsson, E.K., 2009. Applications in mechanical engineering, part B. Springer Hand-
book of Mechanical Engineering. Springer, New York.
Heikkila, K.N., 2004. Environmental impact assessment using a weighting method for alternative
air-conditioning systems. Build. Environ. 39 (10), 1133–1140.
IEA (International Energy Agency), 2013. 2013 Key World Energy Statistics. International Energy
Agency, France.
IPCC (Intergovernmental Panel on Climate Change), 2011. Special Report on Renewable Energy
Sources and Climate Change Mitigation. Cambridge University Press, United Kingdom.
Jørgensen, S.E., Odumb, H.T., Brown, M.T., 2004. Energy and exergy stored in genetic information.
Ecol. Model. 178, 11–16.
Mittal, M.L., Sharma, C., Singh, R., 2012. Estimates of emissions from coal fired thermal power
plants in India. In: 20th Emission Inventory Conference 2012, Tampa, Florida.
Prek, M., 2004. Environmental impact and life cycle assessment of heating and air conditioning
systems, a simplified case study. Energy Build. 36 (10), 1021–1027.
Rosen, M.A., 2004. Energy considerations in design for environment: appropriate energy selection
and energy efficiency. Int. J. Green Energy 1, 21–45.
Rosen, M.A., 2008. Allocating carbon dioxide emissions from cogeneration systems: descriptions of
selected output-based methods. J. Clean. Prod. 16 (2), 171–177.
Rosen, M.A., Sy, E., Gharghouri, P., 1996. Substituting natural gas heating for electric heating:
assessment of the energy and environmental effects in Ontario. Energy Stud. Rev. 8 (2),
143–154.
Strickland, C., Nyober, J., 2002. Cogeneration Potential in Canada: Phase 2. Report for Natural
Resources Canada. MK Jaccard and Associates, Canada.
TCR (The Climate Registry), 2013. 2013 climate registry default emission factors. Table 12.1 U.S.
default factors for calculating CO2 emissions from fossil fuel and biomass combustion, U.S.
Viral, P.S., Debella, D.C., Ries, R.J., 2008. Life cycle assessment of residential heating and cooling,
systems in four regions in the United States. Energy Build. 40, 503–513.
Wikipedia, 2015. Green retrofit. http://en.wikipedia.org/wiki/Green_retrofit (accessed 20.03.15).
Appendix A

GLOSSARY OF SELECTED TERMINOLOGY


This glossary identifies exergy-related terminology from the literature and cat-
egorizes the terms by area. Most exergy terminology has only recently been
adopted and is still evolving. Often more than one name is assigned to the same
quantity and more than one quantity to the same name. Only exergy-related
definitions are given for terms having multiple meanings. The glossary is based
in part on previously developed broader glossaries (Kotas et al., 1987; Kotas,
1995; Kestin, 1980; Dincer and Rosen, 2011, 2013).

GENERAL THERMODYNAMIC TERMS


Control mass A closed system containing a fixed quantity of matter, in which no matter enters
or exits.
Control volume An open system in which matter is allowed to enter and/or exit.
Entropy A measure of disorder, which always increases for the universe.
Heat A form of energy transfer between systems due to a temperature difference. Heat is a flow
quantity, that is, energy in transit. By convention in analysis, heat input to a system is consid-
ered positive, while heat exiting is negative.
Heat capacity Ratio of the heat absorbed in a substance to the resulting increase in temperature.
The change in temperature depends on the heating process, with the most common being con-
stant volume or constant pressure.
Internal energy Sum of all forms of microscopic energy for matter.
Irreversible process A process in which both the system and its surroundings cannot be returned
to their initial state(s) through a subsequent reversible process.
Kinetic energy Energy of a system as a result of a change in its motion relative to a reference frame.
Latent energy Internal energy associated with a phase change of a system.
Potential energy (gravitational) Energy of a system as a result of a change of its elevation relative
to a reference frame in a gravitational field.
Process An action that results in the change in the state of a system.
Property Any characteristic of a system.

367
368 Appendix A

Reversible process A process in which both the system and its surroundings can be returned to
their initial state(s) with no observable effects.
Sensible energy Internal energy of a system associated with a change in the kinetic energies of its
molecules, without phase change.
State The condition of a system specified by the values of its properties.
System A quantity of matter or any region of space (also thermodynamic system).
Work A form of energy transfer. Thermodynamic work can be in various forms, for example,
mechanical, electrical, and magnetic. By convention in analysis, work done on a system is con-
sidered negative, while work done by the system is positive.

EXERGY QUANTITIES
Available energy See Exergy.
Available work See Exergy.
Availability See Exergy.
Base enthalpy The enthalpy of a compound (at To and Po) evaluated relative to the stable com-
ponents of the reference environment (i.e., relative to the dead state).
Chemical exergy The maximum work obtainable from a substance when it is brought from the
environmental state to the dead state by means of processes involving interaction only with
the environment.
Essergy See Exergy. Derived from essence of energy.
Exergy (1) A general term for the maximum work potential of a system, stream of matter, or a heat
interaction in relation to the reference environment as the datum state. Also known as available
energy, availability, essergy, technical work capacity, usable energy, utilizable energy, work
capability, and work potential. (2) The unqualified term exergy or exergy flow is the maximum
amount of shaft work obtainable when a steady stream of matter is brought from its initial state
to the dead state by means of processes involving interactions only with the reference
environment.
Negentropy A quantity defined such that the negentropy consumption during a process is equal to
the negative of the entropy creation. Its value is not defined, but is a measure of order.
Nonflow exergy The exergy of a closed system, that is, the maximum net usable work obtainable
when the system under consideration is brought from its initial state to the dead state by means
of processes involving interactions only with the environment.
Physical exergy The maximum amount of shaft work obtainable from a substance when it is
brought from its initial state to the environmental state by means of physical processes involv-
ing interaction only with the environment. Also known as thermomechanical exergy.
Technical work capacity See Exergy.
Thermal exergy The exergy associated with a heat interaction, that is, the maximum amount of
shaft work obtainable from a given heat interaction using the environment as a thermal energy
reservoir.
Thermomechanical exergy See Physical exergy.
Usable energy See Exergy.
Useful energy See Exergy.
Utilizable energy See Exergy.
Work capability See Exergy.
Work potential See Exergy.
Xergy See Exergy.
Appendix A 369

EXERGY CONSUMPTION, ENERGY DEGRADATION,


AND IRREVERSIBILITY
Degradation of energy The loss of work potential of a system that occurs during an irreversible
process.
Dissipation See Exergy consumption.
Entropy creation See Entropy production.
Entropy generation See Entropy production.
Entropy production A quantity equal to the entropy increase of an isolated system (associated
with a process) consisting of all systems involved in the process. Also known as entropy creation
and entropy generation.
Exergy consumption The exergy consumed or destroyed during a process due to irreversibilities
within the system boundaries. Also known as dissipation, irreversibility, and lost work.
External irreversibility The portion of the total irreversibility for a system and its surroundings
occurring outside the system boundary.
Internal irreversibility The portion of the total irreversibility for a system and its surroundings
occurring within the system boundary.
Irreversibility (1) An effect that makes a process nonideal or irreversible. (2) See Exergy
consumption.

ENVIRONMENT AND REFERENCE ENVIRONMENT


Dead state The state of a system when it is in thermal, mechanical, and chemical equilibrium
with a conceptual reference environment (having intensive properties pressure Po, temperature
To, and chemical potential μioo for each of the reference substances in their respective
dead states).
Environment See Reference environment.
Environmental state The state of a system when it is in thermal and mechanical equilibrium with
the reference environment, that is, at pressure Po and temperature To of the reference
environment.
Ground state See Reference state.
Reference environment An idealization of the natural environment that is characterized by a per-
fect state of equilibrium, that is, the absence of any gradients or differences involving pressure,
temperature, and chemical, kinetic, and potential energy. The environment constitutes a natu-
ral reference medium with respect to which the exergy of different systems is evaluated.
Reference state A state with respect to which values of exergy are evaluated. Several reference states
are used, including environmental state, dead state, standard environmental state, and standard
dead state. Also known as ground state.
Reference substance A substance with reference to which the chemical exergy of a chemical ele-
ment is calculated. Reference substances are often selected to be common, valueless environ-
mental substances of low chemical potential.
Resource A material found in nature or created artificially in a state of disequilibrium with the
environment.
Restricted equilibrium See Thermomechanical equilibrium.
Thermomechanical equilibrium Thermal and mechanical equilibrium.
Unrestricted equilibrium Complete (thermal, mechanical, and chemical) equilibrium.
370 Appendix A

EFFICIENCIES AND OTHER MEASURES


Effectiveness See Second-law efficiency.
Energy efficiency An efficiency determined using ratios of energy. Also known as thermal effi-
ciency and first-law efficiency.
Energy grade function The ratio of exergy to energy for a stream or system.
Exergetic temperature factor A dimensionless function of the temperature T and environmental
temperature To given by (1 To/T).
Exergy efficiency A second-law efficiency determined using ratios of exergy.
First-law efficiency See Energy efficiency.
Rational efficiency A measure of performance for a device given by the ratio of the exergy asso-
ciated with all outputs to the exergy associated with all inputs.
Second-law efficiency A general name for any efficiency based on a second-law analysis (e.g.,
exergy efficiency, effectiveness, utilization factor, rational efficiency, and task efficiency). Often
loosely applied to specific second-law efficiency definitions.
Task efficiency See Second-law efficiency.
Thermal efficiency See Energy efficiency.
Utilization factor See Second-law efficiency.

ENERGY AND EXERGY METHODS


Energy analysis A general name for any technique for analyzing processes based solely on the first
law of thermodynamics. Also known as first-law analysis.
Exergy analysis An analysis technique in which process performance is assessed by examining
exergy balances. A type of second-law analysis.
First-law analysis See Energy analysis.
Second-law analysis A general name for any technique for analyzing process performance based
solely or partly on the second law of thermodynamics.

ECONOMICS AND EXERGY


Exergoeconomics See Thermoeconomics.
Thermoeconomics A technoeconomic method for assessing and designing systems and processes
that combines economics with exergy parameters.

References
Dincer, I., Rosen, M.A., 2011. Thermal Energy Storage: Systems and Applications, second ed. Wiley,
London.
Dincer, I., Rosen, M.A., 2013. Exergy: Energy, Environment and Sustainable Development, second
ed. Elsevier, Oxford, UK.
Kestin, J., 1980. Availability: the concept and associated terminology. Energy-The International
Journal 5, 679–692.
Kotas, T.J., 1995. The Exergy Method of Thermal Plant Analysis. Krieger, Malabar, Florida.
Kotas, T.J., Raichura, R.C., Mayhew, Y.R., 1987. Nomenclature for exergy analysis. In: Moran, M.J.,
Sciubba, E. (Eds.), Second Law Analysis of Thermal Systems. Amer. Soc. Mech. Engineers, New
York, pp. 171–176.
Appendix B

CONVERSION FACTORS

Table B.1 Conversion Factors for Commonly Used Quantities


Quantity SI to English English to SI

Area 1 m2 ¼ 10.764 ft2 1 ft2 ¼ 0.00929 m2


¼ 1550.0 in.2 1 in.2 ¼ 6.452  104 m2
Density 1 kg/m3 ¼ 0.06243 lbm/ft3 1 lbm/ft3 ¼ 16.018 kg/m3
1 slug/ft3 ¼ 515.379 kg/m3
Energy 1 J ¼ 9.4787  104 Btu 1 Btu ¼ 1055.056 J
1 cal ¼ 4.1868 J
1 lbf ft ¼ 1.3558 J
1 hp h ¼ 2.685  106 J
Energy per unit mass 1 J/kg ¼ 4.2995  104 Btu/lbm 1 Btu/lbm ¼ 2326 J/kg
Force 1 N ¼ 0.22481 lbf 1 lbf ¼ 4.448 N
1 pdl ¼ 0.1382 N
Gravitation g ¼ 9.80665 m/s2 g ¼ 32.17405 ft/s2
Heat flux 1 W/m2 ¼ 0.3171 Btu/h ft2 1 Btu/h ft2 ¼ 3.1525 W/m2
1 kcal/h m2 ¼ 1.163 W/m2
1 cal/s cm2 ¼ 41 870.0 W/m2
Heat generation (volume) 1 W/m3 ¼ 0.09665 Btu/h ft3 1 Btu/h ft3 ¼ 10.343 W/m3
Heat transfer coefficient 1 W/m2 K ¼ 0.1761 Btu/h ft2 °F 1 Btu/h ft2 °F ¼ 5.678 W/m2 K
1 kcal/h m2 °C ¼ 1.163 W/m2 K
1 cal/s m2 °C ¼ 41870.0 W/
m2 K
Heat transfer rate 1 W ¼ 3.4123 Btu/h 1 Btu/h ¼ 0.2931 W
Length 1 m ¼ 3.2808 ft 1 ft ¼ 0.3048 m
¼ 39.370 in. 1 in. ¼ 2.54 cm ¼ 0.0254 m
1 km ¼ 0.621 371 mile 1 mile ¼ 1.609344 km
1 yd ¼ 0.9144 m
Mass 1 kg ¼ 2.2046 lbm 1 lbm ¼ 0.4536 kg
1 ton (metric) ¼ 1000 kg 1 slug ¼ 14.594 kg
1 grain ¼ 6.47989  105 kg

Continued
371
372 Appendix B

Table B.1 Conversion Factors for Commonly Used Quantities Continued


Quantity SI to English English to SI

Mass flow rate 1 kg/s ¼ 7936.6 lbm/h 1 lbm/h ¼ 0.000126 kg/s


¼ 2.2046 lbm/s 1 lbm/s ¼ 0.4536 kg/s
Power 1 W ¼ 1 J/s ¼ 3.4123 Btu/h ¼ 0.737562 Ibf ft/s 1 Btu/h ¼ 0.2931 W
1 hp (metric) ¼ 0.735499 kW 1 Btu/s ¼ 1055.1 W
1 ton of refrig. ¼ 3.51685 kW 1 lbf ft/s ¼ 1.3558 W
1 hpUK ¼ 745.7 W
Pressure and stress 1 Pa ¼ 0.020886 lbf/ft2 1 lbf/ft2 ¼ 47.88 Pa
(Pa ¼ N/m2) ¼ 1.4504  104 lbf/in.2 1 lbf/in.2 ¼ 1 psi ¼ 6894.8 Pa
¼ 4.015  103 in water 1 stand. atm. ¼ 1.0133  105 Pa
¼ 2.953  104 in Hg 1 bar ¼ 1  105 Pa
Specific heat 1 J/kg K ¼ 2.3886  104 Btu/lbm °F 1 Btu/lbm °F ¼ 4187 J/kg K
Surface tension 1 N/m ¼ 0.06852 lbf/ft 1 lbf/ft ¼ 14.594 N/m
1 dyn/cm ¼ 1  103 N/m
Temperature T(K) ¼ T(°C) + 273.15 T(°R) ¼ 1.8T(K)
¼ T(°R)/1.8 ¼ T(°F) + 459.67
¼ [T(°F) + 459.67]/1.8 ¼ 1.8T(°C) + 32
T(°C) ¼ [T(°F)  32]/1.8 ¼ 1.8[T(K)  273.15] + 32
Temperature difference 1 K ¼ 1 °C ¼ 1.8 °R ¼ 1.8 °F 1 °R ¼ 1 °F ¼ 1 K/1.8 ¼ 1 °C/1.8
Thermal conductivity 1 W/m K ¼ 0.57782 Btu/h  ft °F 1 Btu/h ft °F ¼ 1.731 W/m K
1 kcal/h m °C ¼ 1.163 W/m K
1 cal/s cm °C ¼ 418.7 W/m K
Thermal diffusivity 1 m2/s ¼ 10.7639 ft2/s 1 ft2/s ¼ 0.0929 m2/s
1 ft2/h ¼ 2.581  105 m2/s
Thermal resistance 1 K/W ¼ 0.52750°F h/Btu 1 °F h/Btu ¼ 1.8958 K/W
Velocity 1 m/s ¼ 3.2808 ft/s 1 ft/s ¼ 0.3048 m/s
1 km/s ¼ 0.62137 mile/h 1 ft/min ¼ 5.08  103 m/s
Viscosity (dynamic) 1 kg/m s ¼ 0.672 lbm/ft s 1 lbm/ft s ¼ 1.4881 kg/m s
(kg/ms ¼ Ns/m2) ¼ 2419.1 lbm/fh h 1 lbm/ft h ¼ 4.133  104 kg/m s
1 centipoise (cP) ¼ 102 poise
¼ 1  103 kg/m s
Viscosity (kinematic) 1 m2/s ¼ 10.7639 ft2/s 1 ft2/s ¼ 0.0929 m2/s
¼ 1  104 stokes 1 ft2/h ¼ 2.581  105 m2/s
1 stoke ¼ 1 cm2/s
Volume 1 m3 ¼ 35.3134 ft3 1 ft3 ¼ 0.02832 m3
1 L ¼ 1 dm3 ¼ 0.001 m3 1 in.3 ¼ 1.6387  105 m3
1 galUS ¼ 0.003785 m3
1 galUK ¼ 0.004546 m3
Volumetric flow rate 1 m3/s ¼ 35.3134 ft3/s 1 ft3/s ¼ 2.8317  102 m3/s
¼ 1.2713  105 ft3/h 1 ft3/min ¼ 4.72  104 m3/s
1 ft3/h ¼ 7.8658  106 m3/s
1 galUS/min ¼ 6.309  105 m3/s
Appendix C

THERMOPHYSICAL PROPERTIES

Table C.1 Thermophysical Properties of Pure Water at Atmospheric Pressure


T (°C) ρ (kg/m3) μ × 103 (kg/m s) ν × 106 (m2/s) k (W/m K) β × 105 (1/K) cp (J/kg K) Pr

0 999.84 1.7531 1.7533 0.5687 6.8140 4209.3 12.976


5 999.96 1.5012 1.5013 0.5780 1.5980 4201.0 10.911
10 999.70 1.2995 1.2999 0.5869 8.7900 4194.1 9.2860
15 999.10 1.1360 1.1370 0.5953 15.073 4188.5 7.9910
20 998.20 1.0017 1.0035 0.6034 20.661 4184.1 6.9460
25 997.07 0.8904 0.8930 0.6110 20.570 4180.9 6.0930
30 995.65 0.7972 0.8007 0.6182 30.314 4178.8 5.3880
35 994.30 0.7185 0.7228 0.6251 34.571 4177.7 4.8020
40 992.21 0.6517 0.6565 0.6351 38.530 4177.6 4.3090
45 990.22 0.5939 0.5997 0.6376 42.260 4178.3 3.8920
50 988.04 0.5442 0.5507 0.6432 45.780 4179.7 3.5350
60 983.19 0.4631 0.4710 0.6535 52.330 4184.8 2.9650
70 977.76 0.4004 0.4095 0.6623 58.400 4192.0 2.5340
80 971.79 0.3509 0.3611 0.6698 64.130 4200.1 2.2010
90 965.31 0.3113 0.3225 0.6759 69.620 4210.7 1.9390
100 958.35 0.2789 0.2911 0.6807 75.000 4221.0 1.7290
Source: Kukulka, D.J., 1981. Thermodynamic and Transport Properties of Pure and Saline Water (M.Sc. thesis). State University of
New York at Buffalo.

373
374 Appendix C

Table C.2 Thermophysical Properties of Air at Atmospheric Pressure


T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

200 1.7458 1.007 132.5 7.59 18.10 10.30 0.737


250 1.3947 1.006 159.6 11.44 22.30 15.90 0.720
300 1.1614 1.007 184.6 15.89 26.30 22.50 0.707
350 0.9950 1.009 208.2 20.92 30.00 29.90 0.700
400 0.8711 1.014 230.1 26.41 33.80 38.30 0.690
450 0.7740 1.021 250.7 32.39 37.30 47.20 0.686
500 0.6964 1.030 270.1 38.79 40.70 56.70 0.684
550 0.6329 1.040 288.4 45.57 43.90 66.70 0.683
600 0.5804 1.051 305.8 52.69 46.90 76.90 0.685
650 0.5356 1.063 322.5 60.21 49.70 87.30 0.690
700 0.4975 1.075 338.8 68.10 52.40 98.00 0.695
750 0.4643 1.087 354.6 76.37 54.90 109.00 0.702
800 0.4354 1.099 369.8 84.93 57.30 120.00 0.709
850 0.4097 1.110 384.3 93.80 59.60 131.00 0.716
900 0.3868 1.121 398.1 102.90 62.00 143.00 0.720
950 0.3666 1.131 411.3 112.20 64.30 155.00 0.723
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; and Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.

Table C.3 Thermophysical Properties of Ammonia (NH3) Gas at Atmospheric Pressure


T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

300 0.6994 2.158 101.5 14.70 24.70 16.66 0.887


320 0.6468 2.170 109.0 16.90 27.20 19.40 0.870
340 0.6059 2.192 116.5 19.20 29.30 22.10 0.872
360 0.5716 2.221 124.0 21.70 31.60 24.90 0.870
380 0.5410 2.254 131.0 24.20 34.00 27.90 0.869
400 0.5136 2.287 138.0 26.90 37.00 31.50 0.853
420 0.4888 2.322 145.0 29.70 40.40 35.60 0.833
440 0.4664 2.357 152.5 32.70 43.50 39.60 0.826
460 0.4460 2.393 159.0 35.70 46.30 43.40 0.822
480 0.4273 2.430 166.5 39.00 49.20 47.40 0.822
500 0.4101 2.467 173.0 42.20 52.50 51.90 0.813
520 0.3942 2.504 180.0 45.70 54.50 55.20 0.827
540 0.3795 2.540 186.5 49.10 57.50 59.70 0.824
560 0.3708 2.577 193.5 52.00 60.60 63.40 0.827
580 0.3533 2.613 199.5 56.50 63.68 69.10 0.817
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.
Appendix C 375

Table C.4 Thermophysical Properties of Carbon Dioxide (CO2) Gas at Atmospheric Pressure
T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

280 1.9022 0.830 140.0 7.36 15.20 9.63 0.765


300 1.7730 0.851 149.0 8.40 16.55 11.00 0.766
320 1.6609 0.872 156.0 9.39 18.05 12.50 0.754
340 1.5618 0.891 165.0 10.60 19.70 14.20 0.746
360 1.4743 0.908 173.0 11.70 21.20 15.80 0.741
380 1.3961 0.926 181.0 13.00 22.75 17.60 0.737
400 1.3257 0.942 190.0 14.30 24.30 19.50 0.737
450 1.1782 0.981 210.0 17.80 28.20 24.50 0.728
500 1.0594 1.020 231.0 21.80 32.50 30.10 0.725
550 0.9625 1.050 251.0 26.10 36.60 36.20 0.721
600 0.8826 1.080 270.0 30.60 40.70 42.70 0.717
650 0.8143 1.100 288.0 35.40 44.50 49.70 0.712
700 0.7564 1.130 305.0 40.30 48.10 56.30 0.717
750 0.7057 1.150 321.0 45.50 51.70 63.70 0.714
800 0.6614 1.170 337.0 51.00 55.10 71.20 0.716
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.

Table C.5 Thermophysical Properties of Hydrogen (H2) Gas at Atmospheric Pressure


T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

100 0.2425 11.23 42.1 17.40 67.00 24.60 0.707


150 0.1615 12.60 56.0 34.70 101.00 49.60 0.699
200 0.1211 13.54 68.1 56.20 131.00 79.90 0.704
250 0.0969 14.06 78.9 81.40 157.00 115.00 0.707
300 0.0808 14.31 89.6 111.00 183.00 158.00 0.701
350 0.0692 14.43 98.8 143.00 204.00 204.00 0.700
400 0.0606 14.48 108.2 179.00 226.00 258.00 0.695
450 0.0538 14.50 117.2 218.00 247.00 316.00 0.689
500 0.0485 14.52 126.4 261.00 266.00 378.00 0.691
550 0.0440 14.53 134.3 305.00 285.00 445.00 0.685
600 0.0404 14.55 142.4 352.00 305.00 519.00 0.678
700 0.0346 14.61 157.8 456.00 342.00 676.00 0.675
800 0.0303 14.70 172.4 569.00 378.00 849.00 0.670
900 0.0269 14.83 186.5 692.00 412.00 1030.00 0.671
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.
376 Appendix C

Table C.6 Thermophysical Properties of Oxygen (O2) Gas at Atmospheric Pressure


T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

100 3.9450 0.962 76.4 1.94 9.25 2.44 0.796


150 2.5850 0.921 114.8 4.44 13.80 5.80 0.766
200 1.9300 0.915 147.5 7.64 18.30 10.40 0.737
250 1.5420 0.915 178.6 11.58 22.60 16.00 0.723
300 1.2840 0.920 207.2 16.14 26.80 22.70 0.711
350 1.1000 0.929 233.5 21.23 29.60 29.00 0.733
400 0.9620 0.942 258.2 26.84 33.00 36.40 0.737
450 0.8554 0.956 281.4 32.90 36.30 44.40 0.741
500 0.7698 0.972 303.3 39.40 41.20 55.10 0.716
550 0.6998 0.988 324.0 46.30 44.10 63.80 0.726
600 0.6414 1.003 343.7 53.59 47.30 73.50 0.729
700 0.5498 1.031 380.8 69.26 52.80 93.10 0.744
800 0.4810 1.054 415.2 86.32 58.90 116.00 0.743
900 0.4275 1.074 447.2 104.60 64.90 141.00 0.740
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.

Table C.7 Thermophysical Properties of Water Vapor (Steam) Gas at Atmospheric Pressure
T (K) ρ (kg/m3) cp (J/kg K) μ × 107 (kg/m s) ν × 106 (m2/s) k × 103 (W/m K) a × 106 (m2/s) Pr

380 0.5863 2.060 127.1 21.68 24.60 20.40 1.060


400 0.5542 2.014 134.4 24.25 26.10 23.40 1.040
450 0.4902 1.980 152.5 31.11 29.90 30.80 1.010
500 0.4405 1.985 170.4 38.68 33.90 38.80 0.998
550 0.4005 1.997 188.4 47.04 37.90 47.40 0.993
600 0.3652 2.026 206.7 56.60 42.20 57.00 0.993
650 0.3380 2.056 224.7 66.48 46.40 66.80 0.996
700 0.3140 2.085 242.6 77.26 50.50 77.10 1.000
750 0.2931 2.119 260.4 88.84 54.90 88.40 1.000
800 0.2739 2.152 278.6 101.70 59.20 100.00 1.010
850 0.2579 2.186 296.9 115.10 63.70 113.00 1.020
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC; Borgnakke, C., Sonntag, R.E.,
1997. Thermodynamic and Transport Properties. Wiley, New York.
Appendix C 377

Table C.8 Thermophysical Properties of Some Solid Materials


Composition T (K) ρ (kg/m3) k (W/m K) cp (J/kg K)

Aluminum 273-673 2720 204-250 895


Asphalt 300 2115 0.0662 920
Bakelite 300 1300 1.4 1465
Brass (70% Cu + 30% Zn) 373-573 8520 104-147 380
Carborundum 872 – 18.5 –
Chrome brick 473 3010 2.3 835
823 – 2.5 –
Diatomaceous silica, fired 478 – 0.25 –
Fire clay brick 478 2645 1.0 960
922 – 1.5 –
Bronze (75% Cu + 25% Sn) 273-373 8670 26.0 340
Clay 300 1460 1.3 880
Coal (anthracite) 300 1350 0.26 1260
Concrete (stone mix) 300 2300 1.4 880
Constantan (60% Cu + 40% Ni) 273-373 8920 22-26 420
Copper 273-873 8950 385-350 380
Cotton 300 80 0.06 1300
Glass
Plate (soda lime) 300 2500 1.4 750
Pyrex 300 2225 1.4 835
Ice 253 – 2.03 1945
273 920 1.88 2040
Iron (C  4% cast) 273-1273 7260 52-35 420
Iron (C  0.5% wrought) 273-1273 7850 59-35 460
Lead 273-573 – – –
Leather (sole) 300 998 0.159 –
Magnesium 273-573 1750 171-157 1010
Mercury 273-573 13,400 8-10 125
Molybdenum 273-1273 10,220 125-99 251
Nickel 273-673 8900 93-59 450
Paper 300 930 0.18 1340
Paraffin 300 900 0.24 2890
Platinum 273-1273 21,400 70-75 240
Rock
Granite, Barre 300 2630 2.79 775
Limestone, Salem 300 2320 2.15 810
Marble, Halston 300 2680 2.80 830
Rubber, vulcanized
Soft 300 1100 0.13 2010

Continued
378 Appendix C

Table C.8 Thermophysical Properties of Some Solid Materials Continued


Composition T (K) ρ (kg/m3) k (W/m K) cp (J/kg K)

Sandstone, Berea 300 2150 2.90 745


Hard 300 1190 0.16 –
Sand 300 1515 0.27 800
Silver 273-673 10,520 410-360 230
Soil 300 2050 0.52 1840
Steel (C  1%) 273-1273 7800 43-28 470
Steel (Cr  1%) 273-1273 7860 62-33 460
Steel (18% Cr + 8% Ni) 273-1273 7810 16-26 460
Snow 273 110 0.049 –
Teflon 300 2200 0.35 –
Tin 273-473 7300 65-57 230
Tissue, human
Skin 300 – 0.37 –
Fat layer (adipose) 300 – 0.2 –
Muscle 300 – 0.41 –
Tungsten 273-1273 19,350 166-76 130
Wood, cross grain
Fir 300 415 0.11 2720
Oak 300 545 0.17 2385
Yellow pine 300 640 0.15 2805
White pine 300 435 0.11 –
Wood, radial
Fir 300 420 0.14 2720
Oak 300 545 0.19 2385
Zinc 273-673 7140 112-93 380
Source: Dincer, I., 1997. Heat Transfer in Food Cooling Applications. Taylor & Francis, Washington, DC;
Incropera, F.P., DeWitt, D.P., 1998. Fundamentals of Heat and Mass Transfer. Wiley, New York.
Index

Note: Page numbers followed by b Carnot, energetic, and exergetic


indicate boxes, f indicate figures and t COPs, 165, 165t B
indicate tables. design parameters, 155, 155t Balance equations
exergy destruction rate, 157 biomass-based system
absorber, 285
A exergy efficiencies, 157, 159f
combustion chamber, 282–283
Absorption chillers relative irreversibilities, 157, 159f
state properties and condenser, 284–285
CHP/CCHP systems, 184 evaporator, 285
double-effect, 206, 207f, 208, 210t thermodynamic data, 156,
157f, 158t generator, 284
function-based subsystems, 210, heat exchanger 1, 283–284
211t Air-to-air heat pumps, 142, 143f, 144
Air-to-water heat pumps, 142–144, solution heat exchanger, 286
heating and cooling sides, solution pump, 286
210–213, 212–213t 143f
Ambient air quality, 6 throttling valve 1, 285
single-effect, 205–206, 207f, 208, throttling valve 2, 286
210t Anova Verzekering Co. Building, 265
Aquifer thermal energy storage (ATES) solar and wind thermal system
thermal energy distribution compressor, 294
subsystem, 211, 212t balances and efficiency, 250–252,
252f condenser 1, 294
Adiabatic mixing, air streams, 83–86, evaporator 1, 295
83f, 84t, 85f buildings, 245–246, 245t
charging and discharging process, expansion valve, 294–295
Air source heat pumps flat plate collector, 292–294
exhaust (ventilation) air, 137–138, 249–250
charging-fluid and discharging- heater, 294
138t, 155 mass, energy, entropy, and exergy
air-handling duct system, 157, fluid flows, 254, 254f
energy losses, 248 rate balances, 302
160f Biomass
assumptions and simplifications, experimental discharge period,
254f, 255 agricultural by-products, 355, 355f
155–156 ambient temperatures, 289–290,
Carnot, energetic, and exergetic feasible commercial and
environment, 246 289–290f
COPs, 165, 165t balance equations
design parameters, 155, 155t hypothetical modification,
254–255 absorber, 285
exergy destruction rate, combustion chamber, 282–283
159–160 linear temperature-time relation,
253 condenser, 284–285
exergy efficiencies, 159–161, 161f evaporator, 285
mass, energy, entropy, and exergy mixing losses, 248
temperature and volumetric flow generator, 284
balance, 159 heat exchanger 1, 283–284
relative irreversibilities, 159–160, rate, 252f, 254
temperature-time profiles, 247, solution heat exchanger, 286
160f solution pump, 286
volumetric flow rate, 157 247f
threshold temperature, 246, 255 throttling valve 1, 285
outside ambient air, 138–139, 138t, throttling valve 2, 286
155 volumetric flow rates and charging
temperature, 252 carbon content, 353–355, 353t,
assumptions and simplifications, 354f, 356f
155–156 working principle, 245–246, 246f 379
380 Index

Biomass (Continued) CHP. See Cogeneration/combined discharging process, 260f, 261


electricity use, 279, 280f heat and power (CHP) energy efficiency, 261–262
energy basis, 354, 355f Chrysler’s New Technology exergy efficiency, 262
energy efficiency Development Center, 266 storing process, 260–261, 260f
heating cycle, 286–287 Coal-fired electricity-generating Combined cooling, heat, and power
system expression, 287 station, 30 (CCHP). See Trigeneration
vapor absorption chiller, 287 Coefficient of performances (COPs), Combustion-based process heating,
exergy destruction rates, 288–289, 132–134, 133f, 147–148, 184, 109–111
289f 184t, 207–208, 287 Community energy systems,
exergy efficiency Cogeneration/combined heat and 170–171
heating cycle, 287 power (CHP) Compressors
system expression, 287–288 advantages, 173–175, 176t HVACR
vapor absorption chiller, 287 conventional separate methods, air, 53–54, 54f
input and calculated process data, 173 energy efficiency, 53–54, 55f
288, 288t definition, 169 exergy efficiency, 53–54, 55f
mass basis, 354, 355f DHC, 205–214 isentropic efficiency, 53–54, 55f
petroleum coke and municipal diesel and natural gas engines, mass, energy, entropy, and exergy
solid waste, 355, 355f 174–175, 174t rate balances, 52
quantity, 357t, 358 disadvantage, 170, 174–175, 176t mechanical efficiency, 53
residential energy use, 279, 280f distributed energy approach, 175 motor efficiency, 53
solid, liquid and gaseous, 351f, 353 energy and exergy efficiencies, total compressor efficiency, 53
specific heat content, 353–354, 180–182, 180t vapor compression refrigeration
353t, 354f environmental emissions impact, system, 51–52, 52f
steam mass flow rate effect, 182 volumetric efficiency, 52
289–292, 291f fuel cells, 174–175, 174t integrated system, 89f, 90–91
system description, 281, 282f fuel cogeneration vs. fuel electrical Conversion factors, 371
thermodynamic analyses, 281 generation Cooling applications. See Cold
unit CO2 emissions, 353, 353t, and electrical heating, 179–180, thermal energy storage (CTES)
355, 356f 180f COPs. See Coefficient of performances
Borehole thermal energy storage and fuel heating, 177–178, (COPs)
(BTES), 234 177f, 178t CTES. See Cold thermal energy storage
Bruce Energy Centre, 112–113, 113f gas turbines, 174–175, 174t (CTES)
heat recovery system, 174–175
C microturbines, 174–175, 174t D
Carbon dioxide allocation methods nuclear cogeneration vs. nuclear Design for environment (DFE)
incremental fuel consumption, electrical generation and fuel clean energy sources, 310
349–350 heating, 178–179, 178t, 179f design level and objectives, 310,
product economic value, 349 operational flexibility, 175 311f
product energy, 346 power generator, 174–175 ERP matrix, 312, 313t
product exergy, 346–347, 347t, thermal energy product, 169–170 factors, 310–312
348f waste heat, 173 increased efficiency, 310
shared emission savings, 350–351 Cold thermal energy storage (CTES) methods and approaches, 312
Carnot heat pump, 147 advantages, 256–259 stages, 310, 311f
Centrifugal chillers chilled-water systems, 256–257 DHC. See District heating and cooling
COP, 207–208 distribution system, 256 (DHC)
function-based subsystems, 210, eutectic salts, 258 District energy systems, 170–171
211t ice systems, 257–258 District heating and cooling (DHC)
heating and cooling sides, mechanical system components, architectural design flexibility, 203
210–213, 212–213t 257 building/community level,
thermal energy distribution specified temperature data, 200–201
subsystem, 211, 212t 262–263, 263t central energy plant, 201
Chemical heat storage systems, thermodynamic analysis cogeneration-based district
222–223, 223t charging process, 259–260, 260f energy, 201
Index 381

chillers (see Absorption chillers; overall system, 287 with HOMER Pro, 327–328,
Centrifugal chillers) vapor absorption chiller, 287 328f
COP, 208 definition, 21 monthly average electricity
design for, 205 solar and wind thermal system production, 329, 329f
energy transfer rates, 209, 210t heating cycle, 302 monthly average heating load,
heat and electricity, 205, 206f heat pump, 295 325t, 327–328
Rosedale power plant, 207–208 overall system, 295–296, 302 photovoltaic (PV) system,
seasonal periods, 206–207, 207t vapor absorption chiller, 295 327–328
comfort and convenience, vapor compression chiller, 302 power output, 329–330, 330f
customers, 202 steady-state processes, 21 SOC, 329, 330f
decreased building capital costs, Energy efficiency ratio (EER), 133f, total net present cost, 328–329,
203 134 328f
ease of operation and maintenance, Energy methods wind
202 building sector, 8–10, 9t averages, 325, 325t
energy generation technologies and efficiency component breakdown,
fuel source, 203, 204t definition, 21 325–326, 326f
fuel flexibility, 202 steady-state processes, 21 HOMER Pro software, 325, 326f
improved energy efficiency, 201 energy consumption and CO2 monthly average electricity
layout of, 203, 203f emissions, 4, 5f production, 327, 327f
linking buildings and industrial environmental concerns, 6–7, 6t power output, 327, 328f
activities, 200–201 vs.exergy, 3–4, 3t SOC, 327, 327f
operation of, 203, 204f FLT, 2–3, 12–13 total net present cost, 325–326,
reduced environmental impact, 201 pollutants, 4–5, 5t 326f
reduced life-cycle costs, 202 residential sector, 8–9, 9f Energy substitution
reliability, 202 second law of thermodynamics, nonrenewable energy sources
Double-effect absorption chiller, 206, 13–14 advantages and disadvantages,
207f, 208, 210t sustainable development, 7–8, 7f 336
Energy retrofits annual CO2 emission reduction,
E complex measures, 324 341t, 342, 343f
Electric process heating, 111–112 diesel annual heating requirements,
Energy analysis with HOMER Pro, 330–331, 331f 337–339, 341t
ATES monthly average electricity classification, 336, 336–337f
balances and efficiency, production rate, 331–332, CO2 emission per unit energy
250–252, 252f 331f production, 337, 339f
charging and discharging, power output, 331–332, 332f CO2 emission per unit fuel mass,
249–250 total net present cost, 331, 331f 337, 340f
charging-fluid and discharging- gasoline, 332, 332–333f energy input, 340–342
fluid flows, 254, 254f homeowners and property fuel characteristics, 336, 338t
experimental discharge period, managers, 322–323 heat contents vs. CO2 emissions,
254f, 255 HOMER Pro microgrid, 324 336–337, 338f, 340f
hypothetical modification, house, characteristics of, 324, 325t specific carbon contents, 337,
254–255 online software analysis tool, 324 339f
linear temperature-time relation, outside and inside design renewable energy sources
253 temperature, 324, 325t advantages and disadvantages,
temperature and volumetric flow performance comparison, 336
rate, 252f, 254 333–334, 334t, 335f annual CO2 emission, 358, 358f
threshold temperature, 255 process, 323 annual heating requirement,
volumetric flow rates and proper and improper operation, 356, 357t
charging temperature, 252 322 biomass (see Biomass)
thermodynamic factor, 225 solar building design, 351–352, 351f
Energy efficiency averages, 325, 325t classification, 336, 336f
biomass-based system component breakdown, CO2 emission reduction,
heating cycle, 286–287 328–329, 329f 356–359, 358–359f
382 Index

Energy substitution (Continued) efficiency log mean temperature difference


energy efficiency and life cycle definition, 21 method, 44–45
carbon dioxide emissions, steady-state processes, 21 ε-NTU method, 45–46
351–352, 352t electricity, 17 Heating applications
residential buildings vs.energy, 3–4, 3t energy efficiency comparison
(see Residential buildings) environmental impact and ecology system X, 241f, 242
types, 351–352, 352t order destruction and chaos system Y, 241f, 242–243
Energy systems, 340f, 343–344, 344f creation, 32–33 exergy efficiency comparison
Environmental accidents, 6 ouroboros, 31–32 advantages, 245
Environmental emissions potential benefits, 34 borehole storage, 245
carbon dioxide allocation methods qualitative depiction, 35, 35f efficiency, 228t, 243
incremental fuel consumption, resource degradation, 33 system X, 243–244, 244t
349–350 waste exergy emissions, system Y, 243–244, 244t
product economic value, 349 33–34 heat storage material and
product energy, 346 exergy consumption, 17 properties, 236–237, 237t
product exergy, 346–347, 347t, FLT, 3, 12–13 inexpensive and good thermal
348f heat exchangers, 46–47 capacity factor, 237
shared emission savings, 350–351 matter flow, 16 iron shot, 237
CHP systems, 345 nonflow exergy, 15–16 rock, 238
exergy analysis, 345–346 open and closed flow balances, thermodynamic analysis
Environmentally responsible product 18–21 ATES (see Aquifer thermal energy
(ERP) matrix, 312, 313t SLT, 3, 13–14 storage (ATES))
Evaporative cooling, 86–87, 87f sustainable development, 35–37, charging process, 239, 239f
Exergetic life cycle assessment 35–37f discharging process, 239f, 240
(ExLCA), 320–321, 321–322f thermal energy, 16–17 energy and exergy efficiency,
Exergy efficiency work, 17 241
biomass-based system storing process, 239–240, 239f
heating cycle, 287 water, 237–238
overall system, 287–288 F Heating season performance factor
vapor absorption chiller, 287 First law of thermodynamics (FLT) (HSPF), 133f, 134–135
definition, 21 energy methods, 2–3, 12–13 Heating, ventilating, air conditioning,
solar and wind thermal system exergy methods, 3, 12–13 and refrigeration (HVACR)
heating cycle, 303 Fuel charge to power (FCP) method, adiabatic mixing of air streams,
heat pump, 296 350 83–86, 83f, 84t, 85f
overall system, 296, 303 air conditioning processes, 64,
vapor absorption chiller, 296 64f
vapor compression chiller, G compressors
302–303 Geothermal heat pumps, 141–142 air, 53–54, 54f
steady-state processes, 21 Ground source heat pump (GSHP), energy efficiency, 53–54, 55f
Exergy methods 132, 141–142, 222 exergy efficiency, 53–54, 55f
benefit of, 3–4, 10–11, 11–12f Ground-to-air heat pumps, 142, 143f, isentropic efficiency, 53–54, 55f
characteristics of, 15 144 mass, energy, entropy, and exergy
conceptual balances, 18 Ground-to-water heat pumps, 142, rate balances, 52
economic analyses, 28 143f, 144 mechanical efficiency, 53
coal-fired electricity-generating motor efficiency, 53
station, 30 H total compressor efficiency, 53
exergoeconomics, 29 Hazardous air pollutant, 6 vapor compression refrigeration
macro/micro economics, 28–29 Heat exchanger effectiveness ε, 45 system, 51–52, 52f
resource degradation, 30 Heat exchangers volumetric efficiency, 52
second law costing, 29 closed heat exchangers, 44, 44f cooling with dehumidification
thermoeconomic, 28–29 exergy efficiency, 46–47 definition, 77–82, 79f
types of, 29 feedwater heater, 47–49, 47–49f, energy/exergy efficiency, 80–82,
waste exergy emissions, 31 48t 81–82f
Index 383

rate balance equations, 79 Heat pumps, 165–166 classification


thermodynamic properties, 81, benefits, 131–132 storage capacity, 232–233, 233t
81t characteristics of, 151–152, 151t storage duration, 231
dead-state properties, 63, 63t chilling materials, 131 temperature range, 231–232,
evaporative cooling commercial and industrial 232t
exergy destruction rate, 65 applications, 107–109, 107f, UTES (see Underground thermal
fan efficiency, 54–56, 56f 108–109t, 151–154, 151t, 153t energy storage (UTES))
heat exchangers (see Heat efficiency measures control system, 221
exchangers) COP, 132–134, 133f CTES (see Cold thermal energy
heating with humidification energy efficiency ratio, 133f, 134 storage (CTES))
process HSPF, 133f, 134–135 electric heater/GSHP, 222
energy and exergy efficiencies, primary energy ratio, 133f, 134 energy analysis
75–77, 77–78f SEER, 133f, 135 ATES systems, 246
mass, energy, entropy, and exergy heat carrier, 135–136 charging, storing, and
rate balance equations, 74–75, heat source/heat sink discharging periods, 227, 228t
74f configurations, 136, 136f conservation of energy
thermodynamic properties, 76, air-to-air, 142, 143f, 144 principles, 227
76f, 76t air-to-water, 142–144, 143f conservation of mass, 227
integrated energy systems ground-to-air, 142, 143f, 144 definition, 227
cooling with dehumidification, ground-to-water, 142, 143f, 144 efficiency, 227, 228t, 229
194, 195f, 196 water-to-air, 142–144, 143f performance evaluations and
energy efficiency, 197–198 water-to-water, 142–143, 143f design, 227
evaporative cooling, 194, 196 preheating feedwater, 131 SLT, 227
exergy destruction, 199–200, primary heat pumps, 136 energy and exergy quantities,
199f, 200t producing hot water, 131 263–264, 263t
exergy efficiency, 197–198 residential applications, 151t, environmental impact
heating and cooling cycle, 152, 153f advantages, 229
193–194 secondary heat pumps, 136 California Energy Commission
heating with humidification, tertiary heat pumps, 136 data, 229–230
194, 196–197 types of heat sources, 132, reducing emissions, 229
input data and assumptions, 136f, 138t exergy analysis
194–196 air (see Air source heat pumps) ATES (see Aquifer thermal energy
schematic diagram, 194, 195f geothermal, 141–142 storage (ATES))
space cooling process, 194, 196 ground source, 141–142 charging, storing, and
space heating, 194, 196–197 solar, 142 discharging periods, 227, 228t
thermodynamic properties, 198, water (see Water source heat conservation of energy
199t pumps) principles, 227
material properties, 63, 63t vapor compression (see Vapor conservation of mass, 227
pump compression heat pump definition, 227
direct lift, 49 system) efficiency, 227, 228t, 229
displacement, and gravity, 49 Heat storage systems performance evaluations and
energy efficiency, 49–50, 51f advantages design, 227
exergy efficiency, 50, 51f cogeneration plants, 223 SLT, 227
frictional heat loss rate, 50 energy purchase shift, 223 UOIT (see Underground thermal
isentropic efficiency, 50, 50f functional integration, 223 energy storage (UTES))
mass, energy, entropy, and exergy increased system reliability, 223 functions, 221
rate balances, 49 increase generation capacity, 223 heat exchanger, 221
sensible cooling (see Sensible Anova Verzekering Co. Building, heating applications (see Heating
cooling) 265 applications)
sensible heating, 64f, 68–72, 72t, applications, 222 Kraft General Foods Headquarters
72–73f chemical, 222–223, 223t Building, 265–266
throttles (see Throttling valve) Chrysler’s New Technology latent, 222–223, 223t
turbines, 60–62, 60–62f Development Center, 266 off-peak power pricing, 222
384 Index

Heat storage systems (Continued) renewable energy sources, use of input data and assumptions,
PCM, 265 biomass, 102 194–196
San Francisco Marriott Hotel, geothermal energy, 102, 104 schematic diagram, 194, 195f
266–267 incentives for, 99–100 space cooling process, 194, 196
sensible, 222–223, 223t solar energy, 102–103 space heating, 194, 196–197
storage material, 221 Integrated heating system, 126–127 thermodynamic properties, 198,
Texas A&M University, 267 assumptions and simplifications, 199t
thermodynamic factors 114–115 multigeneration, 193
efficiency, 225 energy and exergy efficiency, 115, psychrometric processes
energy and exergy, 225 121 atmospheric air, 88
losses, 225 exergy destruction rate, 121–122, compressor, 89f, 90–91
reference-environment 122f condenser, 89f, 91
temperature, 226 heating devices cooling with humidification, 88,
storage duration, 226 boiler, 117 89f
temperature, 225 burner, 116 energy and exergy efficiencies,
thermal stratification, 225–226 furnace, 116–117, 117f 92–95, 94–95f
Heat transfer fluid (HTF), 113–114 heat exchanger, 118, 118f evaporative cooling, 88–89, 89f
HOMER Pro microgrid, 324 HTF, 113–114 exergy destruction, 91, 92t, 92f
HVACR. See Heating, ventilating, input and calculated process data, heating and cooling cycle, 87
air conditioning, and 120t heating with humidification,
refrigeration (HVACR) mass, energy, entropy, and exergy 88–90, 89f
balances, 115 space heating, 88, 89f, 90
I compressor, 119 throttling valve, 89f, 91
Indoor air quality, 6 condenser, 118
Industrial heating and cooling evaporator, 119 K
expansion valve, 119 Kraft General Foods Headquarters
systems Building, 265–266
classifications of, 99, 100f fan, 120
cryogenic processes, 99 merit and component, measure of, L
direct heating systems, 99 121t Land use and siting impact, 6
energy efficiency improvements, parametric analysis Latent heat storage systems, 222–223,
100 ambient temperature, effect of, 223t
exergy analysis, 101 123–124, 123–124f Life cycle assessment (LCA)
exergy conservation and efficiency condenser pressure and ExLCA, 320–321, 321–322f
improvement, 100 temperature, effect of, framework, 313, 314f
fossil fuels, 99–100 124, 125f, 125f goal and scope definition, 313–314
heat pump applications, 107–109, refrigerant and water impact assessment
107f, 108–109t temperature, 124–125, 126f flows and environmental
indirect heating systems, 99 solar thermal energy and heat impacts, 318, 318f
integrated solar and heat pump- pump, 113–114, 114f total CO2 emission, 316–317f,
based system (see Integrated Integrated systems 318–319, 319f
heating system) definition, 171–172 improvement analysis, 319–320
process heating, 99 for HVACR applications inventory analysis
combustion-based process cooling with dehumidification, data acquisition, 314
heating, 109–111 194, 195f, 196 subsystems and components,
data and energy and exergy energy efficiency, 197–198 314–316, 315t
efficiencies, 101, 102t evaporative cooling, 194, 196 time and effort-consuming
electric process heating, 111–112 exergy destruction, 199–200, phase, 314
low-to medium-temperature 199f, 200t total CO2 emissions, 316–317,
range, 104–106, 105t exergy efficiency, 197–198 316–317f
steam process heating, 112–113, heating and cooling cycle, total embodied energy, 315–317,
113f 193–194 316–317f
temperature ranges and energy heating with humidification, total embodied exergy, 316–317,
efficiencies, 101, 101t 194, 196–197 316–317f
Index 385

Linde–Hampson cycle Carnot, energetic, and exergetic petroleum coke and municipal
mass and energy balances, COPs, 165, 165t solid waste, 355, 355f
188–189 design parameters, 155, 155t solid, liquid and gaseous, 351f,
schematic of, 187–188, 188f evaporator system, 162, 163f 353
Logarithmic mean temperature exergy destruction rate, 163–164 specific heat content, 353–354,
difference method (LMTD), exergy efficiencies, 163–164, 164f 353t, 354f
44–45 mass, energy, entropy, and exergy unit CO2 emissions, 353, 353t,
balance, 163 355, 356f
relative irreversibilities, 163–164, biomass quantity, 357t, 358
M 164f building design, 351–352, 351f
Maritime pollution, 6 Ouroboros, 31–32 classification, 336, 336f
Mechanical efficiency, 53 CO2 emission reduction, 356–359,
Motor efficiency, 53
Multigeneration systems, 170, 171f P 358–359f
Primary energy ratio (PER), 133f, 134 energy efficiency and life cycle
Multistage binary power plant carbon dioxide emissions,
mass and energy balances, 188–189 Pump
direct lift, 49 351–352, 352t
schematic of, 187–188, 187f residential buildings
displacement, and gravity, 49
energy efficiency, 49–50, 51f actual energy requirement,
N exergy efficiency, 50, 51f 362–363, 363f
Nonrenewable energy sources frictional heat loss rate, 50 electricity consumption rate,
advantages and disadvantages, 336 isentropic efficiency, 50, 50f 362–364, 363f
annual CO2 emission reduction, mass, energy, entropy, and exergy energy use, 359, 360f
341t, 342, 343f rate balances, 49 and environmental impacts,
annual heating requirements, 360, 361f
337–339, 341t Q parameter values, 360–362, 362t
classification, 336, 336–337f Quadruple effect absorption system solar
CO2 emission per unit energy (QEAS) averages, 325, 325t
production, 337, 339f COPs, 190, 191–193f component breakdown,
CO2 emission per unit fuel mass, mass and energy balances, 328–329, 329f
337, 340f 188–189 with HOMER Pro, 327–328, 328f
diesel schematic of, 185, 186f monthly average electricity
with HOMER Pro, 330–331, 331f production, 329, 329f
monthly average electricity R monthly averages, 325t,
production rate, 331–332, Radiation and radioactivity, 6 327–328
331f Relative irreversibility (RI), 148–150 photovoltaic (PV) system,
power output, 331–332, 332f air source heat pumps 327–328
total net present cost, with circulating air, 157, 159f power output, 329–330, 330f
331, 331f with ventilation, 159–160, 160f SOC, 329, 330f
energy input, 340–342 water source heat pump total net present cost, 328–329,
fuel characteristics, 336, 338t open-loop, 163–164, 164f 328f
gasoline, 332, 332–333f submerged, 162, 162f types, 351–352, 352t
heat contents vs. CO2 emissions, Renewable energy sources wind
336–337, 338f, 340f advantages and disadvantages, 336 averages, 325, 325t
specific carbon contents, 337, 339f annual CO2 emission, 358, 358f component breakdown,
Number of exchanger transfer units annual heating requirement, 356, 325–326, 326f
(NTU), 45–46 357t HOMER Pro software,
biomass 325, 326f
agricultural by-products, 355, monthly average electricity
O 355f production, 327, 327f
Open-loop water source heat pump, carbon content, 353–355, 353t, power output, 327, 328f
155 354f, 356f SOC, 327, 327f
assumptions and simplifications, energy basis, 354, 355f total net present cost, 325–326,
155–156 mass basis, 354, 355f 326f
386 Index

Residential buildings Sensible heat storage systems, design parameters, 155, 155t
actual energy requirement, 222–223, 223t evaporator system, 161, 161f
362–363, 363f Single-effect absorption chiller, exergy destruction rate, 162
electricity consumption rate, 205–206, 207f, 208, 210t exergy efficiencies, 162, 163f
362–364, 363f SLA. See Second law analysis (SLA) mass, energy, entropy, and exergy
energy use, 359, 360f SLT. See Second law of balance, 162
and environmental impacts, 360, thermodynamics (SLT) relative irreversibilities, 162, 162f
361f Solar and wind thermal system
parameter values, 360–362, 362t ambient temperatures, 298–299, T
Resource degradation, 30, 33 298–299f, 303–304, 304–305f Texas A&M University, 267
Rosedale power plant in Edmonton, balance equations Thermal energy storage (TES).
207–208 compressor, 294 See Heat storage systems
condenser 1, 294 Thermophysical properties, 373
S evaporator 1, 295 Throttling valve, 58–59, 58f
Seasonal energy efficiency ratio expansion valve, 294–295 capillary tube, 57
(SEER), 133f, 135 flat plate collector, 292–294 constant pressure or automatic, 57
Second law analysis (SLA) heater, 294 energy efficiency, 58–59, 59f
boiler, 23, 24f mass, energy, entropy, and exergy exergy efficiency, 58–59, 59f
coal-fired boiler, 26, 26t rate balances, 302 float valve, 57
entropy analysis, 22, 22t, 23f energy efficiency integrated system, 89f, 91
exergy analysis, 22, 22t, 23f heating cycle, 302 mass, energy, entropy, and exergy
exergy consumption analysis, 22, heat pump, 295 rate balances, 57
22t, 23f overall system, 295–296, 302 pressure reduction, 56–57
exergy rate balances, 25 vapor absorption chiller, 295 refrigeration systems, 57
first law analysis, 24–26 vapor compression chiller, 302 thermostatic expansion valve, 57
implications of, 27 exergy destruction rates, 296, 297f, Trigeneration
negentropy analysis, 22, 22t, 23f 303, 304f absorption cooling systems,
physical exergy analysis, 22, 22t, 23f exergy efficiency 184–185
Second law of thermodynamics (SLT), heating cycle, 303 advantage, 183–184
3–4, 13–14 heat pump, 296 components, 184
Sensible cooling process overall system, 296, 303 definition, 170
cooling coil, 65–66 vapor absorption chiller, 296 double-effect absorption cooling,
definition, 65 vapor compression chiller, 185
dry bulb temperature, 66 302–303 QEAS (see Quadruple effect
energy/exergy efficiencies generator outlet temperatures, absorption system (QEAS))
ambient temperature, 67–68, 69f 299–300, 299–300f schematic of, 185, 186f
definition, 66–67 input and calculated process data, VHTG (see Very-high-temperature
humid air, 67 296, 297t, 303, 303t generator (VHTG))
relative humidity, 67–68, 69f system description, 292, 293f, Turbines, 60–62, 60–62f
mass, energy, entropy, and exergy 300–301, 301f
rate balance equations, 66 Solar energy
psychrometric chart, 64f, 69 heat pumps, 142 U
thermodynamic properties, 67, 68t industrial heating and cooling, Underground thermal energy storage
Sensible heating process 102–103 (UTES)
definition, 68 Solid wastes and their disposal, 6 applications, 236, 236t
electric resistance heating coil, 68 State of charge (SOC), 327, 327f, 329, aquifer storage, 234
energy/exergy efficiencies, 72f 330f borehole storage, 234
ambient temperature, 72, 73f Steam process heating, 112–113, 113f cavern and pit storage, 234
definition, 70–71 Submerged water source heat pump, characteristics and efficiency, 234,
relative humidity, 72, 73f 155 235t
saturated humid air, 71–72 assumptions and simplifications, feasibility analysis, 235–236
psychrometric chart, 64f, 69 155–156 heat storage technology, 234, 235t
rate balance equations, 70 Carnot, energetic, and exergetic operating cost analysis, 236
thermodynamic properties, 71, 72t COPs, 165, 165t temperature range, 233
Index 387

University of Ontario Institute of open-loop water source heat pump Very-high-temperature heat exchanger
Technology (UOIT), 245 (see Open-loop water source (VHHX), 186–187, 190–192
UTES. See Underground thermal heat pump) Volumetric efficiency, 52
energy storage (UTES) refrigerant, 145
relative irreversibility, 148–150
V submerged water source heat pump W
Vapor compression heat pump (see Submerged water source Waste exergy emissions, 31, 33–34
system, 148–149b heat pump) Water pollution, 6
air source heat pumps (see Air temperature-entropy diagram, 145, Water source heat pumps, 138t
source heat pumps) 145f groundwater, 139–140
Carnot heat pump, 147 Very-high-temperature generator seawater, 140
components, 145 (VHTG) surface waters, 140
energetic COP, 147 binary isobutane power plant, wastewater and effluent, 140–141
energy rate balance, 146 185, 186f Water-to-air heat pumps, 142–144,
entropy rate balances, 146–147 geothermal water leaving, 187–188, 143f
exergetic COP, 147–148 187f Water-to-water heat pumps, 142–143,
exergy destruction rates, 146–147 heat input rate, 187–189, 187f 143f
mass rate balances, 145–146 stream exiting, 186–187

You might also like