Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Exergy Int. J.

1(2) (2001) 6884


www.exergyonline.com

Beyond thermoeconomics? The concept of Extended


Exergy Accounting and its application to the analysis
and design of thermal systems
Enrico Sciubba
Department of Mechanical and Aeronautical Engineering, University of Roma 1, Italy

(Accepted 26 January 2000)

Abstract This paper presents a novel approach to the evaluation of energy conversion processes and systems, based on an extended
representation of their exergy ow diagram. This approach is a systematic attempt to integrate into a unied coherent formalism
both Cumulative Exergy Consumption and Thermo-economic methods, and constitutes a generalisation of both, in that its framework
allows for a direct quantitative comparison of non-energetic quantities like labour and environmental impact (hence the apposition
Extended). A critical examination of the existing state-of-the-art of energy- and exergy analysis methods and paradigms indicates
that an extension of the existing Design and Optimisation procedures to include explicitly non-energetic externalities is indeed
feasible. It appears that it can indeed be successfully argued that some of the issues that are dicult to address with a purely
monetary theory of value can be resolved by Extended Exergy Accounting (EEA in the following) methods without introducing
arbitrary assumptions external to the theory. In this respect, EEA can be regarded as a natural development of the economic theory of
production of commodities, which it extends by properly accounting for the unavoidable energy dissipation in the productive chain.
While a systematic description of the EEA theory is discussed in previous work by the same author, the present paper aims at a more
specic target, and presents a formal representation of the application of EEA to a cogenerating plant based on a gas turbine process.
It is shown that the solution indeed leads to an optimal design, and that its formalism embeds even extended Thermo-economic
formulations. 2001 ditions scientiques et mdicales Elsevier SAS

Nomenclature
ai,j
c
cex
e
E
h
I
Kex
n
O
PF
s
T

elements of the structural matrix A


relative concentration . . . . . . .
specific exergetic cost . . . . . . .
physical exergy of stream . . . . .
total exergy flow . . . . . . . . . .
total enthalpy . . . . . . . . . . . .
input (exergy) flow . . . . . . . . .
exergetic cost factor . . . . . . . .
number of individuals
output (exergy) flow . . . . . . . .
plant factor = hours per year/8760
specific entropy . . . . . . . . . . .
temperature . . . . . . . . . . . . .

.
.
.
.
.
.
.

.
.
.
.
.
.
.

.
.
.
.
.
.
.

%
JJ1
Jkg1
W
Jkg1
W
$MJ1

. . .

. . .
. . .

Jkg1 K1
K

Greek symbols

environmental penalty factor


exergetic conversion efficiency

E-mail address: enrico.sciubba@uniroma1.it (E. Sciubba).

68

chemical potential . . . . . . . . . . . .

Jkg1

1. INTRODUCTION
Already fifty years ago, energy conversion systems
were the target of a detailed analysis based on Second
Law concepts [13]. That analysis showed that the relevant design procedures of the time neglected to recognise that the irreversibility in processes and components
depends on the energy degradation rate and not only
on the ratio between the intensities of the output and
input flows, and that there is a scale of energy quality
that can be quantified by an entropy analysis. In essence,
the legacy of this approach, universally accepted today,
is that the idea of conversion efficiency based solely
on First Law considerations is erroneous and misleading. This method evolved throughout the years into the
so-called availability analysis [4, 5], later properly renamed exergy analysis [6, 7], and it has had a very pro 2001 ditions scientiques et mdicales Elsevier SAS. All rights reserved
S1164-0235(01)00012-7/FLA

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

found impact on the energy conversion system community, to a point that it is difficult today to find a design
standard which does not make direct or indirect use of exergetic concepts in its search for an optimal configuration [8]. The same method has been extended to complex
systems, like an industrial settlement [9], a complete industrial sector [10, 11], and even entire nations [12, 13],
and most recently has been brought to the attention of
energy agencies [14, 15] as a proposed legislative tool
for energy planning and policymaking. Exergy analysis,
though completely satisfactory from a thermodynamic
point of view, has always been regarded as unable to determine real design optima, and therefore its use has been
associated with customary monetary cost-analysis [8, 16,
17]; it was only recently that a complete and theoretically
sound tool, based on a combination of mixed economicand thermodynamic methods and properly named thermoeconomics, was developed to industry standards [7,
8, 11, 1821]. In this approach, efficiencies are calculated
via an exergy analysis, and non-energetic expenditures
(financial, labour and environmental remediation costs)
are explicitly related to the technical- and thermodynamic
parameters of the process under consideration: the optimisation consists of determining the design point and the
operative schedule that minimise the overall (monetary)
cost, under a proper set of financial, normative, environmental and technical constraints.
In spite of a long tradition of contrary opinion, exergy
seems indeed to possess an intrinsic, very strong and direct correlation with economic values: one of the goals
of the extended exergy accounting method (EEA in the
following) is to exploit this correlation to develop a formally complete theory of value based indifferently on an
exergetic- or on a monetary metric (that is, a general valuing or pricing method in which kJkg1 or kJkW1 are
consistently equivalent to $kg1 and $kW1 , respectively): this issue is discussed in detail elsewhere [22],
and it is based on the fundamental idea that, while exergetic and monetary costs may have the same morphology (they represent the amount of resources that must be
consumed to produce a certain output), their topology
(structure) may be different, leading to the possibility of
different optimal design points. It is remarkable that another topic of paramount importance in the engineering
field can be successfully tackled by EEA methods: this
is the environmental issue, taken in its extended meaning of impact of anthropic activities on the pre-existing
environment. A critical analysis of the leading engineering approaches to the environmental issue is reported in
[23]: in that work it is shown that EEA, which proposes
a different quantifier (the extended exergy content) for
the analysis of processes and plants, can be regarded as a

successful combination of the methods put forth by previous researchers [20, 2426].
As discussed in [22, 23], extended exergy accounting
has indeed incorporated some elements of existing methods like life-cycle analysis [28], cumulative exergy analysis [3], emergy analysis [29], extended exergy analysis
[24], complex systems analysis [12, 21, 30], and can thus
be properly considered as a synthesis of the pre-existing
theories and procedures of engineering cost analysis,
from which it has endeavoured to extract the most successful characteristics, as long as they were suitable for
a consistent and expanded formulation based on the new
concept of extended exergy. 1 As a final remark, an explanation of the name chosen for this new methodological approach is in order: the attribute extended refers to
the additional inclusion in the exergetic balance of previously neglected terms (corresponding to the so-called
non-energetic costs, to labour and to environmental remediation expenditures); the word accounting has been
suggested [6, 31] as a reminder that exergy does not satisfy a balance proper, in that the unavoidable irreversibilities which characterise every real process irrevocably destroy a portion of the incoming exergy; it is also a reminder that, as remarked in [9, 32], the exergy destruction is the basis for the formulation of a theory of cost,
because it clearly relates the idea that to produce any output, some resources have to be consumed.
This paper deals with the problem of optimal design. In spite of a vast body of literature that considers
computer-aided design and optimisation of thermal systems as two separate issues, in reality they are intimately
connected in their own very essence [33]: indeed, the outcome of every correctly conducted design act always satisfies two requirements: (a) it performs as specified by
the design data and abides by all constraints, and (b) it
displays the most desirable behaviour under a certain set
of operative conditions. This optimality is not always
expressed by a well-posed (in mathematical sense) objective function: in practice, vaguely formulated optimisation criteria are often at the basis of the design, which
nevertheless cannot be regarded as anything else than an
optimal one, however fuzzy or incompletely identified
this optimum may be. Therefore, in the following considerations we shall assume that the object of the design
1 Notice that the attribute extended is used here in a different
sense than that reported in [24], where the word is employed to
signify the extension of the application of exergy to the calculation of
environmental costs. Here, as described in [22, 23], use is made of a
new physical quantity, called extended exergy, which is the sum of the
physical-, the invested- and the added exergies of a stream, and is the
quantifier employed in the exergy balances.

69

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

activity is to produce an optimal design, so that the accent is shifted on the criterion under which the optimum
is to be searched for, that is, on the formulation of the
objective function.

2. DESCRIPTION OF THE METHOD


2.1. Exergy as a proper quantier for
the evaluation of chemical and
thermal processes
Exergy has several properties that suggest its use as a
convenient measure for an energy-accounting paradigm:
(1) Physical exergy, defined for a stream of matter at
a certain thermodynamic state 1 as

e1 = h1 h0 T0 (s1 s0 )+ (i ci 0 c0 ) [kJkg1 ]
i

(1)
is an extensive state property once the reference state 0
has been selected, 2 and as such its value is uniquely determined by the state parameters; notice that the quantity
defined by equation (1) constitutes the sum of the physical and chemical exergy, as defined for example in [34];
(2) The physical exergy of a stream can be augmented
or decreased by any combination of work-, heat- and
chemical interactions of the stream with other systems,
and, provided these interactions are expressed in exergetic equivalents, the final value taken by the exergy
of the original stream will correctly reflect not only the
quantity of the energetic exchanges (reflected in the variation of its enthalpy), but also their quality (measured by
the variations of the entropy and of the chemical potentials). Since exergy is additive, its cumulative value after the stream has undergone a series of independent thermodynamic transformations represents therefore a proper
measure of the quantity and of the quality of the global
energetic exchanges. As a corollary, it is possible to assign [9] a cumulative exergetic value to a certain product by summing up all the contributions to the different
streams that were used in its fabrication, starting from the
original values of the mineral ores that constituted the initial inputs in the process;
(3) If a proper Earths average chemical reference
state is defined, the chemical portion of the physical
2 For an interesting generalisation of the definition of physical exergy,
that makes it somewhat independent on the choice of a standard
reference environment, see also [35].

70

exergy of a mineral ore is either equal to zero (if the


ore is at the average composition of the Earths crust)
or attains a value exactly computable on the basis of
the ores chemical composition, its physical state and the
Gibbs energy of formation of its constituents. In practical
terms, therefore, the physical exergy of mineral ores and
of fossil fuels is exactly known once their composition
and their thermodynamic conditions at the extraction site
are known;
(4) If an effluent stream of a generic process is required to have a zero impact on the environment, the
stream must be brought to a state of thermodynamic equilibrium with the reference state before being discharged
into the environment. The minimum amount of energy
that must be used to perform this task by means of ideal
transformations is proportional to the physical exergy of
the stream [23]: therefore, the physical exergy of effluents is a correct measure of their potential environmental
impact (the amount considered here is the recycling exergy necessary to the ideal, zero-impact disposal of the
equipment); 3
(5) An invested exergy 4 value can be attached to any
product, and specifically to mechanical, thermal and
chemical equipment: this invested exergy is equal to
the sum of the non-energetic externalities (Labour and
Capital) used in the construction and operation of the
plant in which the product is generated. We maintain here
that a feasible formulation exists to convert both Labour
and Capital into exergetic terms, and that their equivalent
input in any process, referred to in this paper under the
name of added exergy, can be included in the extended
exergy, and would share its properties (additivity, nonconservation in real processes, etc.);
(6) For any product, it is then conceivable to define
an extended exergy as the sum of the physical exergy
and the proper portion of the invested exergy that can
be assigned to the stream under consideration; the environmental impact of the effluents can be included in
the analysis by charging the process with their exergetic
treatment cost, proportional to their corresponding physical exergies. Such an environmental cost is thus reflected both in a higher cumulative exergy content of the
process and in a higher invested exergy (the new components), and therefore in a correspondingly higher extended exergy of the final product.
3 As remarked in [23], one could also consider the recycling exergy
corresponding to a non-zero environmental impact corresponding to the
applicable Environmental Regulations: the numerical value is obviously
different, but the substance of the method is not affected by this change
of reference.
4 E.Yantowskii must be credited with introducing this definition in a
public discussion at a Meeting where ref. [15, 20] were presented.

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

2.2. Exergy algebra for eciency and


cost calculations
Consider a process P (figure 1(a)) in which a certain
well-defined material output is produced: the inputs to
the process will be in general a stream of raw materials
(I1 ), energy supply (I2 ) and human labour (I3 ). The
outputs consist of the desired product (O1 ), of some
energy rejection to the environment (O2 ), of a by-product
(O3 ), and of some waste (O4 ). There could of course be

(a)

several types of raw materials, several types of different


energy inputs, etc.: for our purposes, this rather simple
(but complete for our purposes) description of P shall
suffice. Assume that it is indeed possible to attribute an
exergetic value to each one of the input- and output fluxes
(the procedures proposed by EEA to assign these values
are briefly discussed in Section 3 and, more in detail,
in [22, 23]): then, all the input- and the output exergies
can be lumped together, and the black-box representation
of P takes the form shown in figure 1(b). Since physical
exergy is not conserved in real processes, there will be
an exergy loss, indicated by E . In general, the internal
structure of the process is known to the designer, who can
compute the transfer function of P, i.e., the (matricial)
expression that links the outputs with the inputs [36],
and that in the representation of figure 1(b) takes the
form: 5



 EO 1 

 EI1 



 EO 2 

(2)
EO = EI 
 =  EI2 

 EO 3 

EI3
E 
O4
Notice also that, due to the presence of the physical
exergy destruction E , the algebraic sum of the inputs
and the outputs (EO EI ) may assume only negative
values. The conversion efficiency of P can be computed
as the ratio of the useful output to the sum of the inputs
that concurred to produce it:
EO
P =  1
Ei

(b)

(3)

The exergetic cost of the output is defined as the amount


of input required (used) to generate the output, and is
the reciprocal of the conversion efficiency:

Ei
(4)
cP =
EO 1
If a portion of the energy discarded, say O2 , and of
the waste materials, say O3 , are recycled as inputs of
some other process belonging to the productive structure
of which P is part, the value of the overall process
efficiency will reflect this increased usefulness of the
waste streams:
=

EO1 + EO2 + EO3



Ei

(5)

(c)
Figure 1. (a) A process mass- and energy ow diagram. (b) The
extended exergy ow diagram for the process of gure 1(a).
(c) Introduction of partial output recycling in the process of
gure 1(a).

5 This matrix is identical to the Structural Matrix A used in


thermo-economics [21]. The different notation is adopted here both for
convenience and to remind the reader that EEA is somewhat different
from thermo-economics!

71

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

And the exergetic cost of the output can now be correspondingly computed as the reciprocal of the overall
process efficiency:

Ei
(6)
c(O1 +O2 +O3 ) =
EO1 + EO2 + EO3
If, as it is sometimes the case in complex production cycles, a portion of the output O1 is recycled internally to
P, then the control volume for P should be expanded as
shown in figure 1(c), to encompass the Splitter, i.e., the
physical component Pr that governs the partial recirculation of O1 : if the efficiencies of the sub-processes P
and Pr are known, the overall efficiency of the combined
process (P + Pr ) is given by:
E3
E1 + E2
E4 + E2
Pr =
E3
E4
P (1 Pr )
(P+Pr ) =
=
E1
Pr (1 P )
P =

(8)

eO 1 + eO 3

eO7 + 3i=1 eIi
eO 5
P2 =

eO3 + 7i=4 eIi
eO 6 + eO 7
P3 =

eO1 + eO5 + 10
i=8 eIi

(9)

and the overall efficiency of the conversion from the


resource base to the final product is:

(7)

and the relevant cost for the entire process (P + Pr ) can


be calculated:
c(P+Pr ) =

Pr (1 P )
P (1 Pr )

(10)

where = E2 /E3 is the recycle ratio.


Consider now a chain of technological processes that
represents all of the individual production steps in the
fabrication of some generic product O6 (figure 2): the
inputs can be classified as follows:
Materials and feed stocks: I1 , I4 , and I8
Labour: I3 , I7 , I10
Energy (various forms): I2 , I5 , I8
Inputs from previous fabrication step: I6 , I9
and the corresponding outputs:
Waste materials: O2 , O4 , O8
Waste energy (various forms): O9

Figure 2. Extended exergy ows of a technological chain based


on an extracted resource.

72

Recyclable by-products: O1 , O3 , O5
In equations (2) through (10) the exergetic fluxes have
been expressed in Js1 : it is convenient to normalise
these fluxes with respect to the unit mass flow rate of
useful output. In all cases where there is more than
one product, the mass flow rate (or the energetic rate
in the case of mechanical or electrical work) of the
main output can be taken as normalising quantity. From
now on, therefore, we shall deal with specific exergetic
contents, expressed in Jkg1 . 6 The primary conversion
efficiencies of the single processes are:
P1 =

eO
(P1 +P2 +P3 ) = 10 6

i=1 eIi

(11)
(12)
(13)

(14)

Noticein passingthat these expressions imply that


the fabrication steps that command the higher exergetic
inputs have a stronger influence on the conversion efficiency of the entire chain: this is a known result [9, 16],
and leads to the important consequence that the most
energy-intensive steps in a technological chain should be
given priority when performing a detailed process analysis. For the main purpose of this paper, it is important
to remark that the above expressions can be used to calculate the exergetic conversion efficiency and the corresponding exergetic costs of a given technological chain,
if its structure (as illustrated in figure 2) and the transfer
function of each fabrication step are known.
The procedure can be easily nested, i.e., applied at
different levels of aggregation in a productive structure
[37]. For example, each one of the three processes P1 ,
P2 and P3 shown in figure 2 could be analysed in detail
(disaggregated), and their individual internal structures
are likely to be topologically similar. Or , several technological chains could be joined (aggregated), each one
6 For fluxes of matter. If the output is shaft work or electricity, the
specific exergetic values are expressed in JJ1 . It is clear from the
considerations developed in Section 1, though, that our real goal here is
to attribute an (extended) exergetic value to the unit mass of product, so
that all exergies can in the end be measured in JJ1 .

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

of them being treated as a single process whose transfer function is known: of importance here is the fact that,
since the results are all expressed in homogeneous units,
they can be easily transferred from one aggregation level
to another, resulting in a general and very powerful analysis tool [27].

2.3. The choice of the control volume


The control volume (CV) employed in classical
exergetic analysis is drawn in such a way to include
(and hide) all of the internal exchanges, so that only
the fluxes to- and from the environment and to- and
from other processes are crossing the boundaries: it is
immaterial how far the ideal boundaries of the CV are
from the physical boundaries of the real process. For
EEA, the choice of CV is restricted by two orders of
considerations (for a more detailed discussion, see [22]):
(1) The boundary of the CV must be chosen far
enough from the physical boundaries of the real process
to include a portion of the environment large enough to
allow for the treated effluents to exit the CV boundary in
a state of zero physical exergy (see also footnote 3). This
portion has been called immediate surroundings [8, 34];
(2) If the inputs include unrefined fossil fuels or
minerals in as-mined conditions, the CV must include the
portion of the environment whence the original materials
were extracted, so that their initial (physical) exergetic
value may be computed. This procedure can often be so
bothersome that it becomes unpractical: as an alternative,
one can then directly assign the extended exergetic value
for every input, estimating this initial guess on the
basis an approximate knowledge of the extractionpretreatmenttransportation process (figure 3).

2.4. The time window spanned by an


EEA analysis
EEA is in principle similar to Life Cycle Analysis, 7
and the width of the time window over which the
calculation of the exergy flow diagrams is performed may
affect its outcome. Since time variations of some of the
input parameters cannot be neglected, an EEA analysis
must span the life of the product or plant (figure 4).
It turns out though that there are different time scales
that must be considered when assessing the relevant time
7 Except for the fact that, as clearly represented in figure 4, the net
EEA of any process necessarily leads to negative values.

Figure 3. The correct control volume for EEA analysis.

Figure 4. The time-window for extended exergy accounting.

window over which an EEA analysis can be correctly


applied, and that care should be exercised when applying
the method. For instance, when choosing the time frame
over which mineral ores or fossil fuels are exploited,
it is likely that the invested exergy of their extraction
may vary with time during the projected life of the
plant, reflecting a change in the mining conditions due
to an increased local relative scarcity of the material. Or,
when choosing the biodegradability time scale, often in
practice a conservative (in an economic sense) approach
is preferred to a zero-impact, and most environmental
regulations prescribe a certain time interval within which
the effects of the effluents must be completely buffered
by the immediate surroundings. Thus, extending this
time interval decreases the clean-up costs, but in effect
constitutes a substantial derangement from the zeroimpact assumption; on the other hand, shortening it
excessively may lead to a useless penalty in the exergetic
balance of the process, because it would fail to appreciate
(and exploit) the buffering capacity of the biosphere.

73

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

These two issues are only vaguely related to each other,


and no general rule is apparent for their methodical and
deterministic application: so, the decisions have to be
made on a case-specific basis (but, see also [22]).

3. THE EXERGETIC CONTENT OF AN


INPUT OR OUTPUT STREAM
The idea of adding up all the exergetic contributions
to a productive process is not new to EEA: it is rather
an extension of the Cumulative Exergy Consumption
method systematically described in [9]. In this section,
a detailed account is given of the original contributions
of EEA to the calculation of the exergetic content of the
individual streams.

3.1. Material and energy ows


A material stream is assigned a total (or cumulative)
exergetic content defined (points 5 and 6 in Section 2.1)
as the sum of its raw state exergy (physical exergy) and
of all the net exergetic inputs received, directly or indirectly, in the extraction, preparation, transportation, pretreatment and manufacturing processes. This second part
of the exergetic content may assume different numerical
values for different production chains: so, for instance,
the cumulative exergetic value of back-pressure steam is
different from that of the same steam, cogenerated in a
gas turbine plant. The first component of the total exergetic content, the raw state exergy, is computed with reference to an assumed average composition of the Earths
crust: there is disagreement among researchers on what
exactly this average composition ought to be [4, 10], but
for our purposes here it is only important that the choice,
whatever it is, be used consistently.
If an input or an output consists of an energy flow
(mechanical power, electrical energy, heat transferred
under whatever mode, chemical or nuclear energy, etc.),
EEA simply advocates the use of the corresponding
exergy value for that flux. Notice that to make the
treatment consistent, all exergies must be calculated with
respect to congruent reference conditions.

3.2. Labour and human services


The capability of attaching a properly computed exergetic value to a labour input (taken here to include

74

all service-related, blue- and white collar human activities) is perhaps the most relevant novelty of EEA, but it
represents also a highly controversial issue. Historically,
several different approaches (often formulated in radically different terms from each other) have been debated
among Economists and Industrial Engineers. Currently,
in all practical applications of Engineering Cost Accounting, including Thermoeconomics, labour is accounted for
on a purely monetary basis: but very reasonable alternative approaches have been formulated. Odum [26] proposes to assign labour a value derived by the total average emergetic value of human life in the particular area
where the workers happen to live and work; Industrial
Analysts argue that the equivalent energetic flows due to
labour exchanges between societal sectors (Domestic,
Agriculture, Tertiary and Industrial) are negligible;
Szargut [9] maintains that these contributions cancel out
if the complete picture is considered, because account
must be made for the exergetic costs for the sustenance
of the workers, that are of course borne by the system as
a whole. None of these approaches is completely satisfactory: if a purely monetary value is attached to labour,
market conditions and financial considerations are given
an unjustified advantage over social, technical and environmental issues thatproperly valued (see [38] for
instance)might affect the outcome of optimisation procedures in such a way as to suggest solutions that in their
globality result more advantageous for the well-being of
the people in all affected areas.
The emergetic approach is in principle correct, in that
it attaches to labour an energetic value (not necessarily
linked to a monetary counterpart), and postulates its dependence on local conditions in such a way that the unit
labour cost for a certain sector, in a given location at a
given time, is different for workers deriving their sustenance on different energy sources (different lifestyles).
It suffers though from the inability of emergy to properly
account for the different quality of diverse energetic carriers, and it fails to correctly account for this difference
in quantitative terms, because of its questionable energy
quality scale.
As for the claim about the relatively low incidence of
the energetic equivalent of labour on the energy balance
of a complex system, it can be said that, even if labour
is accounted for at its purely metabolic rate, which is
less than 0.1 kWperson1 , it is not always true that its
contribution can be simply neglected: in many human activities the purely metabolic rate, small as it may be, is of
the same order of magnitude of other production factors.
This is the case of most agricultural activities, but also
of fishing, herding, mining, construction, farming, housecleaning, garbage collecting, and in general of the activ-

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

ities of maintenance workers and technical operators in


all structures related to Service Centres (banks, police
stations, administration offices, etc.). For what highlyintensive industrial processes are concerned, one must
consider that in these processes workers command, on
the average, a much higher power than their bodily output
rate, thanks to a complex chain of technological devices
(controls and machines) that exert a substantial amount
of influence on the large energy fluxes participating to
the process, and thus make their activating energy (i.e.,
the human work input) correspondingly more valuable.
Finally, the classical exergetic approach as expressed
by Szargut appearswith respect to the valuing of
Labourincorrect: for the Society as a whole, the net
contribution does not cancel out. Consider figure 5,
which is a simplification of the structural representation
of a modern industrial society [39], and assume that the
global exergetic flux EL provided by the workforce can
be regarded as being originated in its entirety within the
Domestic Sector, and as flowing entirely into the Production Sector: the overall input (E3 ) into the Domestic Sector must be larger than EL , because it is inconceivable that the conversion efficiency of that Sector is
equal to 100%; so even if the other Sector (Production)
has ideal performance (i.e., in the absolutely unrealistic assumption that it does not destroy any exergy), the
net influx of resources into the system (Domestic + Production) must be higher than the net exergetic outflow.
In end effect, from the point of view of the interaction
society-environment, it is precisely this flux that sustains
the workers, and represents the global equivalent of their
exergy.

Figure 5. Simplied model of a society-environment interaction: the exergetic equivalent of the Labour contribution (E4 )
does not cancel out with the exergetic input into the Domestic
Sector (E2 + E3 ).

EEA proposes to assign labourand in general human servicesin each portion of a Society an exergetic
value computed as the total (yearly averaged) exergetic
resource into that portion input divided by the number of
working hours sustained by it:
eL,Sector =

Ein,Sector
nworkhours,Sector

[Jh1 year1 ]

(15)

3.3. Environmental impact


It is the Authors opinion that the acknowledged
shortcomings of the conventional monetary theory of
value when applied to environmental problems stem from
three basic faults. The so-called neo-classical economic
theory (NCE in the following) is founded on three
assumptions:
(1) that Economy deals with only a portion of reality,
namely the sum of anthropic activities;
(2) that in its original formulation it can treat problems in which there is possibility for unrestrained development (resources are abundant);
(3) that there is absolutely no concern about approaching Earths carrying capacity limits.
Neither of these assumptions is applicable to the
present situation on Earth: for one thing, the importance
of the eco-sphere for the long-range sustainability of anthropic activities is universally recognised; and furthermore, there are unmistakable signals that we are indeed
approaching (though in a broader sense than that expressed by the restricted and possibly erroneous point of
view of resources depletion) 8 the carrying limit of our
planet. These faults in the monetary theory of value have
the following result: when it comes to assess the environmental damage caused by man-made activities, which is
in effect impossible to price (what is the price of biodiversity?), NCE is forced to seek monetary measures of
this damage by introducing either ethical (willingness
to pay, sustainable development) or non-environmentally
relevant (risk assessment) concepts into its essentially
algebraic structure. Consider a generic chemical pollu8 To this regard, it is noteworthy to mention that, as argued very
convincingly in [40], the problem does not lie so much in the material
resources becoming scarce, but rather in the ever increasing cost of their
extraction, pre-treatment and distribution, that is becoming too high in
monetary terms. In EEA terminology, we might say that the extended
exergetic content of the resources is continuously increasing due in a
smaller measure to the depletion of their natural stock and in a larger
measure to our present inability to enact a total resource recycling
process for most of them.

75

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

tant: since it must be disposed of in a biological time, the


technological effort required to quickly reduce it to a set
of components of zero environmental impact has in effect an extremely high monetary price. Consequently, in
line with the current techno-economical paradigm, once
a substance is acknowledged to be (in some sense) harmful, it becomes regulated, i.e., legal upper limits are set
to its free release into the environment. This is equivalent to setting, for the state-of-the-art technology, an
upper price to the clean-up costs for that particular effluent. Moreover, since on the average, for many pollutants, the environmental situation has deteriorated to a
point that the overall balance of the biosphere has already been affected, the cumulative amount of the emissions of a certain type of pollutant over a certain period
of time may reach intolerable levels, so that the only way
to set a non-zero limit is to assess the risks to mankind
in terms of (monetary!) health- and life-expectancy parameters, and decide to set an upper bound to the expenditures in such a way to remain below a certain statistical probability of incurring in that risk. Since there is
no clean technology, this economical burden is transferred both ways in the economy: downwards, affecting
the price of resources that are perceived to be more- or
less clean, and upwards, increasing the price of products seen as less environmental-friendly. But willingness to pay and the attitude towards a sustainable resource exploitation are different in different Countries,
and may well vary in time: as a consequence, the disturbances caused by restricted or relaxed environmental regulations affect regions possibly far away from the point
of origin, and the system becomes intrinsically unfair, because not only the pollution, but the health risks are transferred from a region to the other. The incorporation of
vague and volatile concepts like global risk assessment,
willingness to pay and similar into a deterministic economic treatment, forces in end effect Process Engineers
and Planners to make numerous weakly justified assumptions when performing an Environmental Cost Analysis,
so that the credibility of any process- or plant economic
analysis which entails these costs is negatively affected.
Only recently a solution to this situation has been sought,
by trying to link the monetary structure of the environmental levies to some energetic consideration: this is the
rationale behind the pollution commodity trading and
the exergy tax [14, 15]. These are remedial measures,
though, aimed at a redistribution of the environmental
pressure on a global scale: they do not address the issue
of how high the actual environmental cost is (they take
the currently regulated values as a basis for their calculations).

76

EEA advocates a substantially different approach,


based on an immediate extension of the exergy accounting concept. Consider a process P (figure 6(a)), and assume that its only effluent is a stream which contains hot
chemicals, some of which not present in the environment.
To achieve a zero environmental impact, these chemicals
would have to be brought to both thermal- and chemical
equilibrium with the surroundings, which can be technically achieved in several ways: but in any case, the exergetic cost of the zero-impact (i.e., its extended exergy)
will be proportional to the physical exergy of the effluent; this exergetic cost corresponds to the extended exergy 9 ideally required to cool the effluent to T0 and break
it up into its constituents such that each one of them is
in equilibrium conditions with the surroundings. Real effluent treatment processes will no doubt require a substantially higher exergetic input than ideal ones. A possible representation of such an effluent treating process is
shown in figure 6(b): the additional process Pt requires an
energetic input, possibly some auxiliary materials, labour
and invested exergy, but its output will have a zero physical exergy. These additional exergetic expenditures required by Pt must be charged to the effluent O2 , whose
extended exergy will now be higher than its original one,
because for any real process Pt the net exergetic input
will be higher than the destroyed physical exergy. Therefore, the overall conversion efficiency of the joint process
(P + Pt ) is decreased. There are effluents for which at
least a part of the chemical decomposition reactions takes
place spontaneously, in a short time and in the immediate surroundings of the emitting source: but, as correctly
remarked in [25], in these cases the reactions must draw
on some exergy source within the environment (which can
be a certain particular chemical, oxygen, water, solar radiation, or even a biological system), and this exergy flow
must be accounted for as well.
Using an EEA approach, alternative strategies to deal
with the treatment of effluents can be formulated and assessed, and a discussion is given in [23]. The important
issue here is that EEA allows for a consistent incorporation of the effects of effluent treatment in the extended
exergetic balance of a process, and that it provides an absolute order-of-magnitude estimate of the minimum exergy consumption necessary to achieve zero-impact. Notice that, if an acceptable level of pollutant is specified,
then the minimum exergetic expenditure will be proportional to the difference between the values of the physical exergies of the effluent stream between the point of its
9 I.e., the sum of the physical exergy spent in the clean-up process,
plus the invested exergy (labour and capital) required by the installation
and operation of the effluent clean-up devices.

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

(a)

(b)

(c)
Figure 6. (a) The euent O2 is not a reference conditions, (b) real treatment of O2 . Each one of the nal euents is at its reference
conditions, (c) only a portion of the clean-up is performed by man-made treatment processes: the remaining takes place spontaneously
in the immediate surroundings.

77

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

release and the regulated state point. This last consideration effectively dispenses of the considerable effort required by all methods presently in use to determine what
the tolerable environmental impact limit for a certain
pollutant would be.

3.4. Capital
Parallel to the debate over labour, a similar debate
has been staged over the years about how to include
capital- and in general financial costs in the picture. The
dominant force is today still the above mentioned neoclassical theory of economic price, which uses money
and its time-value as a quantifier for goods, services,
labour, resources, etc. The rationale often invoked by
NCE supporters who oppose the search for an alternative
value paradigm is that, on an average world basis, the
contribution of energy-related costs to the global gross
national product varies between 2 and 9%, and that
therefore money (= Capital) and not energy is the main
quantifier of human activities. This paradigm of money
as a basic metric for energy-related activities is so deeply
rooted into our culture that often even energy balances for
a given industrial field are expressed in monetary price
per unit of product rather than on its energetic (not to
mention its exergetic!) content.
One could though consider a different and opposite
point of view that reverses the issue: namely, that economic systems are eco-systems that function only because
of the energy and material fluxes that sustain human activities. Moreover, these agricultural, industrial and economic activities can only exist as long as they exploit
(use) biophysical resources taken from a reservoir of
non-infinite capacity. Therefore, it is conceivable to reverse NCEs point of view, and value human activities accordingly to a different metric, based on an energy quantifier. This reversal of the scale of values is exactly the
issue raised by EEA, that in addition maintains that exergy be the right quantifier. In this perspective, it appears
clearly that it is not capital that ought to measure the
value of a piece of equipment or of a product by attaching a price tag to it, but exergetic content; and that the
monetary price ought to reflect this new scale of values.
Notice that EEA does not demand an utopic cancellation of monetary prices: it is clear, on the contrary, that
the very structure in which we live and function today
demands for the conservation of the price-tag concept
in everydays economic activities and on a world-wide
scale. What EEA advocates is that this tag be calculated
on the basis of the extended exergetic content (EEC in

78

the following) of a good or service, corrected for environmental impact (as defined in Section 2). It is also immediately clear what the numerical relation between the
EEC and the price of a product ought to be: since the
EEC is expressed in kJunit1 , and the price in $unit1 ,
the conversion factor Kex is the ratio of some measure
of the monetary circulation to the global exergetic input.
The choice of a proper indicator for the monetary flux
is of course somewhat arbitrary: for the purpose of the
present study, we have chosen the absolute measure of
the global monetary circulation (M2), which for each
Country is computed and published by the Central Bank.
When processes are analysed, it is the exergetic equivalent of the capital expenditure that must be inserted in the
balance as an input, and the same applies to the revenues.
It is certain that for some of the processes the substitution of the equivalent exergetic value for the monetary
price will show a discrepancy in the exergy balance: these
adjustments are caused by the present over- or underestimating of the real extended exergetic content of materials, feed stocks, Labour and energy flows. It would be
desirable that the economic and the exergetic value become locally consistent in the long run, meaning that the
two value scales reach a sort of fixed parity: on the other
hand, of course, the very same definition of the exergyequivalent implies that different Countries may have different Kex , due to their different productive and economic
structures and lifestyles.

4. AN EXAMPLE OF APPLICATION OF
THE METHOD
4.1. Extended exergy accounting of
cyclic processes
Consider the gas turbine based cogeneration system
depicted in figure 7 [11, 20, 41]: there are 16 streams
joining eight components in this schematic representation
of the plant. Notice the addition of an effluent treatment
plant for the stack gases. Assuming for the moment that
it is possible to compute the invested exergy content
of all components, the following equations express the
extended exergy accounting of the process:
equations derived from the balances of each individual
component:
Compressor
e2 = e1 + e11 + eC + eC
Air preheater:

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

e4
deCC
eCC
e7
e7
e7
de7 =
de6 +
de8
de9
e6
e8
e9
e7
e7
+
deHRB
deHRB
eHRB
eHRB
e9
e9
e9
de9 =
de6
de7 +
de8
e6
e7
e8
e9
e9
+
deHRB
deHRB
eHRB
eHRB
e10
e10
e10
de10 =
de4
de5 +
deGT
e4
e5
eGT
e10

deGT
eGT
e12
e12
e12
de12 =
de10
de11 +
dePS
e10
e11
ePS
e12

dePS
ePS
e13
e13
e13
de13 =
de12 +
deEL
deEL
e12
eEL
eEL
e15
e15
e15
de15 =
de7 +
dePt
dePt
e7
ePt
ePt

Figure 7. Simplied process scheme of a gas turbine-based


cogeneration plant.

e3 = e2 + e5 e6 + ePR ePR
Combustion chamber:
e4 = e3 + e14 + eCC eCC
Gas turbine:
e10 = e4 e5 + eGT eGT
Power splitter:
e12 = e10 e11 + ePS ePS

(17)

(16)

A similar reasoning applies to the invested exergy of each


component (eight equations):

The variation of each output produced by a variation in


one or more inputs is expressed by a total differential.
Assembling together the eight equations (one for each
stream), one obtains:

eC
eC
eC
eC
de1 +
de2
de11 +
deC
e1
e2
e11
eC
ePR
ePR
ePR
dePR =
de2 +
de3 +
de5
e2
e3
e5
ePR
ePR

de6 +
dePR
e6
ePR
eCC
eCC
deCC =
de3 +
de4
e3
e4
eCC
eCC

de14 +
deCC
e14
eCC
eGT
eGT
deGT =
de4 +
de5
e4
e5
eGT
eGT

de10 +
deGT
(18)
e10
eGT
ePS
ePS
de10 +
de11
dePS =
e10
e11
ePS
ePS
+
de12 +
dePS
e12
ePS
eEL
eEL
eEL
deEL =
de12
de13 +
dePR
e12
e13
ePR

Electrical generator:
e13 = e12 + eEL eEL
Heat-recovery boiler (main flow):
e9 = e6 e7 + e8 + eHRB eHRB
Heat-recovery boiler (by-product):
e7 = e6 + e8 e9 + eHRB eHRB
Gas cleaning unit:
e15 = e7 ePt + ePt

e2
e2
e2
e2
de1 +
de11 +
deC
deC
e1
e11
eC
eC
e3
e3
e3
de3 =
de2 +
de5
de6
e2
e5
e6
e3
e3
+
dePR
dePR
ePR
ePR
e4
e4
e4
de4 =
de3 +
de14 +
deCC
e3
e14
eCC
de2 =

deC =

79

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

eHRB
eHRB
eHRB
de6 +
de7
de8
e6
e7
e8
eHRB
eHRB
+
de9 +
deHRB
e9
eHRB
ePt
ePt
ePt
dePt =
de7
de15 +
dePt
e7
e15
ePt

If the necessary process- and property data are available, the analysis may be further extended by performing
either one of the following steps:

deHRB =

finally, there are eight additional equations for the de , all


of the form:
deC =

eC
eC
eC
de1
de2
de11
e1
e2
e11
eC

deC
eC

(19)

In conclusion, the entire system of equations has the


structure:
A De = 0

(20)

with A = |ai,j | = |ei /ej | and De = |dei |. The coefficient matrix is band-diagonal, and its coefficients are
related by the obvious relation:
am,n =

1
an,m

(21)

Further relations exist between the ai,j -coefficients:


All coefficients for which both i and j are numerical
indexes can be computed by means of thermodynamic
relations, because they represent the extended exergies of
the corresponding streams;
Some terms are 0, or anyhow negligible: for example,
aPR,2 , aC,1 , aCC,3 , aHRB,8 , aGT,5 , etc.
All of the terms for which both i and j are literal
indexes have a known structure: for example, aC,C can
be explicitly computed if eC = f (C ) is known, etc.
In conclusion, the coefficient matrix is easily calculated
and inverted, and the system can be solved, obtaining
the values of the dei as a function of a given set of
input conditions and of design data. Manipulating the
inputs, or the design specifications, or both, it is possible
to inspect the sign- and value variations of the new
set of dei , and proceed towards an extremum: since
the problem is usually complex, it may be advisable to
apply a numerical multi-variable optimisation routine.
It is important to remark that this optimal solution is
obtained by a method that introduces into the equations
considerations related to thermodynamic-, economical-,
and environmental issues: therefore, it can be regarded
as the broadest possible optimisation in engineering
accounting.

80

Modify one or more of the processes Pi constituting the cycle, and recompute the overall conversion
efficiency- and exergetic cost structure. This constitutes
a comparison of different technological scenarios (for
instance, one could substitute an intercooled compressor for a standard one or a new gas turbine for an older
model);
Compare two different cycles that produce the same
main outputs, and assess their relative merits for what
resource conversion is concerned. This amounts to performing a comparison between different production technologies; for example, one could replace the combustion
chamber with a coal gasification process, and compare
the exergetic cost of the generated kW of power and heat.

4.2. A benchmark case: EEA analysis of


a simple gas-turbine based
cogeneration process

The process discussed in the previous section can be


used as a benchmark for testing the feasibility of the extended exergy accounting technique. The practical implementation of the procedure, at least in the case of the example discussed in the following, is in reality somewhat
different from the formally elegant mathematical formulation offered above, but it is definitely more practical:
it is often easier to use a process simulator to perform a
parametric sensitivity analysis of the process with respect
to certain parameters that are known (or thought) to bear a
substantial influence on the process output. In the present
case, the turbine inlet temperature (T5 , TIT in the following) and the compressor delivery pressure (p2 ) were
varied stepwise within a certain industrially acceptable
range, and for each (TIT/p2 ) design point the overall exergetic economics of the process were calculated. Some
problems have been encountered in the construction of a
reliable database: this is a difficulty often encountered in
all Engineering Economic studies, as well in Thermoeconomics. The following assumptions have been made:
(a) All raw materials are assigned a physical exergy
computed on the basis of Szarguts data [10];
(b) The exergetic cost of all equipment is computed
by first calculating the monetary cost via the available
costing tables [8, 42], and then by converting the capital

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

cost into exergetic cost via the exergetic cost factor Kex =
0.055 $ MJ1 10 ;
(c) The fuel exergetic cost has been taken equal to the
sum of the fuel physical exergy, the chemical portion of
which is set equal to the lower heating value;
(d) Labour costs are computed on the basis of industrial monetary estimates, converted to exergetic values by
applying the exergetic cost factor Kex ; where possible, a
direct estimate based on the number of working hours has
been employed, so that the value given by equation (15)
could then be employed;
(e) The extended exergy value of all pure input
streams (air, cooling water) is taken equal to zero;
(f) The environmental costs are computed on the basis of the extended exergy accounting applied to the relevant treatment plant (in this case, only the Desulphuration/Denitrification unit). Since exact process data were
not available, EEenvironmental has been assumed to be
equal to Egas , being a penalty factor larger than 1.
The object of this exercise being that of demonstrating
the feasibility of the procedure, we have elected to sacrifice numerical exactness to congruence, and therefore
have not made any special attempt to verify the congruency of the cost data. It is necessary to recognise here
though that a more exact calculation of the exergetic
costs can become very cumbersome for even the simplest
process. With these limitations in mind, we can now proceed to examine the procedure that led to the optimal
design point:
(1) Mass-, energy and exergy balances were computed using a standard process simulator [33, 36];
(2) The costs of the various components were computed by formulas adapted from [8, 42];
(3) The conversion into exergetic values by means of
the equivalence factor Kex was performed on the basis of
the values of M2 and of Ein calculated for Italy in 1994
[37];
(4) Some additional cost data (mostly, those related
with the fixed expenses, maintenance and overhaul, insurance, etc.) were adapted from [8];
(5) The labour requirements were estimated by analogy with other plants for which the data are in the public
domain. The exergetic costs were computed either by use
of the equivalence factor Kex as under point (3) above,
or by direct man-hour computation were possible. In this
case, the exergetic equivalent of each worker was his apportioned share of primary resources, taken at the value
10 K = M2/E , where M2 is one of the standard indicators of the
ex
in
monetary circulation and Ein is the total yearly exergetic input in the
Country under consideration

based on the data available for Italy in 1994 [37]. This


value can be obtained by equation (15) and amounts to
52.7 MJworkhour1;
(6) The ecological cost penalty factor was assumed
to remain constant for the entire range of the explored
variation of the optimisation parameters (the compression
ratio and the gas turbine inlet temperature TIT): this
approximation can be avoided if a more exact process
analysis is available;
(7) The operating costs were computed under the
further assumption of a fixed-load operation throughout
the year, with a Plant Factor PF = 0.6 (corresponding to
5250 hoursyear1 );
(8) The objective function is the extended exergetic
cost of the global product, i.e., steam and electrical
energy, per unit electrical kW generated. No attempt has
been made here to separately compute the cost of the
steam.
Under the above assumptions, the results are displayed
in figure 8 for three values of the TIT. The following
remarks can be made:
(1) The real efficiency of the process, equal to the
inverse of the extended exergetic cost, takes much lower
values than the ones we are accustomed to: this of course
depends on the fact that the entire process cycle, from
resource extraction to waste disposal, has been taken into
account here. It is this efficiency indeed the one which
our energy projections and models ought to be based on;
(2) The minimum extended exergetic cost is seen to
correspond, for all values of the TIT, to lower pressure
ratios than the corresponding maximum thermodynamic
efficiency. This indicates that the added exergy inputs
charged here to the process (capital- and labour costs,
quasi-life-cycle costs of resources and materials) modify
in fact the balance between the financial charges generated by the installation costs (that grow with p2 ) and the
fuel costs (that behave as the inverse of the efficiency).
In other words, the additional terms introduced into the
efficiency calculation by the extended exergy accounting
technique are relevant to the overall exergetic budget;
(3) With the model parameters adopted here (see table I) both the overall monetary cost of the installed
power (in $kW1 ) and the cost of the generated electricity ($kWh1 ) reflect the values commonly adopted in
cost calculations. This indicates that the results obtained
by the application of extended exergy analysis per se are
relevant, in that a customary cost accounting applied to
the same example presented here would give results in
line with the present state of the art of current energy audits;
(4) The parameters that most influence the cost of the
generated kWh are the financial rate of restitution R, the

81

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

(a)

(b)

(c)

(d)

Figure 8. Results of the EEA analysis of the gas turbine-based cogeneration process depicted in gure 7: (a) capital costs as function
of ; (b) production costs as a function of ; (c) cycle ( ) and EEA ( ) eciencies as functions of ; (d) comparison of the optimal
values of with the two methods.

TABLE I
Process parameters (with reference to gure 7).

Air inlet temperature, T1


Air inlet pressure, p1
Compressor polytropic efficiency, pc
Fuel lower heating value, LHV
Pressure loss in the combustor, %pcc
Turbine polytropic efficiency, pt
Turbine discharge pressure, p5
Steam temperature, T10
Net electrical power output P

300 K
1 bar
0.89
41000 kJkg1
2%p2
0.86
1.1 bar
420 K
110 MW

fuel costs and the environmental penalty factor . This


suggests that it may be interesting to systematically apply
an extended exergy accounting technique both to the
industrial processes that produce the components used
in the plant and to the service sector that generates the
required capital restitution rate;
(5) Under the assumptions adopted here, labour does
not seem to bear a major influence on the overall opti-

82

Water temperature at boiler inlet, T11


Boiler efficiency, b
Gas temperature at DeSox inlet, T12
Min. allowable stack gas temp., T13
Plant book life
Capital rate of return, R
Fuel price
Labour costs (exergy equivalent)
Environmental penalty factor,

350 K
95%
450 K
400 K
20 years
6%
0.2 $ kg1
52.7 MJ(man1 h)1
3

mum. Its direct exergetic equivalent amounts to about 2


3% of the overall exergetic expenditure. Obviously, there
is another hidden contribution of labour to the overall
cost: this is given by the labour equivalent contained in
the extended exergetic cost of each piece of equipment.
On the average, it can be estimated that about 25% of the
exergetic content of each component of the class considered here consists of labour exergy: the boiler and alterna-

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

tor probably averaging a little more, the compressor, gas


turbine and combustion chamber a little less. However,
the contribution is not so small to be neglected without
influencing the overall budget.

5. CONCLUSIONS
The method of Extended Exergy Accounting (EEA)
as described in this paper and in more detail in a series
of companion papers [22, 23, 27], has been applied to
the design optimisation of a cogenerative power plant,
using realistic cost estimates. The exergetic equivalents
of the capital and labour costs have been computed
based on global data available for Italy in 1994. The
results show that EEA is indeed a practical instrument for
performing design optimisation tasks, and that its results
complement and extend those obtainable via a thermoeconomic analysis. Further work is necessary to expand
and to improve the exergetic costing database, which
has clearly a major impact on the practical usefulness of
the method. Also, more work is needed to explore the
possible implications of a generalised application of EEA
to industrial processes and complex systems.

REFERENCES
[1] Bosnjakovic F., Technische Termodynamik, Steinkop, Dresden, 1960.
[2] Fratscher W., Zum Begri des exergetischen Wirkungsgrads, BWK 13 (11) (1961) 486493.
[3] Szargut J., Anwendung der Exergie zur angenherten wirtschaftlichen Optimierung, BWK 23 (12) (1971)
516519.
[4] Ahern J., The exergy method of Energy Systems
analysis, Wiley, New York, 1980.
[5] Moran M.J., Availability Analysis, McGraw-Hill, New
York, 1989.
[6] Gaggioli R.A., Wepfer W.J., Exergy economics, Energy 5 (1980) 823837.
[7] Kotas T., The Exergy Method of Thermal Plant
Analysis, Butterworths, London, 1985.
[8] Bejan A., Tsatsaronis G., Moran M.J., Thermal
Design and Optimization, Wiley, New York, 1996.
[9] Szargut J., Morris D.R., Steward F.R., Exergy Analysis of Thermal, Chemical and Metallurgical Processes,
Hemisphere, Washington, DC, 1988.
[10] Szargut J., Reference level of chemical exergy,
Archiw. Termodyn. 9 (12) (1988) 107117.
[11] von Spakovsky M.R., Frangopoulos C.A., A global
environomic approach for Energy Systems analysis and
optimization, Part 2, in: Proc. ENSEC, Krakow (Poland),
1993, pp. 133144.

[12] Reistad G.M., Available energy conversion and


utilization in the US, ASME J. Engrg. Power 97 (1975) 603
611.
[13] Wall G., Naso V., Sciubba E., Exergy use in the Italian society, Energy 19 (1994) 12671274.
[14] Gong M., Wall G., On Exergetics, economics and
optimization of technical processes to meet environmental
conditions, in: Proc. TAIES 97, Beijing (China), 1997,
pp. 453460.
[15] Hirs G., Exergy loss: a basis for energy taxing,
in: Proc. NATO-ASI Works. Thermodyn. Optimiz. Complex
Energy Systems, Constantza (Romania), 1998.
[16] El-Sayed Y., Gaggioli R.A., The integration of synthesis and optimization for conceptual designs of energy
systems, ASME J.E.R.T 110 (2) (1988) 109113.
[17] El-Sayed Y., Evans R.B., Thermoeconomics and the
design of heat systems, J. Engrg. Power 92 (27) (1970) 27
35.
[18] Evans R.B., A contribution to the theory of ThermoEconomics, M.E. Thesis, UCLA, Dept. of Engineering, 1961.
[19] Frangopoulos C., von Spakowski M., A global environomic approach for Energy Systems analysis and optimization, Part 1, in: Proc. ENSEC 93, Cracow (Poland), pp.
123132.
[20] Tsatsaronis G., Design optimisation using exergoeconomics, in: Proc. NATO-ASI Works. Thermod. Optimiz.
Complex Energy Systems, Constantza (Romania), 1998.
[21] Valero A., Serra L., Lozano M.A., Structural theory
of Thermoeconomics, Proc. ASME AES 30 (1993) 189198.
[22] Sciubba E., Extended exergy accounting: towards
an exergetic theory of value, in: Proc. ECOS 99, Tokyo,
Japan, 1999, pp. 8594.
[23] Sciubba E., Exergy as a measure of environmental
impact, in: Proc. ASME-WAM 99-IMECE, Nashville, TN (USA),
1999, pp. 573581.
[24] Creyts J.C., Carey V.P., Use of extended exergy
analysis as a tool for assessment of the environmental
impact of industrial processes, in: AES-37, ASME, 1997,
pp. 129137.
[25] Makarytchev
S.V.,
Environmental
pollution
evaluationan exergy approach, in: Proc. 1st Internat.
Symp. on Issues in Environm. Pollution, Denver (CO, USA),
1998, pp. 223231.
[26] Odum H.T., Environmental Accounting, Wiley, New
York, 1996.
[27] Sciubba E., A nested black-box exergetic method
for the analysis of complex systems, in: Proc. Adv. in
Energy Studies, P. Venere, Italy, 1998, pp. 471482.
[28] Frankl P., Gamberale M., The methodology of LCA
and its application to the energy sector, in: Proc. Adv. in
Energy Studies, P. Venere, Italy, 1998, pp. 241256.
[29] Odum H.T., Environment, Power and Society, Wiley,
New York, 1971.
[30] M. Tribus, Y. El-Sayed, A specic strategy for the
improvement of process economics through Thermoeconomic analysis, in: Proc. 2nd World Conf. Chem. Engrg.,
Montreal, Canada, 1981, pp. 117129.
[31] Spiegler K.S., Principles of Energetics, Springer,
Berlin, 1983.
[32] OConnor M., Theory of value for open systems
reproduction: the role of energy-based Numeraires in

83

E. Sciubba / Exergy Int. J. 1(2) (2001) 6884

analyses for sustainability, in: Proc. Adv. in Energy Studies,


P. Venere, Italy, 1998, pp. 4976.
[33] Sciubba E., Toward automatic process simulators:
modular design procedures, ASME J.E.R.T. 120 (7) (1998)
18.
[34] Moran M.J., Sciubba E., Exergy analysis: principles
and practice, J. Engrg. G.T. Power 116 (4) (1994) 285290.
[35] Gaggioli R.A., Richardson D.H., Bowman A.J.,
Paulus D.M., Available Energy: IGibbs revisited, IIGibbs
extended, in: Proc. ASME-IMECE 99, AES, Vol. 39, Nashville,
TN (USA), 1999, pp. 285295.
[36] Falcetta M.F., Sciubba E., A computational, modular
approach to the simulation of power plants, Heat Rec. Sys.
& CHP 15 (2) (1995) 131145.
[37] Azzarone F., Sciubba E., Analysis of the energetic
and exergetic sustainability of complex systems, in: Proc.
ASME Conf., AES, Vol. 35, 1995, pp. 161174.

84

[38] Sraa P., Production of commodities by means of


commodities, CUP, 1960.
[39] Sciubba E., Modelling the energetic and exergetic
self-sustainability of societies with dierent structures,
ASME J.E.R.T. 117 (6) (1995) 7586.
[40] Ayres R., Exergy, waste accounting, and life-cycle
analysis, Energy 23 (5) (1998) 355363.
[41] Lego P., Rivero R., de Oliveira S., Schwarzer B.,
Energetic and Economic optimization of industrial systems
compared, Adv. in Thermod. 4 (1991) 4856.
[42] El-Sayed Y., A short Course in Thermo-Economics,
Summer School, Ovidius Univ., Constantza, Romania, July
1999.
[43] Evans R.B., von Spakowsky M.R., The design and
performance optimization of thermal systems, J. Engrg. G.
T. Power 112 (1) (1990) 116122.

You might also like