Sunteți pe pagina 1din 21

Energy 25 (2000) 1097–1117

www.elsevier.com/locate/energy

On the destruction of availability (exergy) due to


combustion processes — with specific application to
internal-combustion engines
*
Jerald A. Caton
Texas A&M University, Department of Mechanical Engineering, College Station, TX 77843-3123, USA
Received 9 August 1999

Abstract

The destruction of availability (exergy) during combustion processes is examined for an adiabatic, con-
stant volume system. This is an analytical examination and did not involve experimental measurements.
The fraction of the fuel’s availability that is destroyed due to the irreversible processes is obtained as a
function of temperature, pressure, and equivalence ratio for octane–air mixtures. In general, the destruction
of the fuel’s available energy due to the combustion process decreases for operation at higher temperatures.
In addition, the effect of equivalence ratio on the destruction of availability is significant and depends on
the particular operating conditions. Specifically, for the conditions of this study, the destroyed availability
due to the combustion process ranged between about 5 and 25% of the original reactant availability. The
implications of these results to combustion processes in internal combustion engines are described.  2000
Elsevier Science Ltd. All rights reserved.

1. Introduction

Combustion processes are known to be irreversible actions that are responsible for the destruc-
tion of availability (also known as exergy) [1–4]. The specific causes of this destruction are
examined in this paper, and the implications to internal-combustion engines are discussed.
Although an extensive literature exists concerning the destruction of availability, only a limited
amount of work has been reported on the specific details for combustion processes (e.g., see
Dunbar and Lior [4]). The insight provided by these specific details is invaluable in understanding
the overall thermodynamics of combustion.

* Fax: 979-862-2418.
E-mail address: jcaton@mengr.tamu.edu (J.A. Caton).

0360-5442/00/$ - see front matter  2000 Elsevier Science Ltd. All rights reserved.
PII: S 0 3 6 0 - 5 4 4 2 ( 0 0 ) 0 0 0 3 4 - 7
1098 J.A. Caton / Energy 25 (2000) 1097–1117

Availability is a direct consequence of the second law of thermodynamics which is a rich and
powerful statement of related physical observations with a wide range of implications with respect
to engineering design and operation of energy conversion systems. For example, the second law
can be used to determine the direction of processes, to establish the conditions of equilibrium, to
specify the maximum possible performance of thermal systems, and to identify those aspects of
processes that are detrimental to overall performance. Relative to combustion processes, the
second law of thermodynamics provides a framework which leads to a more thorough understand-
ing of the energy conversion process, provides a quantitative measure of the capability to produce
useful work, and specifies the magnitudes of the effects of operating parameters on achieving the
goals of high performance and high efficiency devices.
The objective of the current work was to examine in detail the destruction of availability due
to the combustion process. To isolate the effects of combustion on availability, a simplified case
was selected to study. Specifically, the combustion process in an adiabatic, constant volume com-
bustion system was studied. The operating conditions (such as temperatures, pressures and equiv-
alence ratios) selected for this study were representative of internal-combustion engines1. Finally,
the implications of these findings to internal-combustion engines will be described.
The next subsections will review the concept of availability and briefly describe previous work.
This will be followed by sections on the analytical approach, results and discussion, and summary
and conclusions.

1.1. Availability

Availability, also called exergy or essergy (essence of energy) by a number of authors [1–4],
is a thermodynamic property of a system that provides a measure for the maximum useful work
that a given system may attain as the system is allowed to reversibly transition to a thermodynamic
state which is in equilibrium with its environment. In general, the processes of interest are the
thermal, mechanical and chemical processes. An example of the thermal aspect of availability is
a case where the system temperature is above the environmental temperature. By utilizing an ideal
heat engine (such as a Carnot engine), the available energy from the system could be converted to
work until the system temperature equaled the environmental temperature (the remaining energy
is, therefore, the unavailable portion of the energy).
An example of the mechanical aspect of availability is a system which is at a pressure above
the environment. By utilizing an ideal expansion device (such as an ideal turbine), the energy of
the system could be converted to work until the system pressure equaled the environmental press-
ure.
A final consideration is the chemical2 aspect of availability. This aspect considers the useful
energy due to the concentration differences of the various species relative to the related concen-

1
Since the current study is generic, the findings may have application for a number of IC engines including spark–ignition,
compression–ignition, and perhaps even gas turbines. The selection of octane for the fuel, and some of the operating conditions are
probably more applicable to spark–ignition engines, but the general results will have application to other engine types.
2
The chemical aspect of availability by convention refers to the concentration differences between the species in the system and
in the environment. The (chemical) fuel energy is included in the availability terms since the total (chemical and sensible) energy is
used for the thermodynamic properties such as internal energy and enthalpy.
J.A. Caton / Energy 25 (2000) 1097–1117 1099

trations in the environment. In principle, the species of the system could do work as they reach
concentration equilibrium with the same species in the environment. To obtain work from this
concentration (or partial pressure) potential would require appropriate semi-permeable membranes
and highly efficient expansion devices. For species in the system that do not exist in the environ-
ment, these species could be allowed to react such that they form species that do exist in the
environment. Any net reaction energy could be used to produce work [1–4].
The consideration of the species concentration (chemical) component of availability is often
neglected (particularly when considering mobile engine applications) due to the practical difficult-
ies of implementing such a system and the relatively small amounts of work produced. The current
study will also neglect this contribution. Other authors have also recommended neglecting the
species concentration aspect of availability for engine applications [5,6].

1.2. Previous work

The concept of the use of the second law of thermodynamics for performance evaluations
appears to go back to the founding fathers of classical thermodynamics. Hayward [7] includes a
brief historical review of the concept of availability and the related second law topics.
A number of more modern references exist on the subject of availability analysis for engineering
systems [1,8,9]. Dunbar and Lior [4] addressed the explicit subtopic of the destruction of avail-
ability by combustion processes by considering an adiabatic, constant pressure combustion pro-
cess. This would be applicable to combustion processes in furnaces, gas turbines, and other such
devices. The current work is directed at constant volume combustion processes which would be
more applicable to internal combustion engines.
Many engineering systems have been studied using analyses based on the second law. These
previous investigations have studied power plants, cryogenic refrigeration systems, gas turbine
engines, and a number of other devices [1,8,9]. A review of such studies relating to internal-
combustion engines is available elsewhere [10]. The majority of these studies concerning internal-
combustion engines are relatively recent, with the majority of them beginning in the mid-1980s
(e.g., [6,11–15]).

2. Analytical approach

To isolate the effects of combustion on the destruction of availability, the case of a constant
volume, adiabatic combustion system will be examined. This system (constant volume, adiabatic
combustion) has been chosen since any change in availability can be attributed to the combustion
process. This is because other causes of availability changes are not possible. Since the process
is adiabatic, no availability is transferred due to heat transfer. Since the volume is closed, no
availability is transferred due to flows. Since the volume is constant, no boundary work is possible
(and no other form of work is allowed). Therefore, any changes in the availability during a change
of state may be attributed solely to the combustion process. The analytical approach used in this
study follows from related work with engine cycle simulations [16].
1100 J.A. Caton / Energy 25 (2000) 1097–1117

2.1. Assumptions and approximations

The major assumptions and approximations used in the development include the following:

1. the thermodynamic system is the chamber contents (see Fig. 1).


2. the cylinder contents are assumed to be spatially homogeneous and to occupy one zone.
3. the fuel is assumed to be completely vaporized and mixed with the reactant air.
4. the thermodynamic properties (including pressure and temperature) are spatially uniform.
5. the composition is obtained from generally accepted algorithms [5] and the species are assumed
to obey the ideal gas equation of state.
6. the composition of the products is assumed to be that which satisfies chemical equilibrium. In
addition, for comparison, some cases use a standard “frozen” [5] composition for the products.
7. the thermodynamic properties are computed from established formulations [5] based on the
appropriate compositions.
8. the combustion efficiency was assumed to be 100% (i.e., no unburnt fuel).

2.2. First law and second law

Fig. 1 is a schematic of the adiabatic, constant volume system under consideration. For this
simple system, the first law and second law reduce to the following forms:
U1⫽U2 (1)

S2⫺S1⫽s (2)
where U1 and U2 are the initial (reactant) and final (product) internal energy, respectively, S2 is
the total entropy at the end of the process, S1 is the total entropy at the start of the process, and
s is the total entropy generated by any internal irreversibilities. For specified initial conditions,
the final product temperature may be determined (from the solution of Eq. (1)), and then, from the

Fig. 1. Schematic of the adiabatic, constant volume system.


J.A. Caton / Energy 25 (2000) 1097–1117 1101

ideal gas equation of state, the final product pressure may be determined. The solution procedure is
iterative since the final temperature and pressure are not known, and are coupled through the
equilibrium conditions. For this problem, convergence to a final set of consistent values is rapid.
Once the final thermodynamic states are determined, the second law analysis is completed.

2.3. Thermodynamic properties

The thermodynamic properties needed for solving the first law include the instantaneous specific
internal energy, specific gas constant, and molecular mass. In addition, the entropy and availability
are needed to complete the second-law analyses. The thermodynamic properties are either for
mixtures of air and fuel vapor, or for combustion products. The concentrations of the combustion
products may be “frozen” for the lower temperatures, or these concentrations may be based on
an instantaneous determination of chemical equilibrium [17] for higher temperatures. In general,
the determination of the equilibrium composition of the products is one of the more complex and
computationally intensive aspects of this analysis. Complete descriptions of the algorithms used
for determining the compositions are presented in a number of references (e.g., [5,16]).
Once the composition is determined, the individual thermodynamic properties may be determ-
ined. The properties of each species in the mixture are first determined for the given temperature
and pressure by the use of polynomial curve fits [5] to the thermodynamic data [18]. The overall
mixture properties are then determined by suitable expressions. For example, for the specific
internal energy,

u⫽ 冘 i
xiũi (3)

where u is the mixture specific internal energy, xi is the mole fraction of species i and ũi is the
specific internal energy for species i on a per mole basis. For entropy, a similar expression is
used. The specific entropy for each species is

s⫽⫺Ruln 冉冊冘
p
po

i
xi(s̃oi ⫺Ruln xi) (4)

where s is the mixture specific entropy, Ru is the universal gas constant, p is the mixture pressure,
po is the reference pressure, and s̃oi is the temperature dependent portion of the specific entropy
for species i on a per mole basis. Complete descriptions of the thermodynamic properties are
provided elsewhere [5,16].

2.4. Determination of availability

Once the thermodynamic properties are known for a given set of conditions, the determination
of availability is fairly straightforward. In this development, the kinetic and potential energies are
neglected (and can be shown to be negligible). At all times, for the complete system:
a⫽(u⫺uo)⫹po(v⫺vo)⫺To(s⫺so) (5)
1102 J.A. Caton / Energy 25 (2000) 1097–1117

where a is the specific availability (or exergy for closed systems), u is the specific internal energy,
uo, vo and so are the specific internal energy, specific volume and specific entropy for the restricted
(defined below) dead state, respectively, po and To are the pressure and temperature of the dead
state, respectively, v is the specific volume, and s is the specific entropy. The dead state is defined
as the conditions of the environment at a temperature of To and a pressure of po. The dead state
is discussed in more detail below. The term, po(v⫺vo), represents the work completed against the
atmosphere at po and hence is not useful.
Availability may be destroyed by irreversibilities, and hence, is not a conserved property. These
irreversibilities are due to processes such as combustion, heat transfer through a finite temperature
difference, friction, or mixing processes. Between any end states, therefore, the change in the
availability may be related to the relevant processes of flow, heat transfer, and irreversibilities.
For the system under consideration (since there is no flow or heat transfer), the availability balance
reduces to
adest⫽a1⫺a2 (6)
where adest is the availability destroyed by irreversibilities during the process, and a1 and a2 are
the specific availability values at the start and end of the process, respectively. In addition to the
above availability balance (Eq. (6)), the availability destroyed may also be determined from the
change in entropy due to irreversibilities as follows:
adest⫽Tos/m⫽I/m (7)
where s is the change in entropy due to irreversibilities as determined for the entropy balance
described above (Eq. (2)), m is the mass of the system, and I is the availability destruction due
to irreversibilities (this is the common nomenclature) [2]. This approach (Eq. (7)) was used in
the current study to serve as an internal consistency check.

2.5. Restricted dead state

The restricted dead state is defined as the conditions of the local environment. This is chosen
since when a system is in complete equilibrium with its local environment, there is no opportunity
for the system to produce any further useful work and the system is at its “dead” state. Since the
contents of the system are not allowed to mix or react with the environment, this dead state is
not complete and is referred to as “restricted” [2].
The recommended [3] temperature and pressure for the dead state conditions are 298.15 K and
101.325 kPa (1 atm). In addition to the dead state temperature and pressure, the composition of
the dead state must be specified. The selection of the composition of the dead state is important
when considering the chemical availability component. Since the current work and most of the
previous work have neglected this component, the selection of the composition is not important
for the relative values of availability. On the other hand, the composition of the dead state does
affect the absolute values of the availability. This has an implication on the magnitude of these
values and whether they are positive or negative. This work has used standard air (listed in Table
1) for the dead state composition.
J.A. Caton / Energy 25 (2000) 1097–1117 1103

2.6. Energy and availability values for the fuel

For completing the energy and availability balances, values are needed for the energy and
availability of the fuel. For the fuel, the lower heating value (LHV) evaluated for a constant
pressure process is given by [5]:

LHV⫽⫺(⌬H)po,To⫽ 冘P

nihof,i⫺
R
nj hof,j (8)

where ⌬Hpo,To is the change in enthalpy for reactants (R) to products (P) at the reference3 pressure
(po) and temperature (To), and hof is the enthalpy of formation for the species i or j.
By definition, the fuel availability (afuel) is given by the Gibbs free energy4
afuel⫽⫺(⌬G)To,po (9)
and
(⌬G)To,po⫽{(gP)To,po⫺(gR)To,po} (10)
where ⌬Gpo,To is the change in the Gibbs free energy for reactants (R) to products (P) at the
reference conditions. For the octane fuel used in the current study,
(⌬G)To,po ⌬go298 45.789 MJ/kgfuel
⫽ ⫽ (11)
(⌬H)To,po ⌬ho298 44.514 MJ/kgfuel
where ⌬go298 is the change of the Gibbs free energy due to the reaction at the standard conditions,
and ⌬ho298 is the heat of reaction at the standard conditions. The ratio of these numbers is,
(⌬G)To,po
⫽1.0286 (12)
(⌬H)To,po

3. Results and discussion

The above relations for the availability were used to evaluate the effects of temperature, pressure
and equivalence ratio on the destruction of availability due to combustion. For these calculations,
octane–air mixtures were selected to study. Table 1 lists the important characteristics of the fuel.
Results are presented in three sections: (1) basic availability values, (2) effects of temperature
and pressure, and (3) effects of equivalence ratio. These results are followed by a section on the
implications of these findings to internal-combustion engines.

3
Note that the reference conditions for the thermodynamic data are conveniently the same as the conditions for the dead state.
4
Recall that the Gibbs free energy is given by g=h⫺Tos.
1104 J.A. Caton / Energy 25 (2000) 1097–1117

Table 1
Fuel and air parameters

Item Value Used

Fuel Octane
Molecular formula C8H18
Molecular mass (kg/kmol) 114
Lower heating value (MJ/kg) 44.51
Fuel availability (MJ/kg) 45.79
Stoichiometric AF 15.13
“Standard” air composition (mole fractions):
N2 0.7565
O2 0.2029
H2O 0.0313
Ar 0.0090
CO2 0.0003
To (K) 298.15
po (kPa) 101.325

3.1. Availability values

The specific availability was computed for octane–air mixtures for representative conditions.
Fig. 2 shows the specific availability as a function of temperature for reactants and products for
an equivalence ratio of 1.0 for pressures of 100, 1000 and 10,000 kPa. For this figure (for illus-
tration purposes), the product composition was assumed “frozen.”
Since the composition of the dead state used in these computations is standard air, the specific
availability is near zero for the reactants (which are largely air) at a temperature of 298.15 K and
a pressure of 101.325 kPa. The specific availability increases with temperature, and has a modest
dependence on pressure. The effects of pressure on the availability are discussed in more detail
below.
Fig. 3 shows specific availability as a function of temperature for products for an equivalence
ratio of 1.0 for pressures of 100, 1000 and 10,000 kPa. In this figure, the two assumptions (frozen
and equilibrium composition) for the product species were used. As shown, the assumption of
frozen composition for the product species agrees with the equilibrium composition for tempera-
tures below about 2000 K. Above 2000 K, the frozen composition assumption under-predicts the
specific availability since chemical dissociation begins to be important for temperatures above
about 2000 K for these conditions. Also, for the higher pressures, the results based on the equilib-
rium compositions are in better agreement with the results based on the frozen composition. This
is because for the higher pressures dissociation is suppressed, and the composition is closer to
the frozen composition.
Fig. 4 shows the specific availability as a function of pressure for the products for an equival-
ence ratio of 1.0 for four temperatures for the frozen composition assumption. In general, the
specific availability increases slightly with increases in pressure except for the lowest pressures
at the higher temperatures. These results aid in understanding the results in the previous figures
and explain the slight crossover in the curves for the different pressures. Increases in the values
J.A. Caton / Energy 25 (2000) 1097–1117 1105

Fig. 2. Specific availability for reactants and products as a function of temperature for an equivalence ratio of 1.0
for three pressures using the frozen composition assumption.

of availability with increases in pressure are due to decreases in the specific entropy. Furthermore,
decreases in the values of availability with increases in pressure are due to increases in the specific
volume (see Eq. (5)). The net result of these effects on the availability is reflected in Figs. 2 and 3.

3.2. Effects of temperature and pressure

First the adiabatic, constant volume combustion process will be examined for two specific sets
of initial temperatures and pressures. Then, results will be shown for a range of initial tempera-
tures. For applications to internal-combustion engines, the initial temperature and pressures should
be representative of the conditions in the cylinder near the end of compression (at the time of
ignition). Values for these initial temperatures and pressures are dependent on engine load, speed,
residuals, inlet temperature, compression ratio, and other operating and design characteristics. For
typical values of these characteristics, the initial pressure may range from about 500 to 2000 kPa,
and the initial temperature may range from about 500 to 700 K [5].
Solving Eq. (1) (the first law) provides the final product temperature and pressure for any initial
reactant temperature and pressure. With knowledge of the final product temperature and pressure,
the specific availability may be determined as a function of temperature for both the reactants
and for the products. Fig. 5 shows the specific availability as a function of temperature for the
reactants and products for the case of an equivalence ratio of 1.0, and a reactant pressure (pR) of
500 kPa. For these conditions, the final product pressure (pp) was 2974 kPa.
1106 J.A. Caton / Energy 25 (2000) 1097–1117

Fig. 3. Specific availability as a function of temperature for the products for an equivalence ratio of 1.0 for three
pressures, for both the frozen composition assumption and for equilibrium composition.

As shown in Fig. 5, the specific availability for the reactant mixture increases with increases
in reactant temperature. Two curves are shown for the product availability: one for a frozen
composition and one for an equilibrium composition. The specific availability is higher for the
composition based on the equilibrium composition as compared to the composition based on
frozen composition. As noted above, for temperatures below about 2000 K, the two compositions
yield similar values for the availability. For temperatures above about 2000 K, the effects of
dissociation on the product composition become more significant.
Also noted on Fig. 5 are two particular state points for the constant volume, adiabatic process
with the same specific internal energy (see Eq. (1)). The first state point is for the initial reactant
conditions of 500 K and 500 kPa. For this initial state, the final product state (the second state
point) for this process is 2764 K and 2974 kPa for an equivalence ratio of 1.0. These two state
points are denoted in Fig. 5 with diamond symbols, and are connected with a dashed line.
The final specific availability is less than the initial value. Even though this was an adiabatic,
constant volume process, the availability has decreased due to the combustion processes. The
difference is denoted with the symbol, aDAC5. For these two state points, the availability difference
is about 633 kJ/kgmix which results in a percentage of the total reactant availability of 20.2%. In
other words, the availability of the reactants at 500 K and 500 kPa is greater than the availability
of the products at 2764 K and 2974 kPa. This means that the reactants had the potential to do a

5
The subscript “DAC” is for Destroyed Availability by Combustion.
J.A. Caton / Energy 25 (2000) 1097–1117 1107

Fig. 4. Specific availability as a function of pressure for the products for an equivalence ratio of 1.0, for four tempera-
tures for the frozen composition assumption.

greater amount of work than the products. This potential was lost (i.e., destroyed) as the reactants
transformed irreversibly into products.
A second condition is selected now to examine a higher reactant temperature case (2500 K)6.
For this reactant temperature (2500 K) and the reactant pressure of 500 kPa, the solution of the
first law results in a final temperature of 3390 K and a final pressure of 788 kPa. Fig. 6 shows
the results for the specific availability as a function of temperature for the reactants and products
for an equivalence ratio of 1.0 for reactant and product pressures of 500 kPa and 788 kPa, respect-
ively. As with Fig. 5, this figure also shows two curves for specific availability for the products:
one curve is based on equilibrium composition, and one curve is based on frozen composition.
Again, the specific availability is higher for the composition based on the equilibrium composition
as compared to the composition based on frozen composition. For these conditions (and parti-
cularly, for these higher temperatures), the differences between the two compositions is more
significant than for the conditions of Fig. 5.
As shown in Fig. 6, two state points are denoted (by diamond symbols) for two states with
the same specific internal energy (for initial reactant conditions of 2500 K and 500 kPa) for the
constant volume, adiabatic process. These two state points are connected with a dashed line. As
shown, the specific availability does not return to the initial value. For these two state points, the

6
Note that this initial temperature (2500 K) is higher than typical initial combustion temperatures for internal-combustion engines
with today’s technology. This higher temperature was selected, however, so that a large enough range would be investigated to show
the overall trends.
1108 J.A. Caton / Energy 25 (2000) 1097–1117

Fig. 5. The specific availability as a function of temperature for the reactants and products for an equivalence ratio
of 1.0 for reactant and product pressures of 500 kPa and 2974 kPa, respectively.

availability difference is about 367 kJ/kgmix which results in a percentage of the total reactant
availability of 7.6%.
Results such as those presented in Figs. 5 and 6 were completed for a range of initial tempera-
tures. Fig. 7 shows the results for the constant volume, adiabatic combustion of an octane–air
mixture at an equivalence ratio of 1.0 for an initial reactant pressure of 500 kPa. This pressure,
500 kPa, is representative of the pressure at the start of combustion for a range of internal combus-
tion engines [5]. This figure shows the final product temperature (TP) and pressure (pP) as a
function of initial (reactant) temperature for this process. As shown, as the reactant temperature
increases, the product temperature increases in a linear fashion. The final product pressure, on
the other hand, decreases in a non-linear fashion. This decrease is due to the decreasing density
of the mixture as the reactant temperature increases and the reactant pressure remains constant.
Due to the constraint of constant mass, the final product pressure must decrease.
Most importantly, Fig. 7 also shows the percentage of the total reactant availability which is
destroyed due to the combustion process, and this percentage is denoted with the parameter rDAC
which is given by,
aDAC
rDAC⫽ ⫻100% (13)
aR
where aDAC is the specific availability destroyed by the combustion processes (per mass of
J.A. Caton / Energy 25 (2000) 1097–1117 1109

Fig. 6. The specific availability as a function of temperature for the reactants and products for an equivalence ratio
of 1.0 for reactant and product pressures of 500 kPa and 788 kPa, respectively.

mixture), and aR is the specific total availability of the reactant mixture (per mass of mixture). The
specific availability of the reactant mixture includes both a “chemical” and a sensible component:
aR⫽(afuel−mix⫹aSR) (14)
where afuel-mix is the specific availability of the fuel (per mass of mixture), and aSR is the sensible
component of the specific availability of the reactant mixture (per mass of mixture). The specific
availability of the fuel on a per mass of mixture basis is denoted by, afuel-mix, and is given by:
afuel−mix⫽af(AF⫹1) (15)
where af is the specific availability of the fuel (on a per mass of fuel basis), and AF is the mass
air–fuel ratio.
As shown, the percentage of the total reactant availability destroyed by combustion decreases
monotonically from about 20 to 7% as the reactant temperature increases from 500 to 2500 K
for these conditions. In other words, the destruction of the original availability decreases as the
temperature of the combustion process increases. This result can be seen to be a direct conse-
quence of the characteristics of the specific availability as a function of temperature (see Figs. 5
and 6). For the higher temperatures, the state points with the same internal energy have specific
availability values which are closer to each other.
The above results were for a single reactant pressure of 500 kPa. Appendix A contains
additional results for other combinations of reactant temperatures and pressures. These results
1110 J.A. Caton / Energy 25 (2000) 1097–1117

Fig. 7. The final product temperature and pressure, and the percentage of availability destroyed by combustion as a
function of initial (reactant) temperature for a constant volume, adiabatic combustion process.

illustrate further the effects of reactant temperature and pressure on the destruction of availability
due to combustion.
In general, the above results suggest that as the combustion temperature increases, the destruc-
tion of availability decreases. For the assumptions of this study, however, the destruction of avail-
ability does not attain zero even for unrealistically high temperatures. Although higher gas tem-
peratures may minimize the destruction of available energy by combustion [6,7,19], these higher
temperatures may lead to other losses of available energy in practical (actual) engineering systems.
In particular, the higher temperatures may result in higher heat transfer which will remove the
available energy. Also, if not utilized, the higher availability will be expelled with the exhaust
gases. Another consideration would be the potential for higher nitric oxide (NO) formation rates
at these higher temperatures. In any case, these high temperatures and pressures are beyond the
practical limits of today’s designs and materials for combustion devices. Further comments con-
cerning these matters are described in a subsequent section on the implications of these findings
to internal-combustion engines.

3.3. Effect of equivalence ratio

The effects of equivalence ratio are examined next. Fig. 8 shows the specific availability of
the reactants and products as a function of temperature for several equivalence ratios for a reactant
pressure of 500 kPa and a product pressure of 1000 kPa. The specific availability for the reactants
increases with increasing temperature, but as described above, the effects of equivalence ratio are
J.A. Caton / Energy 25 (2000) 1097–1117 1111

Fig. 8. The specific availability of the reactants and products as a function of temperature for several equivalence
ratios for a reactant pressure of 500 kPa and a product pressure of 1000 kPa.

modest for the reactant mixture. As shown, the two lines for equivalence ratios of 0.0 and 1.0
are not too different, and will bracket the regions of interest for the current discussion.
The bottom three lines (with symbols) in Fig. 8 represent the specific availability for the pro-
ducts for equivalence ratios of 0.2, 0.5 and 1.0. As shown, the specific availability increases with
increases in temperature and with decreases in equivalence ratio. The reason for the decrease of
specific availability with decreases in equivalence ratio is that the composition contains decreasing
amounts of combustion products. These observations will be useful in interpreting the following
results for the constant volume adiabatic combustion process.
Fig. 9 shows the specific availability as a function of temperature for the constant volume,
adiabatic combustion of octane–air mixtures for an initial reactant pressure of 500 kPa. In this
figure, the three lines for the specific availability of the products are for three separate final product
pressures (1203, 2037, and 2974 kPa) of the combustion process which are determined by the
three equivalence ratios (0.2, 0.5, and 1.0). The final product states for each of these cases are
designated by solid diamonds and the three are connected by a dashed line. This dashed line is
extended to the reactants. This dashed line is the loci of nearly constant internal energy7. Fig. 9
shows that as the equivalence ratio increases, the specific availability of the products decreases,
and the difference relative to the specific availability of the reactants increases. This means that

7
Since the initial equivalence ratio is different, the initial internal energy is slightly different for each case. For the scale of the
figure, however, these differences are not discernible.
1112 J.A. Caton / Energy 25 (2000) 1097–1117

Fig. 9. The specific availability as a function of temperature for the constant volume, adiabatic combustion of octane–
air mixtures for an initial reactant pressure of 500 kPa and three product pressures.

the absolute value of the destroyed availability, aDAC, on a per mass of mixture basis, increases
as the equivalence ratio increases.
Fig. 10 shows the product temperature (TP), and the specific availability destruction due to
combustion (aDAC-fuel), on a per mass of fuel basis, as a function of equivalence ratio for an initial
reactant temperature and pressure of 500 K and 500 kPa, respectively. As expected, the product
temperature increases as the equivalence ratio increases. On this basis, the destroyed available
energy increases rapidly as the equivalence ratio decreases from stoichiometric. This is due to
two primary causes: first, the irreversible combustion process, and second, the mixing process as
the excess air (for the lean equivalence ratios) mixes with the combustion products.

3.4. Implications to internal combustion engines

The above results from an adiabatic, constant volume combustion system have significant impli-
cations concerning internal combustion engines. For engines, the combustion process does not
occur during an adiabatic, constant volume process, but rather occurs during volume, pressure
and temperature changes with significant heat transfer. Further, the combustion process for com-
pression–ignition engines is highly non-homogeneous, whereas the combustion process for spark–
ignition engines is generally near homogeneous. Even with these differences, the reported results
have direct implications to internal-combustion engines since many aspects of the constant-volume
combustion process are relevant to engines.
J.A. Caton / Energy 25 (2000) 1097–1117 1113

Fig. 10. The availability destroyed by combustion (on a per mass of fuel basis) and the final product temperature as
a function of equivalence ratio for an initial reactant temperature and pressure of 500 K and 500 kPa, respectively.

Although the combustion process in an engine is a complex, time-varying event, some general
observations may be noted based on the above results. As shown above, the combustion process
is responsible for the destruction of a portion of the original availability. The quantitative amount
destroyed is a function of the conditions during combustion. Specifically, one of the most signifi-
cant parameters is the temperature during combustion. The higher the combustion temperatures,
the more of the original availability may be retained. Unfortunately, for realistic temperatures, a
significant amount of the original availability will be destroyed. For the conditions studied here,
about 5–25% of the original availability was destroyed due to the combustion processes. These
numbers are consistent with a number of studies [6,16] on overall engine operation. The lowest
values (苲5%) for the destruction of availability described above were obtained for conditions
which are not typical for today’s engines (i.e., for conditions with impractical high gas
temperatures).
The effect of pressure is much more modest than the effect of temperature on the destruction
of availability. Significant increases in pressure cause only a slight decrease in the destruction of
availability. The effect of equivalence ratio on the destruction of availability is significant. As the
equivalence ratio decreases from stoichiometric, the destruction of availability increases.
There are several important consequences of the above findings to internal-combustion engines.
As mentioned above, although higher combustion temperatures may retain more of the original
availability, these higher temperatures are likely to lead to higher heat losses, higher exhaust gas
1114 J.A. Caton / Energy 25 (2000) 1097–1117

temperatures, and higher nitric oxide formation rates. Higher exhaust gas temperatures suggests
that secondary recovery devices would be needed to utilize the extra available energy in the
exhaust gases [6]. Also, as noted above, the required high pressures and temperatures for the
greatest reductions of the destruction of availability are beyond the practical limits of today’s
designs and materials for internal-combustion engines.
Although combustion processes are irreversible, no practical process is known which will pro-
duce the conditions of the product state (high temperature and pressure) without this loss of
availability. In other words, to utilize the fuel’s availability in conventional heat engines, some
portion of the availability will be destroyed to realize useful high temperature, high pressure gases.
Some authors (see e.g., Clarke [19]; Flynn et al. [6]) have noted that combustion irreversibilities
could be minimized by conducting reactions near chemical equilibrium conditions. Unfortunately,
while equilibrium reactions are reversible, they are not relevant with respect to releasing the
chemical energy from the reactants to form products. In this case, therefore, the attainment of
chemical equilibrium of the combustion products does not satisfy the goal of heat release. Clarke
[19] suggests, however, that a series of carefully selected reactions could be devised which move
the fuel’s availability to a useful form while minimizing the irreversibility. This concept will
require engineering development to be compatible with today’s technology, but the potential bene-
fits would seem to warrant further work on this concept.
Alternatives to the conversion of the fuel availability to high temperature gases include the use
of fuel cells. An ideal, reversible fuel cell converts the fuel’s chemical energy directly to electrical
energy by means of low temperature, controlled chemical reactions. In this fashion, fuel cells
avoid the need for high temperature gases, and therefore, avoid the irreversible, uncontrolled
reactions associated with combustion processes. Unfortunately, at this time, the current fuel cell
technology is not competitive with the internal combustion engine due to mechanical and chemical
difficulties. Also, a fuel cell may not be able to use standard liquid fuels (such as gasoline and
diesel) as readily as today’s engines.

4. Summary and conclusion

An analysis based on the second law of thermodynamics was completed to explore the destruc-
tion of availability due to combustion processes. An adiabatic, constant volume system was studied
for a range of initial (reactant) temperatures, pressures, and equivalence ratios for octane–air
mixtures. For these conditions, the change in availability may be attributed solely to the combus-
tion process. The analysis included the computation of entropy, availability, irreversibilities, and
the related energy, entropy and availability balances. From the balances, destruction of availability
was determined for a range of parameters.
The highest availability was found to exist for the unreacted fuel. This represents a maximum
potential to perform work. When this chemical energy is transformed into thermal energy, some
portion (which depends on the final temperature) of the original availability is destroyed. The
amount of the availability that is destroyed increases for lower final temperatures.
Summary comments of this work follow:

1. For the adiabatic, constant volume combustion process studied, the availability of the products
J.A. Caton / Energy 25 (2000) 1097–1117 1115

is less than for the reactants. This decrease in availability is a loss of potential work, and is
directly related to the irreversibility of the combustion process.
2. As combustion processes are conducted at higher temperatures:
앫 the fraction of the fuel’s available energy which is destroyed decreases
앫 the realization of work from the available energy at these higher temperatures depends on
the application
앫 heat transfer may increase (which will remove available energy from the system)
앫 availability may leave the system with the exhaust gases
3. The effect of changes of the pressure during the combustion processes (all else the same) on
the availability is modest
4. The specific available energy, on a per mass of fuel basis, destroyed by combustion increases
as the equivalence ratio decreases from stoichiometric due to the irreversibilities associated
with both the combustion process and the mixing of excess air with the combustion products.
5. The implications of these results to internal-combustion engines are:
앫 to minimize destruction of the fuel’s available energy due to combustion processes, combus-
tion processes should be conducted at higher temperatures. This may require the use of
devices such as air preheat and in-cylinder materials that can withstand high temperatures.
Of course, conducting combustion at these higher temperatures produces a number of related
issues. Higher combustion temperatures may result in higher exhaust temperatures and higher
levels of availability in the exhaust. Exhaust recovery devices such as turbo-compounding

Fig. 11. The product temperature and availability destruction due to the combustion process as a function of reactant
temperature for three initial reactant pressures.
1116 J.A. Caton / Energy 25 (2000) 1097–1117

may be needed to capture this high availability. And, of course, these higher temperatures
could be a source of increased nitric oxide species.

In general, the use of second law and availability analyses provides deep insight and new perspec-
tives on the destruction of available energy due to combustion processes. Significant implications
of these results relate to destroying the potential for useful work.

Appendix A. Temperature and pressure effects

The explicit effects of reactant temperature and pressure on the destruction of availability due
to combustion will be illustrated next. Fig. 11 shows the final product temperature and the percent-
age of the total reactant availability destroyed by combustion as a function of reactant temperature
for three initial reactant pressures (500, 1000, and 1500 kPa) for an equivalence ratio of 1.0. As
expected, the product temperatures increase proportionally with increases in reactant temperature,
and the destruction of availability due to combustion decreases with increases in reactant tempera-
ture. Also, the availability destruction is not too dependent on the pressure. This weak dependence
on pressure is shown more directly in the next figure.
Fig. 12 shows the product temperature and the percentage of availability destroyed by the
combustion process as a function of reactant pressure for two initial reactant temperatures for an
equivalence ratio of 1.0. As mentioned above, the effect of reactant pressure is modest on both

Fig. 12. The product temperature and the percentage of the total reactant availability destroyed by the combustion
process as a function of reactant pressure for two initial reactant temperatures.
J.A. Caton / Energy 25 (2000) 1097–1117 1117

the final product temperature and on the availability destruction due to combustion. As shown,
the product temperature increases slightly with increases in reactant pressure, and the percentage
of availability destroyed decreases slightly with increases in reactant pressure.

References

[1] Moran MJ. Availability analysis — a guide to efficient energy use. corrected ed. The American Society of Mechan-
ical Engineers: New York, 1989.
[2] Moran MJ, Shapiro HN. Fundamentals of engineering thermodynamics. 3rd ed. Wiley: New York, 1995.
[3] Wark K Jr, Richards DE. Thermodynamics. 6th ed. McGraw-Hill: New York, 1999.
[4] Dunbar WR, Lior N. Sources of combustion irreversibility. Combust Sci Technol 1994;103:41–61.
[5] Heywood JB. Internal combustion engine fundamentals. McGraw-Hill: New York, 1988.
[6] Flynn PF, Hoag KL, Kamel MM, Primus RJ. A new perspective on diesel engine evaluation based on second
law analysis. In: SAE paper no. 840032. Society of Automotive Engineers, 1984.
[7] Hayward RW. A critical review of the theorems of thermodynamic availability, with concise formulations. J Mech
Eng Sci 1974;16(3):160–73.
[8] Sciubba E, Su TM. Second law analysis of the steam turbine power cycle: a parametric study. In: Gaggioli RA,
editor. Computer-aided engineering of energy systems. American Society of Mechanical Engineers, Advanced
Energy Systems Division, AES-vol. 2–3, December 1986.
[9] Adebiyi GA, Russell LD. Second law analysis of alternative schemes for solar-assisted air conditioning. In: Gaggi-
oli RA, editor. Computer-aided engineering of energy systems. American Society of Mechanical Engineers,
Advanced Energy Systems Division, AES-vol. 2–3, December 1986.
[10] Caton JA. A review of investigations using the second law of thermodynamics to study internal-combustion
engines. In: SAE paper no. 2000-01-1081. Society of Automotive Engineers, 2000.
[11] Patterson DJ, van Wylen G. A digital computer simulation for spark-ignited engine cycles. In: SAE progress in
technology, Digital calculations of engine cycles, 1964.
[12] Primus RJ, Flynn PF. Diagnosing the real performance impact of diesel engine design parameter variation (a primer
in the use of second law analysis). In: International Symposium on Diagnostics and Modelling of Combustion in
Reciprocating Engines, 1985:529–38.
[13] Primus RJ, Flynn PF. The assessment of losses in diesel engines using second law analysis. In: Gaggioli RA,
editor. Computer-aided engineering of energy systems, American Society of Mechanical Engineers, Advanced
Energy Systems Division. AES-vol. 2–3, December, 1986.
[14] Alkidas AC. The application of availability and energy balances to a diesel engine. Trans ASME J Eng Gas
Turbines Power 1988;110:462–9 July.
[15] Alkidas AC. The use of availability and energy balances in diesel engines. In: SAE paper no. 890822. Society
of Automotive Engineers, 1989.
[16] Caton JA. Operating characteristics of a spark-ignition engine using the second law of thermodynamics: effects
of speed and load. In: SAE paper no. 2000-01-0952. Society of Automotive Engineers, 2000.
[17] Olikara C, Borman GL. A computer program for calculating properties of equilibrium combustion products with
some applications to IC engines. In: SAE paper no. 750468. Society of Automotive Engineers, 1975.
[18] Stull DR, Prophet H, project directors. JANAF Thermochemical Tables 2nd ed. Report number NSRDS-NBS 37.
Washington, DC: National Bureau of Standards. 1971.
[19] Clarke JM. The thermodynamic cycle requirements for very high rational efficiencies. In: Proceedings of the Sixth
Thermodynamics and Fluid Convention, 1976 Apr 6–8; University of Durham. London, UK: Institute of Mechan-
ical Engineers, 1976, paper no. C53/76.

S-ar putea să vă placă și