You are on page 1of 52

1.

4 Sustainability Dimensions of Energy


Ibrahim Dincer and Calin Zamfirescu, University of Ontario Institute of Technology, Oshawa, ON, Canada
r 2018 Elsevier Inc. All rights reserved.

1.4.1 Introduction 102


1.4.2 Sustainability and Energy Policy 105
1.4.3 Sustainability Indicators 111
1.4.4 Exergy-Based Sustainability Assessment of Energy Systems 119
1.4.5 Sustainability Assessment Based on Exergetic Lifecycle Analysis 122
1.4.6 Clean Energy Solutions for Better Sustainability 130
1.4.7 Case Studies 133
1.4.7.1 Exergosustainability Assessment of a Concentrated Photovoltaic-Thermal System for Residential
Cogeneration 133
1.4.7.1.1 Thermodynamic analysis 135
1.4.7.1.2 Results 137
1.4.7.2 Exergosustainability Assessment of High-Temperature Steam Photoelectrolysis Plant 138
1.4.7.2.1 Thermodynamic analysis 139
1.4.7.2.2 Results 141
1.4.7.3 Exergosustainability Assessment of a Heat Pump Dryer 142
1.4.7.3.1 System description 142
1.4.7.3.2 Results 144
1.4.8 Future Directions 150
1.4.9 Closing Remarks 150
References 151
Further Reading 151
Relevent Websites 151

Abbreviations LT Lifetime, years


ASI Aggregated sustainability indicator NSI Normalized sustainability indicator
COP Coefficient of performance PBP Payback period, years
EcI Eco-indicator PP Performance parameter
EI Environmental impact QP Quality parameter
EPV Energy product value RC Resource consumption
ExCDR Construction exergy expenditure RPC Removal pollution cost
ExIE Exergetic investment efficiency SEI Sustainability efficiency indicator
ExSI Exergetic sustainability index SF Scaling factor
FF Filling factor SI Sustainability index
GF Greenization factor SRW Specific reversible work
IP Improvement potential

Nomenclature E_ Energy rate, kW


A Area, m2 Ex Exergy, kJ
AM Amount of construction material, t _
Ex Exergy rate
Bi Biot number EEx Embodied exergy, GJ/t
c Speed of light constant, m/s fo&m Operation and maintenant factor
C Cost h Planck constant
CC Capital cost, $ I Irradiance, W/m2nm
CIEx Capital internment effectiveness IC Investment cost, $
CM Cost of construction materials, $ Ind Sustainability indicator
CP Cost of pollution, $ J Photocurrent density, A/m2
Cex Exergy-specific price, $/kWh kB Boltzmann constant
cp Specific heat, J/kg K L Length, m
Dp Depletion number LExS Levelized exergy cost savings, $/GJ
e Universal electric charge, C m Mass, kg

Comprehensive Energy Systems, Volume 1 doi:10.1016/B978-0-12-809597-3.00104-8 101


102 Sustainability Dimensions of Energy

n Diode nonideality factor tyear Number of hours in a year, h


n_ Molar flow rate, mol/s U Utility
Q Quantity V Voltage, V
Q_ Heat rate, kW w Weighting factor
SE Specific emission, kg/t W Work, kJ
SExC Specific exergy cost, $/GJ W_ Work rate, kW
t Time, s W Moisture content, kg H2O/kg
T Temperature, K

Greek letters c Exergy efficiency


f Dimensionless moisture content s Standard deviation
F Quantum efficiency s Stefan–Boltzmann constant, W/m2K4
Z Energy efficiency T Transmittance

Subscripts opt Optical


0 Reference state ORC Organic Rankine cycle
a Aperture out Output
abs Absorbed ph Photocurrent
act Actual pr Pollutant removal
b Blackbody PV Photovoltaic
d Destroyed pw Pollutant waste
ex Exergetic r Reflector
g Band gap ref Reference
gt Gas turbine rev Reversible
hp Heat pump tot Total
in Input th Thermal
m Mass transfer w Water
oc Open circuit

Superscripts comb Combined


ch Chemical sep Separated

1.4.1 Introduction

World population continues to grow and energy consumption is increasing. These are the facts documented based on historical
observations. The world energy consumption in 2012 was 13,834 Mtoe, it grew to approx. 14,817 Mtoe in 2016 and is expected to
escalate dramatically with continuing consumption and population increase trends [1]. There is no demonstrable prospect that
concludes that the world population and energy consumption will stop growing. Arguably, energy resources will be depleted and
humans will strive to harvest them. This again cannot be demonstrated. As time passed, more and more diverse resources were
discovered, according to the records. It is clear that future society will be more structured, complex, and faster. It is also clear that
future will consume more fuel and foods than the present, while moving faster and spending more power. Thence, energy will be
consumed at a higher rate than that of today. Facilitating access to water and energy becomes crucial; they are the needs of any
individual, group, or the world. The real challenge of the humankind is to evolve in a sustainable fashion.
Modern society requires many types of services to maintain a good standard of living, having electricity, hot water, space
heating, air conditioning, fuels, various chemicals, and materials, etc. Traditional methods of producing these commodities are
primarily driven by combustion of fossil fuels, which is a major contributor to air pollution. The global population and its
demand for these services are rapidly increasing, so the rising use of fossil fuels is having a major impact on climate change.
We live in a period characterized by accentuated transition efforts from fossil fuel-based economy toward a projected economy
based on sustainable energy. The development of global society passed through the fossil fuel era, when most primary energy was
obtained by combustion of coal, petroleum, and natural gas. At present, there is a larger panoply of primary energy sources
including nuclear power and hydroelectricity; in addition, other forms of renewable energy gain terrain, like wind, solar, biomass
combustion, and biofuels for transportation.
It is projected that the development of a hydrogen-fuel infrastructure for transportation will be needed. In the near future, the
combustion of fossil fuels will be performed in a cleaner fashion such that zero-emission power generation will be achieved at an
Sustainability Dimensions of Energy 103

attractive cost. Hydrogen is a clean energy carrier because it can be generated by low or zero-carbon sources (e.g., nuclear, water,
biomass, solar). Subsequently hydrogen can be converted to electricity, synthetic fuels, or heat with little or no environmental
pollution. Additionally, hydrogen is an important chemical that enters the constituency of many materials used in industry and
society: plastics, foods, pharmaceuticals, fertilizers, metallic materials, construction materials, etc. In cleaner generation of
hydrogen, less pollution is generated from the production of such hydrogen-containing other materials. Therefore, hydrogen itself
can be viewed as a critical commodity for future society, which can replace or at least complement the fossil fuels (oil, coal, natural
gas) that are extensively used at present. In this section, we will follow the historical evolution of the concept of “hydrogen as clean
energy carrier,” emphasizing the main issues and achievements to date.
Consider the history of primary energy that society uses for supplying its needs. There are three types of energies to be
considered, namely, the energy of coal, fossil hydrocarbons (petroleum, oil shale, oil sands, etc. and natural gas), and that of
renewable/sustainable sources that are “zero-carbon” sources. The renewable/sustainable energies are those that do not release
greenhouse gas (GHG) emissions upon their use (although some GHGs are emitted indirectly – e.g., due to the construction of an
energy distribution infrastructure). Typical examples of “zero-carbon” energies are biomass, hydroelectric, wind energy, solar,
nuclear, and others. The International Environmental Agency describes renewable energy as the energy derived from natural
processes using continually replenished sources [1]. The authors compiled data from existing literature to develop the chart shown
in Fig. 1 representing the share of “zero-carbon,” coal, and fossil fuel energy use worldwide since 1800 until the present time. The
chart is extrapolated until 2100 based on a predictive model described in Ref. [2].
Before the first industrial revolution (1800), the world energy supply comprised only “zero-carbon” sources: wood, wind,
hydro, and animal power. Starting with the industrial revolution and the introduction of the steam engine, the consumption of
coal increased sharply. One can speak about a “mechanization” period of around 100 years during which Western civilizations
implemented an industrial society. There was remarkable technical and scientific progress in this period, especially in England and
France.
By 1900, another source of energy started proliferating in society, namely, petroleum. Furthermore, the advances in electric
power generation and the electrical grid propelled a second industrial revolution around 100 years ago, which was called the
“electrification” period. With the help of electricity, process devices like pumps, compressors, conveyers, mixers, etc. became more
effective, which encouraged the sharp development of the process industry. Germany is noted in the beginning of the 20th century
for important inventions in the chemical and process industry. One of the most important processes ever invented is the
Haber–Bosch ammonia synthesis process (1913).
Manufacturing of ammonia by this process became a major user of hydrogen in the world. Further evolution of the process led
to synthesis of urea at an industrial scale, which is a major fertilizer produced from ammonia and carbon dioxide. In the Ruhr
region of Germany, a high development of coal mining and metallurgy occurred at the beginning of the 20th century. Hydrogen
was highly needed by the metallurgical sector. Consequently, by 1938, a hydrogen pipeline of 240 km was installed to deliver
250  106 Nm3/year of hydrogen obtained from coal. Presently in Europe, there are about 1500 km of hydrogen pipelines, while
in the United States, there are about 900 km of hydrogen pipelines.
The geopolitical context in the first half of the 20th century encouraged the development of engines and fuels for transpor-
tation. Gasoline gained in proportion due to its high octane number, its high energy density, and cheap process of extraction from

100
Coal
Fossil hydrocarbons
80 Zero carbon
Energy share (%)

60

Future
40

20

0
1800 1850 1900 1950 2000 2050 2100
Year
Mechanization Electrification Hydrogenation

Fig. 1 Shares of world energy sources since the start of the first industrial revolution, and prospects to 2100.
104 Sustainability Dimensions of Energy

100 12

80 10

Energy share (%)

Meq (kg/kmol)
60 8

40 6

20 Meq (kg/kmol) 4
Carbon (thick line)
Carbon-free
0 2
1800 1850 1900 1950 2000 2050 2100
Year
Fig. 2 Illustration of the historical trend of energy decarbonization.

crude petroleum. Currently, the road transportation sector uses mostly gasoline and diesel and the marine transport industry uses
heavy petroleum or diesel. Hydrogen has found uses in space transport as liquefied rocket fuel.
The intellectual activity and the acquired technological and scientific knowledge created the start of a new 100-year era, which
will follow the “mechanization” and “electrification” eras. This has been often cited as the “hydrogenation” of society. According
to predictions suggested as shown in Fig. 1, during the next 100 years, the “zero-carbon” energy sources will increase in share and a
“hydrogen civilization” will emerge. The rise of a hydrogen civilization is induced by preemptive actions founded on scientific
analyses regarding the negative impact of GHG emissions on climate and the depletion of fossil fuel resources.
The chart in Fig. 2 shows how a basic analysis which predicts the decarbonization of the world energy supply. The plot is made
under the assumption that the average carbon content of fossil hydrocarbons (mainly natural gas and petroleum) is around 1:3,
while 2:3 is the hydrogen content (the carbon to hydrogen ratio for methane is 1:4, while for gasoline it is B1:2). Therefore the
carbon content of the global energy supply is obtained from Fig. 1 by addition of the coal curve to one-third of the fossil
hydrocarbon curve. The noncarbon share of the energy supply is obtained by adding 2/3 of the fossil hydrocarbon curve with the
zero carbon curve. The plot shows how carbon-free sources of energy decreased from 1800 to approx. 1950; and again increased
and are predicted to further increase. Accounting for molecular mass of carbon (12 kg/kmol) and hydrogen (2 kg/kmol) the
averaged molecular mass of the fuel sources (carbon-based + carbon-free) is determined. As shown in the figure, before 1800 the
molecular mass of the equivalent fuel sources has been 2 kg/kmol, illustrating that only negligible fossil carbon sources of energy
were used before the industrial revolution. This equivalent molecular mass increased to a peak of around 11 kg/kmol in the first
half of 20th century, illustrating that the world depended mostly on coal at that time. That has been the prominent era of steam
engine. The curve climbed down, being today at a figure of approx. 6 kmol/kg; this number shows a mix of fuel sources, which
includes mostly petroleum liquids. The prediction to 2100 shows that the equivalent molecular mass of world fuel is around
3 kg/kmol, that is, very close to the hydrogen mass of 2 kg/kmol, an almost fully decarbonized energy supply.
The curve (Meq) shows an interesting feature: it sharply rises during the coal-dependent “mechanization era,” decreases with
fluctuations during the “electrification” era, and steadily decreases during the future “hydrogenation” era. This downward character
toward energy decarbonization is a result of societal will demonstrated during the last 40 years.
Sustainable development is the natural way of development. It is so because to be sustainable is the only way in which
anything persists: sustainable can be translated as “persistent in time.” The quest is how to guide and influence the sustainable
development of the world. Energy sustainability means harmonious and rational development at three levels: power generation,
power distribution, and power utilization (consumption and destruction). Among other aspects, sustainability implies facilitating
the flows of knowledge (know-how), which became crucially important; see also Ref. [31].
At the beginning of the 21st century, a new branch of science emerged, namely, sustainability science, having an inter-
disciplinary character involving the following main disciplines: physics, chemistry, biology, medicine, social and economic sci-
ences, and engineering. One of the goal of sustainability science is to model the complex interactions between society,
economy, and environment, also accounting for resource depletion. One crucial aspect is the provision of theoretical foundations
and tools for sustainability assessment. Science became more and more the dominant driving force influencing the development
of society. Due to science and the establishment of the information era, a global self-consciousness is developed – the so-called
noosphere – that made humans aware of possible catastrophic scenarios related to the potentially unsustainable development of
technology.
National and international institutions were created in the last quarter of the 20th century to promote sustainable development
of society. As such, the World Commission on Environment and Development acted within the United Nations for a period of
Sustainability Dimensions of Energy 105

5 years in mid-1980s and defined formally the concept of sustainable development and showed that societal development and the
environment are interrelated for any foreseeable time horizon. Sustainable development has been defined by the United Nations
as “development that meets the needs of the present without compromising the ability of future generations to meet their own
needs.”
Due to its wide character, spanning from environment to society, sustainability is difficult to assess based on quantifiable
indicators. Some efforts to categorize the tools for sustainability assessment are reported by Ness et al. [3]. In the past, attempts
were made to formulate empirical indicators for sustainability assessment, such as “environmental sustainability index,” “well-
being index,” “human development index,” etc. More recently, product-related assessment methods were proposed based on
product material flow analysis, lifecycle assessment, and thermodynamic analysis. Among those, the sustainability, environmental
and energy assessment based on exergy analysis is noted for furthering the goal of more efficient resource use with reduced waste
rejection into the environment.
Many factors contribute to achieving sustainable development. One of the most important is the requirement for a supply of
energy resources that is fully sustainable. Increased efficiency is also considered critically important. Kyoto Protocol, in 1997,
envisaging to control GHG concentrations in the atmosphere, is intimately related to the goal of sustainable development
achievement. Availability and utilization of energy resources are highly related to the level of sustainable development.
Sustainable energy supplies are required for sustainable development. These type of resources must be available for a rea-
sonably foreseeable future at an affordable cost and with reasonable access, without causing negative societal and environmental
impacts (EIs). Therefore better design of energy systems must be developed with increased efficiency and reduced fossil fuel
consumption such that (1) less primary energy resources are used, and (2) reduced pollutants, such as GHG (i.e., GHGs, such as
CO2, CH4), acid gases (SO2, NOx responsible for acid precipitation), and chlorofluorocarbons (CFCs, responsible for destroying
the stratospheric ozone) are emitted. Considering the significance of energy conversion, power generation, distribution, and
utilization in the global economy, any improvement of designs toward efficient and environmentally benign drying technology
will help the sustainable development of society. In this respect, exergy analysis plays a major role, as reported in [4], because it
connects sustainability with energy and the environment.
In this chapter, the sustainability aspects of energy systems are discussed. There are various aspects to be analyzed in this respect.
We will discuss in this chapter the theoretical assessment tools and indices for sustainable development and their impacts in shaping
energy policies and strategies, which ultimately influence the progress of society toward better sustainability. The sustainability
concept and its assessment methods are introduced with a focus on exergy-based assessment. Some case studies are provided as well.

1.4.2 Sustainability and Energy Policy

Sustainability refers to the feature of a system to be sustainable. If we refer to human-made (engineered) systems, a system to be
sustainable must be conceived such that it does not permanently damage the environment, and it does not consume excessive
natural resources, rather it operates effectively for a sufficiently long period of time without jeopardizing the ability of future
generations to meet their needs of natural resources and clean environment.
To be truly sustainable, an energy system must meet the following criteria: (1) minimal or no negative environmental or social
impact; (2) no natural resource depletion; (3) able to supply the current and future population’s energy demand; (4)
equitable and efficient; (5) air, land, and water protection; (6) little or no net carbon or other GHG emissions; and (7) safety today
without burdening future generations. There are of course several other criteria to consider, such as abundance, local availability,
cost effectiveness, reliability, safety, and environmental friendliness.
The march toward a global sustainable society is importantly influenced by policy and decision makers. Fig. 3 shows how
policy and decision makers influence the way in which the world’s future is shaped toward sustainability provided that appropriate
theoretical tools for sustainability assessment are made available and green technologies are developed. Behind any policy and

Assessment tools
(exergy, sustainability
index, etc.)

Present: Policy and Future:


transition period from decision sustainable energy
fossil fuel economy makers society

Technological
development

Fig. 3 The role of assessment tools and indicators in shaping the future sustainable society.
106 Sustainability Dimensions of Energy

decision, there must be a rationale support, which constrains the decision spectrum. Apart from various ideologies that do exist,
the status of technological development and the existent theoretical assessment and prediction tools play a major role in influ-
encing policy making and high-level decisions.
When shaping sustainable energy policies and strategies, it is important to understand the benefit of clean energy technologies
at one hand, and the effects of fossil fuel-based systems on the other. The kinds of energy technologies with zero or minimum
negative EI (e.g., through associated pollution) – that is, more environmentally benign and more sustainable – are in general
named green energy. Considering the benefits of green energy, sustainability of green energy supply and progress is assumed to be
a key element in the interactions between nature and society. An essential increase in the scale and pace of policy instruments and
their effectiveness is required to change course toward a sustainable path. It is important to possess theoretical tools to quanti-
tatively assess green energy development and:

• help elaborate rationale strategies and policies;


• help understand main concepts and issues about green energy use and sustainability aspects;
• develop relations between green energy use and sustainability development;
• encourage the strategic use and conservation of the green energy sources;
• provide the methods for energy security, implementation and development;
• increase the motivation on the implementation of green energy strategies for better energy supply;
• give an idea to reduce the negative EIs by considering the possible green energy strategies; and
• form a scientific platform to discuss the possible green energy strategies for sectoral use.

The need of policy is justified by the fact that free markets must be enforced to meet the needs of vulnerable groups, reduce
environmental pollution, and ensure the energy security. At upper level, strategies are needed; where a strategy represents a plan or
method toward a goal (in our discussion “the goal” means progress toward a global society based on sustainable energy). Once a
strategy is established, the policy, which represents the course of actions to implement the strategy, must be elaborated. Any policy
will specify “policy instruments,” which represent specific measures taken to implement a policy. Examples of policy instruments
are:

• imposing efficiency standards;


• setting public procurement policies;
• imposing appliance labeling norms;
• obligation to buy or supply energy from renewable sources; and
• supporting research and development in demonstration projects.

Some examples of promising policies are given here as follows: phasing out subsidies for fossil fuel-based energy, restructuring
the energy sector, supporting energy sector innovation, promoting energy efficiency, financing rural energy, providing decen-
tralized options, and improving access to modern and efficient cooking fuel. Some other rule of thumb aspects that influence
elaborations of green energy strategies and policies can be summarized as follows:

• Competition among sustainable energy technologies may be hampered by market distortions that give advantages to certain
players.
• Improving energy efficiency requires less investment than does new generation.
• Financing large energy projects discourages smaller renewable energy projects.
• Commercialization of green energy technologies is not occurring quickly enough to meet the challenges of sustainability.
• For significant sustainable progress, a critical mass of public support is needed.
• It is important to liberalize energy markets in order to offer an opportunity to change.

The sustainable development is at the confluence of energy resources sustainability, environmental sustainability, economic
sustainability, and social sustainability as shown in Fig. 4. Green energy resources and technologies are a key component of
sustainable development for three main reasons. Firstly, they generally cause less EI than those of other energy sources. The variety
of green energy resources provides a flexible array of options for their use. Secondly, they cannot be depleted. If used carefully in
appropriate applications, green energy resources can provide a reliable and sustainable supply of energy almost indefinitely.
Thirdly, they favor system decentralization and local solutions that are somewhat independent of the national network, thus
enhancing the flexibility of the system and providing economic benefits to small isolated populations. Also, the small scale of the
equipment often reduces the time required from initial design to operation, providing greater adaptability in responding to
unpredictable growth and/or changes in energy demand.
The major considerations involved in the development of green energy technologies are illustrated in Fig. 5 and include social
and EIs, commercialization, technical aspects, and economic factors. Apart from these considerations, one can identify a number
of parameters (factors) that are important in establishing green energy strategies and policies. They include public information,
environmental education, innovation stimulation, promotion of technologies, financing, and very important elaboratory eva-
luations tools and techniques.
Green energy technologies are expected to play a key role in sustainable energy scenarios for the future. The foremost factor that
will determine the specific role of green energy and technologies will likely be energy demand. Therefore, in order to compensate
Sustainability Dimensions of Energy 107

Social Environmental
sustainability sustainability

Sustainable
development

Energy and
Economic
resources
sustainability
sustainability

Fig. 4 Interdependence of the factors affecting sustainable development. Modified after Dincer I, Rosen MA. Thermodynamic aspects of
renewable and sustainable development. Renew Sustainable Energy Rev 2005;9:169–89.

Increase of green energy


and technology utilization

Social and Commercialization


environmental impact Research and
Social benefits, global development,
peace, environmental information technology,
impact, higher living incentives, training,
standard, clean air and education,
environment, etc communication, etc.

Technical aspects Economic factors


Availability, grid Investments, generation
connection, technology costs, lower operation
level and use, cost, lower cost energy
technological recovery, lower cost
innovations, advanced transportation,
technologies, etc. externalities, etc.

Fig. 5 Considerations involved in development of green energy technologies. Modified after Dincer I, Rosen MA. Thermodynamic aspects of
renewable and sustainable development. Renew Sustainable Energy Rev 2005;9:169–89.

the energy requirement, it will be possible to produce green energy from renewable energy sources, such as hydraulic, solar, wind,
geothermal, wave, biomass, etc. Green energy technologies are largely shaped by broad and powerful trends that have their roots in
basic human needs. Wastes (convertible to useful energy forms through, e.g., waste-to-energy incineration facilities) and biomass
fuels are also usually viewed as sustainable/green energy sources.
In conjunction with this, the increasing world population requires the definition and successful implementation of green
energy technologies. Briefly, the important parameters and their interrelations as outlined in Fig. 6 are definitely required to carry
out the best green energy program and select the most appropriate green energy technology/technologies for sustainable
development.
The green energy and technologies can be utilized for many applications as shown in Fig. 7. Thus it can be said that green
energy and technologies are expected to play a key role in sustainable energy scenarios for the future. The foremost factor that will
determine the specific role of green energy and technologies will likely be energy demand. Therefore, in order to compensate the
energy requirement, it will be possible to produce green energy from renewable energy sources, such as hydraulic, solar, wind,
geothermal, wave, biomass, etc. If so, the green energy and technologies can be utilized for many applications as shown in the
figure. Thus it can be said that green energy and technologies, which are abundantly available, can help to:

• provide a more environmentally benign and more sustainable future;


• increase energy security;
• facilitate or necessitate the development of new, clean technologies;
• reduce air, water, and soil pollution and the loss of forests;
108 Sustainability Dimensions of Energy

Public and
Successful green Agencies, training,
government Public
energy program facilities
channels

Professional support Energy utilization


Sustainable green
Media support Environmental
energy program
Public support impact

Environmental
Public information/
Communication education and
awarness
training

Essential factors for green


energy technologies

Innovative green Promoting green Evaluation tools and


Financing
energy strategies energy resources techniques

Effective sustainable Short and long term


green energy policies for Implementation of
program implementation green energy
Monitor each step
systems and
and evaluate the
technologies locally
data findings
and globally for
obtained
sustainable
Environmentally
Public relations, development
benign green
training, counseling
energy programs

Fig. 6 Main factors influencing sustainable energy strategies and policies. Modified after Midilli A, Dincer I, Ay M. Green energy strategies for
sustainable development. Energy Policy 2006;34:3623–33.

• reduce energy-related illnesses and deaths; and


• reduce or stop conflicts among countries regarding energy reserves, etc.

Therefore green energy and related technologies are needed to ensure global stability by reducing the harmful effects of fossil-
based energy consumption. Thus the importance of green energy in reducing the world’s problems and achieving a sustainable
energy system should be emphasized considering the sustainable green energy strategies, and a transition to green energy
economy, should be encouraged, and developed countries, in particular, should increase investments in green energy and
technologies.
In order to develop and publicize the sustainable green energy technologies in a developed or less developed country, the
following important green energy strategies should be taken into consideration:

• industrial and technological support for transition to green energy technologies;


• control of the projection and analysis of green energy sources;
• governmental and public support for green energy economy;
• production, consumption, distribution, conversion, management, and marketing of green energy;
• research, development, and application of sustainable green energy technologies;
• availability, productivity, and reliability of green energy and technologies; and
• design and fabrication of green energy-based environmental and ecological applications.
Sustainability Dimensions of Energy 109

Application areas of green energy


and technologies

• fuel cells • fuel cells


• gas turbines Applications for • IC engines
Vehicle
• hydrogen plants power generation • combustion
applications
• efficiency
• heating improvement
• cooling • defense industry
• cooking Domestic • transport
• air conditioning applications
• pumping, etc.,
• power generation
• ship engines
Navigation
• defense
applications
Industrial • communication
• ammonia synthesis
applications • transportation
• fertilizer production
• tourism
• oil distillation
• pollution control
• petrochemical production
• energy storage
• metallurgical applications
• ni and Fe production
• energy storage • gas turbines
• flammable mixtures • jet engines
• electronic industry Aeronautics • defense industry
• glass and fiber production application • rockets
• nuclear reactors • antimissile
• power generation systems • space industry
• energy storage

Fig. 7 Applications of green energy technologies. Modified after Midilli A, Dincer I, Ay M. Green energy strategies for sustainable development.
Energy Policy 2006;34:3623–33.

Energy conservation is vital for sustainable development, and for the best benefit of present and future generations, should be
implemented by all possible means. A secure supply of energy resources is generally agreed to be a necessity but not a sufficient
requirement for development within a society. Furthermore, sustainable development demands a sustainable supply of energy
resources that, in the long term, is readily and sustainably available at a reasonable cost, and can be utilized for all required tasks
without causing negative societal impacts. Supplies of such energy resources as fossil fuels (coal, oil, and natural gas) and uranium
are generally acknowledged to be finite; other energy sources, such as sunlight, wind, and falling water, are generally considered
renewable and therefore sustainable over a relatively long period of time.
Here, we look at the renewable energy resources and energy conservation. While not all renewable energy resources are
inherently clean, there is such a diversity of choices that a shift to renewables carried out in the context of sustainable development
could provide a far cleaner system than would be feasible by tightening controls on conventional energy. Furthermore, by being
naturally site-specific, they favor power system decentralization and locally applicable solutions more or less independent of the
national network. It enables citizens to perceive positive and negative externalities of energy consumption. Consequently, a small
scale of equipment often makes the time required from initial design to operation short, providing greater adaptability in
responding to unpredictable growth and/or changes in energy demand.
The exploitation of renewable energy resources and technologies is a key component of sustainable development. There are
three significant reasons for this as follows:

• They have much less EI compared to other sources of energy because there is no energy source with zero EI. There are such a
variety of choices available in practice that a shift to renewables could provide a far cleaner energy system than would be
feasible by tightening controls on conventional energy.
• Renewable energy resources cannot be depleted unlike fossil fuel and uranium resources. If used wisely in appropriate and
efficient applications, they can provide reliable and sustainable supply energy almost indefinitely. In contrast, fossil fuel and
uranium resources are finite and can be diminished by extraction and consumption.
• They favor power system decentralization and locally applicable solutions more or less independent of the national network,
thus enhancing the flexibility of the system and the economic power supply to small isolated settlements. That is why many
different renewable energy technologies are potentially available for use in urban areas.

Taking the aforementioned reasons into consideration, the relations between energy conservation and sustainability are finally
presented as in Fig. 8. As suggested in the figure, energy resources and their utilization are intimately linked to sustainable
development. For societies to attain or try to attain sustainable development, much effort must be devoted not only to discovering
sustainable energy resources, but also to increasing the energy efficiencies of processes utilizing these resources. Under these
110 Sustainability Dimensions of Energy

Energy sustainability Sustainability program


• availability of energy • sustainable energy strategies
• energy management • innovative energy strategies
• energy production and consumption • sustainable energy programs
• energy conservation and distribution • short-long term energy policies
• productivity of energy • programs for clean energy

Energy
conservation and
sustainability

Economic sustainability Environmental sustainability


• availability of energy • political support
• energy management • reliability/knowledge investments
• energy production and consumption • environmental planning
• energy conservation and distribution • marketing innovations
• productivity of energy • environmental control mechanisms

Fig. 8 Linkages between energy conservation and sustainable development.

circumstances, increasing the efficiency of energy-utilizing devices is important. Due to increased awareness of the benefits of
efficiency improvements, many institutes and agencies have started working along these lines.
Many energy conservation and efficiency improvement programs have been and are being developed to reduce present levels of
energy consumption. To implement these programs in a beneficial manner, an understanding is required of the patterns of “energy
carrier” consumption, for example, the type of energy carrier used, factors that influence consumption, and types of end-uses.
Environmental concerns are an important factor in sustainable development. For a variety of reasons, activities that continually
degrade the environment are not sustainable over time, for example, the cumulative impact on the environment of such activities
often leads over time to a variety of health, ecological, and other problems. A large portion of the EIs in a society are associated
with its utilization of energy resources.
Ideally, a society seeking sustainable development utilizes only energy resources that cause no EI (e.g., that release no emissions
to the environment). However, since all energy resources lead to some kind of EI, it is reasonable to suggest that some (not all) of
the concerns regarding the limitations imposed on sustainable development by environmental emissions and their negative
impacts can be in part overcome through increased energy efficiency. Clearly, a strong relation exists between energy efficiency and
EIs because, for the same services or products, less resource utilization and pollution is normally associated with increased energy
efficiency.
The share of energy R&D expenditures going into energy conservation grew greatly since 1976, from 5.1% to 40.1% in 1990
and 68.5% in 2002. This indicates that within energy research and development, research on energy conservation is increasing in
importance. When R&D expenditures on energy conservation are compared with expenditures for research leading to protection of
the environment in the 2000s, the largest share was spent on environment research. In fact, it is not easy to interpret the current
trends in R&D expenditures because energy conservation is now part of every discipline from engineering to economics. A marked
trend was observed since the mid-1970s, in that expenditures for energy conservation research grew significantly both in absolute
terms and as share of total energy R&D. They also grew more rapidly than environmental protection research, surpassing it in the
early 1980s. Therefore if R&D expenditures reflect long-term concern, there seems to be relatively more importance attached to
energy conservation as compared to environmental protection. In addition to the general trends discussed above, consider the
industrial sector and how it has tackled energy conservation.
The private sector clearly has an important role to play in providing finance that could be used for energy efficiency invest-
ments. In fact, governments can adjust their spending priorities in aid plans and through official support provided to their
exporters, however, they can only indirectly influence the vast potential pool of private sector finance.
Many of the most important measures to attract foreign investors include reforming macroeconomic policy frameworks,
reforming energy market structures and pricing, banking reform, debt recovery programs, strengthening the commercial and legal
framework for investment, and setting up judicial institutions and enforcement mechanisms. These are difficult tasks that often
involve lengthy political processes. Thus Fig. 9 presents a series of important factors that can contribute to improving energy
conservation in the real life. Although there are a large number of practical solutions to environmental problems, three potential
solutions are given priority as follows:
Sustainability Dimensions of Energy 111

Improvement factors of energy conservation

Goal and scope definition, energy management program, funding


Understanding energy
opportunities, consumer-consultant relations, local and industrial
conservation concepts
energy improvements.

Assessment of energy conservation opportunities, cost-effectiveness


Understanding key of energy utilization, simple-payback of energy conservation, energy
energy-related concepts conservation measurements, energy saving potentials, and individual
energy conservation.

Energy consumption and sources, energy costs, energy recovery


Current energy usage possibilities, energy saving, energy production capacity, energy
conversation, and energy-related industries.

Energy conservation in maintenance and operations, energy saving


Energy-related operation and cost effectiveness, staff education, energy education cost, energy
and maintenance control recovery units and techniques, and application of energy conservation
measurements.

Energy systems and equipments survey, current energy use analysis,


Assessment of energy
recognition of energy saving in local and industrial applications,
conservation
encourage of energy saving, social and technical investment for energy
opportunities conservation, and completion of cost/benefit energy conservation.

Fig. 9 Improvement factors of energy conservation for better sustainability.

• energy conservation technologies (efficient energy utilization);


• renewable energy technologies; and
• cleaner technologies.

Among these technologies, we pay special attention to energy conservation technologies and their practical aspects and EIs.
Each of these technologies is of great importance and requires a careful treatment and program development. Considering the
above priorities to environmental solutions, the relevant technologies are summarized in Fig. 10.

1.4.3 Sustainability Indicators

The sustainability indicators are constructed based on the drivers-pressures-state-impact-responses (DPSIR) model of the society,
as shown in Fig. 11. In this model, the drivers are the developments in society, economy, and the environment. These devel-
opments (or changes) exert pressures on the sustainability, and as a consequence, the sustainability state changes in some
direction. This eventually leads to foreseeable impact on sustainable development: it improves or degrades the sustainability. The
society can respond to this change by taking actions to reverse the impacts (see the double arrow in the figure). In addition, the
society can send feedback to drivers (i.e., to impose educated changes in the society, economy, and environment). The responses
can also act directly on the pressures and on the state of sustainability by applying adequate measures if possible. The responses
must be effective, because their effectiveness affects directly the sustainable development. Two simplified DPSIR models exist,
which apply to sustainability assessment, namely, drivers-state-responses (DSR) or pressure-state-response (PSR).
Any conceivable sustainability indicator has some degree of subjectivity because it is not actually possible to objectively determine
and fully understand the interrelationships between social change, environment, energy, and economic development. The independent
formulation of figures of merit for energy systems and formulation of environmental, economic and social indicators is more facile. The
112 Sustainability Dimensions of Energy

Environmental information services


Energy and Research and
environment development centers
management for environmental
centers problems
Potential solutions to environmental problems

Renewable Energy Clean


energy conservation environment
technologies technologies technologies

Green Pollution
buildings control
technologies

Energy
efficiency Greenhouse gas
technologies reduction
technologies
Waste and soil
management Recycling
technologies technologies

Fig. 10 Linkages between possible environmental and energy conservation technologies.

Drivers Pressures State Impact Responses

Fig. 11 Drivers-pressures-state-impact-responses (DPSIR) model for sustainability assessment.

United Nations developed more than 130 indicators for sustainability assessment of which 30% are social indicators, 17% are
economic indicators, 41% are environmental indicators, and the remaining 12% represent institutional indicators. Approximately 33%
of sustainability indicators are of the driving force (drivers) kind; 39% are state indicators, and 18% are response-kind indicators.
Table 1 gives the significant indicators that can be used to elaborate aggregate sustainability indexes (SIs). In principle, the
indexes can be elaborated according to the purpose of the analysis. Some specific aspects on elaboration and adoption of a holistic
sustainability assessment methods are discussed as follows:

• Sustainability assessment should integrate ecological conditions assessment of human and life systems habitat with economic
development, social wellbeing, and with equity and disparity of current population and future generations.
• Obtain a consensus on adoption of a time horizon sufficiently long to be in accordance with ecosystem time scale and
anticipated society time scale.
• Sustainability indicators and indices must be able to link sustainability categories to society goals.
• Sustainability assessment should integrate the following category of models: economic models, stress–response models,
multiple capital models, social–economic–environmental models, and human–ecosystem wellbeing models.
• Definition of a reference environment (or state) is required for determining the sustainability change.
• Monetary assessment of environmental damage should be included in sustainability assessment.
Fig. 12 represents a generic model of an energy system, which is useful for sustainability assessment. The energy system is a
generic one. It may be a power plant where chemical energy of a fuel is converted to electric power. Other energy conversions can
be nuclear-to-power, wind-to-power, hydraulic pressure-to-hydropower, solar-to-power, and so on. Beside energy conversion for
power generation, the energy system can also include power distribution.
There are two categories of inputs into the energy system, as shown in the model from the figure: the resources and the process
enablers. The process enablers are of three kinds: the human operators, the energy system (including all machines and technical
Table 1 Categories and types of indicators influencing the sustainability assessment

Kind Drivers indicators State indicators Responses indicators

Category

Social • unemployment rate • poverty gap index • gross domestic product (GDP) on education
• population growth rate • income inequality index • childhood immunization
• adult literacy rate • school life expectancy • infrastructure expenditure per capita
• motor fuel consumption per capita • house price per income • health expenditure per capita
• loss rate due to natural disasters • floor area per person • hazardous chemicals in foods
Economic • per capita GDP • proven mineral reserves • environmental protection expenditure
• investment share of GDP • fossil fuel reserves • funding rate on sustainable development
• annual energy Consumption • lifetime (LT) of energy reserves • amount of funds on sustainability
• natural resource consumption (RC) • share of renewable energy • percentage of new funds on sustainability
• capital goods imports • manufacturing added value • funding grants on technology
Environmental • water consumption per capita • ground water reserves • wastewater treatment expenditures
• generation of wastes • monthly rainfall index • natural resource management
• ozone-depleting potential (ODP) substances emission • desertification rate • air pollution mitigation expenditures
• emission of greenhouse gas (GHG), SO2, NOx • pollutants concentration • waste management expenditures
• energy use in agriculture • acute poisoning • number of restricted chemicals
Institutional • number of scientists per capita • policy on sustainable development

Sustainability Dimensions of Energy


• number of engineers per capita • environment protection programs
• internet access per capita • GDP share of R&D expenditures
• telephone line access • number of R&D personnel per capita
• other information channels • ratification/implementation of agreements
Source: Reproduced from UN. Indicators of sustainable development: guidelines and methodology. 3rd ed. New York, NY: United Nations.

113
114 Sustainability Dimensions of Energy

Process enablers:
Operators
• safety aspects
• health impact
• education aspects
Energy system related
• system type/capacity
• life cycle
• construction related
emissions

Environment
• temperature
• pressure
Input resources:
• composition
Construction materials and Process outputs:
labour Exergy
• metallic materials • electric power
• concrete, wood • process heat
• advanced materials • synthetic fuels
Energy • materials
Oxidant (if applies)
system
• humidity Recyclable or disposable
• temperature streams
• flow rate • rejected heat
Energy input • emissions
• work (or electric power) • expelled water/moisture
• heat Wastes:
Pollutants emissions
• GHGs
• NOx and SO2
• other pollutants
Energy wastes
• heat wasted
• waters heating due to
heat rejection at power
generation

Material wastes
• wasted moisture
• wasted process material
• indirect garbage
• landfill wastes resulting
from energy system
scrapping

Fig. 12 Sustainability assessment model of energy systems based on material and energy balances. GHG, greenhouse gas. Modified after Linke
B, Das J, Lam M, Ly C. Sustainability indicators for finishing operations based on process performance and part quality. Procedia CIRP
2014;14:564–9.

components), and the surrounding environment (characterized by certain temperature, pressure, and composition). With respect
to the human operators, the sustainability indicators must mainly quantify safety, health, and educational aspects. Related to the
energy system, the sustainability indicators must consider the lifecycle and the specific environmental emissions and impact
related to the system construction.
The input resources are of three kinds: construction materials and labor for system fabrication; oxidant (if applicable), for
example, combustion air; and energy (power). For each of the inputs relevant indicators, quantities, and qualities can be deter-
mined. For example, the construction materials are characterized by type, specific texture, shape, and dimensions. The energy input
is required in the form of work (power) and heat (heating rate).
The outputs are the actual process outputs and the process wastes. The main output of an energy system is typically the
generated power. Three types of wastes can be considered: environmental pollutants, energy wastes, and material wastes. The
pollutant emissions are due mainly to combustion processes directly or indirectly associated with the respective energy conversion
system. Also, the emission of landfill gas due to drying system scrapping at the end of the lifetime (LT) is a pollutant waste. The
Sustainability Dimensions of Energy 115

Technological
knowhow and
advancements

Present:
Planned future:
Fossil fuel-based Policy and
sustainable
economy in decision making
energy society
transition

Sustainability Energy and exergy efficiency,


assessment sustainability index, greenization
indicators factor, other performance indicators

Fig. 13 The role of sustainability assessment indicators in shaping policies for a planned sustainable energy society.

energy wastes can be observed in the form of heat rejected into the environment. This heat can be rejected directly by heat transfer
at the system boundary. Another way of energy waste is through warm water released in lakes, rivers, or seas where power plants
are installed. It is demonstrated that the cooling water for power plant condensers affect the aquatic environment by changing the
local temperature, which often negatively impacts the life systems. Noise is another way of energy waste with EI. The material
wastes include also landfill material generated at product scrapping phase.
Apart from various ideologies that do exist, the status of technological development and the existent theoretical assessment and
prediction tools play a major role in influencing policy making and high-level decisions. Some examples of energy policies and
strategies include encouraging the expansion of energy- and exergy-efficient systems, expanding the use of renewable energy, and
widening clean fossil fuel combustion technologies. Behind any policy and decision there must be a rationale support, which
constrains the decision spectrum. Fig. 13 illustrates how sustainable development strategies can be shaped with the help of
assessment of and advancement in technology.
The implications of the limited nature of energy and other resources have led in part to significant efforts in energy-utilization
efficiency improvement, and resource recycling and reuse. For example, refuse and other solid wastes are now often used to
supplement fuel supplies. Recycling (resource recovery) extends the LTs of many of natural resources, and is often profitable and
usually beneficial environmentally. Energy efficiency and conservation can postpone shortages of energy resources, reduce
environmental damage, and provide economic benefits. The efficient use of energy is of particular importance to developing
countries, as it can forestall the need for very large capital investments. Enormous potential exists through improvements in energy
efficiency and conservation for decreasing the world’s total energy consumption, and thereby the effects of energy consumption on
the environment. Often such improvements require a myriad of small changes in consumption patterns.
Increased energy efficiency reduces energy-related EIs, such as those listed in the Chapter Environmental Dimensions of Energy
(00103) (e.g., environmental damage due to the process of extracting energy resources from the ground, and the competition for
water between hydropower and such other uses as agriculture and recreational activities, and associated damage to water quality).
In addition, improved efficiency enhances the reliability of future energy supplies, and improves the longevity of energy supplies.
The potential for energy efficiency is significant during both energy production and consumption (e.g., the 30% of oil in a reservoir
that is extracted from onshore wells could be improved upon using secondary recovery techniques, such as water flooding, and
thermal stimulation). The following measures are essential for reducing the EI of energy systems:

• Increasing the efficiency of power generation systems.


• Energy conservation measures.
• Use of more environmentally benign energy sources.

When shaping sustainable energy policies and strategies, it is important to understand the benefit of clean energy technologies
on one hand, and the effects of fossil fuel-based systems on the other. The kinds of energy technologies with zero or minimum
negative EI (e.g., through associated pollution), that is, which are more environmentally benign and more sustainable, are in
general named renewable energy. Considering the benefits of renewable energy, sustainability of renewable energy supply and
progress is assumed to be a key element in the interactions between nature and society. An essential increase in the scale and pace
of policy instruments and their effectiveness is required to change course toward a sustainable path. It is important to possess
theoretical tools to quantitatively assess renewable energy development, and:

• help elaborate rationale strategies and policies,


• help understand main concepts and issues about renewable energy use and sustainability aspects,
• develop relations between renewable energy use and sustainability development,
116 Sustainability Dimensions of Energy

• encourage the strategic use and conservation of the renewable energy sources,
• provide the methods for energy security, implementation and development,
• increase the motivation on the implementation of renewable energy strategies for better energy supply,
• give an idea to reduce the negative EIs by considering the possible renewable energy strategies, and
• form a scientific platform to discuss the possible renewable energy strategies for sectoral use.

The need of policy is justified by the fact that free markets must be enforced to meet the needs of vulnerable groups, reduce
environmental pollution, and ensure the energy security. Upper level strategies are needed, where a strategy represents a plan or
method toward a goal (in our discussion “the goal” means progress toward a global society based on sustainable energy). Once a
strategy is established, the policy, which represents the course of actions to implement the strategy, must be elaborated. Any policy
will specify “policy instruments,” which represent specific measures taken to implement a policy. Examples of policy instruments are:

• Imposing efficiency standards.


• Setting public procurement policies.
• Imposing appliance labeling norms.
• Obligation to buy or supply energy from renewable sources.
• Supporting research and development in demonstration projects.

One of the most important aspects is thorough evaluation of the costs of reducing CO2 emissions. From a developing-country
perspective, the discussion of costs and benefits has to take into account the need for policies promoting rapid economic growth.
Achieving such a balance between economic development and emissions abatement requires the adoption of domestic policies
aimed at improving the efficiency of energy use and facilitating fuel switching, and the implementation of international policies
enabling easier access to advanced technologies and external resources.
It is expected that some countries will offer certain tax reductions for those businesses that promote renewable energy
technologies, especially because these technologies are characterized by low or zero CO2 emission. Some possible solutions
include cleaning fossil fuels before combustion; burning them more cleanly, for example, using fluidized bed technology for coal;
using renewable energies; switching to hydrogen economy; implementing thermal energy storage technologies; promoting efficient
public transport; using more fuel-efficient vehicles; etc.
An example of development of a sustainability assessment framework is given here from Ontario Hydro in 1997 (currently Ontario
Power Generation). According to Ref. [6], there are five categories for sustainability assessment established by Ontario Hydro, namely:
(1) energy and resource use efficiency, (2) environmental integrity, (3) renewable energy, (4) financial integrity, and (5) social integrity.
For the first category of energy and resource use efficiency the following indicators are considered in a compounded manner:

• power consumption and transmission losses as a percentage of sales


• fuel conversion efficiency
• water withdrawals
• fuels and commodities consumption
• internal energy recovery and savings
For the environmental integrity category, the monitored emission rates are used for all activities including design, development,
construction, commissioning, operation, decommissioning, and material management. Here are the considered individual sus-
tainability indicators for this category:

• GHG emissions
• ozone-depleting substances emissions
• acid gas releases
• waste management indicator
• radioactivity levels of generated wastes
• hazardous wastes emissions
• reportable spills
• compliance violations
• environmental expenditures

Regarding the specific indicators for renewable energy use, the following indicators are considered:

• energy share generated from renewable sources


• energy generated from wind power
• energy generated from solar power

The consistent generation of cash flow is considered by Ontario Hydro with the help of the financial integrity category for
sustainability assessment. Here, the following specific indicators are considered:

• net income
• interest coverage
Sustainability Dimensions of Energy 117

• debt ratio
• total cost of the energy unit
The interaction of Ontario Hydro with the communities and with its own employees determines the social integrity category for
sustainability assessment. These specific indicators are selected such that innovation and greater employee involvement in sus-
tainability is encouraged; thence, the indicators are as follows:

• employee accident severity


• corporate citizenship program
• employee productivity
• payments in lieu of taxes
• number of public fatalities
• aboriginal grievances
• severity of environmental complaints and their number
The Ontario Hydro corporate developed two composite indices that assess sustainability on long-term targets and are based on
the above five categories of specific indices. These are (1) resource use efficiency composite indicator focusing on inputs (e.g., fuels,
water) and (2) environmental performance indicator focused on outputs (pollutants and wastes emissions). The composite
indicators are constructed such that they take positive values.
The reference value is set to zero at the level of year 1995. The value of 100 was predicted to be reached by the composite
indicator in year 2000, while in year 2002 it reached the value of 190. This case study shows how a corporation can assess
sustainability internally, which helps adapt its strategy to achieve its goals. In this respect, an adoption of a clear definition of the
development vision is crucial. It is noteworthy that the sustainability must be related to both the quantity and the quality. Some
sustainability indices are introduced next based on past works of Midilli et al. [5] and Dincer and Zamfirescu [7]. Those are defined
as follows:

• Ecological footprint (EF) analysis is an accounting tool enabling the estimation of resource consumption (RC) and waste
assimilation requirements of a defined human population or economy in terms of corresponding productive land use.
• Sustainable process index (SPI) is a means of measuring the sustainability of a process producing goods. The unit of measure is
square meter (m2) of land. It is calculated from the total land area required to provide raw materials, process energy (solar
derived), infrastructure (including energy generation production facilities), and waste disposal.
• Sectorial impact ratio (Rsi) is based on the provided financial support of public, private, and media sectors for transition to
green energy-based technologies, and depends on the total green energy financial budget as a reference parameter.
• Technological impact ratio (Rti): this parameter quantifies the provided financial support for research and development,
security, and analysis of green energy-based technologies. This parameter depends on the total green energy financial budget as
a reference parameter.
• Practical application impact ratio (Rpai): this parameter quantifies the provided financial support for design, production,
conversion, marketing, distribution, management, and consumption of green fuel from green energy sources and also depends
on the total green energy financial budget.

Another largely used sustainability indicator based on mass and energy balances is the eco-efficiency indicator representing the
ratio between value in outputs and the EI. If one denotes with eco-indicator (EcI), and with energy product value (EPV), and with
EI associated to the EPV, then the EcI is expressed mathematically as follows:
EPV
EcI ¼ ð1Þ
EI
Linke et al. [8] introduced an assessment indicator for sustainability denoted as the “sustainability efficiency indicator” (SEI)
defined by the ratio between a selected performance parameter (PP) and quality parameter (QP) for a specified RC; this has the
mathematical formula as follows:
PP
SEI ¼ ð2Þ
QP  RC
The SEI indicator can be extended for drying systems in various ways, such as in the form of SEs per final moisture and energy
input, where the PP is the SEs, the QP is the final moisture content, and the RC is the energy input.
Sustainability indicators of various types can be aggregated into one single figure of merit for sustainability, denoted as an
aggregated sustainability indicator (ASI). The aggregated indicators for sustainability can be formed by various possible methods
of normalization, ranking, weighting, and scaling. In order to be possible to aggregate the individual indicators, these have to be
normalized first. The normalization can be done based on a reference sustainability indicator. Various methods of normalization
were proposed in past studies, such as in Refs. [9–11]. Denote Indref an arbitrary selected reference sustainability indicator. This
indicator can be the maximum of the dimensionless indicators, defined as follows:
Indref ¼ max fIndi ji ¼ 1; 2; …; ng ð3Þ
where n is the number of dimensionless sustainability indicators Indi to be aggregated.
118 Sustainability Dimensions of Energy

Accordingly, the normalization can be done based on the reference value as follows:
Indi
NSIi ¼ ð4Þ
Indref
The sustainability indicator can be also normalized based on the relative alleviation from the average, as follows:
Indi  Ind
NSIi ¼ ð5Þ
Indi
where Ind represents the average of dimensionless indicators.
Instead of the average value, a target value of the indicator can be used in Eq. (5). A variation for the normalization given by Eq.
(5) is as follows, where the weighted average is used instead of the arithmetic average as follows:
Indi  Ind s
NSIi ¼ ð6Þ
si
where si represents the standard average of Indi.
The normalization can be also done by scaling as follows:
Indi  minfIndi jiA f1; ngg
NSIi ¼ ð7Þ
max fIndi jiA f1; ngg  minfIndi jiA f1; ngg
Based on the normalized sustainability indicator (NSI) various aggregation possibilities exist such that an aggregated
(or compounded) sustainability indicator can be obtained. A typical approach is that of a weighted average, determined as
follows:
P
wi NSIi
ASI ¼ P ð8Þ
wi

where ASI is the aggregated sustainability index and wi are the weighting factors, which can be chosen or determined based on
various considerations; in principle, the weighting factors can be determined as follows:
NSIi
wi ¼ P ð9Þ
NSIi
The use of weighting factor injects an element of subjectivity on the ASI. Based on [11] the weighting factors can be
defined according to trade-offs between different criteria, each criteria being represented by a normalized indicator. The trade-
offs are generally established by the decision maker and therefore the vision of decision maker influences the process. Three
types of decision maker archetypes can be envisioned, namely, the individualist, the egalitarian, and the hierarchist.
The individualist decision maker is not interested on the inter- and intragenerational equity in relation to sustainability. She/he
will favor criteria related to production performance to the detriment of criteria related to environmental benefit. The egalitarian
decision maker is concerned with the inter- and intragenerational equity, therefore he/she adopts a wider view with attention to
the long-term effects and EI. The moderate approach is that of the hierarchist decision maker who will compromise and negotiate a
balanced approach on sustainability.
Making energy systems more sustainable is the synonym of increasing their performance and efficiency, while enhancing the
share of renewable energy supply. Greenization effort is a key solution to achieve sustainable energy systems by trying to make
them more efficient, more environmentally benign, and more cost effective. An ASI for energy systems will then account for the
relative improvement toward greenization with respect to a baseline case. Therefore a greenization factor (GF) has been proposed
in Ref. [12]. This factor has been related to the ASI as shown in [13]:
ASI  ASIref
GF ¼ ð10Þ
ASI
In Eq. (10), the following inequality is assumed: ASI4ASIref. Therefore a GF of zero signifies no improvement with respect to
the baseline system. When ASI becomes significantly larger than ASIref, the GF defined according to Eq. (10) tends to be 1. Fig. 14
shows the greenization process for improved sustainability of energy systems. The trend must be a reduction of dependence on
polluting conventional sources and an increase of the ability to convert renewable energy sources. In the same time, the efficiency
of the energy system must increase in time. Furthermore, due to an increased use of green energy sources and an increased system
efficiency, the environmental pollution can be reduced and minimized for the same net output.
This requires an ultimate modification and retrofitting of the existing energy systems. Increasing the energy system efficiency is
another key for a greener system. For steam power plants, this can be achieved by optimizing the operating parameters and the
components of the system. This would include better turbine designs with higher efficiencies, and heat exchangers with higher
effectiveness and lower temperature differences, and the selection of the boiler. Integrated energy systems for multigeneration
purposes are a promising approach for achieving greener and sustainable energy systems through increasing the operating
efficiency of the system.
The technology changes of integrated power generation systems have evolved so fast during the last decades. Cogeneration,
mainly combined heat and power (CHP), extended for trigeneration by producing cooling as a useful commodity of the system.
Sustainability Dimensions of Energy 119

Green Greener
Conventional
Green + conventional
(polluting) energy Green energy sources
energy mix
sources

Less fossil
fuel-based
energy

Useful Useful Useful


energy energy energy
Greenized, Green, highly
Reference
more efficient efficient energy
energy system
energy system system

Less
Polluting No or miror
polluting
effluents polluting
effluents
effluents

Less polluted Minimally polluted


Polluted environment
environment environment

Fig. 14 Greenization of energy systems for increased sustainability.

Further commodities have been recently considered for multigeneration energy systems. Besides electric power, heating, and
cooling, other useful outputs include fresh water, domestic hot water, hydrogen, syngas, ammonia, and other commodities. For
such systems, it is expected to perform at higher efficiencies and lower EI, which results in greener systems. These systems can be
further greenized when considering renewable energy sources to fuel the system.

1.4.4 Exergy-Based Sustainability Assessment of Energy Systems

Exergy analysis offers a basis for sustainability assessment because exergy is a measure of abatement of the system subjected to the
analysis from the environment. Exergy analysis is of major importance in assessment of sustainability, because exergy-based
efficiency of systems and processes represent a true measure of imperfections. It also indicates the possible ways to improve the
energy systems and to design better ones. Destruction of exergy must be reduced as much as possible. Assessment of the exergy
destruction offers the opportunity to quantify the EI and the sustainability of any energy system Ness et al. [3] categorizes exergy as
one of the emerging methods for sustainability assessment.
A precursor of the exergy-based sustainability assessment is regional and sectorial exergy analysis. Wall [14,15] presented exergy
analysis for Japan and the United States, respectively. Utlu and Hepbasli [16] presented a sectorial exergy assessment for Turkey.
Sectorial exergy assessment of transportation sector is comparatively studied for six countries in Ref. [7]. In a sectorial exergy
assessment, a geopolitical region or the whole world is analyzed based on energy and exergy method accounting for the involved
mass, energy, and exergy balances. In this analysis the thermodynamic system can be a sector of activity, such as industrial,
commercial, transportation, agricultural, utility, etc.
Fig. 15 shows a model for sectorial exergy assessment. The exergy efficiency of a sector will then be defined as the ratio between
the exergy delivered to the user and the exergy input. One of the most important sectors is the electric utility, which is responsible
for power generation and distribution. The electric utility sector does not include the electricity produced by industrial estab-
lishments for their own use. The electric utility sector also includes electrical power generated by desalination plants. Desalination
plants produce electricity as a byproduct with a high efficiency, that is, about 46% as compared to the average efficiency of a power
station, which is 28%. Countries with desert regions, such as Saudi Arabia, use water desalination to produce drinkable water. In
Saudi Arabia, the overall efficiency of the utility sector, calculated by Ref. [17], is 31.75%. In Ref. [7] the averaged world exergy
efficiency of the electric utility sector is determined to be approximately 25%.
Sectoral exergy assessment has been expanded to cover environment and sustainable development by Ref. [18]. It is
shown that exergy is at the confluence of energy, environment, and sustainable development. The need to understand the
linkages between exergy and energy, and the EI, has become increasingly significant. Less exergy destruction implicitly leads to
reduced EI. Further, as energy policies increasingly play an important role in addressing sustainability issues and a broad
range of local, regional, and global environmental concerns, policy makers also need to appreciate the exergy concept and its
120 Sustainability Dimensions of Energy

Energy loss/
exergy destroyed

Thermodynamic macro-
system
(activity sector, geopolitical
region, country, etc.)
Energy, exery
inputs Useful energy/
(primary energy exergy
sources/
associated
exergy)
Fig. 15 Thermodynamic analysis at macroscale based on energy and exergy.

ties to these concerns. The environmental impact assessment (EIA) methods for energy systems can be classified in four
categories:

• Environmental tools, including impact assessment, and EFs.


• Thermodynamic tools; performance indicators; energy, exergy, and material flux analyses.
• Sustainability tools, including lifecycle assessment, SPI, exergetic sustainability index (ExSI). exergetic improvement potential
(IP), GF, etc.
• Risk assessment.

EIA is an environmental tool used in assessing the potential EI of a proposed activity. The derived information can assist in
making a decision on whether or not the proposed activity will pose any adverse EIs. The EIA process assesses the level of impacts
and provides recommendations to minimize such impacts on the environment.
Risk assessment can estimate the likelihood of potential impacts and the degree of uncertainty in both the impact and the
likelihood it will occur. Once management has been informed about the level of risk involved in an activity, the decision of
whether such a risk is acceptable or not can be subsequently made.
Bejan [19] studied the application of entropy generation minimization principles to formulate energy policy aspects. The
primary purpose of Bejan’s work was threefold: (1) to employ the art of entropy generation minimization at a level as yet
unexplored; (2) in the process, to present a unified framework with a sound theoretical basis for making and analyzing energy
policy proposals; and (3) to demonstrate the potential benefits of a dialog between different disciplines, academic or professional,
by presenting the result of an ongoing one between an “accountant” and an “engineer.” Further developments led to the use of
exergy destruction within energy systems as a quantifier of EI. Therefore exergy-based methods can be developed to help policy
making.
The relative magnitude of exergy destruction within a system or process with respect to energy input suggests the meaning of a
resource depletion factor. Connelly and Koshland [20] proposed such a factor denoted by them as “depletion number” and
defined as follows:

Exd
Dp ¼ ð11Þ
Exin

Assume that the analyzed system is from the domain of industrial ecology, that is, it comprises a number of combined
technologies and industrial fluxes that operate as a whole. In this case, the exergy efficiency of the integrated system can be
evaluated on the base of depletion number of each independent component. In an abstract manner, the exergy efficiency of
ðcombÞ
industrial ecology systems is illustrated in Fig. 16. For combined technologies, the depletion number Dp is lower than that for
ðsepÞ
separate technologies Dp , which is expressed by

_ comb
Ex _ comb
Ex
ð1Þ ð2Þ
DðsepÞ
p1 p2
¼ DP þ comb DP ð12Þ
p _ comb
Ex p1
_ p2
þ Ex comb _ p1 þ Ex
Ex _ comb
p2

_ comb and Ex
where Ex _ comb are the rates of output exergy flows for products 1 and 2, respectively.
p1 p2
The application of exergy method to an industrial ecology analysis can be done by calculating the exergy flows of every stream
of matter and energy and associating depletion numbers to every independent technology or process. Further, the depletion
Sustainability Dimensions of Energy 121

Exergy rate of
output streams
Exergy rate of Exergy rate of
output streams output streams Exp(comb)
Exp(1) Exp(2)

Combined technology

Technology 1 Technology 2 Technology 1 Technology 2


Depletion number Depletion number Depletion number Depletion number

Dp(1) Dp(2) Dp(1) Dp(comb) Dp(2)

Exergy rate of Exergy rate of


input streams input streams Exergy rate of
(1) (2)
ExIn input streams
ExIn
(comb)
Exin

(A) (B) (C)


Fig. 16 Depletion number of separate and combined technologies.

number of the separate and combined technologies is calculated and compared in order to quantify the benefit of technology
integration from both an energetic and ecologic point of view.
The depletion factor is related to the exergy efficiency of the system as follows:

c ¼ 1  Dp ð13Þ

Let us consider an energy source, such as a fuel. When the fuel is used in an engine its internal energy is converted into work.
The maximum energy conversion potential equals the chemical exergy of the fuel. Better converter is the same as higher exergy
efficiency or smaller depletion number, according to Eq. (13). A SI can be introduced as the reciprocal of the depletion number.
Such a SI can be determined for any type of energy system. Mathematically, the definition of SI is written as follows:

SI ¼ 1=Dp ð14Þ

Another criterion to assess a renewable energy system is the IP defined as follows:


 
Exout
IP ¼ Dp 1  ð15Þ
Exin
In addition, the GF introduced previously according to Eq. (10) can be expressed alternatively using exergy destruction as
follows:
Exd;ref  Exd
GF ¼ ð16Þ
Exd;ref
where Exd is the lifecycle exergy destruction of the studied system and Exd,ref is the exergy destruction of the reference system.
Both the system and the reference system process generate the same amount of product. But, the energy, exergy, environmental,
and cost parameters of the two systems differ. Observe from Eq. (16) that if Exd-0 then GF-1, meaning that the system tends to
be greener. If xd ¼ Exd,ref, no greenization effect is observed, both the reference and the studied system having similar EI.
The GF can be also defined based on an EI indicator. If one denotes EI, then the GF for an energy system becomes:
EIref  EI
GF ¼ ð17Þ
EIref
In the above equation, the EI must be specified for two cases: the reference system and the greenized system. Depending on the
specific problem analyzed, various types of EI factors may be formulated. For many energy systems, specific GHG emissions can be
used as EI factor.
Fig. 17 shows the representation of a generic energy system for sustainability analysis. The scope of the analysis is the entire
lifespan of the system. The energy from the source is converted to a useful form, such as power or exergy-carrying synthetic fuel.
The produced exergy corresponds to stream 5 in the figure. A part of this stream is then used to maintain the system in operation
for the LT span and to construct a newer system, ready for operation at the end of the lifespan of the old system. The net exergy
122 Sustainability Dimensions of Energy

System boundary for exergo-


sustainability assessment

Source
exergy
input Overall system for 5 7 Exergy output to user
the lifecycle
1
Newly constructed
6 system
System construction process
(exergy is destroyed)

Exergy destroyed
at system
construction
Exergy destroyed Materials
at system operation
Fig. 17 Representation of an energy system for exergosustainability assessment.

Sustainable
development

Environmental
Efficiency impact

Exergy

Energy Environment
Energy and
material
balances

Fig. 18 Representation of the exergy at the confluence of energy, environment, and sustainable development. Reproduced from Dincer I,
Zamfirescu C. Sustainable hydrogen production. New York, NY: Elsevier; 2016.

given to the used equals to the difference between stream 5 and stream 6. If the exergy output in 7 is positive (if there is output)
then an ExSI can be defined as follows:
Ex7
ExSI ¼ ð18Þ
Ex5

Other approaches based on the second law of thermodynamics led to proposal of sustainability indicators, such as renewability
or resources, toxicity of generated emissions, input of used materials, recoverability of the products at the end of their use, and
technological efficiency [21].

1.4.5 Sustainability Assessment Based on Exergetic Lifecycle Analysis

Exergy, EI, and economic factors are well interrelated with sustainability. Besides, these energy and material balances and system’s
efficiency are important. Fig. 18 shows a representation of the intimate connection between exergy and sustainability and other factors.
As shown by Dincer [22], exergy efficiency can be correlated with SI and the EI index. It is also correlated with the lifecycle
sustainability index (LCSI). Here, the difference between SI and LCSI consists of the fact that SI refers to the system utilization
Sustainability Dimensions of Energy 123

phase only, while LCSI considers the system lifecycle including the construction and scrapping phases. An exergetic lifecycle
analysis can be conducted to determine the lifecycle exergy destruction and from here to assess the system sustainability.
By using exergy destruction as a measure to quantify environmental effects associated with emissions and resource depletion,
the EIA is made based on purely physical principles. 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 damage. The discharged wastes in the environment and their impacts can be quantified by accounting for
exergy destructions. It is known that the exergy destroyed (and wasted) by the system is of two kinds:

• internal exergy destruction, which represents the lost opportunity to perform work; the EI and rejected wastes due to all
upstream processes (e.g., power generation) can be related to the internal exergy destruction;
• lost exergy, or exergy destruction at system interaction with its surroundings, which is related to the discharged wastes by the
process itself. In principle, all discharged wastes by the system can be recovered to use their exergy and reduce the EI; however,
this action may be very expensive and generally is not undertaken in practice, except when it is justified economically or
enforced by sustainability policies.

The economic factor is also an important one to consider in sustainability. Economy is related to wellbeing, and it connects to
value and utility. In the real world, the economic factor has one of the most important influences in selecting any technical design.
Ultimately, a business case must be presented in terms of investment cost (IC) and profitability, and this guides the system
development. It is worth noting that the cost is a very volatile factor; therefore, any technical-economic analysis is not absolute, but
subjected to regional and temporal economic constraints.
In economics, the theory of value attempts to explain the correlation between value and price of traded goods and services. In
today’s economy, the trades are normally made by paying a price in money as a standardized currency for payments of goods,
services, and debts. The price must reflect the value of the trade, which is related to costs and profitability. The theory of value
offers an ideological basis for quantification of the benefit from a traded good or service. This helps assigning a price to a value.
Three theories of value received much attention in the last century.
The first is the power theory of value, which states that political power and the economy (which is constrained by laws of trade)
are so highly interlaced that prices are established based on an internal hierarchy of values of the society rather than on a
production and demand balance. The second is the labor theory of value, which states that the value is determined by the labor
developed to produce the good or the service including the labor spent to accumulate any required capital for the production
process. The third is the utility theory of value, which quantifies the value of tradable goods and services based on their utility.
There is no direct way of measuring the utility as a representation of the preferences for trading various services or goods because
this depends on subjective factors of human individuals, such as wishes and wants. However, the utility can be observed indirectly
though the price that is established by trading activity. The price is determined by the balance between marginal utility and
marginal cost. Here, the term marginal means an infinitesimal change. Let us assume that a quantity Q of products are traded; if
one denotes U a quantified utility of the product, then the utility marginality is the derivative dU/dQ; also, if the production cost is
C, then the marginal cost would be dC/dQ. The price according to the utility theory of value is spontaneously established such that
dU dC
¼ ð19Þ
dQ dQ
Eq. (19) explains why gold price is much higher than that of water. The marginality of gold cost (the change in cost for an
infinitesimal change of quantity) is obviously much higher than the marginality of water cost. The utility of water is much higher
than the utility of gold, but the scarcity of gold is high, whereas the availability of water is high. This means that the marginality of
water utility is much smaller than the marginality of gold utility, and therefore the marginalities of the costs of water and gold
must be on the same relationship.
As Georgescu-Roegen [22] points out, if a theory quantifies value through a conservable quantity then it fails. For example, if
when applying the theory of labor one assumes that any type of labor can be valued through the amount of mechanical work (or
energy) deployed to do it, then, this leads to misappropriations (which are in fact common) because energy does conserve and
value does not. Value degrades or augments. Another example is that if, when applying the theory of utility, one assumes that the
utility is expressed in terms of mass (amounts, quantities) of a precious metal (gold), then this is a fallacy because although finite,
mass is conserved (we do not consider here any nuclear reactions).
The true values that humans and also any other living organisms appreciate are the sources of low entropy or high exergy. These
quantities do not conserve. When used, a low entropy source is converted in a high entropy waste by living organisms, which in
the meantime perform their activity. Equivalently, one says that the high exergy sources are degraded by the Earth systems (living
species, natural cycles, etc.): exergy degrades from the source to waste. The net exergy absorbed by the Earth, consequently, is
gradually destroyed, but during this destruction, it manages to drive the Earth’s water, wind, and other natural systems, as well as
life on Earth. Because it does not conserve, a source of low entropy can be used only once and never reused. The same stands for
exergy: once destroyed it cannot be reused. Exergy is also related to the surrounding environment as it accounts for its temperature,
pressure, and species concentration. Therefore due to these attributes, exergy can be used in establishing a theory of value. In fact,
exergy represents the part of energy that is useful to society and therefore it has economic value.
Furthermore, once the economic value of exergy is expressed in terms of currency, then it can effectively be used for exer-
goeconomic and exergoenvironmental analyses. Various methods can be approached to price exergy for analyses purpose. It is
124 Sustainability Dimensions of Energy

important to determine sound methods to set the prices and the costs in relation to exergy content. This in fact requires
formulation of a theory of costs based on exergy. It has been suggested that when analyzing a thermal system it is reasonable to
distribute costs in relation to outputs and accumulations of exergy. With regards to the prices of physical resources (fuels,
materials), these also must be set in a tight relation with the resource exergy content, such as to foster resource saving and effective
technology.
Let us do a simple attempt to quantify exergy in terms of monetary currency. In doing this, one considers the main fuels in a
society. The energy content is taken as the lower heating value (LHV) of the fuel and the exergy content is the chemical exergy of
each fuel. Table 2 gives the specific energy and exergy content of fuels considered in this brief analysis. The price per unit of mass of
each fuel is also given. When the price is divided to chemical exergy, then the exergy-specific price Cex is obtained. Table 2 is
constructed for Canada; however, the methodology presented subsequently is general.
In our approach, a country or region must be considered first. Then, the primary energy sources are inventoried. For Canada,
the following primary energy sources can be considered: coals, refined natural gas, natural gas liquids, crude oil derivate, hydro,
nuclear, and biomass derivate (here, wind and solar are neglected, as they are not highly represented). Further, the method for
exergy price estimation goes as follows:

• Cost of each fuel type is obtained from the market and expressed in dollars per kilogram. For the case of hydro and nuclear, this
step is skipped.
• Based on the available statistics, the consumed energy fraction (CEF) for each type of fuel is determined. In Table 2 the CEF is
obtained from the previous work by Dincer and Zamfirescu [7], chapter 17. The CEF represents the fraction of specified primary
energy source from the total energy consumed from primary sources.
• The LHV and specific chemical exergy for categories of fuels are averaged. For example, the mean LHV for natural gas liquids is
an average of the LHVs for the LPG, methanol, and ethanol.
• The specific price of fuel is averaged for each fuel category. For example, the mean fuel price Cf for fuels obtained from crude oil
(crude oil derivate) is the average of prices of gasoline, diesel, kerosene, and fuel oil.
• The quality factor for each category of fuel is determined as indicated in the table, that is, by the ratio between averaged specific
chemical exergy and LHV. The quality factor for hydropower and nuclear energy is 1.
• The exergetic price Cex for each fuel category is determined as shown in the table, by the ratio between the averaged fuel cost and
specific chemical exergy. The exergetic price of hydropower results from the specific price of electric power divided by 0.8, as it is
fair to assume that the exergy efficiency of the hydro power plant is 80%. The exergetic price of nuclear energy is determined
from the price of electric power divided by 0.31 based on the fact that, as shown in Dincer and Zamfirescu [23], the exergy
efficiency of CANDU power plants is 31.3%, in average. The cost of electric power in Canada is taken at an average cost of
¢11/kWh.
P
• The consumed exergy factor for each fuel category is calculated with CExFi ¼ CEFigi/ (CEFigi), where i is an index representing
each type of the fuel. These represent weighting factors.
P
• The averaged price of exergy results as a weighted average, Cex ¼ CExF i  Cex;i .
• The exergy price for Canada, based on the exergy of the primary resources, is 8.4 ¢/kWh or 23.3 $/GJ. This compares well with
the average electricity price of ¢11/kWh.

Once a price of exergy is determined, further models can be created to establish a costing scheme for other items of interest in
an exergosustainability analysis. Nonenergetic costs, such as labor, material supply, environment remediation expenditure, inci-
dental expenditures, etc. can be priced using exergy content as a basis for cost accounting. The economic value of system outputs
can be also allocated based on exergy.
The practical connection between exergy price, wasted exergy, and EI can be discovered by correlating recorded emission data at
the level of a region or globally with the chemical exergy of emitted pollutants. Here, some exergy destruction versus emission data
correlation is presented for Ontario. The general aspects of environmental policy in Ontario can be described as follows:

• in Ontario, the Environmental Protection Act by the Ministry of Environment exists, giving the legislation on environmental
quality and air pollution limits, which are conceived such that human health and the ecosystem are not endangered;
• the potential of a substance to impact the environment is evaluated using a set of 10 parameters:
○ transport
○ persistence
○ bioaccumulation
○ acute lethality
○ sublethal effects on mammals
○ sublethal effects on plants
○ sublethal effects on nonmammalian animals
○ teratogenicity
○ mutagenicity/genotoxicity
○ carcinogenicity
• an aggregated indicator is determined based on the 10 impact parameters (above), referred as the point of impingement (PoI),
which is determined based on the best known available pollution control technology;
Table 2 Calculation table for exergy price based on primary energy

Fuel Data input Calculated item


            $ $
LVH MJ ex ch MJ
Cf $ CEF(%) LHV MJ ex ch MJ Cf $ g Cex CEF  g CExF(%) CExF  Cex
kg kg kg kg kg kg GJ GJ

Coal 24.0 15.0 0.15 8.3 24.0 15.0 0.15 0.62 10.0 0.052 5.1 0.51
Refined natural gas 50.7 52.4 0.18 30.8 50.7 52.4 0.18 1.03 3.43 0.318 31.3 1.07
Natural gas liquids LPG 46.0 54.9 1.43 2.9 31.6 35.6 0.93 1.13 26.1 0.033 3.3 0.85
Methanol 19.9 22.4 0.47
Ethanol 28.8 29.5 0.88
Crude oil derivate Gasoline 43.5 47.7 1.74 45.1 42.4 45.5 1.37 1.07 30.1 0.484 47.7 14.3
Diesel 42.8 44.2 1.78
Kerosene 43.1 49.1 0.94
Fuel oil 40.1 41.1 1.03
Hydro N/A N/A N/A 7.4 N/A N/A N/A 1 34.7 0.074 7.3 0.16
Nuclear N/A N/A N/A 5.0 N/A N/A N/A 1 89.6 0.050 4.9 6.6
Biomass derivate Whole tree 19.7 22.1 0.2 0.5 15.6 17.8 0.05 1.14 2.8 0.005 0.4 0.14
Wood pellets 14.6 18.5 0.2
Wood chips 10.0 11.0 0.2
Pine wood 18.9 24.8 0.5
Sawdust 8.0 8.5 0.2
Straw 14.5 16.5 0.3
Rice straw 14.1 15.9 0.8
Waste paper 17.7 20.1 0.1

Sustainability Dimensions of Energy


Biogas 22.5 ch 23.2 2.0
Equations used: g ¼ ex Cex ¼ Cf
ch
CExF ¼ PCEF g
ðCEF gÞ
Averaged price of exergy:
8.4 ¢/kWh
LHV ex
Averaged price of electricity: 11.0 ¢/kWh

Source: Reproduced from Dincer I, Zamfirescu C. Drying penomena: theory and applications. New York: Wiley; 2016.
Abbreviation: N/A ¼ not applicable.

125
126 Sustainability Dimensions of Energy

• the methodology denoted removal pollution costs (RPCs) is applied to correlate the exergy of the waste stream with the cost of
removing pollutants from the waste stream prior to discharge into the surroundings. The cost for waste emissions is evaluated
as the total fuel cost per unit fuel exergy multiplied by the chemical exergy per unit fuel exergy, and divided by the exergy
efficiency of the pollution removal process;
• in Canada, the environmental pollution costs (EPCs) are estimated based on qualitative and quantitative evaluations of the
pollution cost to the society for compensation and correction of the environmental damage and to prevent harmful discharges
into the environment. Table 3 gives the EPCs of Ontario pollutants;
• the average composition of volatile organic compound (VOC) emissions in Ontario are approximated as given in Table 4;
• the average composition of particulate matter (PM) emissions in Ontario are given in Table 5;
• the fuel cost for three types of fossil fuels, namely, coal, no. 6 fuel oil, and natural gas are in average value of CN$ 2013 as
follows:

Table 3 Estimations of exergetic cost of atmospheric pollutants for Ontario

Pollutant Environmental pollution M (kg/ Molar environmental exch exPC Point of impingement
cost (EPC) ($/kg) kmol) pollution cost (MJ/kmol) ($/MJ) (PoI) (mg/m3air)
(MEPC) ($/kmol)

CO2 0.0402 44 1.77 19.60 0.090 56,764


CH4 1.2998 16 20.80 831.66 0.025 N/A
NOx 3.8944 38 148.18 72.4 2.047 500
SO2 3.551 64 227.26 310.99 0.731 830
CO 5.5074 28 154.21 275.00 0.561 6000
Volatile organic compounds (VOCs) 0.603 44 26.53 1233.00 0.021 N/A
Particulate matter (PM) 5.5074 68 374.50 436.00 0.860 N/A

Sources: Reproduced from Carpenter S. The environmental cost of energy in Canada. In: Sustainable energy choices for the 90’s. Proceedings of the 16th annual conference of the
solar energy society of Canada, Halifax, NS; 1990. p. 337–42; De Gouw JA, Warneke C, Stohl A, et al. Volatile organic compounds composition of merged and aged forest fire plumes
from Alaska and western Canada. J Geophys Res 2006;111:D10303; Pellizzari ED, Clayton CA, Rodes CE, et al. Particulate matter and manganese exposures in Toronto, Canada.
Atmos Environ 1999;33:721–34, Ontario Regulation 419/05.
Note: Costs are in 2013$ based on the Canadian consumer price index.

Table 4 Approximated average volatile organic compound (VOC) composition and characteristics in Ontario

VOCs Formula M (kg/kmol) y (kmol/kmol) exch (MJ/kmol) Point of impingement (PoI) (mg/m3air)

Methanol CH3OH 32 0.020 612 12,000


Acetonitrile CH3CN 41 0.823 1169 180
Acetaldehyde CH3CHO 44 0.001 1063 500
Acetone CH3COCH 58 0.111 1636 48,000
Acetic acid CH3COOH 60 0.025 780 2,500
Butanone C2H5COCH3 72 0.015 2755 250
Toluene C6H5CH3 92 0.005 3771 2,000

Source: Reproduced from De Gouw JA, Warneke C, Stohl A, et al. Volatile organic compounds composition of merged and aged forest fire plumes from Alaska and western Canada. J
Geophys Res 2006;111:D10303, Ontario Regulation 419/05.

Table 5 Approximated average particulate matter (PM) composition and characteristics in Ontario

Particulate matter (PM) Formula M (kg/kmol) y (kmol/kmol) exch (MJ/kmol) Point of impingement (PoI) (mg/m3air)

Lead Pb 207 0.053 249.2 10


Cadmium Cd 112 0.097 298.4 5
Nickel Ni 59 0.185 242.6 5
Chromium Cr 52 0.210 584.4 5
Copper Cu 63 0.173 132.6 100
Manganese Mn 55 3.115E-7 487.7 7.5
Vanadium V 51 0.214 721.3 5
Aluminum Al 27 0.006 795.7 26
Calcium Ca 40 0.054 729.1 14
Magnesium Mg 24 0.008 626.9 60

Source: Reproduced from Pellizzari ED, Clayton CA, Rodes CE, et al. Particulate matter and manganese exposures in Toronto, Canada. Atmos Environ 1999;33:721–34, Ontario
Regulation 419/05 (2013).
Sustainability Dimensions of Energy 127

○ average coal: $1.411/GJ LHV


○ average no. 6 fuel oil: $1.864/GJ LHV
○ average natural gas: $3.899/GJ LHV
• Table 6 gives the EPC and RPC for the main fossil fuels in Ontario.

A simplified method to estimate the cost of pollutant removal from the waste stream is based on the exergy efficiency of
pollutant removal. According to Rosen and Dincer [27], this exergy efficiency is in the range of 1%–5%. Therefore once the exergy
destroyed due to pollutant discharge is known, the required exergy to remove the pollutant from the waste stream can be
calculated. Furthermore, the average price of exergy can be estimated for any geopolitical region; for Canada, it is approximated as
Cex ¼ 8:4 ¢/kWh ¼ 2.3 ¢/MJ. When the exergy required to remove the pollutants is multiplied by exergy price, the removal
pollutant cost RPC is obtained. Therefore one has:

RPC ¼ Cex cpr Exd;pw ð20Þ

where Cex is the exergy price, cpr is the exergy efficiency of pollutant removal from the waste stream, Exd,pw is the exergy destroyed
due to pollutant waste in the environment.
Assuming an average exergy efficiency of pollutant removal from waste stream of 3%, the RPC can be estimated as 0.7 $/GJ. The
pollution removal cost of power generation can be roughly estimated based on statistical data that allow for the estimation of the
exergy destructions. Table 7 gives the rough estimate of RPC associated with Canadian power generation. The cost of pollution
(CP) associated with the construction of power generation facilities, reparations, and maintenance is not included. As given, the
total RPC for power generation is B1870 mill $ and the overall exergy efficiency of the power generation sector is 51%. The RPC
becomes $2.5/MWh generated power.
Lifecycle pollutant emission from power generation technologies is determined in Table 8 based on multiple literature data
sources [24–26]. The amounts of atmospheric pollutants are given with respect to the gigajoule of source exergy. Using the data
from Table 7, weighted average pollutant emissions are obtained for the Canadian power generation mix. The averages are given in
kilogram of pollutant per megawatthour of power generated; in order to convert from source exergy basis to generated power basis,
the average Canadian exergy efficiency of power generation is used. Then, the exergy-based EPC for power generation is calculated
for each pollutant in dollars per megawatthour; the Canadian average of EPCex is $17.8/MWh. Therefore the cost of pollutant
removal from the waste stream is much lower to society than the cost of pollutant emission.
Materials used for system construction bring associated embodied energy (EE) and pollution. Table 9 gives the EE, specific
pollution, and exergetic pollution cost with various construction materials. The EE represents the amount of energy spent to
produce one ton of material. The specific emission (SE) represents the mass of pollution (here kilogram CO2 equivalent emitted in

Table 6 Environmental pollution costs (EPC) and removal pollution costs (RPC) for fossil fuels in Ontario

Pollutant EPC ($/GJfuel exergy) RPC ($/GJfuel exergy)

Coal No. 6 fuel oil Natural gas Coal No. 6 fuel oil Natural gas

CO2 5.2662 5.7352 7.7184 3.35 2.7604 1.7822


CH4 5.1724 1.1524 0 0.8174 0.1608 0
NOx 0.0469 0.03082 0.03082 0.938 0.469 0.3886
CO 2.1038 0.03082 0.3484 7.5442 0.0938 0.4556
SO2 0.4958 0.402 0 2.5594 1.5544 0

Source: Reproduced from Rosen MA, Dincer I. Exergy analysis of waste emissions. Int J Energy Res 1999;23:1153–63.
Note: Monetary values are in CN$ 2013.

Table 7 Pollution removal cost for power generation in Canada

Primary energy Exergy input (PJ) Power output (PJ) Exergy destruction (PJ) Removal pollution cost (RPC) (mill $) c (%)

Nuclear 1073 339 734 514 32


Hydro 1698 1358 340 238 80
Coal 630 378 252 176 60
Fuel oil 62.5 23 39.5 28 37
Natural gas 254 112 142 99 44
Natural gas liquids 33.6 8 25.6 20 24
Diesel fuel 8.3 3.2 5.1 4 38
Biomass 66 19 47 33 29
Secondary sources 1629 543 1086 760 33
Overall 5454 2783 2671 1870 51

Source: Reproduced from Dincer I, Zamfirescu C. Sustainable energy systems and applications. New York, NY: Springer; 2011 [Chapter 17].
128 Sustainability Dimensions of Energy

Table 8 Lifecycle emissions into the atmosphere for power generation technologies (kg/GJ)

Technology Kilogram per gigajoule fuel exergy

CO2 CH4 NOx SO2 CO Volatile organic Particulate matter (PM)


compound (VOCs)

Coal fired power plants 274 73 0.180 0.400 1.374 0.251 8.463E-6
Fuel oil fired power plants 176 47 0.115 0.200 0.876 0.160 5.439E-6
Natural gas fired power plants 112 30 0.073 0.150 0.558 0.102 2.557E-6
Photovoltaic (PV) power generation 44 12 0.045 0.090 0.038 0.003 1.336E-6
Wind power generation 33 9 0.033 0.070 0.028 0.003 1.024E-6
Hydro power 9 2 0.006 0.015 0.044 0.008 0.268E-6
Nuclear power generation 48 13 0.032 0.065 0.241 0.044 1.488E-6
Canada averages
CO2 CH4 NOx SO2 CO VOCs PM Total ($/MWh)
Kilogram pollutant per megawatthour power 42 11 0.029 0.060 0.194 0.035 1.288E-6
EPCex ($ per megawatthour power) 1.7 14.7 0.1 0.2 1.1 0.02 7.1E-6 17.8

Sources: Reproduced from Carpenter S. The environmental cost of energy in Canada. In: Sustainable energy choices for the 90’s. Proceedings of the 16th annual conference of the
solar energy society of Canada, Halifax, NS; 1990. p. 337–42; De Gouw JA, Warneke C, Stohl A, et al. Volatile organic compounds composition of merged and aged forest fire plumes
from Alaska and western Canada. J Geophys Res 2006;111:D10303; Pellizzari ED, Clayton CA, Rodes CE, et al. Particulate matter and manganese exposures in Toronto, Canada.
Atmos Environ 1999;33:721–34; Rosen MA, Dincer I. Exergy analysis of waste emissions. Int J Energy Res 1999;23:1153–63.
Abbreviation: EPCex ¼ exergetic environmental pollution cost.

Table 9 Embodied energy (EE), specific emission (SE) and exergetic pollution cost of construction
materials (CM)

Material EE (GJ/t) SE (kg CO2/GJ) EPCex ($2013/GJ)

Concrete 1.4 24 70.5


Iron 23.5 11 29.3
Steel 34.4 11 29.3
Stainless steel 53 62 29.3
Aluminum 201.4 10 29.9
Copper 131 57 60.1
Fiberglass 13 62 66.0

Source: Reproduced from Rosen MA, Dincer I. Exergy analysis of waste emissions. Int J Energy Res 1999;23:1153–63.
Abbreviation: EPCex ¼ exergetic environmental pollution cost; SE ¼ specific greenhouse gas (GHG) emissions.

Table 10 Embodied exergy (EEx) and the environmental pollution


costs (EPCs)

Material EEx (GJ/t) EPC ($/t)

Concrete 1.3 102


Iron 21.1 687
Steel 31 1009
Stainless steel 47.7 1553
Aluminum 181.3 6023
Copper 117.9 7873
Fiberglass 11.7 858

the atmosphere) per gigajoule of energy used in the fabrication process of the material. Concrete, copper, and fiberglass have the
highest EPC among the listed materials. The highest EE is due to aluminum fabrication, which as it is known, requires an energy-
intensive electrochemical process. Using an averaged quality factor, the embodied exergy (EEx) can be obtained from the EE in the
construction materials. Table 10 gives the EEx and EPCs for the respective construction materials.
Using the concepts introduced here, the exergy destruction can be utilized to determine the EI of a system in various manners as
follows:

• The removal of pollution cost can be approximated by multiplying the exergy destruction with the exergy price estimated for a
specific region or country.
• The rate of exergy input into the system can be used to determine the system physical size.
Sustainability Dimensions of Energy 129

• From the system physical size, the mass amounts of materials required for system construction are determined.
• The amount of each construction material correlates with embedded energy required for its extraction and with GHG emissions
and EPC (see Table 9).
• The lifecycle total exergy input into the system is equal to the exergy required for system construction, system operation, and
system salvage.
• Based on the lifecycle, total exergy input, and the exergy efficiency of power and heat generation system, the exergy destruction
can be determined at power and heat generation.
• The emissions and EPC due to lifecycle total exergy supply are determined based on the exergy destruction and exergetic EPC of
the power and heat generation subsystem.
• The exergy destruction of the system itself allows for determination of pollutant wastes and EPC for operation during the entire LT.
• If a percent of materials recycling is provided, then the wasted energy and emissions of scrapped system can be determined.
• The total pollutant emissions and EPC result from the summation of the terms associated to power and heat generation from
primary sources, system manufacturing, system operation, and system scrapping.

Scaling factors (SFs) are important for predicting the cost of a scaled-up system in correlation with its size (or production
capacity). Typically, for many process equipment and chemical plants, the SF is of the order of 0.6. Note that the system size and
capacity can be correlated to the produced exergy or consumed exergy. This means that the amount of construction materials
required to build an energy system can be approximated based on a scaling law correlated to the input exergy. Therefore the
amount of construction materials (AM) required to build the system can be modeled as follows:
_ 0:6
AM ¼ SF Ex ð21Þ
in
_ in is the exergy input.
where SF is the scaling factor and Ex
The SF depends on the type of the equipment (or plant) and the type of the material. EEx and the amount of materials can be
used to estimate the cost of materials as mapped on an exergy value. The following formula can be used to estimate cost of
construction materials (CM):
CM ¼ EEx  AM  SExC ð22Þ
where EEx is the embodied exergy in one ton of materials (cf. Table 10) and specific exergy cost (SExC), which cf. Table 2, which is
8.4¢/kWh or 23.3$/GJ.
In addition, the cost associated with the environmental pollution can be specified for each material with the help of the
indicator called EPC given in dollar per ton of pollutant. Table 10 gives the EPC associated to the fabrication of some materials.
Accordingly, the CP due to the use of a specified construction material can be determined as follows:
CP ¼ AM  EPC ð23Þ
The total cost of material and total CP are added to determine the capital cost (CC) in a fair way, which accounts for the
sustainability aspect, as follows:
CC ¼ CM þ CP ð24Þ
Assume that the money to cover CC is covered by capital bonds (CBs) offered by the government due to policy on alternative
energy and invested capital (IC) sourced both from private funds and government. Denote r the fraction of CB from total CC. Then
the IC becomes:
IC ¼ ð1  r ÞCC ð25Þ
The operation and maintenance cost can be modeled as a factor (fo&m) of the cost associated with exergy destruction. Denote
with payback period (PBP) in years. Then, the total O&M cost for the duration of PBP is given as follows:
_ d SExC fo& m PBP
Co& m ¼ 3600tyear Ex ð26Þ
The total cost spent to have and to run the system during the PBP, meaning the sum of IC plus the cost of operation and
maintenance, is
Ctot ¼ IC þ Co& m ð27Þ
Assume that the energy system produces the amount of exergy Exprod during the whole PBP. The value of the produced exergy is
then equal to ExprodSExC. If the total cost is smaller than the value of the produced exergy, namely, CtotoExprodSExC, then the
difference of ExprodSExC  Ctot represents exergy cost savings. Therefore one can define levelized exergy cost savings (LExS) as follows:
Ctot
LExS ¼ SExC  ð28Þ
Exprod
where LExS is measured in dollars per gigajoule.
In Eq. (28) the LExS is a parameter that accounts for EI due to system construction (through EEx and EPC associated with
materials fabrication) and also accounts for the costs due to operation and maintenance (through the factor fo&m, which can
incorporate the CP). Therefore LExS as defined is a parameter that quantifies the sustainability. If LExS is positive, then the system
may be sustainable; however, if LExS calculated with Eq. (28) results negative, then the system can be considered as
nonsustainable.
130 Sustainability Dimensions of Energy

In order to estimate the extent to which the system is sustainable, the system model shown in Fig. 17 will be used. We base the
sustainability assessment on the exergy method; therefore this is denoted as an exergosustainability assessment, and it is part of the
exergetic lifecycle assessment. The exergy balance equation must be written for the whole LT of the system. As a rule of thumb, the
LT is three times longer than PBP. If there is net exergy output in state 2 (Fig. 17), then the system is sustainable because in that
case, the system is able to produce sufficient exergy to construct itself and pay for its emissions. Otherwise, if net exergy in #5 is
zero or negative, the system is not sustainable. As can be deduced from the figure, the net exergy amount Ex7 can be approximated
as follows:
Ex7 ¼ Ex5  Ex6 ð29Þ
The term Ex6 in Eq. (29) represents the exergy amount embedded in construction materials plus the exergy required to remove
the pollutants emitted during the construction process from the atmosphere. This exergy can be related to the CC with the help of
the SExC. One obtains the following:
CC
Ex6 ¼ ð30Þ
SExC
An ExSI can be defined showing the following ratio of exergies
Ex7
ExSI ¼ ð31Þ
Ex5
The ExSI is subunitary. For a system to be sustainable, the ExSI must be positive and close to 1. The exergy amount delivered to
the user results from combining Eqs. (29) and (31) as follows:

_ 5  CC
Ex7 ¼ 3600LT tyear Ex ð32Þ
SExC
Therefore the exergy SI becomes:
CC
ExSI ¼ 1  ð33Þ
_ 5 SExC
3600LT tyear Ex

1.4.6 Clean Energy Solutions for Better Sustainability

In this section, clean energy solutions to achieve better sustainability are discussed. The possible sustainable energy solutions
include mainly renewables and hydrogen, which form the basis for clean energy systems. Those clean energy solutions are revised
here comparatively and ranked according to their outputs. In this respect, the work of Dincer and Acar [28,29] is taken as reference.
The ranking of the energy sources is based on technical, economical, and environmental criteria, which are then compounded to
obtain aggregated indicators for ranking.
Clean energy systems have the potential to do the following: (1) reduce emissions by taking advantage of renewable and
cleaner sources; (2) lower energy input requirements; (3) increase system efficiencies by expanding useful outputs (i.e., multi-
generation); and (4) reduce emissions and waste by recovering energy. In Refs. [28,29], the renewable energies are qualified as
sustainable energy based on dispatchability, geographical diversity, predictability, and control criteria. Indicators were used to
quantify the sustainability of renewable energies based on those four criteria. Each indicator has been set for a scale from 1 to 10.
Then an aggregated indicator has been calculated as the average of particular indicators. The results are shown in Fig. 19 in the
form of a chart.
The dispatchability indicator is ranked at maximum value (rank 10) when the power generator can be loaded from zero to full
capacity without significant delay. Geographical diversity indicator quantifies the degree to which siting of the technology may
mitigate variability and improve predictability, without substantial need for additional network. The technology with 100%
mitigation potential has a rank 10 assigned for the geographical diversity. The predictability indicator quantifies the accuracy to
which plant output can be predicted at relevant time scales. The control indicator shows technology capability of active control
and response during normal situations (steady state, dynamic) and during network fault situations. In terms of dispatchability,
biomass and geothermal have the highest performance, while ocean and wind have the lowest dispatchability. Wind has the
highest geographical diversity; on the other hand, it has very low predictability. In terms of control, biomass, geothermal, and
hydropower provide better performance. Overall, biomass and geothermal are closest to the ideal case, and wind shows the
poorest performance.
Electric power production from renewable energy is of significant interest in today’s world, both in centralized and decen-
tralized systems. In Dincer and Acar [28], a comparative assessment table is compiled giving the annual generation (TWh), capacity
factor, mitigation potential (GtCO2), energy requirements (in kilowatt-hour thermal per kilowatt-hour electric), specific GHG
emissions (gCO2/kWh), and production cost (US¢/kWh) (Table 11).
NSIs were then calculated based on each parameter. For the parameters that needed to be maximized, i.e., annual generation,
capacity factor, and mitigation potential, Eq. (7) multiplied by factor 10 is used to determine the normalized sustainability
potential. For parameters that need to be minimized, i.e., energy requirements, specific GHG emissions, and production cost, the
Sustainability Dimensions of Energy 131

10

Normalized sustainability indicator


8
Biomass
Geothermal
6 Hydropower
Ocean
Solar
4
Wind

2
y

l
tro
lit

ity hi

lit
bi

bi
rs ap

on
ha

ta
ve r

C
di eog

ic
tc

ed
pa

Pr
is
D

Fig. 19 Normalized sustainability indicators (NSIs) for various power generation technologies.

Table 11 Normalized sustainability indicators (NSIs) for renewable energies

Energy source NSIAG NSICF NSIMP NSIER NSIGHG NSIPC

Coal 10 9 0 1 0 10
Oil 1 8 0 9 9 0
Gas 5 6 0 8 6 0
Nuclear fusion 4 10 4 0 2 0
Biomass 0 6 2 10 0 0
Geothermal 0 9 5 0 0 1
Hydro (large scale) 4 4 5 0 2 1
Hydro (small scale) 0 5 3 0 0 3
Ocean 0 2 6 1 1 6
Solar (photovoltaic (PV)) 0 0 2 7 1 10
Solar (CSP) 0 2 2 1 0 6
Wind 0 1 10 0 0 0

Source: Reproduced from Dincer I, Acar C. A review of clean energy solutions for better sustainability. Int J Energy Res 2015;39:585–606.
Abbreviations: AG, annual generation; CF, capacity factor; CSP, concentrated solar power; ER, energy requirements; GHG, specific greenhouse
gas emissions; MP, mitigation potential; PC, production cost.

following equation is used to determine the normalized indicator:


max fIndi jiA f1; ngg  Indi
NSIi ¼ 10 ð34Þ
max fIndi jiA f1; ngg  minfIndi jiA f1; ngg
The obtained NSIs are given in Table 10. These results show that large-scale hydro and nuclear options have the highest annual
generation, and solar concentrated solar power (CSP) has the lowest. In terms of capacity factor, nuclear and geothermal give the
closest to ideal case results, while solar PV has the poorest performance. Mitigation potentials show that wind gives the ideal
results, and biomass has the least among the selected options. Geothermal and hydro have the ideal energy requirements, and
biomass has the poorest performance. Hydro has the lowest emissions, while solar technologies have the highest. When it comes
to production costs, nuclear, wind, and biomass have the best performance, while ocean and solar technologies have the highest
production costs per kWh electricity (Table 12).
The aggregation of the NSIs has been done in Dincer and Acar [28] by averaging. The aggregated indicator is used to rank the
sustainable technology. The bar chart shown in Fig. 20 compares the ASI for the renewable energies. Nuclear has the highest
ranking compared with renewables because it is already seen as a mature technology. In 2012, nuclear contributed 10.9% of the
total global electricity generation, while this number is 21.2% for all renewables combined. Nuclear also has a capacity factor of
86%, which is among the highest of all technologies and a competitive-levelized cost between 4 and 7 US¢/kWh.
Wind is the second strongest option with an annual growth rate around 34%. Wind technology is simple, and it is mature in
developed countries. Although wind energy is a small industry, it is competitive. Hydropower contribution percentage to overall
132 Sustainability Dimensions of Energy

Table 12 Normalized sustainability indicators (NSIs) for nonatmospheric emissions of energy sources

Energy Land use Water consumption Quality of discharge Ground contamination Biodiversity

Coal High 0 High 0 Moderate to high 1.6 Low to high 3.3 High 0
Gas Moderate 3.3 Low 6.6 Zero to high 5 Low 6.6 Low 6.6
Nuclear Moderate 3.3 High 0 High 0 High 0 Moderate to high 1.6
Biomass Low to high 3.3 Moderate 3.3 Moderate 3.3 Low 6.6 High 0
Geothermal Low 6.6 Zero 10 Low 6.6 Zero 10 Low 6.6
Hydro (with storage) High 0 Moderate3.3 Moderate 3.3 Moderate 3.3 Moderate 3.3
Hydro (run of river) Low 6.6 Low 6.6 Zero 10 Zero 10 Low 6.6
Ocean Low 6.6 Zero 10 Zero 10 Zero 10 Low 6.6
Solar (photovoltaic (PV)) Low to high 3.3 Zero to low 8.3 Low to high 3.3 Zero 10 Zero 10
Wind Moderate 3.3 Zero 10 Zero 10 Low 6.6 Low 6.6

Source: Reproduced from Dincer I, Acar C. A review of clean energy solutions for better sustainability. Int J Energy Res 2015;39:585–606.

8
7.06
6.49 6.44 6.57

6 5.4

4.17
ASI

4
3.14
2.66
2.3
2

0
al

)
on

e)

d
)

an

)
s

PV
le

SP

in
as

al
si

ca

ce

W
er

r(
sc
om

C
fu

ls

O
th

r(
la
ar

e
Bi

al
eo

So

la
rg
le

So
(la
G
uc

(s
N

ro

ro
yd

yd
H

Fig. 20 Aggregated sustainability indicators (ASIs) for nuclear and renewable energies. CSP, concentrated solar power; PV, photovoltaic.

renewable electricity generation is expected to decrease as geothermal, solar, wind, biomass, and ocean electricity generation
technologies evolve.
Various EIs can be considered when assessing the sustainability of energy technologies. Beside the atmospheric pollution, the
following EIs are relevant: land use, solid waste and ground contamination, biodiversity, water consumption, and quality of discharge.
Table 12 gives a sustainability assessment of energy technologies ranked with normalized indicators, which quantify nonatmospheric
pollution. The following ranks are assigned to the values of the indicators: zero (10), low (3.3), medium (6.6), and high (0).
When compared with the other options presented in the table, solar (PV and thermoelectric) has the lowest nonair impact.
However, the water quality/discharge issue should be addressed. Coal has the highest EI, which is expected. In regard to nuclear
power, radioactive waste and contamination appear to be major concerns as they need careful treatment and handling. Another
concern may be high water consumption in nuclear power plants. Land use of hydropower and adverse impact of biomass on
biodiversity should also be addressed in order to make them more sustainable. The NSIs for nonatmospheric emissions of energy
sources are aggregated by averaging. The aggregation results are shown in Fig. 21, where the energy technologies are ranked. Ocean,
geothermal, and wind result as the best sustainable technology according to nonatmospheric pollution-based indicators. The worst
options in this respect are coal and nuclear energy.
Besides power generation, the other major use of energy is for heating and cooling applications. The demand of heating and
cooling worldwide is significant. Conventional heating uses fossil fuel combustion. However, many new applications relate to
renewable energies for heating and cooling. Residual heat from industry is a beneficial source for nonconventional heating and
cooling applications. Renewable energy systems for heating and cooling use solar, geothermal, and biomass as sources. Ocean
thermal energy can potentially contribute to dispatching heating and cooling demands. Heat pumps, district heating, bathing/
swimming, pond heating, drying, refrigeration, HVAC, and industrial heat requirements are some of the current methods of
heating/cooling use. The cooling can also be produced by renewable energy-based absorption cooling.
The renewable heating and cooling technologies were ranked for sustainability assessment according to their typical capacity
(MW), IC, capacity factor, and system LT. Normalized indicators were obtained for these ranking criteria. The ASI for those
technologies is determined as an average of the NSIs. The results are shown in Fig. 22. The ideal ranking is 10. It shows that
geothermal district heating, geothermal pond-based heating, and biomass steam turbine cooling heating and power are the best
Sustainability Dimensions of Energy 133

8.64
7.96 7.96

Non-atmospheric pollution ASI


8 7.3
6.98

6 5.62

4 3.3
2.64

2
0.98 0.98

0
l

as

ar

al

e)

an

V)

d
oa

er
as

in
rm
le

ag

(P
G

ce
riv
C

W
m
uc

or

ar
o

th

of
N

st
Bi

l
eo

So
n
ith

(ru
G

(w

ro
ro

yd
yd

H
H
Fig. 21 Nonatmospheric pollution-based aggregated sustainability indicators (ASIs) for energy sources. PV, photovoltaic.

Solar domestic hot water


Geothermal heat pumps
Geothermal ponds
Geothermal greenhouse heating
Geothermal district heating
Geothermal building heating
Biomass anaerobic digestion CHP
Biomass steam turbine CHP
Biomass municipal solid waste CHP
Biomass domestic heating
0 1 2 3 4 5 6 7 8
ASI
Fig. 22 Aggregated sustainability indicator (ASI) for renewable heating technologies. CHP, combined heat and power.

sustainable technologies. Overall, the averaged ASI for biomass sources is the best with 5.2/10 ranking, followed closely by
geothermal with 4.9/10 and solar heating with 2.3/10.

1.4.7 Case Studies

1.4.7.1 Exergosustainability Assessment of a Concentrated Photovoltaic-Thermal System for Residential Cogeneration


In this case study, a novel concentrated photovoltaic (PV) thermal system that combined PV power generation with a special
Rankine engine is to be assessed. Fig. 23 shows a system description. The system has several compound parabolic concentrators
(CPCs) of through type installed on the southward face of a residence. The CPC concentrates light both on a vapor generator and
small-area PV module. The vapor generators produce high-pressure cyclohexane vapors in a calandria (thermosiphon) loop that is
part of an organic Rankine cycle (ORC). The system generates power through concentrated PV and the ORC jointly, and heat for
water heating through the ORC condenser.
The following emerging energy technologies are integrated in this system: concentrated PV with CPC as hybridized with an
organic vapor generator, cyclohexane ORC with thermo-mechanical solar energy storage and cogeneration. The hybridized CPC
system is detailed in Fig. 24. It captures sunlight under a half acceptance angle (yc ¼ 60 degree) and an aperture Aa. All light
captured at this angle will reach the receiver surface of aperture Ar and never escape back. Part of the light is absorbed directly by a
copper tube coated in black. The tube is placed in a glass shell, which is vacuumed. Inside the tube, saturated cyclohexane vapors
are generated in form of bubbles. A part of the light falls on the PV modules installed at the back. The modules are covered with
low-band reflection coating.
134 Sustainability Dimensions of Energy

Southward
roof
2
Compund
parabolic 3
concentrator Vapor
accumulator
11
1

5 4

Thermo-syphon loop

6
Regenerator
7
11
Condenser

Hot water tank


(cogeneration)

9
10
8

Fig. 23 Concentrated photovoltaic (PV) thermal system for residential cogeneration.

Direct
radiation

Aa

Concentrated PV module
covered with low band
reflecting coating

c

Black-coated copper
tube with vacuumed
glass shell (vapor
APV/2
generator)

Fig. 24 Hybridized compound parabolic concentrator (CPC) for concentrated photovoltaic (PV) and vapor generation (cross-sectional cut).

The PV coating is of dielectric type and reflects all light with wavelength longer than 900 nm. This radiation is eventually
absorbed by the vapor generator (mainly) after repeated reflection on the CPC, as it cannot escape back through the aperture. The
vapor accumulator is well insulated thermally as it keeps (all day including overnight) hot vapors at 1201C under pressure. The
Sustainability Dimensions of Energy 135

Table 13 Materials amount correlation and environmental parameters

Material Scaling factor (SF) Specific emission (SE) (kgpollutant/tmaterial)

SECO2 SECH4 SENOx SESO2 SECO SEVOC SEPM

Concrete 1E-5 2.1E þ 1 5.3E þ 0 1.4E-2 2.8E-2 9.2E-3 1.7E-2 6.4E-7


Iron 1.3E-4 5.8E-1 1.5E-1 4.0E-4 8.1E-4 2.7E-4 4.7E-4 1.8E-8
Steel 6.6E-4 3.9E-1 1.0E-1 2.7E-4 5.5E-4 1.8E-4 3.2E-4 1.2E-8
Stainless steel 2.6E-4 1.4E þ 0 3.8E-1 1.0E-3 2.0E-3 6.7E-4 1.2E-3 4.4E-8
Aluminum 4.5E-4 6.1E-2 1.6E-2 4.2E-5 8.8E-5 2.8E-5 5.1E-5 1.9E-9
Copper 1.5E-4 5.4E-1 1.4E-1 3.7E-4 7.5E-4 2.5E-4 4.5E-4 1.7E-8
Fiberglass 3E-5 5.9E þ 0 1.5E þ 0 4.1E-3 8.2E-3 2.7E-3 4.9E-3 1.8E-7

Abbreviations: PM, particulate matter; VOC, volatile organic compound.

pressurized vapor in #3 generates work by passing through an expander. A part of the expanded vapors are extracted at an
intermediate pressure in #5, cooled to 581C as in #7 and condensed in a coil, in #7–8 immersed in a water tank that stores water
at approximately 451C for service (kitchen, bathroom). A part of the expanded vapors is extracted to low pressure in #4, and
cooled at 301C in #6 and then condensed at a condenser temperature of B201C. The condenser rejects the heat in a ground-coil,
buried at a depth of B2 m, such that the temperature remains constant throughout the year.
Assumptions:
The following assumptions are made for this case study:

• For determination of exergy input Ex_ 1


○ total area concentrator exposed to Aa ¼ 100 m2
○ total area of PV arrays APV ¼ 1 m2
○ the total annual irradiance is of Elight ¼ 1.8 MWh/m2
○ the total number of sufficient sunshine is tyear ¼ 4000 h per year
○ the sunlight spectrum can be assimilated to that of a blackbody at Tsun ¼ 6000K
○ the reference temperature is taken as T0 ¼ 298K
• For calculating the photocurrent
○ the transmittance of the PV coating T l is zero for lo200 nm and for l4900 nm
○ the average quantum efficiency of the PV cell is Fe ¼ 0.8
○ the optical efficiency of concentrator is Zopt ¼ 0.8
○ the band gap temperature Tg ¼ 11,000K
○ the diode nonideality factor ni ¼ 1.5
○ the temperature of the PV cell is TPV ¼ 901C
○ the resistance of PV array is Rs ¼ 1E-5 O
• For calculation of the ORC
○ thermal efficiency of solar concentrator Zth ¼ 0.7
○ energy efficiency of the ORC ZORC ¼ 0.25
○ the temperature of heated water T4 ¼ 501C
○ supply temperature of water Tw ¼ 151C
○ amount of hot water produced per day mw ¼ 2000 kg
○ specific heat of water is cp ¼ 4285 J/kgK
• For the economic analysis
○ SExC ¼$23/GJ
○ fraction of CBs from total CC, r ¼0.3
○ the operation and maintenance fraction fo&m ¼ 1E-9
○ the PBP ¼5 years
• For the exergosustainability analysis
○ the system LT ¼ 20 years

Table 13 gives the SFs for the materials used by the system, as well as the specific pollution due to fabrication processes using
various materials.

1.4.7.1.1 Thermodynamic analysis


Here, the thermodynamic analysis aims ultimately to quantify the irreversibilities, that is, to determine the exergy destruction.
Exergy balance equation is applied for the overall system, which is described as shown in Fig. 25. The only input is the exergy from
the sunlight, while the outputs delivered are the power (produced by the ORC engine and PV) plus the exergy associated to the
heating of water. The exergy balance equation for the overall system states that exergy input is equal to the exergy delivered to the
136 Sustainability Dimensions of Energy

PV-power
2 Exergy to
Sunlight exergy input ORC-power 5
3 the user
1 Overall system
Heating
4
Exergy
Reference
destroyed
environment at T0

Fig. 25 Description of exergy balance for the overall energy system. ORC, organic Rankine cycle; PV, photovoltaic.

Table 14 Parameters related to the photovoltaic (PV) array modeling

Parameter Equation
 
T
Saturation current density J0 ¼ ð1:5E 9Þexp  TPVg
 
Dimensionless open circuit voltage J
voc ¼ ni ln 1 þ Jph0
Open circuit voltage Voc ¼(kBTPV)/(e)voc
 
voc lnðvoc 0:72Þ R J A
Filling factor (FF) FF ¼ voc þ1 1  s Vphoc PV
The concentrated light irradiance incident on the PV surface E A R1
IPV ¼ Zoptic t AlightsTr 4 0 T l Il;b dl
year a ð sun Þ

user plus exergy destroyed. The exergy balance is:


_ 1 ¼ Ex
Ex _ 5 þ Ex
_ d ð33Þ

where the exergy delivered to the user is


_ 5¼W
Ex _ 2þW
_ 3 þ Ex
_ 4 ð34Þ
with W_ 2 being the power generated by PV, W _ 3 the power generated by ORC, and Ex _ 4 is the exergy associated to the water heating
process.
_ 1 , is:
Therefore, the exergy rate of the light falling on the heliostat mirrors, input Ex
 
_ 1 ¼ Elight 1  T0 Aa
Ex ð35Þ
tyear Tsun
_ 2 one accounts for the fact that when the concentrated sunlight of partial spectrum falls on the
For the calculation of W
photocathode a photonic current is generated as follows:
e Z 1
Elight Ar
Jph ¼ Zoptic   Fe lT l Il;b dl ð36Þ
hc tyear Aa sTsun4
0

where sTsun4
approximates the irradiance of sun, Fe is the average quantum efficiency, T l is the transmittance of the PV coating,
Zoptic is the optical efficiency of the concentrator, and Il,b is the spectral irradiance of blackbody at temperature Tsun. The power
developed by the PV array is given by
_ 2 ¼ FF V oc Jph
W ð37Þ
Table 14 gives the quantities that must be calculated for the PV array. Note that kB is Boltzmann constant kB ¼ 1.381E-23 J/K
and e is the universal electric charge e¼ 1.602E-19 C. Furthermore, one defined the thermal efficiency of solar
concentrator as being the ratio between heat absorbed Q _ a and light energy on the absorber E_ light;abs . This efficiency is given as
follows:

Q_ abs
Zth ¼ ð38Þ
_Elight;abs

Moreover, the energy efficiency of the ORC is defined as the ratio between the network generated and heat absorbed by vapor
_ abs . Mathematically, one writes:
generator Q
W_3
ZORC ¼ ð39Þ
_ abs
Q
The energy balance for the photonic radiation on the solar concentrator states that the light energy rate passing through the
aperture must be equal to the light energy rate absorbed on the PV array plus the light energy rate on the thermal absorber (that is
Sustainability Dimensions of Energy 137

the vapor generator); this statement is expressed as follows:


Elight
Aa ¼ IPV APV þ E_ light;a ð40Þ
tyear
Based on the thermal efficiency of the solar concentrator the above equation becomes:
_ abs
Q Elight
¼ Aa  IPV APV ð41Þ
Zth tyear
Furthermore, based on the ORC efficiency, the above equation can be manipulated to determine the power generated by the
ORC engine, as follows:
 
W_ 3 ¼ ZORC Zth Elight Aa  IPV APV ð42Þ
tyear
The exergy associated with the water heating delivered to the user is given as follows:
 
_ 4 ¼ mw cp ðT4  Tw Þ 1  T0
Ex ð43Þ
24  3600 T4

From Eqs. (24), (29), and (30), the total exergy delivered to the user becomes:
   
_ 5 ¼ voc  lnðvoc  0:72Þ 1  eRs Jph APV kB TPV voc Jph þ ZORC Zth Elight Aa  IPV APV
Ex
voc þ 1 kB TPV voc e tyear

 
mw cp ðT4  Tw Þ T0
þ 1 ð44Þ
24  3600 T4

1.4.7.1.2 Results
A simple engineering equation solver (EES) code is created calculate the integrals required to determine the values of Jph, voc, and
IPV for the PV array. The code is listed as given in Table 15. The main results with the code are the following: Jph ¼ 7949 A/m2,
voc ¼27.21, and IPV ¼ 21,976 W/m2.
The following results are obtained based on the assumed data for the case study analyses presented above. The light exergy
input becomes
 
_ 1 ¼ Elight 1  T0 Aa ¼ 42;765 W
Ex
tyear Tsun
Total exergy output is calculated as follows:
   
_ 5 ¼ voc  lnðvoc  0:72Þ 1  eRs Jph APV kB TPV voc Jph þ ZORC Zth Elight Aa  IPV APV
Ex
voc þ 1 kB TPV voc e tyear

 
mw cp ðT4  Tw Þ T0
þ 1 ¼ 9506 W
24  3600 T4
The total exergy destroyed results as follows:

Ex _ 1  Ex
_ d ¼ Ex _ 5 ¼ 33;259 W
_ 1 ¼ 42;765 W as given in Table 16. Also given the assumptions r ¼ 0.3, fo&m ¼1E-9,
The pollutant emissions are calculated for Ex
PBP ¼5 years, the total cost CC corresponding to the PBP can be determined with Eq. (26) in which the IC is obtained from

Table 15 Engineering equation solver (EES) code to calculate the photovoltaic (PV) array parameters

$units J K R_s¼1E-5
//Given eta_optic¼0.8
E_light¼ 1.8E6 "Wh/m2" Phi¼0.8
t_year¼4000"h" Tau¼0.9integr¼integral(lambda/1e6*Phi*Tau*Eb(T_sun,lambda),lambda,0.2,0.9)
T_sun¼ 6000 J_ph¼eta_field*e#/h#/c#*A_a/A_PV*E_light/t_year/sigma#/T_sun̂4*integr
A_a¼100"m2" J_0¼(1.5E9)*exp(-T_g/T_PV)
A_PV¼1 v.oc¼n_i*ln(1 þ J_ph/J_0)
T_g¼11000 I_PV¼integral(Phi*Tau*Eb(T_sun,lambda),lambda,0.2,0.9)*E_light/t_year/sigma#/T_sun̂4*A_a/A_PV
n_i¼1.5
T_PV¼90 þ T_zero#
138 Sustainability Dimensions of Energy

Table 16 The results regarding the atmospheric pollutants

Material Amount AM (t) Pollutant emissions, Epollutant (kgpollutant) ¼AM  SEpollutant

ECO2 ECH4 ENOx ESO2 ECO EVOC EPM

Concrete 1.20E-02 2.5E-01 6.4E-02 1.7E-04 3.4E-04 1.1E-04 2.0E-04 7.7E-09


Iron 1.56E-01 9.1E-02 2.3E-02 6.2E-05 1.3E-04 4.2E-05 7.3E-05 2.8E-09
Steel 7.93E-01 3.1E-01 7.9E-02 2.1E-04 4.4E-04 1.4E-04 2.5E-04 9.5E-09
Stainless steel 3.12E-01 4.4E-01 1.2E-01 3.1E-04 6.2E-04 2.1E-04 3.7E-04 1.4E-08
Aluminum 5.41E-01 3.3E-02 8.6E-03 2.3E-05 4.8E-05 1.5E-05 2.8E-05 1.0E-09
Copper 1.80E-01 9.7E-02 2.5E-02 6.7E-05 1.4E-04 4.5E-05 8.1E-05 3.1E-09
Fiberglass 3.60E-02 2.1E-01 5.4E-02 1.5E-04 3.0E-04 9.7E-05 1.8E-04 6.5E-09
Total (kg) 1.4E þ 00 3.7E-01 9.9E-04 2.0E-03 6.6E-04 1.2E-03 4.4E-08

Abbreviations: AM, amount of construction material; PM, particulate matter; VOC, volatile organic compound.

Eqs. (21–24) and the cost of operation and maintenance is obtained from Eq. (25). Thence, for the levelized exergy savings by
using the system becomes:
_ d SExC fo& m PBP 
ð1  r ÞCC þ 3600tyear Ex $
LExS ¼ SExC  109 ¼ 4:36
_ 5
3600PBP tyear Ex GJ

Given system LT ¼20 years, the ExSI is written as


CC
ExSI ¼ 1  ¼ 0:76
_ 5 SExC
3600LT tyear Ex

Therefore the system is sustainable since only 100%  76%¼24% of produced exergy can be used to build the system itself and
to compensate pollution, whereas the remaining 76% is delivered to the user.
In order to encourage the use of renewable energy a policy plan can be established as follows:
1. To encourage research and development by providing focused grants aiming at reducing exergy destruction Ex _ d . This will
lead to an increase of savings on exergy and therefore to better sale of the system. Also, the ExSI will improve. Assume
that this measure aims at 5% reduction of exergy destruction. Then the expected effect of this measure is written as
_ d -0:95Ex
Ex _ d.
2. To improve the system’s economic performance by increasing the CBs (which applies a higher fraction r representing the ratio
between CBs and CC). This will again increase the exergy savings obtained with the system. Assume that one increases the CBs
with 10%; then r-1.1r.
3. To apply higher carbon tax to the coal, petroleum, and natural gas power generators such that the SExC increases in the
jurisdiction. This will lead to an improved ExSI and better savings with the system. Consider the application of carbon tax to
induce SExC in the jurisdiction higher with 5%, then SExC-1.05SExC.

With the considered policy measured, the new values for LExS and ExSI become, respectively:

_ d fo& m PBP 
ð1  1:1r ÞCC þ 3600tyear 0:95Ex $
LExS ¼ 1:05SExC  109 ¼ 6:34
_ 5
3600PBPtyear Ex GJ

CC
ExSI ¼ 1  ¼ 0:78
_ 5 1:05SExC
3600LT tyear Ex

In conclusion, the policy measures are necessary to encourage the system sales since the exergy savings as well as the ExSI
increase.

1.4.7.2 Exergosustainability Assessment of High-Temperature Steam Photoelectrolysis Plant


In this case study, a solar hydrogen production system is considered in which a heliostat field concentrates the sunlight on the
surface of a high-temperature photoelectrochemical cell for steam electrolysis, having a light-exposed photocathode. Liquid water
is pumped atop of the tower and H2 and O2 gases are generated. The solar radiation is concentrated on photoelectrochemical cells;
those are assembled in modules to form arrays.
The cell has a transparent glass with a concave shape to increase the strength as the gas inside is at B20 atm. A mixture of 90%
steam and 10% H2 is fed, and, due to the photoelectrochemical reaction occurring there, a mixture of 90% H2 and 10% steam is
extracted at approximately 10001C. The permeable photocathode has a rough surface (3D, “volumetric” configuration) and it is
doped with cheap metallic electrocatalysts and CuO, Cu2O semiconductors that operate as photosensitizers (excite electrons into
the conduction band when photons are absorbed).
Sustainability Dimensions of Energy 139

Assumptions:

• The following assumptions are made for determination of Ex _ 1:


○ total reflecting area heliostat mirrors Ar=915E3 m2
○ total aperture area Aa=725 m2
○ the total annual irradiance is of Elight=2.0 MWh/m2
○ the total number of sufficient sunshine hours is tyear=4000 h per year
○ the sunlight spectrum can be assimilated to that of a blackbody at Tsun=6000K
○ the reference temperature is taken as T0=298K
● The following data for calculating the photocurrent are used:
○ the heliostats have an aluminum reflective surface. The spectral reflectance of aluminum is approximated as follows
8
< 0:92 for λo700 nm
>
Rλ ¼ 0:88 for λo700 nm and λo900 nm
>
:
0:98 for λ4900 nm
○ the spectral quantum efficiency Fe,λ=0.9 for λ≤580 nm and Fe,λ=0.6 for λ≤1033 nm and Fe,λ=0 for λ41033 nm.
○ the field efficiency is given as ηfield=0.7
● The following data are assumed for calculation of the exergy content of produced hydrogen:
○ the chemical exergy of hydrogen exch H2 ¼ 236 MJ/kmol
● The following data are assumed for economic analysis:
○ the SExC=$23/GJ
○ fraction of CBs from total CC, r=0.3
○ the operation and maintenance fraction fom=1E-9
○ the PBP=20 years
● For the sustainability analysis, the following data are assumed:
○ The system LT=40 years
● The assumed values of the SF and the SEs for the case study are given in Table 17.

1.4.7.2.1 Thermodynamic analysis


Thermodynamic analysis aims to determine the exergy destruction. Exergy balance equation is applied for the overall system is
described as shown in Fig. 26. The inputs are the sunlight and the water; and the only useful output is hydrogen (carrying its
chemical exergy). The exergy balance determines the exergy destroyed, which accounts for any irreversibility plus the release of
oxygen in the atmosphere (as a waste from the process).

Table 17 Materials amount correlation and environmental parameters

Material SF SE (kgpollutant/tmaterial)

SECO2 SECH4 SENOx SESO2 SECO SEVOC SEPM

Concrete 1381 2.1E þ 1 5.3E þ 0 1.4E-2 2.8E-2 9.2E-3 1.7E-2 6.4E-7


Iron 24 5.8E-1 1.5E-1 4.0E-4 8.1E-4 2.7E-4 4.7E-4 1.8E-8
Steel 236 3.9E-1 1.0E-1 2.7E-4 5.5E-4 1.8E-4 3.2E-4 1.2E-8
Stainless steel 118 1.4E þ 0 3.8E-1 1.0E-3 2.0E-3 6.7E-4 1.2E-3 4.4E-8
Aluminum 2 6.1E-2 1.6E-2 4.2E-5 8.8E-5 2.8E-5 5.1E-5 1.9E-9
Copper 5 5.4E-1 1.4E-1 3.7E-4 7.5E-4 2.5E-4 4.5E-4 1.7E-8
Fiberglass 1 5.9E þ 0 1.5E þ 0 4.1E-3 8.2E-3 2.7E-3 4.9E-3 1.8E-7

Abbreviations: PM, particulate matter; SE, specific emission SF, scaling factors; VOC, volatile organic compound.

Sunlight exergy input Hydrogen


1 exergy output
Overall system 3
Water exergy input
2

Exergy
Reference destroyed
environment at T0

Fig. 26 Description of exergy balance for the overall hydrogen production system.
140 Sustainability Dimensions of Energy

The exergy balance equation for the overall system states that exergy input is equal to the exergy output plus exergy destroyed.
The exergy input is:
Ex _ 3 þ Ex
_ in ¼ Ex _ d ð45Þ
_ d is the destroyed exergy rate and Ex
where Ex _ in is the rate of input exergy, which is given as follows:

Ex _ 1 þ Ex
_ in ¼ Ex _ 2 D Ex
_ 1 ð46Þ
_ 2 ) is negligible with respect to the exergy rate of light input: (Ex
In Eq. (46) the exergy rate carried by water (Ex _ 1 ). Therefore the
exergy balance equation becomes:
Ex _ 3 þ Ex
_ 1 ¼ Ex _ d ð47Þ
_ 1 , is:
Furthermore, the exergy rate of the light falling on the heliostat mirrors, input Ex
 
_ 1 ¼ Elight 1  T0 Ar
Ex ð48Þ
tyear Tsun
When the concentrated sunlight falls on the photocathode a photonic current is generated. The photonic current density is
given by:
e Z 1
Elight Ar
Jph ¼ Zfield   lFe;l Rl Il;b dl ð49Þ
hc tyear Aa sTsun4
0

where sTsun4
approximates the irradiance of sun, Fe,l is the spectral quantum efficiency, Rl is the reflectance of heliostat mirrors,
Zfield is the efficiency of the heliostat field, Il,b is the spectral irradiance of blackbody at temperature Tsun.
The molar flow rate of produced hydrogen results from the photonic current:
Jph Aa
n_ H2 ¼ ð50Þ
zF
where z¼2, F¼ 96,486,700 C/kmol (Faraday’s constant).
The exergy rate carried by the generated hydrogen becomes:

_ 3 ¼ n_ H2 exch ¼ Jph Aa exch


Ex ð51Þ
H2 H2
zF
H2 ¼ 118 MJ/kmol is the chemical exergy of hydrogen.
where exch
The total hydrogen produced in kilogram for the PBP results as follows:
Jph Aa
mH2 ¼ 7200tyear PBP ð52Þ
zF
Fig. 27 shows the system representation for sustainability analysis, as based on the exergetic lifecycle assessment. If there is net
exergy output on hydrogen in state 2, then the system may be sustainable because in that case, the system is able to produce sufficient
exergy to construct itself and pay for its emissions. Otherwise, if net exergy in #5 is zero or negative, the system is not sustainable.
As deduced from the figure, the net exergy amount Ex5 can be approximated with:
Ex5 ¼ Ex3  Ex4 ð53Þ
The term Ex4 represents the exergy amount embedded in construction materials plus the exergy required to remove the
pollutants emitted during the construction process from the atmosphere. This exergy can be related to the CC with the help of the
SExC. One has:
CC
Ex4 ¼ ð54Þ
SExC

Source
exergy
input 3 5
Overall system for Exergy output to user
the lifecycle
1
Newly constructed
4 system
System construction process
(exergy is destroyed)

Exergy destroyed
at system
construction
Exergy destroyed Materials
at system operation

Fig. 27 System model for exergosustainability analysis of H2 production system.


Sustainability Dimensions of Energy 141

Therefore the net exergy amount delivered in #5 becomes

_ 3 CC
Ex5 ¼ 3600LT tyear Ex ð55Þ
SExC
The ExSI can be defined in a similar way as suggested previously in Eq. (30). Here, using the current notations, the ExSI is
written as follows:
Ex5 CC
ExSI ¼ ¼1 ð56Þ
Ex3 _ 5 SExC
3600LT tyear Ex

1.4.7.2.2 Results
The calculations of the integral of Eq. (49) for the photocurrent density requires the elaboration of a simple code. The sequence of
code given in Table 18 can be used to calculate the photonic current density. This code is written in EES and required conversion
from microns to m for the wavelength l. Using the code, the following value is obtained, namely, Jph ¼ 75,061 A/m2.
The following results are obtained based on the assumptions for the case study analyses presented above. The light exergy input
becomes
   
_ 1 ¼ Elight 1  T0 Ar ¼ 2E6 1  298 915E3 ¼ 434;777;500 W
Ex
tyear Tsun 4000 6000

Given Jph ¼75,061 A/m2, one calculates the exergy carried by hydrogen as follows:

_ 3 ¼ Jph Aa exch ¼ 75; 061  725  118E6 ¼ 66;552;888 W


Ex
zF H2
2  96; 486; 700
The total exergy destroyed becomes:

Ex _ 1  Ex
_ d ¼ Ex _ 3 ¼ 434; 77; 500  30;521;184 ¼ 368;224;612 W
_ 1 ¼ 434;777;500 W as given in Table 19. The CC calculations are given in Table 20. Denote
The pollutants are calculated for Ex
LH2P as the levelized hydrogen price for selling. The following economic balance states that the revenues from hydrogen selling
during the PBP balance the total cost:
mH2 LH2 P ¼ Ctot

Table 18 Engineering equation solver (EES) code to calculate the photocurrent density

$units K J
A_r¼915E3
A_a¼725
E_light¼2E6
t_year ¼4000
Int_a ¼0.92*0.6*integral(Lbd_3a/1e6*Eb(T_sun,Lbd_3a),Lbd_3a,0.2,0.58)
Int_b ¼0.92*0.6*integral(Lbd_3b/1e6*Eb(T_sun,Lbd_3b),Lbd_3b,0.58,0.7)
Int_c ¼0.88*0.3*integral(Lbd_3c/1e6*Eb(T_sun,Lbd_3c),Lbd_3c,0.7,0.9)
Int_d ¼0.98*0.3*integral(Lbd_3d/1e6*Eb(T_sun,Lbd_3d),Lbd_3d,0.7,0.9)
J_ph¼(Int_a þ Int_b þ Int_c þ Int_d)*eta_field*e#/h#/c#/sigma#/T_sun̂4*A_r/A_a*E_light/t_year

Table 19 The results regarding the atmospheric pollutants for H2 production system

Material Amount AM (t) Pollutant emissions, Epollutant (kgpollutant) ¼AM  SEpollutant

ECO2 ECH4 ENOx ESO2 ECO EVOC EPM

Concrete 210,452,416 4.3E þ 9 1.1E þ 9 3.0E þ 6 6.0E þ 6 1.9E þ 6 3.6E þ 6 1.3E þ 2


Iron 3,657,392 2.1E þ 6 5.5E þ 5 1.5E þ 3 2.9E þ 3 9.7E þ 2 1.7E þ 3 6.6E-2
Steel 35,964,352 1.4E þ 7 3.7E þ 6 9.7E þ 3 2.0E þ 4 6.5E þ 3 1.2E þ 4 4.4E-1
Stainless steel 17,982,176 2.6E þ 7 6.7E þ 6 1.8E þ 4 3.6E þ 4 1.2E þ 4 2.1E þ 4 7.9E-1
Aluminum 304,783 1.9E þ 4 4.9E þ 3 1.3E þ 1 2.7E þ 1 8.6E þ 0 1.5E þ 1 5.7E-4
Copper 761,957 4.1E þ 5 1.1E þ 5 2.8E þ 2 5.8E þ 2 1.9E þ 2 3.4E þ 2 1.3E-2
Fiberglass 152,391 9.0E þ 5 2.3E þ 5 6.2E þ 2 1.3E þ 3 4.2E þ 2 7.4E þ 2 2.7E-2
Total (kg) 4.4E þ 9 1.1E þ 9 3.0E þ 6 6.1E þ 6 2.0E þ 6 3.6E þ 6 1.4E þ 2

Abbreviations: AM, amount of construction material; SE, specific emission; PM, particulate matter; VOC, volatile organic compound.
142 Sustainability Dimensions of Energy

Table 20 Capital cost (CC) calculations for H2 production system

Material CM ($) CP ($) Total ($)

Concrete 6,292,527 21,466,146 27,758,674


Iron 1,774,932 2,512,628 4,287,560
Steel 25,642,583 36,288,031 61,930,614
Stainless steel 19,728,245 27,926,319 47,654,565
Aluminum 1,270,913 1,835,706 3,106,619
Copper 2,066,198 5,998,884 8,065,082
Fiberglass 41,009 130,752 171,760
Total 56,816,407 96,158,467 CC¼152,974,874

Abbreviations: CM, cost of construction material; CP, cost of pollution.

The total cost Ctot corresponding to the PBP can be determined with Eq. (26) in which the IC is obtained from Eqs. (21–24)
and the cost of operation and maintenance is obtained from Eq. (25). Based on the above equations and the assumptions made
and the estimated CC the levelized hydrogen price LH2 P is determined as follows:
_ d SExC fo& m PBP
ð1  r ÞCC þ 3600tyear Ex $
LH2 P ¼ Jph Aa
¼ 1:31
7200 kg
zF tyear PBP

Given the system LT ¼ 40 years, the ExSI becomes:


CC
ExSI ¼ 1  ¼ 0:73
_ 3 SExC
3600LT tyear Ex
In order to encourage the use of hydrogen, a policy plan can be established as follows:
1. Encourage research and development by providing focused grants aiming at reducing exergy destruction, Ex_ d . This will lead to a
decrease of levelized hydrogen price and therefore to better revenue from sales. Also, the ExSI will improve.
Assumption: Encourage research aiming at 5% reduction of exergy destruction, then Ex _ d
_ d -0:95Ex
2. Try to improve the system’s economic performance by increasing the CBs (which applies a higher fraction r representing the
ratio between CBs and CC). This will again increase the exergy savings obtained with the system.
Assumption: Increase the CBs with 10%, then r-1.1r
3. Apply higher carbon tax to the coal, petroleum, and natural gas power generators such that the SExC increases in the jur-
isdiction. This will lead to an improved ExSI and better revenues with the system.
Assumption: Apply carbon tax to induce SExC in the jurisdiction higher with 5%, then SExC-1.05SExC

Due to the considered policy measures, the new values for the levelized hydrogen price for sale LH2P and the ExSI
become:
_ d fo& m PBP
ð1  1:1r ÞCC þ 3600tyear 0:95Ex $
LH2 P ¼ J A
¼ 1:22
7200 phzF a tyear PBP kg

CC
ExSI ¼ 1  ¼ 0:74
_ 3 1:05SExC
3600LT tyear Ex
In conclusion, based on the system assessment, the system is sustainable since only 27% of produced exergy is used to build
itself and compensate pollution. However, for better economic success a system improvement may be beneficial. Once policy
measures are applied by encouraging research through grants, increase in investment CBs, and application of carbon taxation to
power generators, the levelized hydrogen price becomes compatible with the price of hydrogen produced by conventional
methods and the ExSI increases to 0.83.

1.4.7.3 Exergosustainability Assessment of a Heat Pump Dryer


In this case study, an industrial Douglas fir wood chips drying process is considered as the reference system for exergy sustainability
assessment. The reference system is taken from the previous work of Coskun et al. [30]. The reference system uses combustion gases
for drying. An improved system is considered in which the combustor is replaced with a heat pump that recirculates air as drying
agent, whereas the air has similar parameters as the flue gases. It is assumed that the improved system is connected to the Ontario
regional grid.

1.4.7.3.1 System description


The reference industrial drying system uses a directly heated rotary kiln with an average capacity of 93 t/h moist material. The kiln
is supplied with preheated wood chips, which after being dried are separated from the flue gases in cyclones. The system also
Sustainability Dimensions of Energy 143

includes a local gas turbine power plant with cogeneration. The core process, which is the drying process in the rotary kiln, is
described in Fig. 28.
In the reference system configuration, no humid air is recycled; therefore, all material and exergy in state 4 is wasted. The overall
reference drying system configuration is described for a lifecycle exergosustainability assessment in Fig. 29, whereas the state point
parameters are given in Table 21. The following remarks are made about the reference drying system:
● State point parameters 9, 11, 12, 14, and 15 were calculated based on balance equations under the following assumptions; the
calculated values are shown with italic font in the table
○ mixing processes are ideal (no exergy destruction);
○ there is negligible heat loss in the combustor; and
○ the energy efficiency of the gas turbine is 20% (including the generator losses).

● The reversible work rate of 893 kW is assumed for the drying process, as in Ref. [30].
● Because no chemical reactions occur, the chemical exergy of wood is not considered; in addition, the industrial process is
conducted to process wood as construction material and not firewood.
● The balance of plant require auxiliary power of 670 kW.

1 4 State Description T (K) P (kPa)  (g/kg) (kg/kg) m (kg/s) E (kW) Ex (kW)


1 Drying air 739 101.325 72 N/A 81.53 90,385 20,256
2 Moist material 288 101.325 N/A 0.84 14.09 1322 0
2 5
Rotary kiln dryer 3 Power input N/A N/A N/A N/A N/A 765 765
4 Humid air 403 101.325 217 N/A 81.53 88,610 11,964
3 6 5 Dry product 363 101.325 N/A 0.006 14.09 3479 323
6 Destroyed exergy (heat losses) 309 N/A N/A N/A N/A 383 10
7 Destroyed exergy (internal irreversibility) N/A N/A N/A N/A N/A N/A 8724
7

Fig. 28 Wood chips drying process in an industrial rotary kiln dryer.

22

21 Fuel 23 Wastes/
Fuel sources
processing pollution

17

Power
Exergy 16 20 Wastes/
generation
resources pollution
(grid)

18 19
Reference
Materials system
sources
Improved
24 25 system

Materials processing and system


manufacture

26 27

Wastes/ System scrapping and 28 Wastes/


pollution materials recycling pollution

Fig. 29 Model for lifecycle operations for drying system manufacture, scrapping, and fuel production.
144 Sustainability Dimensions of Energy

Table 21 State points description and parameters for the reference drying system

State Description T (K) P (kPa) o (g/kg) ṁ (kg/s) Ė(kW) Ėx (kW)

8 Air leak in 298 101.325 7.2 2.23 691 0


9 Flue gas 749 100 74 79.3 89,694 20,256
10 Flue gas 803 100 60.6 62.4 71,834 7,722
11 Air input 298 101.325 7.2 61.1 2,651 0
12 Fuel (nat. gas) 298 101.325 N/A 1.26 69,183 65,410
13 Flue gas 533 101.325 123.1 16.9 17,890 12,534
14 Fuel (nat. gas) 298 1200 N/A 0.31 17,176 16,239
15 Air input 298 101.325 7.2 16.47 714 0

● The site consumes 2 MW of power for housekeeping and other industrial uses.
● The fuel for gas turbine and combustor is natural gas.

1.4.7.3.2 Results
We now calculate the energy efficiencies and the exergy destroyed by each component and the overall system. The exergy destroyed
by the rotary kiln unit is the sum of the exergy wasted as humid air (stream 4); the exergy of the warm, discharged wood; the exergy
wasted as heat loss (stream 6); and the internal irreversibility (stream 7). The exergy input into the kiln is the sum of exergies for
drying air, moist material, and power. Therefore from Fig. 28 one has:
_ d;kiln ¼ 11;964 þ 323 þ 10 þ 8724 ¼ Ex
Ex _ in;kiln ¼ 20;256 þ 0 þ 765 ¼ 21;021 kW

The exergy efficiency of the process is determined based on given reversible work as follows:
W_ rev 893
ckiln ¼ ¼ ¼ 0:0425
_Exin;kiln 21;021
The exergy destroyed in the combustor can be calculated based on the data from Table 21 and it is:
Ex _ 11 þ Ex
_ d;comb ¼ Ex _ 12  Ex
_ 10 ¼ 57;687 kW

The exergy destroyed by the gas turbine is determined as follows:


Ex _ 14 þ Ex
_ d;gt ¼ Ex _ 15  Ex
_ 13 ¼ 270:2 kW

The total exergy destruction becomes:


Ex _ d;gt þ Ex
_ d ¼ Ex _ d;comb þ Ex
_ d;kiln ¼ 78;979 kW

whereas the total exergy input is


Ex _ 12 þEx
_ in ¼ Ex _ 14 ¼ 81;649 kW

In order to operate the plant consumes 1.57 kg/s natural gas. The annual operation hours are assumed 260 h. The
system LT is 25 years. The total amount of natural gas consumed is 36,738 t or 1.91 PJ of exergy. The overall plant produces two
useful outputs: the reversible work for drying and generated power for other process (2 MW); therefore the exergy efficiency
becomes:
893 þ 2000
c¼ ¼ 0:0354
81; 649
The SI of the process is calculated based on Eqs. (12) and (13) as follows:
1 1
SI ¼ ¼ ¼ 1:03
1c 1  0:0354
The system sustainability is low as indicated by a low value of the SI (a value close to 1). There should be
room for improvement. A more profound analysis must therefore be made, considering the whole lifecycle of the system. We will
start this analysis by an attempt to estimate the sizes of system components and from here determine the amounts of materials
needed.
First, the drying time must be determined. We take the moisture diffusivity for Douglas fir as D¼32.4 E-10 m2/s. Assume the
wood chips have a slab form with an average half width L¼ 0.0002 m. Assume that the equilibrium moisture content is 10% of the
final moisture content (this is a reasonable assumption for practical systems). Therefore, W e ¼ 0:1W 5 ¼ 0:0006 kg/kg, and the
final dimensionless moisture content becomes:
W6  We 0:006  0:0006
ff ¼ ¼ ¼ 0:0064
W2  We 0:84  0:0006
Sustainability Dimensions of Energy 145

Under the assumption that Bim4100 and using a transient drying model for the slab, the following value is obtained for the
drying time:
 
pL2 4
tdry ¼  ln Ff ¼ 45s
4D p

Assume that inside the kiln the particles move at the periphery such that the particles’ velocity makes an angle of p/6 with air
velocity. The dimensions of the kiln can then be found such that the gas flow regime is turbulent. We approximated that if Uair ¼20
m/s, then the required kiln diameter to accommodate the 82 kg/s of air flow rate (as given in the table shown in Fig. 28) is 3.3 m
and Reynolds number is 890,000. Furthermore, the wood velocity component in the axial direction becomes Uwood ¼ 0.36 m/s.
Therefore the length of the dryer is estimated as

Lkiln ¼ Uwood tdry ¼ 17 m

If the kiln is made of stainless steel of 8 mm thickness, then the mass of steel is mssteel ¼ pdkilnLkiln ¼ 10.7 t. A supporting
structure of carbon steel will be required, say 80% lighter; mcsteel ¼ 3.4 t. A concrete platform must be casted below the kiln with an
approximate size of Lkiln  dkiln  0.3m-mconcreteD40 t.
The combustor generates B62 kg/s hot gases at 803K, which is equivalent to 141 m3/s. With a combustion residence time of 5
ms (practical value) and a combustion zone of 0.5 m, the combustor can be approximated as a cylinder with 1.3 m diameter and 5
m height (length), placed vertically; therefore the required metal mass (carbon steel) is 15 t including the supports. The concrete
required for the foundation is 3 t.
The gas turbine is made of stainless steel with aluminum blades and has a connected generator that comprises mainly copper
and iron. The flow rate of air at gas turbine suction is 15 m3/s. The estimated materials are 13 t stainless steel, 2 t carbon steel, 0.5 t
aluminum, 2 t iron, 4 t copper for the generator, and 9 t concrete. The construction materials data is summarized in Table 22 where
the embedded energy, SEs, and EPC are given.
In order to perform the lifecycle assessment, the model presented in Fig. 29 is elaborated. Primary exergy resources are
consumed and destroyed to generate the power required for the essential operations during its LT. A part of the consumed power is
used for fuel processing (stream 17) in which the stream 22 of refined natural gas is produced and distributed to supply the
reference drying system for the LT. Another part of electrical grid exergy is consumed for materials processing and system
manufacture, stream 18; and, another part for system scrapping and materials recycling, stream 19. In all processes, wastes and
pollution are emitted to the environment.
Assume that 10% of scrapped materials are recycled. Therefore the embedded energy for materials processing and manufacture
can be approximated with 90% of 2689 GJ given in Table 22; this is 2420 GJ. The SEs will be of 106.7 t GHG with an EPC of
$88,194. The exergy consumed to refine and distribute the natural gas represents a small fraction of the fuel exergy. Based on data
from lifecycle assessment presented in Dincer et al. [10] for the system lifecycle, the exergy used to extract, refine, and distribute the
fuel is estimated to 9950 GJ as given in Table 23.
The total exergy consumed for fuel processing, materials extraction, manufacturing, and system scrapping becomes
9950 þ 2420 þ 135¼ 12,505 GJ. Based on the averaged exergy efficiency of the Canadian grid, the consumed exergy resource is
12; 505/0.51 ¼ 24,520 GJ (stream 16 in Table 23). Based on calculations shown in Table 8, the EPC for the Canadian grid is 17.8
$/MWh or 4.9 $/GJ; thence, the EPC associated to power production for materials, manufacturing, fuel processing, and system
scrapping is 4.9  24,520 ¼ $120,146.
The amount of dried wood chips is 14.9 kg/s  (25  260  3600)s ¼ 3.487E5 t, which are to be valorized on the market. The
total expenditure in the fuel is 1.91E6 GJ t  (3.889  52/50) $/GJ ¼ $3.68E6, where the fraction 52/50 is the ratio of chemical
exergy to the LHV of natural gas. The total exergy destruction during system operation is 78.979 MW  (25  260) h ¼513.3
GWh ¼1.85E6 GJ. The total EPC for the LT is, according to data given in Tables 23, $15.825E6 of which 2% (totaling $325,672)
represents the pollution costs associated with materials extraction, manufacturing, system scrapping, and recycling, fuel processing,
and distribution; note that the exergy destruction associated with this EPC is 18.3 GWh¼ 0.07E6 GJ.
Therefore the total LT exergy destruction of the reference drying system is Exd,ref ¼ 1.92E6 GJ. The total exergy input to sustain
the LT operation and the system construction and scrapping is from Table 23: Exlc,in ¼ 24,250 þ 1.91E6 ¼ 1.934E6 GJ. The total

Table 22 Pollution parameters of construction materials for drying system

Material m (t) EE (GJ) SE (kg GHG) EPCex ($2013)

Concrete 52 73 1,752 5,146


Iron 2 47 517 1,377
Carbon steel 20 688 7,568 20,158
Stainless steel 23.7 1256 77,827 36,800
Aluminum 0.5 101 1,010 3,020
Copper 4 524 29,868 31,492
Total 2689 118,542 97,993

Abbreviations: EE, embodied energy; EPC, environmental pollution cost; GHG, greenhouse gas; SE, specific emission.
146 Sustainability Dimensions of Energy

Table 23 Embedded exergy and environmental pollution costs (EPCs) for reference drying system lifecycle

Stream Description Ex (GJ) EPC ($)

16 Exergy extracted from primary sources for power generation 24,520 N/A
17 Exergy consumed for fuel processing 9550 N/A
18 Exergy consumed for materials processing and system manufacture 2420 N/A
19 Exergy consumed for system scrapping and materials recycling 135 N/A
20 Environmental waste stream for grid power generation for lifetime (LT) N/A 120,146
21 Primary fuel sources extracted (natural gas) 1.91E6 N/A
22 Refined natural gas distributed to consumption point 1.91E6 15.5E6
23 Waste emissions of pollutants due to fuel processing N/A 77,332
24 Materials extracted (ores) for system manufacturing N/A N/A
25 Embedded energy in the constructed system 2420 N/A
26 Polluting wastes due materials processing and system manufacture N/A 88,194
27 Recycled materials 134 N/A
28 Wastes and pollution due to system scrapping and recycling N/A 40,000

reversible work is Wlc,rev ¼ 873 kW  (25  260) h¼ 20,428 GJ. The sustainability of the reference drying system can be assessed
based on the following parameters:
● Thermodynamic parameters – i.e.,
○ Total lifecycle exergy destruction, Exd,ref ¼ 1.92E6 GJ.
○ Lifecycle exergy efficiency is clc ¼ Wlc,rev/Exlc,in ¼ 0.020428/1.934 ¼ 0.01.
○ Specific reversible work (SRW) ¼ Wlc,rev/Exd,ref ¼ 0.020428/1.92 ¼ 0.011.

● Exergoeconomic parameter – i.e.,


○ Specific exergetic capital investment (ExCI) defined as the ratio of exergy invested in system construction
Exinv ¼9950 þ 2420 þ 135 ¼12,505 GJ and the amount of reversible work needed to dry the product ExCI ¼Exinv/Wlc,rev ¼
12,505/20,428 ¼ 0.612.
○ Exergetic investment efficiency (ExIE) defined here as the ratio between exergy expenditure for investment and exergy
destroyed, ExIE ¼ Exinv/Exd,ref ¼ 12,505/1.92E6 ¼ 0.0065.
○ Capital investment effectiveness defined here by the ratio between exergy input into the drying system for operation during
the LT and the exergy consumed for materials extraction and system manufacturing, capital internment effectiveness
(CIEx) ¼1.91E6/36,490 ¼ 52.2.

● EI parameter – i.e.,
○ Lifecycle EPClc ¼ $15.825E6.
○ EPC for system construction and scrapping EPCcs ¼ $128,194.
○ Construction exergy expenditure (ExCDR) to lifecycle exergy destruction ratio ExCDR¼ 0.0081.

The sustainability can be assessed based on an ASI. Here, we formulate the ASI based on exergy destruction using
thermodynamic, economic, and environmental assessment parameters. In this respect, one notes that the social benefit of the
system is proportional to the sum of the reversible work, the exergy investment in the system construction, and the exergy
expenditure equivalent to EPC for system construction and scrapping. The benefit of the investment in the reversible work is
in fact that the dry product has market and societal value. The benefit exergy expenditure for system construction is in fact the
drying system as a good, capable for production. The benefit on exergy investment to compensate for the environmental
pollution at system construction and scrapping represents a general societal good consisting of a better environment. The sum
of these three terms must be compared with the lifecycle exergy destruction in order to form the ASI. Therefore the equation
for ASI becomes:

ASI ¼ SRW þ ExIE þ ExCDR ð57Þ

where SRW is the specific reversible work, ExIE is the exergetic investment efficiency, ExCDR construction exergy expenditure
to lifecycle exergy destruction ratio.
For the reference system, the ASI becomes

ASIref ¼ 0:011 þ 0:0065 þ 0:0081 ¼ 0:0256

This aggregate index must be compared to that for the improved system. The reference system can be improvement with respect
to sustainability provided that more sustainable energy sources are used as input and in addition the system irreversibilities are
reduced by providing a better design. In the reference system, the energy source is from natural gas. If instead of combusting
natural gas, a heat pump is used, supplied from the Ontario regional grid, then the associated EI will be substantially reduced
because in Ontario the grid power is relatively clean, as in the energy mix much hydro and nuclear do exist. Also, when a heat
Sustainability Dimensions of Energy 147

pump is used, no drying agent is expelled to the environment; therefore, heat is regenerated internally and less energy supply is
required. This explains why a heat pump dryer will have a higher exergy efficiency than that of the reference dryer.
We assume that the drying process as described in Fig. 28 remains the same. Therefore, the same reversible work is required,
W_ rev ¼ 839 kW, for the same production rate of 14.09 kg/s dry product. The exergy input required to drive the process must be the
same as for the reference case, namely, Ex_ in;kiln ¼ 21; 021 kW.
The heat pump concept is relatively simple and it allows for moisture extraction from humid air prior to heating. Fig. 30 shows
the heat pump system for drying air recirculation and moisture removal. In state 4, a flow rate of 81.53 kg/s of humid air with
403K and 217 g/kg humidity ratio enters the heat pump system and it is split in two fractions, #8 and #13. Once expanded to a
vacuum pressure, water condensates both in states 9 and 14. However, water from state 9 is gravitationally separated whereas the
moisture from state 14 continues to flow together with the air stream. Since water is separated under vacuum, a pump is used to
pressurize it to atmospheric pressure in 11.
A transcritical carbon dioxide heat pump heats the drying air from state 15 to state 16 where it is recompressed to atmospheric
pressure and delivered to the required parameters in state 1: temperature 739K and humidity ratio of 7.2 g/kg. The carbon dioxide
heat pump uses the surrounding medium (e.g., a lake) to draw heat at reference temperature T0 ¼ 298K and evaporate the working
fluid from state 17 (low vapor quality two-phase mixture) to state 18 (saturated vapor). The vapor is superheated to state 19 using
internal heat regeneration, and further compressed so that it is able to deliver heat by heat transfer to the drying air in processes
20–21. Typical mass, energy, and exergy balances are written for each component from Fig. 30 and solved to determine the state
points. The heat pump system state points descriptions and thermodynamic parameters are given in Table 24. The net required
work input for the heat pump system can be calculated as follows:

_ net;in ¼ E_ 1  E_ 16 þ E_ 11  E_ 10 þ E_ 20  E_ 18 þ E_ 9  E_ 8 þ E_ 14  E_ 13 ¼ 12;866 kW
W

The heat input for the heat pump is

_ in ¼ E_ 18  E_ 17 ¼ 30;763 kW
Q

The exergy of the heat input is zero because the system boundary is set at T0. The heat output neglecting heat losses is

_ out ¼ Q
Q _ in þ Wnet;in ¼ 43;629 kW

The coefficient of performance (COP) of the heat pump becomes:

Q_ out 43;629
COP ¼ ¼ ¼ 3:39
_
W net;in 12;866

13 Drying air
output 1
4
Humid
14
air input
8
16
15
12 20

9
Transcritical
carbon
dioxide
heat pump
19
21
10
11

Legend
Drying air 22
Removed water 18
Carbon dioxide 17
Work
Heat

Fig. 30 Heat pump for improved wood chips drying system.


148 Sustainability Dimensions of Energy

Table 24 State descriptions and thermodynamic parameters for the heat pump system

State Description T (K) P (kPa) o (g/kg) x (kg/kg) E_ (kW)

1 Drying air (output) 739 101.325 72 N/A 59,359


4 Humid air (input) 403 101.325 217 N/A 59,231
8 Humid air 403 101.325 217 N/A 39,578
9 Saturated humid air 321.7 44.32 217 N/A 33,296
10 Liquid water (vacuum) 321.7 44.32 N/A 1 30,635
11 Liquid water 321.7 101.325 N/A 1 30,636
12 Dry air 321.7 44.32 0 N/A 2,661
13 Humid air 403 101.325 217 N/A 19.653
14 Saturated humid air 321.7 44.32 217 N/A 16,533
15 Cold drying air 321.7 44.32 72 N/A 19,194
16 Preheated drying air 594.2 44.32 72 N/A 45,014
17 Two-phase vapor-liquid CO2 283 4485 N/A 0.019  25,628
18 Saturated CO2 vapor 283 4485 N/A 1  7,731
19 Superheated CO2 vapor 526.3 4485 N/A N/A 18,729
20 Hot supercritical CO2 614.2 10,133 N/A N/A 26,651
21 Cooled supercritical CO2 381.7 10,133 N/A N/A 831.3
22 Subcooled CO2 liquid 297.9 10,133 N/A N/A  25,628

Because the heat is delivered at temperature T1 ¼ 739K, the exergetic COP becomes:
 
_ out 1  T0  
Q T1 43;629 1  298
COP ex ¼ ¼ 739
¼ 2:0
W _ net;in 12;866
The exergy balance equation allows for determination of the exergy destruction, which is
 
_ 1þW
Ex _ out 1  T0 þ Ex
_ net;in ¼ Q _ d;hp ¼ 3707 kW
_ d;hp -Ex
T1
The total exergy destruction of the drying process coupled to the heat pump is equal to the sum of exergy destruction by the kiln
and the exergy destruction by the heat pump:
Ex _ d;hp þ Ex
_ d ¼ Ex _ d;kiln ¼ 3707 þ 21;021 ¼ 24;728 kW

The total power required for the drying process is equal to the power consumed by the blowers and kiln rotation system plus
the power consumed by the heat pump:
_ act ¼ 765 þ 12;866 ¼ 13;631 kW
W
The exergy efficiency of the dryer becomes:
_ rev
W 839
c¼ ¼ ¼ 0:061
_
W act 13;631
The efficiency is improved 1.7 times with respect to the exergy efficiency of the reference system. Furthermore, the SI of the
improved system becomes:
1 1
SI ¼ ¼ ¼ 1:065
1c 1  0:061
Taking into account the additional 2 MW required onsite, the total power consumed by the grid by the improved system is
W_ in ¼ 15;866 kW. For a lifecycle of 25 years and 260 h of annual operation the total electric energy consumed from the grid
becomes Win ¼ 371,265 GJ. To this energy, the amount of energy required for system construction and scrapping must be added.
The improved system no longer has the combustor and the gas turbine. However, it is fair to assume for a rough estimation that
the amount of materials, such as copper, aluminum, and stainless steel, used for the gas turbine, combustor, and the electric power
generator are now used to construct the heat pump, which itself includes similar components as for the reference system: electrical
motors, compressors, and turbines.
Therefore Table 22 remains unchanged for the improved system. However, in Table 24 the streams 17, 21, 22, and 23 do not
exist because only electrical power is demanded by the improved system. The lifecycle operations of the improved system are
described as shown in Fig. 29. Table 25 gives the embedded exergy and EPCs for improved drying system.
The sustainability assessment parameters are calculated for the improved system in a similar manner as for the reference system.
The sustainability for the two systems are compared as shown by the results given in Table 26. The total exergy destruction for the
LT is given by the sum of the exergy destruction by the system itself and the exergy destruction at power generation, which is
373,828(1  0.51) þ 24,728¼ 207,904 GJ, where 0.51 is the exergy efficiency of the grid. The exergy input for the lifecycle is given
Sustainability Dimensions of Energy 149

Table 25 Embedded exergy and environmental pollution costs (EPCs) for the improved drying system lifecycle

Stream Description Ex (GJ) EPC ($)

16 Exergy extracted from primary sources for power generation 373,820 N/A
17 Electric energy consumed by the system for operation (for lifetime (LF)) 371,265 N/A
18 Exergy consumed for materials processing and system manufacture 2420 N/A
19 Exergy consumed for system scrapping and materials recycling 135 N/A
20 Environmental waste stream for grid power generation for LF N/A 1.848E6
24 Materials extracted (ores) for system manufacturing N/A N/A
25 Embedded energy in the system constructed and delivered 2420 N/A
26 Polluting wastes due to materials processing and system manufacture N/A 88,194
27 Recycled materials 134 N/A
28 Wastes and pollution due to system scrapping and recycling N/A 40,000

Table 26 Sustainability comparison of reference and improved dryer system

Parameter type Parameter Drying system

Reference Improved

Thermodynamic Exd;ref 1.92E6 GJ 0.21E6


clc 0.01 0.055
SRW 0.011 0.098
Exergoeconomic ExCI 0.612 0.612
ExIE 0.0065 0.060
CIEx 52.2 145
Environmental EPClt $15.825E6 $1.98E6
EPCcs $128,194 $128,194
ExCDR 0.0081 0.065
Sustainability ASI 0.0256 0.223
GFEx;d 0 0.89
GFASI 0

Abbreviations: ASI, aggregated sustainability indicator; CIEx, capital internment effectiveness; EPC, environmental
pollution cost; ExCDR, construction exergy expenditure; ExCI, specific exergetic capital investment; ExIE,
exergetic investment efficiency; GF, greenization factor; SRW, specific reversible work.

in Table 25, Exlc,in ¼ 373,820 GJ. The total reversible work amount is previously calculated as Wlc,rev ¼ 20,428 GJ. Thence, the
lifecycle exergy efficiency becomes clc ¼ 0.055 showing 5 times improvement with respect to the reference case.
The SRW becomes SRW ¼ Wlc,rev/Exd,ref ¼ 20,428/207,904 ¼0.098. The specific exergetic capital investment remains unchanged
because the investment and the reversible work are not affected by the system change, and the same investment has been assumed
for both systems. Also, the EPC for system construction does not change. However, the ExIE changes to ExIE ¼Exinv/
Exd,ref ¼12,505/207,904 ¼ 0.060. Finally, the ASI for the improved system is determined by Eq. (47) and becomes:
ASI ¼ 0:098 þ 0:060 þ 0:065 ¼ 0:223
Further comparison of the improved and reference system is done using the GF. Using Eq. (15) the exergy destruction based GF
becomes:
Exd;ref  Exd 1:92  0:21
GF ex ¼ ¼ ¼ 0:89
Exd;ref 1:92
If the ASI is used for the GF, then from Eq. (10) one obtains:
ASI  ASIref 0:223  0:0256
GF ASI ¼ ¼ ¼ 0:88
ASI 0:223
In this case study, the exergosustainability assessment of a drying system is demonstrated. Although the analysis is approxi-
mated, it clearly demonstrates the main steps to follow in order to assess the system. The scope of the analysis is extended to the
entire LT, which considers three phases: materials extraction and system construction (including fuels processing), system
operation, and system scrapping and materials recycling. For all phases, the embedded exergy, the EPC, and exergoeconomic costs
can be determined such that eventually the overall sustainability can be assessed by an aggregated index. Furthermore, if the
system is improved or another system is comparatively assessed for sustainability, the GF can be used, which is a simple method to
quantify the system improvement toward greenization. The example shows that application of heat pumps, when clean grid power
is locally available, is beneficial.
150 Sustainability Dimensions of Energy

1.4.8 Future Directions

New ways of thinking at energy sustainability emerged in recent years. Bejan [31] brings to attention that sustainability requires to
sustain the flow of useful energy (or exergy) throughout the inhabited by humans. The problem has wide implications because this
global flow of exergy must be controlled wisely by human such that it maintains propel all basic and more elaborated activity. For
example, a proper distribution of water – a basic need – in arid areas implies a proper distribution of exergy resources and power
generation. Distribution of knowledge requires information transfer through communication network, which are also related to
the way in which exergy and power generation is distributed over the globe.
Sustainability of energy/exergy supply is such an important. The application of Constructal law – see Bejan [31] – in sus-
tainability shows real promises toward discovering better designs (configurations) in power generation and use, transportation,
technology, knowledge, wealth, and government. For instance in Bejan [31] [Figure 1.2] the energy use is shown in a linear
dependence in a double-logarithmic plot. This demonstrates that the world speeds up toward more energy consumption, anot less.
There is struggle in this, because as time passes more energy resources are involved in humans sustainability, and these resources
must be wisely administrated. The answer from the Constructal law view point is that hierarchy is emergent in generating,
distributing, and consuming power globally. The new aspect consists of realizing that sustainability of energy sources is rooted in
sustainability of humans species. Sustainability emerges though better designs, while the concepts of design, better and sus-
tainability are revealed as belonging to physics.
Novel development in sustainability of energy systems led to consideration of “smart energy systems,” a term coined in Dincer
and Acar [32]. An energy system that uses technologies and resources that are adequate, affordable, clean, and reliable is denoted
as “smart energy system.” These systems can be developed and assessed for efficiency, environmental performance, energy, and
material resources. Generation of multiple and diverse products is possible with smart energy systems supplied from a single
energy sources. The multigeneration is shown to be beneficial for decreasing the environmental impact per unit of product and
increasing the efficiency, thus enhancing the sustainability.
It is important to portray the “energy to sustainability” for future in a mathematical form as “3S þ 2S ¼ Sustainability” as
illustrated in Fig. 31. The figure shows that the energy picture covers 3S approach, including source-system-service, meaning that a
source is always needed to provide energy input to a system which will produce services (useful commodities, such as electricity,
heat, cooling, hot water, drying air, hydrogen, ammonia, fuels, and fresh water). This, of course, makes the 3S, and 2S will be
added to this, as illustrated in Fig. 31, since there is a need for energy storage between the source and system, depending on the
source and system, especially for renewables (such as solar energy: which is available during daytime and not available at night. So,
the storage becomes an essential need), and between system and source due to fact that the production becomes more than the
demand at most of the times. So, there is a need to do storage to offset the mismatch between demand and supply. This overall
makes the equation 3S þ 2S what really brings sustainability. So, it is extremely important to dwell on these concepts and apply
correctly as well as address correctly.

1.4.9 Closing Remarks

In this chapter, sustainability aspects of energy are reviewed with a focus on development of assessment tools and exergy methods
for sustainability. The role of sustainability indicators on sustainability assessment and policy making is described. There is a large
interdependence of factors affecting sustainability of energy systems and technologies, such as environmental, economic, social
factors. The linkage of sustainability with energy conservation is also presented. The DPSIR model is very relevant, as shown, for
elaborating sustainability indicators focused on varied aspects.
Normalization of sustainability indicators can be made based on reference indicators, such as the maximum or average
values or based on the alleviation from standard average of indicator variance. Once normalized indicators are determined, those
can be aggregated to form a single SI used to rank and assess the sustainability of any specific energy technology. There are various

Source System Service

Storage Storage

Sustainability
Fig. 31 Dimensions of sustainability under 3S (Source–System–Service) with 2S (Storage) options.
Sustainability Dimensions of Energy 151

ways to consider weighting factors to aggregate the normalized indicators. The weighting factors can be defined according to
trade-offs between different criteria, each criteria being represented by a normalized indicator. The trade-offs are generally
established by the decision maker and therefore the vision of decision maker influences the process. Three types of decision maker
archetypes can be envisioned, namely, the individualist, the egalitarian, and the hierarchist. An ASI for energy systems will then
account for the relative improvement toward greenization with respect to a baseline case. Therefore a GF has been proposed in the
literature.
Exergosustainability assessment of energy systems considers the entire lifespan of the system and investigates the flows of exergy.
It is then possible to define an ExSI that accounts for system construction and for alleviation of any environmental pollution. SFs and
pollution cost become important in the analysis. Three exergosustainability case studies are presented to illustrate the method.

References

[1] IEA. International Energy Agency technical report. Key world energy statistics. Available from: http://www.iea.org/publications/freepublications/publication/KeyWorld2014.pdf;
2014.
[2] Bareto L, Makihira A, Riahi K. The hydrogen economy in the 21st century: a sustainable development scenario. Int J Hydrogen Energy 2009;28:267–84.
[3] Ness B, Urbel-Piirsalu E, Anderberg S, Olsson L. Categorising tools for sustainability assessment. Ecol Econ 2007;60:498–508.
[4] Kanoglu M, Dincer I, Cengel YA. Exergy for better environment and sustainability. Environ Dev Sustain 2009;11:971–88.
[5] Midilli A, Dincer I, Ay M. Green energy strategies for sustainable development. Energy Policy 2006;34:3623–33.
[6] Hardi P, Zdan TJ. Assessing sustainable development: principles in practice. Winnipeg, MB: The International Institute for Sustainable Development; 1997.
[7] Dincer I, Zamfirescu C. Sustainable energy systems and applications. New York, NY: Springer; 2011.
[8] Linke B, Das J, Lam M, Ly C. Sustainability indicators for finishing operations based on process performance and part quality. Procedia CIRP 2014;14:
564–9.
[9] Singh RK, Murty HR, Gupta SK, Dikshit AK. An overview of sustainability assessment methodologies. Ecol Indic 2009;9:189–212.
[10] Dincer I, Rosen MA, Zamfirescu C. Economic and environmental comparison of conventional and alternative vehicle options. In: Pistoia G, editor. Electric and hybrid
vehicles. New York, NY: Elsevier; 2010.
[11] Hacatoglu K, Dincer I, Rosen MA. A new model to assess the environmental impact and sustainability of energy systems. J. Clearer Prod 2015;103:211–8.
[12] Dincer I, Zamfirescu C. Potential options to greenize energy systems. Energy 2012;46:5–15.
[13] Dincer I, Zamfirescu C. Sustainable hydrogen production. New York, NY: Elsevier; 2016.
[14] Wall G. Exergy conversion in Japanese society. Energy 1990;15:435–44.
[15] Wall G.. Exergy use in the Swedish society 1994. In: TAIES’97 Thermodynamic analysis and improvement of energy systems conference, Beijing, China; 1997.
[16] Utlu Z, Hepbasli A. A review and assessment of the energy utilization efficiency in the Turkish industrial sector using energy and exergy analysis method. Renew
Sustainable Energy Rev 2007;11(7):1438–59.
[17] Dincer I, Rosen MA. Exergy: energy, environment and sustainable development. Oxford: Elsevier; 2013.
[18] Rosen MA, Dincer I. Exergy as the confluence of energy, environment and sustainable development. Exergy, Int J 2001;1:3–13.
[19] Bejan M. Energy policy, in entropy generation through heat and fluid flow. New York, NY: Wiley; 1994.
[20] Connelly L, Koshland CP. Two aspects of consumption: using an exergy-based measure of degradation to advance the theory and implementation of industrial ecology.
Resour Conserv Recycl 1997;19:199–217.
[21] Delwuf J, Van Lagenhove H. Integtaing industrial ecology principles into a set of environmental sustainability indicators for technology assessment. Resour Conserv Recycl
2005;43:419–32.
[22] Dincer I. Exergy as a tool for sustainable drying systems. Sustain Cities Soc 2011;1:91–6.
[23] Dincer I, Zamfirescu C. Advanced power generation systems. New York, NY: Elsevier; 2014.
[24] Carpenter S. The environmental cost of energy in Canada. In: Sustainable energy choices for the 90’s. Proceedings of the 16th annual conference of the solar energy
society of Canada, Halifax, NS; 1990. p. 337–342.
[25] De Gouw JA, Warneke C, Stohl A, et al. Volatile organic compounds composition of merged and aged forest fire plumes from Alaska and western Canada. J Geophys Res
2006;111:D10303.
[26] Pellizzari ED, Clayton CA, Rodes CE, et al. Particulate matter and manganese exposures in Toronto, Canada. Atmos Environ 1999;33:721–34.
[27] Rosen MA, Dincer I. Exergy analysis of waste emissions. Int J Energy Res 1999;23:1153–63.
[28] Dincer I, Acar C. A review of clean energy solutions for better sustainability. Int J Energy Res 2015;39:585–606.
[29] Dincer I, Acar C. Review and evaluation of hydrogen production methods for better sustainability. Int J Hydrog Energy 2015;40:11094–111.
[30] Coskun C, Bayraktar M, Oktay Z, Dincer I. Energy and exergy analyses of an industrial wood chips drying process. Int J Low-Carbon Technol 2009;4:224–9.
[31] Bejan A. The physics of life: Evolution of everything. St. Martin’s Press; 2016.
[32] Dincer I, Acar C. Smart energy systems for better sustainability. Applied Energy 2017;194:225–35.

Further Reading
Dincer I, Zamfirescu C. 2016. Drying phenomena. Chichester: Wiley; 2016.
Georgescu-Roegen N. 1986. The entropy law and the economic process. East Econ J 1986;12:3–25.

Relevent Websites

https://energy.gov/management/spo/sustainability-performance-office
Office of Management.
http://ontario-sea.org/?gclid=CjwKEAjwr_rIBRDJzq-Z-LC_2HgSJADoL57HsFAKGqK7VMZrkEiwccY4q_97DkE8DP5Q4ChtcsTnAxoC9Yvw_wcB
Ontario Sustainable Energy Association.
152 Sustainability Dimensions of Energy

https://www.siemens.com/global/en/home/company/topic-areas/sustainable-energy.html
Siemens.
https://www.asme.org/engineering-topics/articles/sustainability/the-ammonia-economy
The American Society of Mechanical Engineers.
http://www.exergoecology.com/
The Exergoecology Portal.
http://cbey.yale.edu/programs-research/defining-sustainability-indicators-and-metrics
Yale Center for Business and the Environment.

You might also like