Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Physica A 462 (2016) 421–430

Contents lists available at ScienceDirect

Physica A
journal homepage: www.elsevier.com/locate/physa

Econophysics and bio-chemical engineering


thermodynamics: The exergetic analysis of a municipality
Umberto Lucia
Dipartimento Energia, Politecnico di Torino, Corso Duca degli Abruzzi 24, 10129 Torino, Italy

highlights
• Econophysics is very powerful to support municipality policies.
• Bioeconomics plays a fundamental role for new perspectives in economy.
• Biochemical engineering thermodynamics highlights the role of irreversibility and exergy.
• A link between econophysics and biochemical engineering thermodynamics is obtained.

article info abstract


Article history: Exergy is a fundamental quantity because it allows us to obtain information on the useful
Received 16 April 2016 work obtainable in a process. The analyses of irreversibility are important not only in the
Received in revised form 24 May 2016 design and development of the industrial devices, but also in fundamental thermodynam-
Available online 21 June 2016
ics and in the socio-economic analysis of municipality. Consequently, the link between en-
tropy and exergy is discussed in order to link econophysics to the bio-chemical engineering
Keywords:
thermodynamics. Last, this link holds to the fundamental role of fluxes and to the exergy
Entropy
Entropy generation
exchanged in the interaction between the system and its environment. The result consists
Exergy in a thermodynamic approach to the analysis of the unavailability of the economic, pro-
Constructal law ductive or social systems. The unavailability is what the system cannot use in relation to its
Irreversible biochemical engineering internal processes. This quantity result is interesting also as a support to public manager
thermodynamics for economic decisions. Here, the Alessandria Municipality is analyzed in order to highlight
Econophysics the application of the theoretical results.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction

In the analysis of the thermodynamic behavior of open systems, irreversible processes represent one of the fundamental
topics of investigation in thermodynamics. Indeed, the studies on irreversibility are important not only in the design and
development of the industrial devices [1], but also in the studies related to biomedical applications [2,3]. The studies on
irreversibility are very interesting in thermodynamics because they allow us to write the dissipation terms by using entropy
generation and, consequently, to express the usual inequality by means of equations [1,3–9].
In 1824, Carnot [10] introduced an ideal engine operating on a reversible cycle without dissipation. This engine converts
the absorbed heat in work and, apparently, it has no irreversibility. Carnot proved that [10,11]:
1. All ideal engines operating between the same two thermal baths (thermal reservoirs) of temperature T1 and T2 , with
T1 > T2 , have the same ideal efficiency ηC = 1 − T1 /T2 .
2. Any other engine, operating between the same temperature, has an efficiency η such that it is always η < ηC .

E-mail address: umberto.lucia@polito.it.

http://dx.doi.org/10.1016/j.physa.2016.06.119
0378-4371/© 2016 Elsevier B.V. All rights reserved.
422 U. Lucia / Physica A 462 (2016) 421–430

Carnot’s conclusions represent the existence of a definite limit for any conversion of the heat into the kinetic energy
and work [10,11]. Just to develop the analysis of dissipative processes, Clausius [12,13] introduced a new thermodynamic
quantity, named entropy [12,13].
Science and technology are considered fundamental to the growth and socio-economic development of countries; indeed,
technological development has impact on income distribution, economic growth, employment, trade, environment and
industrial structure [14,15]. The 1992 Earth Summit stipulated that countries at national level as well as governmental and
non-governmental organizations at international level should develop indicators of sustainable development in order to
support countries in making decisions on sustainable development [14].
At all levels, the role of science and technology is fundamental; scientific knowledge and technologies are the basis
to challenge economic, social and environmental problems in order to avoid unsustainable conditions. The analysis of
technological processes can be developed using a thermodynamic approach for the whole system and for all its interactions
both internal to the process and external to the environment and society. The results consist of a quantitative evaluation
of the flows of matter and energy which occur in the system and of the consumption rate of the available resources. This
information can represent a fundamental support to policy planning and resource management [15].
In order to evaluate the technological level and the advanced level of industrial processes some indicators must
be considered. Every company applies different production processes, which cause different carbon emissions or
environmental impact, therefore as regards the environmental effects, the process itself results more important than the
product obtained. In order to analyze both the environmental impact and the technological level acquired by the countries
several indicators can be introduced. These can be defined as in Ref. [15] ‘‘an aggregate, a quantitative measure of the impact
of a ‘community’ on its surroundings (environment)’’. It implies that:
1. The ecological indicators must be applicable to any ‘‘community’’;
2. They are aggregated because it cannot be limited to a single individual;
3. They consider only the effects produced on the environment that surrounds the community under examination.
From this definition it follows that the community and the environment must be considered as two separate, but interacting
systems [15]. The consequent properties of the environmental indicators can be summarized as follows:
1. They must be evaluated using unambiguous and reproducible methods under a well defined set of fundamental
assumptions;
2. They must be expressed by a numeric expression whose results can be ordered in an unambiguous way;
3. They must be calculated on the basis of intrinsic properties of the community and of the environment;
4. They must be normalized in order to compare different communities or environments;
5. They must be defined on the basis of the accepted laws of thermodynamics.
Sciubba analyzed in detail a lot of indicators and pointed out their limits in Ref. [15]. Here his results are summarized
according to the most used environmental indicators:
1. MTA (Material Throughput Analysis or Material Inventory Analysis): it is an indicator based on the assumption that the
lifestyle of a community can be measured by the global equivalent material flow used to produce the commodities on
which it thrives. The method involves highly disaggregated accounting of the material inputs/outputs and it requires
detailed knowledge on production processes. Moreover, it does not use the second law of thermodynamics;
2. EEn (Embodied Energy): it is an indicator which obtains a direct measure of environmental impact. The amount of energy
used to construct a product, in terms of resources and work done, is evaluated, but it does not include any measure of
the quality of the energy flows considered;
3. The tranformity: in the Emergy Analysis the energy accounting is considered, but the fundamental assumption is that
the only input form of energy is the solar radiation. All other flows of matter and energy are related to equivalent solar
energy necessary to obtain them. This evaluation is carried out using a proper set of coefficients, the transformities. It
does not include any measure of the different quality of the energy flows.
Now, we must consider that energy is a thermodynamic property which characterizes any state of a thermodynamic
system in relation to a reference state. It is a conserved quantity in relation to the universe (the system and its environment),
and its total value is always constant [16]. So, its physical meaning is related to its variation, which is the value and the causes
of the useful work. Consequently, any change in a system is no more than a transition between two states.
On the other hand, the exergy of a system is the maximum shaft work obtainable by the system in relation to its specified
reference environment, which is considered infinite, in equilibrium and it is specified by fixing its temperature, pressure
and chemical composition [16]. The exergy was introduced [17] by Carnot [10], even if only implicitly, as a fundamental
concept of the thermodynamics [18–22]. But, Gibbs [21] was the first to define explicitly the available work, even if also
Tait and Lord Kelvin have defined a quantity similar to Gibbs’ one [22], but they did not improve the concept [17], while the
Gibbs approach was the basis of a great number of developments [23–27].
In 1889, Gouy and, in 1905, Stodola, independently, proved that the lost exergy in a process is proportional to the entropy
generation [25–29]. Gouy’s results were used in a great number of analyses of irreversibility [30–69,15,70].
Exergy is a quantity defined in relation to a reference, the environment [16,71,72]. It is defined as the maximum amount
of work obtainable by a system as it comes to equilibrium with its reference environment. So, it represents a measure of the
U. Lucia / Physica A 462 (2016) 421–430 423

ability of a system to cause changes, due to its non complete stable equilibrium, in relation to the reference environment.
Consequently, we highlight that [16,71]:
1. The exergy of a system in complete equilibrium with its environment is null;
2. Exergy does not follow any conservation law;
3. A system carries exergy proportional to the level of disequilibrium with its environment;
4. Any loss of energy quality results in consumption of exergy.
Exergy is a quantity that allows the engineers to design system with the aim of not only obtaining the highest efficiency at
a least cost under the actual technology, economic and legal conditions, but also considering ethical, ecological and social
consequences; indeed,
1. It allows the evaluation of the impact of energy resource utilization on the environment;
2. It allows the evaluation of more efficient energy-resource use, and of the locations, types, and magnitudes of wastes and
losses;
3. It is an efficient technique to evaluate if it is possible to design more efficient energy systems by reducing the inefficiencies
in existing technologies.
Then, many other approaches have been developed [73–91,86,92,93] in order to describe the internal needs of the open
systems, their optimal paths in their phase space, based on the minimum momentum, i.e. on the least time. But, it is not
always possible to obtain information from some complex systems, due to their partial accessibility [94–101]. So, for such
systems, the use of the traditional approaches is difficult, with the only result of obtaining specific and not generic models.
Now, we wish to highlight that any effect in Nature is always the result of the dynamic balances of the interactions between
the systems and their environment [102–109]. So, the exchange of energy drives some behavior of natural systems, i.e. their
evolution is driven by the decrease of their free energy in the least time [110–120].
But, all the real systems operate on irreversible thermodynamic processes which take place in a finite proper time [108],
its process lifetime. Consequently, Carnot’s results, if analyzed by using the entropy generation approach can be explained
in the following way: the systems must use part of the heat absorbed by a thermal reservoir just to maintain their internal
processes, so they may not convert all the absorbed energy. These exergy flows are no more than the heat exchanged with
the required second thermal reservoir of a Carnot engine [121]. Consequently, the exergy flow, between the system and its
environment, results in a fundamental quantity for the analysis of the open systems.
In this paper, some considerations on the thermodynamics of open systems will be developed, based on the analysis
of interactions between the systems and their environments. The consequence is to introduce the unavailability analysis
of socio-economic systems, based on the bio-chemical engineering thermodynamics [1,2]. It will be highlighted that the
consequences of these interactions consist in the entropy generation related to mass, energy, ions and chemical flows
across the boundary of the system. Moreover, the flows will be considered a sort of communication between system and
environment. Consequently, it is easy to develop observations of the environment, so the analysis of the entropy generation
of the system can be carried out by evaluations of the effects of the interaction system-environment on the environment
itself. To do so in Section 2 the link between first and second laws with the exergy is discussed, in Section 3 some examples
of complex systems are discussed and in Section 4 the results are discussed to open new perspectives to the second law
analysis of municipality, applying it to an interesting case.

2. The first and second law in terms of exergy

In accordance with the first law of thermodynamics for open systems, any change in the energy of the system can be
expressed in terms of the transfer of
1. flows of matter across the system boundary, which bring internal, kinetic, chemical and other forms of energy;
2. heat across the system boundary;
3. performance of work developed by or on the system.
So the power developed by a system, dE /dt, results [1,6–9,13,49,93]:
dE    
= ṁi hi + ek,i + ep,i + ech,i + Q̇j − Ẇ (1)
dt i j

where h is the specific enthalpy, ek is the specific kinetic energy, ep is the specific potential energy, ech is the specific chemical
energy, ṁ is the mass flow, Q̇ is the heat power, Ẇ is the mechanical power, the suffixes i and j are related to the matter flows
across the boundary of the system, using the positive sign for the incoming flows and the negative sign for the outcoming
ones.
Any process, interaction, cycle, etc. occur in a specific time τ , which can be considered the lifetime of this phenomenon.
Consequently, during any process or interaction the energy variation of an open system results:
τ τ
dE
  
dt = ṁi hi + ek,i + ep,i + ech,i dt + (2)
 
1E = Qj − W
0 dt i 0 j
424 U. Lucia / Physica A 462 (2016) 421–430

where Q is the heat exchanged and W is the work developed. As a consequence of this energy variation, the following
entropy variation, 1S, of the system occurs:
τ τ
dS
   Qj
1S = dt = ṁi si dt + + Sg (3)
0 dt i 0 j
Tj

where dS /dt is the rate of entropy variation, T is the temperature, s is the specific entropy and Sg is the entropy variation
due to irreversibility, named entropy generation [93,108]. This last term represents the degradation of energy [1], i.e. the
energy dissipated during any process for friction, viscosity or any other cause of irreversibility.
Now, it is possible to combine the two equations by multiplying Eq. (3) per the environmental temperature T0 and
subtracting from Eq. (2), obtaining:
τ τ
d 
  
T0
(E − T0 S ) dt = ṁi hi + ek,i + ep,i + ech,i − T0 si dt +
 
1− Qj − W − T0 Sg (4)
0 dt i 0 j
Tj

which can be written in a better form by highlighting the work lost:


 τ  
T0
Wt = ṁi (ei − T0 si ) dt + 1− Qj − T0 Sg − ∆ (E + p0 V − T0 S ) (5)
i 0 j
Tj

where e = h + ek + ep + ech , Wt is the work developed in any process, while p0 is the pressure of the environment to the
system border and V is the volume of the system. If there is not any irreversibility:
Sg = 0 (6)
and the consequent maximum work results:
 τ  
T0
Wt ,max = ṁi (ei − T0 si ) dt + 1− Qj − ∆ (E + p0 V − T0 S ) . (7)
i 0 j
Tj

Following Carnot [10], the efficiency of a reversible cycle is the upper bound of the efficiency of any other heat engine
operating between the same thermostats [93], and Eq. (7) represents the maximum value of the work for such approach. But,
the result of Carnot has always been experimentally verified. Many studies [122–125] have been developed to improve the
evaluation of the efficiency of real machines in relation to the consideration that real processes occur in finite-size devices
and in finite-time, with irreversibility. But, any attempt to improve the Carnot’s results has always confirmed them both
theoretically and experimentally. So, we can state that Carnot’s result represents the general law that there exists a limit
for the conversion rate of heat into the mechanical energy and also that this limit is a fundamental characteristic of natural
processes.
But, the physical question is why any process, ideal included, cannot convert all the energy absorbed into useful energy,
and if this result is also valid for molecular machines. Now, in order to try to obtain an answer for these questions, we develop
some considerations in a general way, and then we apply the results to some examples.
First, considering Eqs. (5) and (7) we can obtain the physical meaning of the entropy generation:
Wt ,max − Wt Wλ
Sg = = (8)
T0 T0
where Wλ is the work lost for friction, viscosity and any other irreversibility and the entropy generation is the entropy
variation caused by these irreversible processes. Moreover, the relation (8) represents a mechanical way to evaluate
entropy generation, while the relation (5) represents a thermodynamic way to evaluate it. Now, introducing the following
definitions [93]:
τ
1. The flow exergy due to mass flow, Jex = 0
ṁ (e − T0 s) dt;
 
T0
2. The exergy transfer due to heat transfer, ExQ = 1 − T
Q;
3. The accumulation of non-flowexergy, B = (E + p0 V − T0 S );
4. The chemical exergy, Exch = l nl (µl − µl0 ) with n molar number;

Eqs. (5) and (7) become:


 
Wt = Jex,i + ExQ ,j − T0 Sg − 1B (9)
i j
 
Wt ,max = Jex,i + ExQ ,j − 1B (10)
i j

known as efficiency equations for the real and the ideal (no irreversibility) case. From these relations the work is expressed
in terms of exergy and exergy flows; consequently, the work lost by irreversibility can be considered as the exergy lost by
U. Lucia / Physica A 462 (2016) 421–430 425

the system, named also energy or unavailability Aλ of the system, which means that the entropy generation expresses the
unavailability of the system in relation to its environmental temperature T0 :
Wλ Aλ 1
Sg = = = (Exin − Exout − W ) (11)
T0 T0 T0
which represents the theorem of Gouy–Stodola [25–29,126] and where in means total inflow exergy and out total outflow
exergy. So, this consideration allows us to state that the entropy generation can be considered a sort of ‘exergy footprint ∆’
of the processes inside the system, into the environment. Consequently, the entropy generation can be explicitly obtained
as follows:
 τ  τ
ṁi hi + ek,i + ep,i − T0 si dt + ṅl νl gl⊕ − ex⊕
ch,l dt
   
T0 S g =
i 0 l 0

τ
 d
 
T0
− 1− Qj − Wt − (E − T0 S ) dt (12)
j
Tj 0 dt
⊕ ⊕ c
where g is the molar specific Gibbs function of reaction, exch = i=1 νi gi − exch,i is the chemical exergy, ⊕ means
 
standard conditions and ν is the stoichiometric coefficient. But, the entropy generation is always positive (in real processes)
or null (in the ideal processes), consequently:
τ τ τ
d
  
(E − T0 S ) dt ≥ ṁi hi + ek,i + ep,i − T0 si dt + ṅl νl gl⊕ − ex⊕
  
ch,l dt

0 dt i 0 l 0

 T0

− 1− Qj − W t (13)
j
Tj

which allows to state that the variation of the accumulation of non-flow exergy is always generated by the flow exergy
due to mass flows or chemical flows and reactions and by the exergy transfer due to heat transfer and by the conversion
of energy in work (mechanical flows). This statement represents a fundamental link between the exergy approach and the
constructal law [127].
As a consequence of the previous considerations, in engineering thermodynamics and thermoeconomics, the concept
of exergy result is useful to evaluate the available energy for conversion from a reservoir with a reference to the ambient
environmental temperature, representing the thermodynamic quality of the energy of a system.
Now, considering the relation (11) we can define a parameter useful to quantify the technological level of a process
related just to the unavailability, named exergy unefficiency, as:
Aλ T 0 Sg
ελ = = . (14)
Exin Exin
We propose this quantity as useful to evaluate the technological maturity of a production system or a production sector
in a country, because it allows us to obtain information on the losses of processes. The less the value of the unavailability
percentage the more the industrial process is efficient in terms of energy use.

3. Thermoeconomic application of exergy lost

In this section an application of the previous result is developed. Here we apply the usual approach of engineering
accounting, linking the prices of components to their operating parameters and to their exergetic efficiency, and pricing
not the unit mass, but the specific exergy content of material or energy, by following the approach used in Ref. [128], which
consists in subdividing the country system into the following sectors:
1. Extraction, which includes mining and quarrying, oil and natural gas, refining and processing;
2. Conversion, which comprises heat and power plants;
3. Agriculture, forestry, fishery and related industries;
4. Industry, manufacturing industry except food industry and oil refineries;
5. Transportation services;
6. Tertiary sector, services other than transportation;
7. Domestic sector, households.
All fluxes between these sectors and between the surroundings and sectors within the system are being considered; each
one of them is characterized as follows:
1. Resources: primary (fossil fuels, solar, wind, minerals, metals, geothermal, hydraulic) and secondary (products from
petroleum refining, mineral and metal working) and electric energy;
2. Natural resources: agricultural products, wood, natural fibers, livestock, fish, game;
426 U. Lucia / Physica A 462 (2016) 421–430

3. Products: products and services generated by industry, tertiary and transport sectors;
4. Trash fluxes: organic and inorganic waste materials, deposited in the environment;
5. Discharge: combustion gases, thermal discharge including radiated heat, heat and mass spread in the environment;
6. Human work.
Then, we develop the combination of exergetic and economic analysis by the definition of:
1. production cost of a system;
2. costs associated to losses;
3. performance of Sectors or elements.
The control volume on balances, here considered, is the district of Alessandria. In relation to the data available, related only
to the year 2004, we can consider only the flows of exergy from energy resources, neglect of the flows of products and
services [128]. Alessandria is an Italian district of Piedmont region, which covers a surface of 3560 km2 , with a population
of around 440,613 people, including 190 municipalities, among which the administrative center: the municipality of
Alessandria. We analyze the just the municipality of Alessandria, which covers 204 km2 with a population of 93,922 people.
We consider the following exergy flows only related to the city management in order to obtain information from the energy
management of the city administration:
1. The exergy inflow from the tertiary sector: it is distinguishing trait mainly from consumption of building heatings,
water systems and electrical appliances, from electricity you obtain low temperature heat, from fuels, which consists
of Electricity 712 TJ (85% for low temperature heat) and Fuels 559 TJ, with a total amount of 1271 TJ;
2. The exergy outflow from the tertiary sector uses: Electricity 289 TJ (low temperature heat 182 TJ, other uses 107 TJ) and
Fuels 148 TJ, with a total amount of 437 TJ;
3. The exergy inflow from the residential sector: the consumptions of this branch are mainly for residential use for
residential lightening, heating, etc.: Electricity 309 TJ and Fuels 2825 TJ, with a total amount of 3134 TJ;
4. The exergy outflow from the residential sector uses: Electricity 125 TJ (low temperature heat 79 TJ, other uses 46 TJ), and
Fuels 992 TJ, with a total amount of 1117 TJ;
5. The exergy inflow from the public transport: this sector receives in input fuel and in output mainly produces mechanical
power: Electricity 14 TJ and Fuels 29 TJ, with a total amount of 43 TJ;
6. The exergy outflow from the public transport: Electricity 0 TJ, and Fuels 10 TJ, with a total amount of 10 TJ;
7. The exergy inflow from the private transport: this sector receives in input fuel and in output mainly produces mechanical
power: Electricity 0 TJ and Fuels 2230 TJ, with a total amount of 2230 TJ;
8. The exergy outflow from the private transport: Electricity 0 TJ, and Fuels 652 TJ, with a total amount of 652 TJ;
with a total exergy inflow of 6678 TJ and exergy outflow 2216 TJ and an exergy lost of 4462 TJ. By using the relation (23) we
can evaluate the exergy unefficiency as 0.668. Now, we can consider possible policy decision of the city administration, as
follows:
1. introduce district heating: it would reduce
i. the exergy inflow for tertiary sector of electricity uses to 107 TJ for the electricity and to 0 TJ for the fuels;
ii. 90% of fuels for the exergy inflow and outflow from the residential sector;
2. improve the public transportation: it would:
i. improve, for public sector, the exergy inflow to 100 TJ and the exergy outflow to 34 TJ;
ii. decrease, for private sector, the exergy inflow to 1400 TJ and the exergy outflow to 409 TJ.
The result of these two energy management decisions is to reduce the exergy unefficiency to 0.338, with a better energy
management. We can also compare this result with an economic indicator, the equivalent primary resource value for the
work-hour EI, defined as:
Exin
EI = (15)
nh nw
where nh is the number or work hour and nw the number of workers. This quantity indicates the exergetic cost necessary to
support the work hours and to generate capital flows. For Alessandria city, the number of workers in 2004 was 21,289, while
the work hours per worker were 1819, so the indicator EI result 172 MJ/work hour when the exergy unefficiency is 0.668
and 52 MJ/work hour when the exergy unefficiency is 0.338. We can highlight how EI is reduced to one third by halving ελ .

4. Conclusions

The entropy concept and its production in non-equilibrium processes can be considered the bases of the modern
thermodynamics and statistical physics. In this paper a link between entropy generation and exergy has been reviewed.
This paper has suggested a review of the exergy concept developing some considerations on the fundamental link
between entropy, exergy and fluxes. It has been suggested to introduce the second law analysis to complex systems, in
particular to natural systems. This approach follows the general idea of Georgescu-Roegen [129] of using the classical
U. Lucia / Physica A 462 (2016) 421–430 427

thermodynamics to economics, but the approach here developed is completely different, because here we consider the
interaction between the system and its surrounding based on the analysis of fluxes [2,3,110,130].
Exergy is a fundamental quantity to obtain information on the useful work obtainable in a process. The related analyses of
irreversibility are important both in the design and development of the industrial processes and in the analysis of the socio-
economic systems, municipalities included. Consequently, the link between entropy and exergy is also the link between
the econophysics and the irreversible bio-chemical engineering thermodynamics. Last, this link holds to the fundamental
role of fluxes and to the related exergy exchanged in the interaction between the system and its surrounding. The result
obtained is a thermodynamic approach to the analysis of the unavailability of the economic, productive or social systems.
Here, unavailability is considered what the system cannot use in relation to any internal process. This quantity result is
interesting also as a support to public manager for economic decisions. The use of the exergy unefficiency could have
been useful to highlight the possible difficulties and which were the sectors to be improved well in advance. Indeed, as
a consequence of the analysis developed, the public transport results as the sector with more opportunity of development.
The use of this indicator could have avoided the economic difficulties, by pointing out the best and useful sector of the public
investments.
We must highlight that exergy is a quantity evaluated in relation to a reference environment. Its reference temperature
is traditionally taken to be T0 = 298.15 K (25 °C), but it would be more realistic to use a time and space variation of
temperature, even if it could not allow the comparison of engineering achievements efficiency. So, a common temperature
is useful to develop comparisons, but it is not correct from a fundamental point of view. Reference pressure is traditionally
taken to be p0 = 101 325 Pa (1 atm), but, again, it would be more realistic to use a time and space variation of pressure.
Consequently, in order to develop an econophysical analysis and economic comparisons, the reference temperature and
pressure can be considered, without any error in the evaluations.
Last, starting from the previous comments a new quantity for the econophysical analysis is introduced and used in
an application to the analysis of Alessandria city. The analysis of Alessandria is very interesting because the Alessandria
municipality was declared in financial difficulty by the Court of Auditor in 2012 [131,132]. Today, the Alessandria public
administrators have solved the economic difficulties in the last four years.
The result obtained represents an irreversible bio-thermo-economic development of fundamental results obtained in
engineering and thermoeconomics [133–148]. Our result highlights that in some sectors there exists a connection. The public
and private transportation are interconnected; indeed, if the Municipality improves the public transportation and the cycle
ways, it follows that people are supported in their moving around the city with the consequence of a lower air pollution.
Moreover, some services, like district heating, can support people in their house services, with a global improvement of the
efficiency, and decreasing of air pollution, too. On the other hand, other sectors are independent of one another; indeed,
transportation and house heating have no connection. The fundamental result consists in a proof of the fundamental role
of the municipality in improving the efficiency of the system by developing the public services. Indeed, the consequence of
any public service consists in the increase of the efficiency of the energy management and of decreasing the pollution in
the city. The related consequence is also an increase of the local wealth of the people, and of the municipality itself, which
allows the wealth of the citizens. Indeed, it emerges that the real basis of the wealth is not the continuous growth, but the
maintaining of the purchasing power of people. Comparing this statement with the life sciences, it is possible to highlight
how a continuous growing cell system inside the human body is named cancer. So, the continuous increase of the economic
systems is not important, but it is fundamental to guarantee citizens services and wealth for well-living. The approach here
introduced results useful to support municipality in their decision making.

References

[1] R.C. Tolman, P.C. Fine, On the irreversible production of entropy, Rev. Modern Phys. 20 (1948) 51–77.
[2] U. Lucia, Bioengineering thermodynamics: An engineering science for thermodynamics of biosystems, Int. J. Thermodyn. 18 (2015) 254–265.
[3] U. Lucia, Bioengineering thermodynamics of biological cells, Theor. Biol. Med. Model. 12 (1–16) (2015) 29.
[4] T. De Donder, P. Van Rysselberghe, Affinity, Standford University Press, Menlo Park (USA), 1936.
[5] H.C. Weber, Thermodynamics for Chemical Engineers, John Wiley & Sons, New York, 1939.
[6] J.H. Keenan, Thermodynamics, John Wiley & Sons, New York, 1941.
[7] C. Eckart, The thermodynamics of irreversible processes. I. The simple fluid, Phys. Rev. 58 (1940) 267–269.
[8] C. Eckart, The thermodynamics of irreversible processes. II. Fluid mixtures, Phys. Rev. 58 (1940) 269–275.
[9] P.W. Bridgman, The Nature of Thermodynamics, The Cambridge University Press, Teddington (UK), 1941.
[10] S. Carnot, Rèflexion sur la puissance motrice du feu sur le machine a dèvelopper cette puissance, Bachelier Libraire, Paris, 1824.
[11] U. Lucia, Carnot efficiency: Why? Physica A 392 (17) (2013) 3513–3517.
[12] R. Clausius, The Mechanical Theory of Heat—with its Applications to the Steam Engine and to Physical Properties of Bodies, John van Voorst, London
(UK), 1865.
[13] B.H. Lavenda, Thermodynamics of Irreversible Processes, Dover, Mineola (USA), 1993.
[14] P. Stoneman, The Economic Analysis of Technology Policy, Clarendon Press, Oxford (UK), 1987.
[15] E. Sciubba, Exergy-based ecological indicators: a necessary tools for resource use assessment studies, Termotechnica 2 (2009) 11–25.
[16] I. Dincer, Y.A. Cengel, Energy, entropy and exergy concepts and their roles in thermal engineering, Entropy 3 (2001) 116–149.
[17] E. Sciubba, G. Wall, A brief commented history of exergy. From the beginnings to 2004, Int. J. Thermodyn. 10 (2007) 1–26.
[18] B.P.E. Clapeyron, Sur la Puissance de le Chaleur Paris. 1834, Dover Publishing, New York (USA), 1960, (In French). English Edition: E. Mendoza
Reflections on the motive power of fire by S. Carnot and other papers.
[19] W.J.M. Rankine, Miscellaneous Scientific Papers, Charles Griffin and Company, London (UK), 1881.
[20] W. Thomson (Lord Kelvin), Collected Papers in Physics and Engineering, Cambridge University Press, Cambridge (UK), 1912.
428 U. Lucia / Physica A 462 (2016) 421–430

[21] J.W. Gibbs, A method of geometrical representation of the thermodynamic properties of substances by means of surfaces, Trans. Conn. Acad. Arts.
Sci. 3 (1876) 108–248.
[22] P.G. Tait, Sketch of Thermodynamics, Edinburgh University Press, Edinburgh (UK), 1868.
[23] P. Duhem, Sur la stabilite de l’equilibre en thermodynamique et les recherches de J. W. Gibbs au sujet de ce probleme, Procés-verbaux des séances
de la Soc. des Sciences, Phys. et Natur. de Bordeaux (1904) 112–130 (in French).
[24] C. Carathéodory, Untersuchungen uber die grundlage der thermodynamik, Math. Ann. 67 (1909) 355–386 (in German).
[25] G. Gouy, Sur les transformation et l’équilibre en thermodynamique, C. R. Acad. Sci. Paris 108 (1889) 507–509.
[26] G. Gouy, Sur l’énergie utilisable, J. Physique 8 (1889) 501–518.
[27] A. Stodola, Steam and Gas Turbines, Vol. 2 (L.C. Loewenstein Trans.) Van Nostrand, New York (USA), 1945.
[28] P. Duhem, Sur les transformations et l’équilibre en thermodynamique. Note de M.P. Duhem, C. R. Acad. Sci. Paris 108 (1889) 666–667.
[29] G. Gouy, Sur l’énergie utilisable et le potentiel thermodynamique. Note de M. Gouy, C. R. Acad. Sci. Paris 108 (1889) 794.
[30] E. Jouget, Remarques sur la thermodynamique des machines motrices, Rev. Mec. 19 (1906) 41 (in French).
[31] E. Jouget, Le théoreme de M. Gouy et quelques-unes de ses applications, Rev. Méc. 20 (1907) 213–238 (in French).
[32] E. Jouget, Theorie des Moteurs Thermiques, Gauthier-Villars, Paris, 1909 (in French).
[33] G.A. Goodenough, Principles of Thermodynamics, H. Holt Pub., New York, 1911.
[34] W.L. De Baufre, Analysis of power-plant performance based on the second law of thermodynamics, Mech. Eng. ASME 47 (1925) 426–428.
[35] M. Born, Kritische betrachtungen zur traditionellen darstellung der thermodynamik, Phys. Z. 22 (1921) 218–224 (in German).
[36] M. Born, Kritische betrachtungen zur traditionellen darstellung der thermodynamik, Phys. Z. 22 (1921) 249–254 (in German).
[37] M. Born, Kritische betrachtungen zur traditionellen darstellung der thermodynamik, Phys. Z. 22 (1921) 282–286 (in German).
[38] G. Darrieus, Définition du rendement thermodynamique des turbines a vapeur, Rev. Gen. Electr. 27 (1930) 963–968 (in French).
[39] G. Darrieus, L’evolution des centrales thermiques et la notation d’energie utilisable, Sci. Indust. 15 (1931) 122–126 (in French).
[40] G.V. Lerberghe, P. Glansdorff, Le rendement maximum des machines thermiques, Publ. Ass. Ing. Mines 42 (1932) 365–418 (in French).
[41] J.C. Maxwell, Theory of Heat, first ed., Longmans Green and Co, London (UK), 1871.
[42] H. Lorenz, Beitrage zur beurteilung von kuhlmaschinen, Zeit. Vereins Deutsch Ing. 38 (1894) 62–68 (in German).
[43] H. Lorenz, Beitrage zur beurteilung von kuhlmaschinen, Zeit. Vereins Deutsch Ing. 38 (1894) 98–102 (in German).
[44] H. Lorenz, Beitrage zur beurteilung von kuhlmaschinen, Zeit. Vereins Deutsch Ing. 38 (1894) 124–130 (in German).
[45] H. Lorenz, Die beurteilung der dampfkessel, Zeit. Vereins Deutsch Ing. 38 (1894) 1450–1452 (in German).
[46] J.H. Keenan, A steam chart for second law analysis, Mech. Eng. ASME 54 (1932) 195–204.
[47] J.H. Keenan, A.H. Shapiro, History and exposition of the laws of thermodynamics, Mech. Eng. ASME 69 (1947) 915–921.
[48] J.H. Keenan, Availability and irreversibility in thermodynamics, Br. J. Appl. Phys. 2 (1951) 183–192.
[49] J.H. Keenan, G.N. Hatsopoulos, Principles of General Thermodynamics, John Wiley & Sons, New York, 1965.
[50] J.H. Keenan, E.P. Gyftopoulos, G.N. Hatsopoulos, The fuel shortage and thermodynamics - the entropy crisis, in: M. Macrakis (Ed.), Proc. MIT Energy
Conf, the MIT Press, Cambridge (USA), 1974.
[51] E.H.F. Bosnjakovic, Solar collectors as energy converters, in: Studies in Heat Transfer, Hemisphere Publ. Corp, Washington (USA), 1979, pp. 331–381.
[52] R. Emden, Why do we have winter heating, Nature (1938) 141.
[53] G. Wall, The exergy conversion in the society of Ghana, in: The 1st International Conference on Energy and Community Development. Athens, 10–15
July, 1978.
[54] G. Wall, The Use of Natural Resources—A Physical Approach Report, SMR51–24, International Centre for Theoretical Physics (ICTP), Trieste, Italy,
1978.
[55] G. Wall, Exergy—a useful concept (Ph. D. thesis), Physical Resource Theory Group, Chalmers University of Technology, S-412 96 Göteborg, Sweden,
1986.
[56] G. Wall, Thermoeconomic optimization of a heat pump system, Energy 11 (1986) 957–967.
[57] G. Wall, Exergy conversion in the swedish society, Resour. Energy 9 (1987) 55–73.
[58] G. Wall, Exergy flows in a pulp and paper mill, in a steel plant and rolling mill, in: Proc. IV International Symposium on Second Law Analysis of
Thermal Systems, Rome, 25–29 May 1987, pp. 131–140.
[59] G. Wall, Thermoeconomic optimization of a single stage heat pump system, in: Proc. IV International Symposium on Second Law Analysis of Thermal
Systems, Rome, 25–29 May 1987, pp. 89–95.
[60] G. Wall, Exergy flows in industrial processes, Energy 13 (1988) 197–208.
[61] G. Wall, Exergy conversion in the Japanese society, Energy 15 (1990) 435–444.
[62] G. Wall, Exergy needs to maintain real systems near ambient conditions, in: S.S. Stecco, M.J. Moran (Eds.), Florence World Energy Research
Symposium, 28 May-1 June 1990, Florence, Italy. A Future for Energy, Pergamon, New York, 1990, pp. 261–270.
[63] G. Wall, On the optimization of refrigeration machinery, Int. J. Refrig. 14 (1991) 336–340.
[64] G. Wall, Energy, society and morals, J. Hum. Values 3 (1997) 193–206.
[65] E. Sciubba, Beyond thermoeconomics? the concept of extended exergy accounting and its application to the analysis and design of thermal systems,
Exergy 1 (2001) 68–84.
[66] E. Sciubba, Modelling the energetic and exergetic self-sustainability of societies with different structures, J. Eng. Res. Technol. 117 (1995) 75–86.
[67] E. Sciubba, Artificial intelligence applications to the synthesis of thermal processes, in: Proc. Conf. on 2-nd Law Analysis of Energy Systems, Roma
1995.
[68] E. Sciubba, Beyond thermoeconomics? The concept of extended-exergy accounting and its application to the analysis and design of thermal systems,
Exergy Int. J. 1 (2000) 68–84.
[69] E. Sciubba, Exergo-economics: thermodynamic foundation for a more rational resource use, Int. J. Energy Res. 29 (2005) 613–636.
[70] E. Sciubba, Exergy as a direct measure of environmental impact, (S.M. Aceves, S. Garimella, R. Peterson, Eds) Proc. ASME-AES 39, 1999, pp. 573–581.
[71] U. Lucia, E. Sciubba, From Lotka to the entropy generation approach, Physica A 392 (2013) 3634–3639.
[72] Y.A. Cengel, M.A. Boles, Thermodynamics: An Engineering Approach, fourth ed., McGraw-Hill, New York (USA), 2001.
[73] E.G. Gyftopoulos, G.P. Beretta, Thermodynamics. Foundations and Applications, Dover Publications, Mineola (USA), 2005.
[74] I. Prigogine, Modération et transformations irréversibles des systèmes ouverts, Bull. Cl. Sci. Acad. R. Belg. 31 (1945) 600–606.
[75] I. Prigogine, Étude Thermodynamique des Phenomènes Irréversibles, Desoer, Liège, Belgium, 1947.
[76] I. Prigogine, Introduction to Thermodynamics of Irreversible Processes, Thomas, Springfields (USA), 1955.
[77] I. Prigogine, Structure, dissipation and life, in: M. Marois (Ed.), Theoretical Physics and Biology, North Holland Pub. Co., Amsterdam (The Netherlands),
1969, pp. 23–52.
[78] P. Glansdorff, I. Prigogine, Thermodynamic Theory of Structure, Stability, and Fluctuations, Wiley-Interscience, New York (USA), 1971.
[79] G. Nicolis, Stability and dissipative structures in open systems far from equilibrium, in: I. Progogine, S.A. Rice (Eds.), in: Advances in Chemical Physics,
vol. XIX, Wiley-Interscience, New York (USA), 1971, pp. 209–324.
[80] D. Kondepudi, I. Prigogine, Modern Thermodynamics, From Heat Engine to Dissipative Structures, Wiley, New York (USA), 1999.
[81] I. Prigogine, Time, structure, and fluctuations, Science 201 (4358) (2001) 777–785.
[82] H. Ziegler, Thermodynamik und rheologische probleme, Ing. Arch. 25 (1957) 58–70.
[83] H. Ziegler, Chemical reactions and the principle of maximal rate of entropy production, Appl. Math. Phys. ZAMP 34 (1928) 832–844.
[84] H. Ziegler, Introduction to Thermomechanics, North-Holland, Amsterdam (The Netherlands), 1983.
[85] H. Ziegler, C. Wehrli, On a principle of maximal rate of entropy production, J. Non-Equilib. Thermodyn. 12 (1987) 229–243.
[86] A. Bejan, Shape and Structure, from Engineering to Nature, Cambridge University Press, Cambridge (UK), 2000.
[87] A. Bejan, Entropy Generation through Heat and Mass Fluid Flow, Wiley & Sons, New York (USA), 1982.
U. Lucia / Physica A 462 (2016) 421–430 429

[88] A. Bejan, Entropy Generation Minimization, CRC Press, Boca Raton (USA), 1995.
[89] A. Bejan, Method of entropy generation minimization, or modeling and optimization based on combined heat transfer and thermodynamics, Rev.
Gen. Therm. 35 (1996) 637–646.
[90] A. Bejan, A. Tsatsatronis, M. Moran, Thermal Design and Optimization, Wiley & Sons, New York (USA), 1996.
[91] A. Bejan, Convection Heat Transfer, Wiley, New York (USA), 2000.
[92] A. Bejan, S. Lorente, The constructal law and the thermodynamics of flow systems with configuration, Int. J. Heat Mass Transfer 47 (2004) 3203–3214.
[93] A. Bejan, S. Lorente, The constructal law of design and evolution in nature, Philos. Trans. R. Soc. B 365 (2010) 1335–1347.
[94] A. Bejan, Advanced Engineering Thermodynamics, John Wiley, New York (USA), 2006.
[95] C. Tsallis, Possible generalization of Boltzmann-Gibbs statistics, J. Stat. Phys. 52 (1988) 479–487.
[96] C. Tsallis, Some comments on Boltzmann-Gibbs statistical mechanics, Chaos Solitons Fractals 6 (1995) 539–559.
[97] R. Dewar, Information theory explanation of the fluctuation theorem, maximum entropy production and self-organized criticality in non-equilibrium
stationary states, J. Phys. A: Math. Gen. 36 (2003) 631–641.
[98] Q.A. Wang, Incomplete information and fractal phase space, Chaos Solitons Fractals 19 (3) (2004) 639–644.
[99] V. García-Morales, J. Pellicer, Microcanonical foundation of non extensivity and generalized thermostatistics based on the fractality of the phase
space, Physica A 361 (2006) 161–172.
[100] K.G. Denbigh, Note on entropy, disorder and disorganization, British J. Philos. Sci. 40 (1989) 323–332.
[101] D.G. Denbigh, The many faces of irreversibility, British J. Philos. Sci. 40 (1989) 501–518.
[102] V. Bertola, E. Cafaro, A critical analysis of minimum entropy production theorem and its application to heat and fluid flow, Int. J. Heat Mass Transfer
51 (2008) 1907–1912.
[103] U. Lucia, Irreversibility entropy variation and the problem of the trend to equilibrium, Physica A 376 (2007) 289–292.
[104] U. Lucia, Statistical approach of the irreversible entropy variation, Physica A 387 (14) (2008) 3454–3460.
[105] U. Lucia, Probability, ergodicity, irreversibility and dynamical systems, Proc. R. Soc. Lond. Ser. A Math. Phys. Eng. Sci. 464 (2008) 1089–1184.
[106] U. Lucia, Irreversibility, entropy and incomplete information, Physica A 388 (2009) 4025–4033.
[107] U. Lucia, Maximum entropy generation and κ−exponential model, Physica A 389 (2010) 4558–4563.
[108] U. Lucia, Thermodynamic paths and stochastic order in open systems, Physica A 392 (18) (2013) 3912–3919.
[109] U. Lucia, Stationary open systems: a brief review on contemporary theories on irreversibility, Physica A 392 (5) (2013) 1051–1062.
[110] U. Lucia, Exergy flows as bases of constructal law, Physica A 392 (24) (2013) 6284–6287.
[111] V. Sharma, A. Annila, Natural process—Natural selection, Biophys. Chem. 127 (2007) 123–128.
[112] V. Sharma, V.R.I. Kaila, A. Annila, A protein folding as an evolutionary process, Physica A 388 (2009) 851–862.
[113] A. Annila, S. Salthe, Physical foundations of evolutionary theory, J. Non-Equilib. Thermodyn. 35 (2010) 301–321.
[114] A. Annila, All in action, Entropy 12 (2010) 2333–2358.
[115] A. Annila, S. Salthe, Cultural naturalism, Entropy 12 (2010) 1325–1343.
[116] T. Grönholm, A. Annila, Natural distribution, Math. Biosci. 210 (2007) 659–667.
[117] V.R.I. Kaila, A. Annila, Natural selection for least action, Proc. R. Soc. A 464 (2009) 3055–3070.
[118] P. Tuisku, T.K. Pernu, A. Annila, In the light of time, Proc. R. Soc. A 465 (2009) 1173–1198.
[119] A. Annila, The 2nd law of thermodynamics delineates dispersal of energy, Int. Rev. Phys. 4 (2010) 29–34.
[120] T. Hartonen, A. Annila, Natural networks as thermodynamic systems, Complexity 18 (2012) 53–62.
[121] T.K. Pernu, A. Annila, Natural emergence, Complexity 17 (2012) 44–47.
[122] U. Lucia, Entropy and exergy in irreversible renewable energy systems, Renewable Sustainable Energy Rev. 20 (2013) 559–564.
[123] A. Durmayaz, O.S. Sogut, B. Sahin, H. Yavuz, Optimization of thermal systems based on finite-time thermodynamics and thermoeconomics, Prog.
Energy Combust. Sci. 30 (2004) 175–217.
[124] F.L. Curzon, B. Ahlborn, Efficiency of a Carnot engine at maximum power output, Amer. J. Phys. 43 (1975) 22–24.
[125] C. Wu, L. Chen, J. Chen (Eds.), Recent Advances in Finite Time Thermodynamics, Nova Science Publishers, New York (USA), 1999.
[126] R.S. Berry, V. Kazakov, S. Sieniutycz, Z. Szwast, A.M. Tsirlin, Thermodynamic Optimization of Finite-Time Processes, Wiley, New York (USA), 2000.
[127] R.H. Haywood, A critical review of the theorems of thermodynamics availability, with concise formulations, J. Mech. Eng. Sci. 16 (1974) 160–173.
[128] E. Sciubba, S. Bastionani, E. Tiezzi, Energy and extended exergy accounting of very large complex system with an application to the province of siena,
Int. J. Environ. Manage. 86 (2008) 372–382.
[129] N. Georgescu-Roegen, The Entropy Law and the Economic Process, Harvard University Press, Cambridge (USA), 1971.
[130] U. Lucia, Considerations on non equilibrium thermodynamics of interactions, Physica A 447 (2016) 314–319.
[131] Linee di Pianificazione Energetica della Provincia di Alessandria, Energia e Territorio, Maggio 2007 (in Italian).
[132] http://www.comune.alessandria.it/flex/cm/pages/ServeBLOB.php/L/IT/IDPagina/8424.
[133] T.H. Ko, K. Ting, Optimal Reynolds number for the fully developed laminar forced convection in a helical coiled tube, Energy 31 (2006) 2142–2152.
[134] M.R. Hajmohammadi, G. Lorenzini, O. Joneydi Shariatzadeh, C. Biserni, Evolution in the design of V-shaped highly conductive pathways embedded
in a heat-generating piece, J. Heat Transfer 137 (6) (2015) 061001.
[135] M.R. Hajmohammadi, S. Poozesh, A. Campo, S.S. Nourazar, Valuable reconsideration in the constructal design of cavities, Energy Convers. Manage.
66 (2013) 33–40.
[136] M.R. Hajmohammadi, A. Campo, S.S. Nourazar, A.M. Ostad, Improvement of forced convection cooling due to the attachment of heat sources to a
conducting thick plate, Trans. ASME, J. Heat Transfer 135 (2013) 124504-1.
[137] M.R. Hajmohammadi, S. Poozesh, M. Rahmani, A. Campo, Heat transfer improvement due to the imposition of non-uniform wall heating for in-tube
laminar forced convection, Appl. Therm. Eng. 61 (2013) 268–277.
[138] M.R. Hajmohammadi, E. Shirani, M.R. Salimpour, A. Campo, Constructal placement of unequal heat sources on a plate cooled by laminar forced
convection, Int. J. Therm. Sci. 60 (2012) 13–22.
[139] M.R. Hajmohammadi, O.J. Shariatzadeh, M. Moulod, S.S. Nourazar, Phi and Psi shaped conductive routes for improved cooling in a heat generating
piece, Int. J. Therm. Sci. 77 (2014) 66–74.
[140] M.R. Hajmohammadi, H. Eskandari, M. Saffar-Avval, A. Campo, A new configuration of bend tubes for compound optimization of heat and fluid flow,
Energy 62 (2013) 418–424.
[141] M.R. Hajmohammadi, S. Poozesh, S.S. Nourazar, Constructal design of multiple heat sources in a square-shaped fin, J. Process Mech. Eng. 226 (2012)
324–336.
[142] M.R. Hajmohammadi, S. Poozesh, R. Hosseini, Radiation effect on constructal design analysis of a T-Y-shaped assembly of fins, J. Therm. Sci. Technol.
7 (2012) 677–692.
[143] M.R. Hajmohammadi, V.A. Abianeh, M. Moezzinajafabadi, M. Daneshi, Fork-shaped highly conductive pathways for maximum cooling in a heat
generating piece, Appl. Therm. Eng. 61 (2013) 228–235.
[144] M.R. Hajmohammadi, M. Rahmani, A. Campo, O. Joneydi Shariatzadeh, Optimal design of unequal heat flux elements for optimized heat transfer
inside a rectangular duct, Energy 68 (2014) 609–616.
[145] M.R. Hajmohammadi, S.S. Nourazar, A. Campo, S. Poozesh, Optimal discrete distribution of heat flux elements for in-tube laminar forced convection,
Int. J. Heat Fluid Flow 40 (2013) 89–96.
430 U. Lucia / Physica A 462 (2016) 421–430

[146] M.R. Hajmohammadi, M. Reza Salimpour, M. Saber, A. Campo, Detailed analysis for the cooling performance enhancement of a heat source under a
thick plate, Energy Convers. Manage. 76 (2013) 691–700.
[147] H. Najafi, B. Najafi, P. Hoseinpoori, Energy and cost optimization of a plate and fin heat exchanger using genetic algorithm, Appl. Therm. Eng. 31
(2011) 1839–1847.
[148] A. Houberechts, Thermodynamique technique, vol. 1–2, Vander, Librairie Universitaire, Louvain (France), 1962 (in French).

You might also like