Thermodynamic Requirements For Maximum Internal Combustion Engine Cycle Efficiency. Part 1: Optimal Combustion Strategy

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/245395230

Thermodynamic requirements for maximum internal combustion engine


cycle efficiency. Part 1: Optimal combustion strategy

Article  in  International Journal of Engine Research · December 2008


DOI: 10.1243/14680874JER01508

CITATIONS READS

64 491

3 authors, including:

Kwee-Yan Teh
Shanghai Jiao Tong University
21 PUBLICATIONS   231 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

POD Outlier Analysis: Cause and Effect View project

All content following this page was uploaded by Kwee-Yan Teh on 14 June 2016.

The user has requested enhancement of the downloaded file.


International Journal of Engine Research
http://jer.sagepub.com/

Thermodynamic requirements for maximum internal combustion engine cycle efficiency. Part 1:
Optimal combustion strategy
K-Y Teh, S L Miller and C F Edwards
International Journal of Engine Research 2008 9: 449
DOI: 10.1243/14680874JER01508

The online version of this article can be found at:


http://jer.sagepub.com/content/9/6/449

Published by:

http://www.sagepublications.com

On behalf of:

Institution of Mechanical Engineers

Additional services and information for International Journal of Engine Research can be found at:

Email Alerts: http://jer.sagepub.com/cgi/alerts

Subscriptions: http://jer.sagepub.com/subscriptions

Reprints: http://www.sagepub.com/journalsReprints.nav

Permissions: http://www.sagepub.com/journalsPermissions.nav

Citations: http://jer.sagepub.com/content/9/6/449.refs.html

>> Version of Record - Dec 1, 2008

What is This?

Downloaded from jer.sagepub.com by guest on October 11, 2013


449

Thermodynamic requirements for maximum internal


combustion engine cycle efficiency.
Part 1: optimal combustion strategy
K-Y Teh*, S L Miller, and C F Edwards
Department of Mechanical Engineering, Stanford University, Stanford, California, USA

The manuscript was accepted after revision for publication on 26 July 2008.

DOI: 10.1243/14680874JER01508

Abstract: This is the first of a two-part study that examines, from the exergy management
standpoint, the fundamental thermodynamic requirements for maximizing internal combus-
tion (IC) engine cycle efficiency. The optimal cycle is shown to comprise three distinct engine
architectural elements – reactant preparation, combustion, and work extraction from the
products – each of which can be analysed separately. This study shows, based on dynamical
system optimization, that it is the equilibrium thermodynamics (specifically, the constant-
internal energy–volume (UV) product state at the end of combustion) and not chemical kinetics
(i.e. reactions taking place during combustion) that ultimately dictates the amount of exergy
destroyed due to combustion. The strategy for minimizing this destruction term reduces to
carrying out reactions at the highest possible internal energy state – following what may be
called the ‘extreme state’ principle – so as to minimize the corresponding constant-UV entropy
change from reactants to equilibrium products. The extreme state principle remains unaltered
when system inhomogeneity (from fuel vaporization and mixing with air) and heat loss are
accounted for. Based on this optimal combustion strategy, the companion paper examines the
remaining elements of the engine cycle (reactant preparation and work extraction) so as to
improve overall cycle efficiency.

Keywords: combustion, entropy generation, exergy destruction, internal combustion engine


efficiency, optimal engine cycle, thermodynamics

1 INTRODUCTION engine efficiencies and what is possible given the


laws of thermodynamics. Furthermore, an important
Engines convert energy in a resource into work. In source of the inefficiency is still not well understood.
the case of an internal combustion (IC) engine, the
fuel–oxidizer mixture (fuel being the chemical energy
resource) that undergoes chemical reactions also
functions as the working fluid that produces expan- 1.1 Reactive engine efficiency
sion work. IC engines dominate the road transporta- The absolute maximum work extractable from a
tion sector, which accounted for 1.66103 million resource, set by the laws of thermodynamics, is
tonnes of oil equivalent – or 21 per cent – of global called the ‘resource exergy’. In order to achieve this
energy consumption (end-use) in 2004 [1]. The maximum-work limit it is required that the engine:
attendant CO2 emission was 4.66103 million tonnes, (a) processes the resource reversibly (e.g. when
or 17 per cent of CO2 emissions worldwide [2]. Yet converting reactants to products), and (b) interacts
there exists a sizable gap between state-of-the-art IC with the environment via reversible transfers (e.g.
when bringing in oxygen from the environment to
*Corresponding author: Sandia National Laboratories, Combus- oxidize the fuel). It follows that exergy is a function
tion Research Facility, PO Box 969, MS9052, Livermore, CA of the thermodynamic state of the resource as well as
94551-0969, USA. email: kteh@sandia.gov its environment.

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


450 K-Y Teh, S L Miller, and C F Edwards

For example, given some amount of fuel, the choice of LHV as the basis for defining fuel
corresponding fuel exergy (denoted X) is conversion efficiency is empirical, without consid-
X eration of the second law [3]. Therefore, it is not
X ~UzP0 V {T0 S{ mj,0 nj N impossible for maximum-work limits to exceed 100
j per cent by this efficiency measure.) In contrast,
where U, V, S, and N refer to the internal energy, today’s IC engines have peak brake thermal efficien-
volume, entropy, and moles of fuel respectively; P0, cies ranging from low 30s to mid 40s [3]. Clearly,
T0, and mj,0 are the ambient pressure, temperature, there is substantial room for efficiency improvement
and chemical potential respectively for each species and thus the potential for significant reductions in
j present in the environment; nj is the (signed) fuel resource consumption and CO2 emissions.
stoichiometric coefficient for species j in converting Before reviewing past thermodynamic analyses of
one mole of fuel to environmental products the causes for this efficiency gap – the subject of the
X next section – it is worth noting the distinction
ðFuel speciesÞ? nj ðenvironmental species jÞ between a reactive engine and a heat engine. For the
j latter system, supplied with energy Q in the form of
In addition to the fuel exergy, other maximum- heat transfer from a thermal resource at some high
work limits can be defined based on the thermo- temperature T . T0, only a fraction of the energy can
dynamic constraints imposed upon the engine to be converted to work W, such that
which the fuel is supplied. For instance, part of IC  
engine operation (between intake valve closing and T0
Max W ~ 1{ Q
exhaust valve opening) is usually modelled as an T
adiabatic process for a closed system. Under these
conditions, the second law dictates that the combus- yielding the heat engine Carnot efficiency, 1 2 T0/T,
tion product final entropy must be greater than, or at which is a function of the thermodynamic state for
best equal to, the initial entropy of the reactants in the thermal resource (temperature T) and that of the
the system. Therefore, the corresponding maximum surroundings (temperature T0). The maximum-work
engine work output is the difference in internal limits listed in Table 1 are, in a similar manner,
energy U between reactants and products at the functions of the thermodynamic state for the
same entropy S. As regards notation, the operator Dr chemical resource, i.e. the fuel, and that of the
will be used here to refer to state change from surroundings. It is important to recognize that the
reactants (subscripted ‘R’) to products (‘P’). In the
peak temperature that a combustion engine reaches
case just described, if the product gas mixture is also
during operation does not figure in the maximum-
taken to be at the same volume V as the initial
work limit calculations, since it is not a heat engine.
reactant mixture, the maximum-work limit will then
be denoted

jDr U jS,V ~jUP ðS, V Þ{UR ðS, V Þj


1.2 Exergy analyses of IC engines
As another example, fuel cells are often modelled as
isothermal, isobaric flowing systems, i.e. operating Exergy is a useful concept for evaluating the
at some fixed temperature T and pressure P. The performance of various energy systems and, through
maximum work that can be extracted from its balance equation, identifying and quantifying
the reactants supplied to such a system equals the inefficiencies in the underlying thermodynamic
Gibbs’ free energy difference |DrG|T,P between the processes [4]. From this perspective, the challenge
reactants and products at T and P. of building an efficient energy system becomes a
Table 1 lists these maximum-work limits (some- matter of managing the resource exergy supplied to
times called ‘second-law limits’) for several fuels. it to minimize exergy losses and, in the case of an
The limits are comparable with the respective fuel engine, maximize exergy outflow in the form of
lower heating values (LHVs, equal to |DrH|T,P at work. The present paper offers, from this exergy
standard-state T 5 25 uC, P 5 1 atm). The theoretical management viewpoint, a discussion as to what
limits on the efficiency of engines supplied with comprises the optimal IC engine thermodynamic
these fuels, measured as work produced per unit cycle, i.e. one that maximizes the engine efficiency.
LHV (also called the ‘thermal’ or ‘fuel-conversion Exergy analysis (sometimes referred to as ‘second-
efficiency’), range between 89 and 107 per cent. (The law analysis’) of IC engines has been the subject of

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 451

Table 1 The maximum-work limits and lower heating values for some fuels (in MJ/kg fuel)
{
*
Fuel species Chemical formula X (fuel exergy) |DrU|S,V |DrG|T,P |DrH|T,P (LHV)
Hydrogen H2 117.2 107.0 111.4 120.0
Methane CH4 51.9 49.8 49.8 50.0
Ethanol C2H5OH 29.6 28.4 28.4 27.7
Propane C3H8 48.8 47.0 47.1 46.3
Isooctane C8H18 47.5 45.8 45.9 44.7

*All species are taken as ideal gases. Reactions are with stoichiometric air at 25 uC, 1 atm.
{
The environment is taken as 25 uC, 1 atm, 363 ppm CO2, 2 per cent H2O, 20.5 per cent O2, balance N2.

many studies; see, for example, review articles [5] exhaust sensible energy for use) as well as engine
and [6]. These studies consistently show that four insulation (to reduce heat transfer losses), by placing
terms dominate the exergy balance as applied to these thermal strategies in the context of reducing
the engine cylinder (excluding subsystems for the respective exergy transfer terms [8–10]. For
compounding): (a) exergy destruction due to instance, the effects of insulation – both direct
combustion; (b) exergy transfer associated with (reduced exergy outflow with heat transfer) and
heat loss; (c) exergy in the exhaust of the combus- indirect (higher in-cylinder temperature, leading to
tion products; (d) work due to gas expansion. By increased compression work, throttling losses, and
contrast, exergy destruction terms due to mixing exhaust exergy) – have been described in detail by
and pumping losses are typically one or two orders Primus et al. [8].
of magnitude smaller. Table 2 shows representative By comparison, exergy destruction associated with
values for the net transfer and destruction terms combustion – a chemical transformation process – is
over one engine cycle taken from studies by Caton less well understood. (It has been said that the term
[7] and Primus et al. [8]. These studies indicate that ‘remains the most obscure and difficult availability
maximizing the expansion work – and thus engine destruction to cope with’ [6].) Some trends with this
efficiency – requires careful management of the destruction term have been noted previously. For
remaining three dominant exergy destruction and instance, it decreases with increasing adiabatic
transfer terms. flame temperature (brought about by compressing
This observation is not new, since the various the reactant, heating it, or increasing its equivalence
elements of the exergy management strategy have ratio towards stoichiometric) for combustion at
been discussed in the past, albeit at varying levels of constant pressure as well as at constant volume
detail and understanding. Several studies have [11–13]. On the other hand, many important para-
provided insights into the concepts of turbocharging meters of the combustion process, including its
and compounding (to recover a portion of the duration, temporal profile, and ignition timing, were
found to have only a slight impact on the amount of
exergy destroyed [14–16]. The present paper pro-
vides a thermodynamic explanation that unifies
Table 2 Representative results from exergy analyses these seemingly disparate results. Its implications
of spark-ignition (SI) and compression-
on the exergy management approach to raising IC
ignition (CI) engines in the literature [7, 8]
engine efficiency will also be discussed here and in
SI engine* CI engine{ the companion paper [17].
Exergy balance terms (% fuel exergy) (% fuel exergy)
Destruction due to 20.6 15.9
combustion
Net transfers with: 1.3 Paper outline
heat loss 23.0 21.4
exhaust 24.7 20.4 In the next section, an explanation is given of
expansion
{
work 29.7 39.1 terms that contribute to the exergy balance equa-
Balance 2.0 3.2
tion as applied to IC engines, and the equation is
*Based on a V-8 engine operating at part-load (1400 r/min, 27 kW) simplified for the case of an adiabatic, homoge-
[7].
{
Based on a naturally aspirated, direct-injection CI engine at
neous engine model. This model isolates combus-
2100 r/min, 185 kW [8]. tion as the only mechanism for exergy destruction,
{
For the SI engine, the balance includes destruction due to inlet which is to be minimized. While the exact mathe-
mixing and exergy of unburned fuel (distinct from the product
exhaust exergy term). For the CI engine, it includes destruction matical treatment of this optimization problem is
due to intake and exhaust valve throttling. presented elsewhere [18], its thermodynamic im-

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


452 K-Y Teh, S L Miller, and C F Edwards

plications on IC engine efficiency are elaborated in as typified by the results shown in Table 2. The
section 3 of this paper. Sections 4 and 5 clarify the balance equation makes evident the fact that these
effects on the optimal (i.e. minimum exergy exergy terms are intimately related. For instance, in
destruction) strategy when the stipulations that the engine insulation study discussed in reference
the system be homogeneous and adiabatic are [8], not all the exergy retained by reducing engine
lifted. heat loss is extracted in the form of work, i.e. the
In Part II of this study [17], a separate set of total exergy outflow $dXout is not fixed. Instead, a
thermodynamic requirements for optimal expansion substantial fraction is left in the exhaust stream
(i.e. maximum work extraction) is discussed and (indicated by high Xf), recoverable only via com-
contrasted with the conditions imposed by the optimal pounding.
combustion process. Given these twin sets of pre- Now consider a homogeneous, adiabatic IC engine
requisites, the companion paper then examines model that takes into account all significant effects
several resource preparation strategies – reactant except for heat loss and gas mixing. Excluding heat
compression, dilution, heating, and cooling – so transfer and other spatial inhomogeneities isolates
that the ensuing reaction (combustion) and work combustion as the sole source of exergy destruction
extraction processes can be carried out in an optimal and simplifies the model for mathematical analysis.
fashion. (The effects of reactant mixing and heat transfer on
the exergy destruction minimization strategy are
examined later in sections 4 and 5.) Under these
2 THE EXERGY BALANCE AND ITS stipulations, boundary work is the only permissible
MANAGEMENT mode of exergy transfer into or out of the system, so
that the net transfer is
Consider as a thermodynamic system the in-cylinder
fuel–air mixture during the closed portion of IC dXin {dXout ~{ðP{P0 Þ dV ~dUzP0 dV ð2Þ
engine operation between intake valve closing and
exhaust valve opening. A balance equation describ- where the first law for a closed, adiabatic system
ing the evolution of exergy of the working fluid may dU 5 2P dV has been incorporated in the expres-
be written as sion.
Heat and matter transfers are absent in this model,
ðf ðf as are the associated entropy flows. Hence, any
Xf {Xi ~ dX ~ dXin {dXout {dXdest ð1Þ entropy change within the system can be attributed
i i to the entropy generated during combustion, and
thereby the exergy destroyed (scaled by ambient
where Xi is the initial exergy of the system, dX is its temperature T0)
differential exergy change, the dX terms are differ-
ential amounts of exergy transferred in or out, or dXdest X mj
destroyed by irreversible (i.e. entropy-generating) ~dSgen ~dS~{ dNj ð3Þ
T0 j
T
processes, and Xf is the final working fluid exergy
that remains at the end.
Equation (1) indicates how the engine expansion Based on the first equality, often attributed to Gouy
work output – a component of the exergy outflow and Stodola, the terms ‘exergy destruction’ and
term $dXout – is affected by other aspects of its ‘entropy generation’ will be used interchangeably
operation. In this closed-system formulation, the for the remainder of this paper. The last equality
remainder of the outflow term can be entirely relates the system entropy increase to changes in its
ascribed to heat loss to the cooling jacket, while chemical composition due to combustion. It comes
the exergy associated with engine exhaust is sepa- from combining the first law with the Gibbs
rately accounted for by the final exergy term Xf. equation. Equations (2) and (3) reduce the balance
Working fluid compression constitutes an exergy equation (1) to
investment, given by the inflow term $dXin, whereas
exergy destruction, given by the term $dXdest and Xf {Xi ~ðUf {Ui ÞzP0 ðVf {Vi Þ{T0 ðSf {Si Þ ð4Þ
largely due to combustion-generated entropy, can
never be recovered. These aspects of IC engine Various aspects of the exergy management strategy
operation dominate the exergy balance equation (1), revolving around this balance equation will be

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 453

discussed in the following sections and the compa- composition N (a vector of mole numbers) of a
nion paper. closed, homogeneous system at any point in time
constitute a natural set of variables sufficient to
describe fully its thermodynamic state at that
3 STRATEGY FOR MINIMUM EXERGY instant. The corresponding constant-UV equili-
DESTRUCTION DUE TO COMBUSTION brium, denoted by the subscript ‘eq’, is then the
state that the system would have reached had it
As the review of past second-law analyses (section undergone chemical reactions to completion while
1.2) indicates, exergy destruction due to combustion being isolated with internal energy and volume fixed
remains a concept that is not well understood, even at instantaneous values U, V, and with its atomic
though it accounts for as much as a quarter of the composition unchanged. In other words – and this is
fuel exergy supplied to an IC engine. In fact, it can be the pivotal and recurring concept in the present
shown that this destruction term is minimized (and discussion – the constant-UV equilibrium state is
engine efficiency increased) by the strategy of always the attractor for the actual system at any
constant-UV equilibrium entropy minimization. point in its thermodynamic state space (U, V, N), in
The reader is referred to reference [18] for a detailed accordance with the second law.
mathematical treatment of the subject, which is The entropy function S(U, V, N) is used to quantify
based on the homogeneous, adiabatic IC engine this property for a closed system. The function is
model introduced above and formulated as a maximized at the constant-UV equilibrium state
dynamical system optimal control problem.
 
The thermodynamics underlying the optimal Seq U, V , N eq ¢SðU, V , N 0 Þ
solution is presented in this section. The discussion
provides a framework for dealing with issues for the set {N9} of all possible mixture compositions
associated with the homogeneity and adiabatic with the same atomic composition, including the
assumptions – the implementation challenges as actual, instantaneous composition of the system
well as opportunities with fuel–air mixing in the N [ {N9}. Applying the maximum-entropy principle to
former case, the problem of heat transfer losses in equation (3), which is restated here on a time-rate
the latter case – which are topics for the next two basis, yields the inequality displayed below.
sections.
1 dXdest dSgen dS X mj
~ ~ ~{ fj ðU,V ,N Þ¢0 ð5Þ
T0 dt dt dt j
T
3.1 Thermodynamics of the optimal solution
The optimal solution for minimum exergy destruc- Here, fj 5 dNj/dt is the net production rate of species
tion due to combustion in a piston engine can be j due to chemical reactions taking place in-cylinder.
summarized as For gas-phase (i.e. combustion) kinetics, these rates
are determined based on the mixture’s thermody-
(a) rapid compression of the reactant mixture when
namic state (U, V, N). (The equality holds only at
cylinder pressure is below the corresponding
equilibrium, such that Smj,eq fj (U, V, Neq) 5 0, which
constant-UV equilibrium pressure (and while
reduces Gibbs’ equation to Teq dSeq 5 dUeq + Peq
chemical reactions proceed);
dVeq.)
(b) rapid expansion of the products when cylinder
For a closed, adiabatic system, equation (5) shows
pressure exceeds its corresponding constant-UV
that its rate of entropy generation due to combustion
equilibrium value.
at every point in time is dictated solely by its
In thermodynamic terms, the optimal solution above thermodynamic state at that instant. The equation
reduces to a strategy that minimizes the constant- provides no indication as to how entropy generated
UV equilibrium entropy of the system of in-cylinder within the system can be minimized via boundary
gases. work, which is the only permissible mode of energy
This equilibrium corresponds to the combustion interaction. This is the rationale for introducing the
product state in the limit of instantaneous (and constant-UV equilibrium concept: a strategy invol-
therefore constant volume), complete reaction. For a ving compression and expansion work (namely, the
more precise definition – crucial to understanding optimal solution summarized above) can be devised
the strategy outlined above – it is first noted that the to minimize the entropy of this equilibrium state.
internal energy U, volume V, and chemical species More importantly, the exact same strategy would

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


454 K-Y Teh, S L Miller, and C F Edwards

also minimize entropy generation from the actual brium entropy would then be negative
thermodynamic system (i.e. the in-cylinder gas {1
 
mixture), since the system must satisfy the require- dSeq ~Teq Peq {P dV v0
ment of complete reaction – to avoid quenching or
incomplete combustion – and thus necessarily 2. Gas expansion, dV . 0. Gas expansion (work
approach equilibrium at the end of the engine cycle. extraction) is optimal at each instant when the
The optimal strategy can be explained by con- system pressure exceeds its constant-UV equili-
sidering changes in the volume and internal energy brium value, as that also leads to a decrease in
variables. For the actual adiabatic system, these equilibrium entropy: dV . 0 and Peq 2 P , 0 imply
changes are related by the energy balance {1
 
dU 5 2P dV. It also means, by definition, that dSeq ~Teq Peq {P dV v0
variables Veq and Ueq for the corresponding con-
stant-UV equilibrium attractor would change by the The strategy of equilibrium entropy minimization
same amount. Therefore is best depicted on the internal energy/entropy
diagram, as shown in Fig. 1. It is based on the
dVeq ~dV numerical example used in reference [18] for the
optimal combustion of a stoichiometric propane/air
dUeq ~dU~{P dV ~{P dVeq
ð6Þ mixture with 50 per cent exhaust gas recirculation
(EGR). (Appendix 2 includes additional comments
This means that Gibbs’ equation at equilibrium can on the illustrative examples used in this paper.) The
be written as figure includes the optimal U, S trajectory as well as
its constant-UV equilibrium counterpart, showing
  that the thermodynamic state of the reacting gas is
Teq dSeq ~dUeq zPeq dVeq ~ {PzPeq dV ð7Þ
indeed being steered such that the constant-UV
Equation (7) makes evident the strategy for mini- equilibrium entropy is always decreasing.
mizing entropy of the constant-UV equilibrium state These trajectories typify combustion of common
via boundary work. fuel–air mixtures. Since the combustion process is
exothermic, the mixture pressure would initially be
1. Gas compression, dV , 0. Gas compression – lower than its corresponding constant-UV equili-
constituting a marginal investment in the form brium value, Pi , Peq,i. This condition calls for gas
of work (dV , 0, dU 5 2P dV . 0) – is optimal at compression such that the equilibrium entropy
each instant when the system pressure is below decreases. Chemical reactions are typically negligi-
the corresponding constant-UV equilibrium pres- ble during the initial reactant compression process,
sure, Peq 2 P . 0. The resulting change in equili- but as the mixture temperature and pressure

Fig. 1 The equilibrium entropy minimization strategy for reducing exergy destruction due to
combustion. (a) Optimal U, S and corresponding U, Seq trajectories. (b) Trajectories near
chemical equilibrium

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 455

increase, chemical kinetics accelerate until the Combustion kinetics affect the instantaneous mixture
mixture autoignites. The optimal U, S trajectory – composition, temperature, and thus pressure, which in
process i–B̄ in Fig. 1(a) – reflects this; it starts off turn determine the optimal switching point, P 5 Peq at
largely isentropic while internal energy U increases, B̄. Notwithstanding these effects of detailed kinetics,
but curves to the right as the reaction proceeds with the optimal solution means that it is the equilibrium
an attendant increase (i.e. generation) of entropy S. thermodynamics (specifically, the constant-UV pro-
Upon autoignition, the optimal trajectory shows duct states at the end of combustion) and not chemical
significant entropy generation with small internal kinetics (i.e. reactions taking place during combustion)
energy increase, which indicates that the reactants that ultimately dictates the amount of exergy de-
are converted to products very rapidly in compar- stroyed. (This explains the very mild dependence of
ison to the rate of compression. the entropy generation term on various parameters of
The pressure difference Peq 2 P . 0 diminishes as the combustion process, as mentioned in the intro-
chemical reaction (and compression) proceeds, and duction, i.e. section 1.2, references [14] to [16].) After
reaches zero at point B̄. This point is close to all – and to reiterate our main conclusion – the strategy
chemical equilibrium, signifying that combustion is of minimizing entropy Seq for the constant-UV
near completion at this juncture (92 per cent equilibrium to which a closed system of reacting gases
combustion efficiency). It should be noted that the is attracted will also, by virtue of the requirement for
criterion P 5 Peq does not imply equilibrium has complete reaction, minimize the actual entropy gen-
been reached. (If a system is at equilibrium, then its eration due to combustion of the gas mixture.
pressure P 5 Peq. The converse, however, need not
be true.) In fact, the in-cylinder pressure proceeds to
overshoot its equilibrium value by a small margin 3.2 Compression ratio constraint
past point B̄. Using the ideal gas model Figure 1 illustrates the well-known fact that a fuel–
    air mixture cannot be compressed with no limit (or,
SNj RT SNj,eq R Teq
PwPeq [ w ð8Þ in equivalent terms, its internal energy cannot be
V Veq raised indefinitely) without combustion taking place,
which accounts for the fundamental constraint on
where V 5 Veq. Given the complex nature of finite-rate, compression ratio of premixed/homogeneous-
detailed combustion chemistry, it is not inconceivable charge, spark-ignition engines. At any rate, engine
for the actual gas composition and temperature to geometry imposes a separate, practical constraint on
differ from the constant-UV equilibrium values such the minimum allowable clearance volume and it
that, upon multiplication, (SNj)T exceeds (SNj,eq)Teq. defines the engine compression ratio. Nevertheless,
For example, SNj . SNj,eq may indicate excess radical it is important to recognize that the approach for
species yet to recombine to yield stable products at minimizing entropy generation due to combustion is
equilibrium; T . Teq may be due to slow, endothermic unaltered even when this clearance volume con-
reactions (e.g. NO formation). straint is included. After all, the requirement that the
Following the optimal strategy, inequality (8) calls product gases reach equilibrium at the end of the
for gas expansion such that the constant-UV equili- engine operating cycle is unchanged. Hence, the
brium entropy decreases. Since combustion is strategy of minimizing entropy Seq for the constant-
effectively complete, entropy generation during UV equilibrium attractor also remains valid.
expansion is typically miniscule. This is illustrated As pointed out in the previous section, the initial
by process B̄–C̄9 and the corresponding equilibrium pressure Pi of common fuel–air mixtures is below
U, Seq trajectory in Fig. 1. The calculation was that of its constant-UV equilibrium counterpart Peq,i
terminated at point C̄9, where the complete-reaction since combustion is exothermic. Compression of the
criterion (97 per cent combustion efficiency) was fuel–air mixture then results in entropy reduction for
reached. After that point, a dashed line is appended its corresponding constant-UV equilibrium state.
to indicate isentropic expansion of the combustion This compression strategy applies until the engine
products to the initial system volume (point C̄). compression ratio is reached.
It should be emphasized that combustion and the It is reasonable to assume that Peq 2 P . 0 at this
consequent entropy generation are accounted for phase of the optimal engine operating cycle. Had the
throughout process i–B̄–C̄9. (This includes the case clearance volume constraint been absent, the equili-
where combustion begins and approaches comple- brium attractor entropy minimization strategy would
tion, all while the volume is still being reduced.) have called for further gas compression (thus reverting

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


456 K-Y Teh, S L Miller, and C F Edwards

to the optimal solution presented in section 3.1). chemical equilibrium, which signifies near-complete
Given the constraint, however, the strategy is to reaction (98 per cent combustion efficiency). An
maintain the gas volume (which, under adiabatic isentropic line (dashed) is appended to indicate
conditions, implies fixed attractor entropy dSeq 5 0) combustion product expansion to the initial system
rather than expand the gas (which leads to attractor volume (point C).
entropy increase dSeq . 0). Chemical kinetics will Not surprisingly, the reactant compression pro-
accelerate during this constant-volume phase, until cess is largely isentropic as the fuel–air mixture does
the in-cylinder pressure overshoots the corresponding not autoignite before compression to the clearance
constant-UV equilibrium pressure, after which the volume (point A). Similarly, exergy destruction
optimal strategy is, again, gas expansion. during expansion of the combustion products is
To summarize, in a typical IC engine subject to a also minuscule, since the optimality criterion for
compression ratio constraint, the optimal operating expansion, P . Peq, is only satisfied when the system
cycle that minimizes the entropy generation due to is close to chemical equilibrium. As a result, entropy
combustion is: generation during combustion at the clearance
(a) rapid compression of the reactant mixture to volume dominates the exergy destruction term. In
the minimum allowable volume set by the other words, the optimal operating cycle for the
engine compression ratio (while chemical reac- homogeneous, adiabatic IC engine model is approxi-
tions proceed); mately an ideal fuel–air cycle, consisting of the
(b) maintenance of the clearance volume until the following sequence of processes:
cylinder pressure rises to its corresponding con-
(a) isentropic compression of the reactant mixture
stant-UV equilibrium value (due to combustion);
– with its chemical composition frozen – to the
(c) rapid expansion of the products.
engine clearance volume;
For example, Fig. 2 shows the U, S diagram for (b) constant-volume combustion, converting the
optimal combustion of the same homogeneous reactants into products at thermodynamic
mixture considered previously (stoichiometric pro- equilibrium;
pane/air, 50 per cent EGR), albeit constrained by 13:1 (c) isentropic expansion of the equilibrium product
compression ratio. (Without the compression ratio mixture.
constraint, the mixture would ignite upon compres-
sion near a volume ratio of 19:1; see Fig. 1.) Again, the Depending on the assumption, the expansion may
optimal U, S and equilibrium U, Seq trajectories illus- end at some prescribed final volume (e.g. symmetric
trate the equilibrium entropy minimization strategy expansion to the initial volume, akin to the Otto
during compression (process i–A in Fig. 2(a)) as well cycle) or pressure (e.g. expansion to the atmospheric
as expansion (B–C9 in Fig. 2(b)). Point C9 lies close to or exhaust pressure, akin to the Atkinson cycle), with

Fig. 2 The strategy of equilibrium entropy Seq minimization remains unaltered upon
consideration of the IC engine compression ratio constraint. (a) Optimal U, S and
corresponding U, Seq trajectories. (b) Trajectories near chemical equilibrium

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 457

corresponding implications on the maximum ex- equilibrated) products at the end of the process. In
pansion work output from each cycle. each case, the system thermodynamic properties are
This is an important result that warrants amplifi- fully defined by a fundamental relation of the form
cation. It means that at any given compression ratio,
piston engine efficiency is maximized when com- U~U ðS, V , N Þ ð9Þ
bustion occurs at the minimum allowable volume.
This is a far stronger assertion than what is already In physical terms, this equation describes the
well known – that efficiency is highest for the functional interdependence between the internal
constant-volume combustion cycle as compared energy U, entropy S, and volume V as the gas
with the constant- and limited-pressure cycles (also mixture is manipulated thermally (heated or cooled)
known as the diesel and dual cycles) [3] – since the or mechanically (compressed or expanded). In
optimal solution presented above is obtained upon differential terms, this functional dependence takes
consideration of all possible engine operating cycles. the form of Gibbs’ equation
The optimization results presented above,
founded upon the equilibrium attractor entropy
minimization strategy given by equation (7), also dU~T dS{P dV ð10Þ
discount the notion that simultaneous work transfer
and chemical reactions could enhance the efficiency in which the term Smj dNj associated with chemical
of combustion engines. Instead, the chemical reac- transformation does not appear. In the case of the
tions are ‘separable’ from the mechanical (compres- frozen reactants, this is because the mole number Nj
sion and expansion) processes at optimality. Each of of each chemical species is fixed, so the change dNj
the three processes enumerated in the optimal is always zero. (The system is said to be in restricted –
sequence above correlates to one of three essential i.e. thermal and mechanical, but not chemical –
elements of engine architecture: preparation, reac- equilibrium.) In the case of the products, the sum
tion, and extraction of work from the energy vanishes under the assumption of chemical equili-
resource (i.e. the fuel) supplied to the engine. The brium. The fundamental relation (9) for U can
concept of separability allows each of these archi- therefore be represented graphically as a two-
tectural elements to be analysed individually, while dimensional manifold of S and V.
ascribing the entropy generated over the entire For example, the manifolds for the frozen reac-
operating cycle to just the reaction stage of the cycle. tants considered thus far (stoichiometric propane–
air, 50 per cent EGR) and the corresponding
equilibrium combustion products are shown in
3.3 Exergy destruction due to constant-UV Fig. 3. In both cases, the mixture’s internal energy
combustion increases with increasing entropy and decreasing
volume, at slopes equalling the mixture temperature
The discussion thus far points to the equilibrium T and pressure P, respectively. Isotherms are shown
thermodynamic states for products of combustion as on the figure as dashed lines. The reactant isotherms
the most crucial element in the entropy generation coincide with the constant-U contours under the
minimization scheme. In the case of an IC engine with assumption of ideal gas behaviour. The product
a compression ratio constraint, it is noted that the isotherms diverge at high internal energy and
magnitude of the entropy generation term is very volume owing to dissociation.
close to the entropy difference between the reactants
A common reference state and set of reference
and its constant-UV equilibrium products at the same
species from the NASA thermochemical database are
clearance volume. This approximation reduces the
used to account for the energy associated with
question of exergy destruction minimization from a
chemical bonds for both the reactant and product
more challenging problem involving reaction kinetics
species [19]. Graphically, this means that both
and other non-linear dynamics, to a substantially
manifolds can be plotted on the same set of U, S,
simpler problem involving only the thermodynamics
and V axes, as shown in Fig. 4. Elements of the
of the reacting gas mixture at two points, i.e. its initial
volume-constrained fuel–air cycle are also shown in
reactant and final product states. The present section
this figure. They include:
analyses this thermodynamic problem.
Consider a thermodynamic system composed of (a) the constant-entropy curves on the manifolds,
an unreacted (i.e. chemically frozen) fuel–air mixture for isentropic reactant compression and pro-
at the start of combustion, and fully reacted (i.e. duct expansion;

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


458 K-Y Teh, S L Miller, and C F Edwards

Fig. 3 The fundamental relation manifolds U 5 U(S, V) for (a) the frozen mixture of fuel
(propane), air, and EGR; and (b) the equilibrium products of combustion

(b) the transition (or ‘jump’) across the manifolds increase, or equivalently, destroys less exergy. The
at the same volume and internal energy, for magnitude of this destruction term is proportional to
constant-volume combustion; the distance between the two manifolds along the S
(c) the constant-UV equilibrium trajectory on the axis. Its contour map on the U, V bounding plane is
product manifold, to which the actual system shown in Fig. 5. The reactant isentrope (at entropy
trajectory is attracted. S 5 Si) and isobar (at pressure P 5 Pi 5 1 atm) passing
through the initial mixture state are also included in
The projection of these lines onto the U, S bounding the figure for reference. For example, the figure
plane closely approximates the trajectories shown in shows that complete combustion of the reactants
Fig. 2. after a 20:1 isentropic compression destroys 15 per
The strategy for minimizing exergy destruction cent of the fuel exergy supplied to the engine. This
associated with combustion is evident in Fig. 4, amounts to a two-fifths reduction from the 26 per
which shows the reactant and product manifolds cent exergy destruction incurred if the reacting
approaching each other at high internal energy. mixture is not compressed.
Therefore, the reacting mixture needs to be steered
towards a thermodynamic state with higher energy,
such that subsequent combustion (i.e. the jump to
the product manifold) incurs a smaller entropy

Fig. 5 The contour map for entropy generation due to


constant-UV combustion of a propane/
air mixture with EGR. The corresponding
Fig. 4 Optimal combustion cycle trajectory on the U, fractions of fuel exergy XC3 H8 destroyed are in
S, and V axes parentheses

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 459

It is important to recognize that the amount of Returning to the subject of exergy destruction by
exergy destroyed is strongly dependent on the constant-volume combustion, its functional depen-
internal energy of the fuel–air mixture just prior to dence on the state variables U and V can be
combustion. Of far less significance is the volume at established analytically based on the reactant and
which combustion occurs, since the contours are product fundamental relations in differential form, i.e.
sloped only slightly along the V axis. This means that Gibb’s equation (10). Consider the adiabatic combus-
the strategy for minimizing exergy destruction is less tion of some quantity of reactants at fixed volume. The
a matter of reducing the reactant mixture volume, resulting exergy destruction term is given by
and more a matter of raising its internal energy. For
Xdest ~T0 |½SP ðU, V Þ{SR ðU, V Þ
the adiabatic engine model under discussion, this
strategy takes the form of reactant compression
simply because boundary work is the only permis- where both entropy terms are evaluated at the same
sible mode of energy transfer into the closed system; volume V and internal energy U. Any change in exergy
this need not be the case in general. For instance, if destruction can therefore be related, via equation (10),
heat transfer is permitted, Fig. 5 indicates that the to changes in these two variables
same two-fifths reduction in exergy destroyed (from dXdest
~dSP {dSR
26 to 15 per cent of fuel exergy supplied) can also be T0
attained by preheating (isobarically) the reactant    
1 1 P P PR
mixture to 1450 K, rather than by compressing it ~ { dUz { dV ð11Þ
TP TR TP TR
(isentropically) through a 20:1 volume ratio. The
excessive temperatures upon preheating or com-
Equation (11) indicates that the conclusions drawn
pression are artifacts of the high initial reactant
based on the contour map in Fig. 5 can be general-
temperature used in the illustrative examples [18],
ized as follows.
retained here merely for consistency. The numbers,
while well past the autoignition point for most fuels, 1. Change in internal energy, dU. For exothermic
do not obviate the more general conclusion that reactions (which is the case for combustion of
exergy destruction can be reduced by internal energy common fuel–air mixtures), the product gas tem-
increase by either mode of transfer. This is not to perature TP is greater than that of the reactants TR,
say that the two approaches are equivalent in the so (TP21 2 TR21) is negative. Therefore, raising the
thermodynamic sense; for example, entropy transfer internal energy of the reactants (dU . 0) necessarily
also takes place during reactant heating but is absent reduces the exergy destroyed during the constant-
during adiabatic compression. The attendant issues volume combustion process that ensues (dXdest , 0).
are discussed further in the companion paper [17]. 2. Change in volume, dV. Variation of the destruc-
tion term with volume is less clear-cut, and
depends on the chemical system under consid-
eration. Assuming that the reactant and product
gas mixtures are ideal, then


PP PR R 
{ ~ SNj,P {SNj,R ð12Þ
TP TR V

noting that VP 5 VR 5 V for constant-volume com-


bustion. If the combustion process leads to a net
increase in total mole number (i.e. mole-producing
SNj,P 2 SNj,R . 0), reducing the volume at which it
takes place (dV , 0) will reduce the exergy de-
stroyed (dXdest , 0). This is the case for most fuel–
air systems, as exemplified by the propane–air
system considered thus far (with positively sloped
Fig. 6 The contour map for entropy generation due to
exergy destruction contours in the V axis direction
constant-UV combustion of stoichiometric hy-
drogen/air mixture. The corresponding frac- in Fig. 5). The converse – exergy destruction
tions of fuel exergy XH2 destroyed are in decreases when combustion occurs at a larger
parentheses volume – is true if mole number decreases during

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


460 K-Y Teh, S L Miller, and C F Edwards

the process (i.e. mole-destroying SNj,P 2 SNj,R , 0), the thermodynamic analysis presented thus far for
typified by hydrogen combustion. Figure 6 shows minimum exergy destruction. This is clearly the case
the corresponding map of exergy destruction with the switch to fuels with improved antiknock
contours (negatively sloped along the V axis). properties, since the results in section 3 are applic-
able to all fuel species.
While the slope of the contour lines may be fuel
The case for direct-injection strategies requires
dependent, the basic message is unaltered: the
further explication. Obviously, the assumption of
reactants need to combust at a high internal energy
mixture homogeneity no longer holds; some exergy
state in order to minimize the exergy destroyed.
is destroyed owing to fuel vaporization and mixing
Comparing the two contour maps in Figs 5 and 6, it
with air. Nevertheless, the thermodynamic require-
may be concluded that the internal energy invest-
ment for exergy destruction minimization in this
ment is best made in the form of reactant compres-
instance can be deduced using reasoning that
sion if the combustion process is mole producing
(i.e. along the isentrope in Fig. 5), and via reactant mirrors that presented earlier for premixed systems.
heating if the reaction is mole destroying (along the First, it is observed that prior to mixing, fuel and
isobar in Fig. 6). However, the two approaches to air can be considered as separate components in a
raising reactant internal energy have very different non-premixed system of reactants. Hence, the
second-law implications on the work extraction internal energies, entropies, and volumes – all are
stage of the engine cycle, as alluded to above and extensive variables – of the two components are
elaborated in the companion paper [17]. simply additive to yield variables U, S, and V for the
aggregate system. This also applies to multi-zone
modelling of IC engines, by summing the variables
for all discrete zones of the model. In the absence of
4 EFFECTS OF REACTANT MIXING AND
the homogeneous mixture assumption, however, the
IGNITION PHASING
set of variables U, S, and V are no longer related by
the reduced form of Gibbs’ equation (10) during the
The optimal solution for minimum entropy generation
mixing process. In fact, the entropy S of the
due to combustion (sections 3.1 and 3.2) is exact only
under the assumption that the gas mixture is homo- aggregate system at any given internal energy U
geneous. The subsequent analysis based on reactant and volume V is not unique. Rather, it depends on
and product manifolds (section 3.3) further presumes parameters such as the fuel injection duration,
that the mixture remains chemically frozen prior to extent of fuel–air mixing, etc. However, the value of
combustion. However, these assumptions are incom- S will coincide with the homogeneous frozen
patible with practical implementation of the optimal reactant mixture entropy upon complete mixing.
combustion strategy – as noted earlier, the internal Stated another way, the fundamental relation
energy of a premixed system cannot be raised indefi- manifold UR(S, V) for premixed reactants introduced
nitely without combustion taking place. Autoignition in section 3.3 (e.g. Fig. 3(a)) serves as the attractor
sets a bound on the minimum entropy generation that manifold for the state trajectory of the composite
such a system must incur, as illustrated in Fig. 1. system as mixing proceeds, in the same way that
Further reduction in entropy generation beyond the equilibrium product manifold UP(S, V) (e.g.
this limit, while possible, would require delaying Fig. 3(b)) is the thermodynamic attractor for the
ignition of the reacting mixture to enable even state trajectory as combustion proceeds. Both
higher compression of reactants. Ignition delay can manifolds are shown in Fig. 7, along with one
be accomplished by modifying the chemical kinetics constructed for air and 50 per cent EGR plus a
of the system, e.g. by using fuel blends with higher stoichiometric amount of propane at 25 uC, 1 atm
octane ratings. Ignition timing can also be controlled prior to mixing. The two reactant manifolds are
via fuel injection in non-premixed or stratified- almost indistinguishable from each other, whereas
charge, compression-ignition engines. In this case, the product manifold is displaced further away along
air (the oxidizer) is compressed separately to a high the S axis, indicating that entropy generated due to
internal energy state. Combustion does not com- reactant-to-product transformation is far larger than
mence until fuel is introduced into the system, after that associated with fuel mixing. Figure 7 suggests
which the reaction rate is primarily determined by that the strategy to carry out chemical reactions at
the mixing process between fuel and air. high internal energy states, for the purpose of
It is asserted here that, for the most part, these minimizing the subsequent exergy destruction due
dynamic strategies for ignition phasing do not affect to combustion, is readily extensible to non-premixed

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 461

systems. Table 3 illustrates this quantitatively by diesel operation, autoignition occurs near the inter-
showing that mixing entropy Sgen,mix often accounts phase regions – at stoichiometric or richer equivalence
for less than 2 per cent of the exergy content for most ratios – between the fuel jet and the surrounding air.
fuels. The entropy change Sfg due to liquid fuel Jet momentum and turbulence induce further mixing
vaporization also tends to be small, e.g. 2.4 and of the reactants and control the subsequent rate of
0.7 kJ/(kg fuel)21 K21 (2.4 and 0.5 per cent fuel combustion. Therefore, the preparation and reaction
exergy) for ethanol and isooctane at their respective steps should not be separately analysed.
boiling points. By comparison, the exergy destroyed Despite these differences, the current authors
due to constant-volume combustion (also listed in assert that the optimal combustion strategy need
Table 3) ranges from 10 to 25 per cent of fuel exergy. not be altered. The reason traces back to equations
There are two plausible objections to the explana- (6) and (7), restated below
tion thus far with regards to the thermodynamics of
delayed combustion phasing, via diesel-like opera- dUeq ~dU~{P dV ~{P dVeq ð6Þ
tion with direct or indirect fuel injection and  
compression-ignition, for minimum combustion- Teq dSeq ~dUeq zPeq dVeq ~ {PzPeq dV ð7Þ
generated entropy. First, the order of magnitude
argument becomes less tenable when higher levels These equations remain applicable for an adiabatic,
of EGR are involved. For instance, Fig. 7 does not non-premixed system at pressure P. The energy
take into account entropy of mixing between a fresh balances for different components (fuel, air, reac-
charge of air and 50 per cent EGR, which as Fig. 8 tant, and product mixtures at any given ratio) add up
indicates can approach one-fifth of entropy genera- to yield the overall balance dU 5 2P dV in equation
tion due to combustion if the exhaust is hot. (6) involving internal energy U and volume V of the
The second objection pertains to the assumption aggregate system. (Again, the exact same argument
that the gases are completely mixed as part of the extends to multi-zone engine modelling by summing
reactant preparation process, separate from the sub- the energy balances for all the zones.) Gibbs’
sequent chemical reaction step. A reasonable argu- equation (7) also remains valid since, regardless of
ment may be made that conventional diesel combus- whether the system is homogeneous or not, it is
tion is substantially different from this scheme. In constrained at the end of combustion by a unique
equilibrium product manifold. Therefore, the opti-
mal strategy – whether to compress or expand –
continues to be determined by comparing the
pressures P and Peq, and still reduces to minimizing
the equilibrium attractor entropy Seq.

5 EFFECTS OF LOSSES DUE TO ENGINE HEAT


TRANSFER

Aside from the chemical kinetic (autoignition)


limitation considered previously, materials issues
impose another set of constraints on the actual
execution of the strategy for exergy destruction
minimization. In particular, the melting points of
advanced materials today (e.g. near 1500 K for
nickel-based superalloys, or 1800 K for coated car-
bon–carbon composites [20]) are still far below the
combustion temperature of most fuel–air systems,
especially if the reactions are to take place at high
internal energy states. This material constraint
necessitates active cooling of the IC engine and
Fig. 7 Manifold for a composite system of air and 50
per cent EGR before mixing with stoichiometric removes a significant fraction of energy from
propane at 25 uC, 1 atm, shown relative to the the working fluid. Consequently, as much as
well-mixed frozen reactant and chemically one-quarter of the exergy supplied to the system is
equilibrated product manifolds lost to the engine coolant. In other words, the

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


462 K-Y Teh, S L Miller, and C F Edwards

Table 3 Entropy generated, Sgen, due to fuel–air mixing and to constant-UV combustion, as well as the associated
exergy destruction Xdest (entropy values are in the units of kJ/(kg fuel)21 K21)
Xdest Xdest Xdest Xdest Xdest
ð250 CÞ{ ð1000 KÞ ð1:1Þ { ð20:1Þ ð50:1Þ
Fuel species* Sgen,mix (% Xfuel) Sgen,mix (% Xfuel) Sgen (% Xfuel) Sgen (% Xfuel) Sgen (% Xfuel)
Hydrogen 8.5 (2.2) 14.7 (3.8) 66.9 (17.1) 38.6 (9.9) 31.7 (8.1)
Methane 1.7 (1.0) 3.0 (1.8) 41.0 (23.5) 27.6 (16.6) 24.1 (13.9)
Ethanol 0.7 (0.7) 1.6 (1.6) 25.2 (25.5) 17.8 (18.0) 15.8 (15.9)
Propane 0.8 (0.5) 1.9 (1.2) 40.4 (24.9) 28.1 (17.7) 25.9 (15.2)
Isooctane 0.4 (0.2) 1.5 (0.9) 40.6 (25.6) 28.8 (18.2) 25.6 (16.1)

*All species are taken as ideal gases. Reactions are with stoichiometric air.
{
Entropy generation due to mixing of fuel at 25 uC with stoichiometric air at the same temperature or at 1000 K (superscripted as such), and
the equivalent fraction of fuel exergy destroyed.
{
Entropy generation due to constant-UV combustion of reactants (initially at 25 uC, 1 atm) before, as well as after, compression
(superscripted 1:1, 20:1, and 50:1 respectively), and the equivalent fraction of fuel exergy destroyed.

Ðt
combustion process is far from adiabatic, which runs where tif dQ is the total heat loss from the engine PE
counter to the dynamical system model analysed over the time interval, while SR,i and SP,f denote,
above. respectively, the initial reactant and final product
If it is assumed that all the exergy transferred to the entropies within the engine. The requirement of
coolant is destroyed via coolant heat transfer to the complete combustion has thus been incorporated
surroundings, and not recovered via compounding, into the expression.
the overall exergy destruction for the piston engine Not surprisingly, equation (13) shows that the
Total
(denoted PE) and its cooling system (CS) is then exergy destruction Xdest consists of both a chemical
component and a heat loss component. Lowering
Total PE CS
dXdest ~T0 dSgen zT0 dSgen the overall destruction term therefore requires care-
    ful management of both combustion and heat
dQ 1 1
~T0 dSz zT0 { dQ transfer processes. For one thing, it is no longer a
T T0 T straightforward matter of decreasing the final pro-
duct entropy SP,f, since that may be accomplished at
or, when integrated over some time interval [ti, tf]
the expense of enhanced heat loss from the engine.
ð tf Total
In fact, to solve the problem of Xdest minimization
Total
 
Xdest ~T0 SP,f {SR,i z dQ ð13Þ would entail specifying the heat transfer process
ti
which is, in thermodynamic terms, a path-depen-
dent (i.e. non-unique) function. This is not unlike
the challenge, mentioned in the previous section, of
determining mixing entropy in a unique, path-
independent manner. As an alternative to solving
the problem analytically (possible under very re-
strictive conditions [21]), here a numerical approach
to solving the related problem of expansion work
maximization for a non-adiabatic piston engine was
examined in [22]. In this study, the engine compres-
sion ratio was fixed, which sets a hard bound on the
maximum internal energy level to which the in-
cylinder gas mixture can be raised via compression.
The optimized solution is characterized by a slight
reduction in heat loss during the compression
process, which leads to combustion at a marginally
higher internal energy state.
Fig. 8 Comparison between the entropy generated In other words, to the extent that heat loss can be
due to mixing of stoichiometric propane and Total
kept low, Xdest minimization reverts to the question
air (both at 25 uC, 1 atm) with varying amounts
of EGR (at 25 uC or 600 K, 1 atm), and the
of how to transform reactants to products in the
entropy generated during combustion of most efficient (i.e. with the least loss) manner, and
the resulting reactant mixture (after 20:1 the thermodynamic results of previous sections once
compression) again apply. Operating a high-efficiency combustion

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 463

engine necessarily entails carrying out the combus- much stronger than the well-known result – that the
tion process at the highest possible energy state, constant-volume combustion cycle has higher effi-
such that the constant-UV entropy change from ciency than the constant- or limited-pressure com-
reactants to products is minimized. After all, relaxing bustion cycles – because the assertion is based upon
the requirement of adiabatic engine operation does optimization and consideration of all possible
not invalidate the second-law principle that a closed engine cycles.
combustion system that is not in equilibrium always Two important implications of the optimal solu-
seeks to maximize its constant-UV entropy. tion are that:
There are many approaches to reducing engine
heat loss, such as wall insulation and other low-heat- (a) the reactant and product thermodynamic states
rejection concepts, lowering of the combustion at the start and the end of combustion, as
chamber surface-to-volume ratio (as in large diesel opposed to chemical kinetics or other non-
engines), etc. A thorough review of these concepts is linear dynamics during combustion, determine
beyond the scope of this paper, and the reader is the minimum exergy destruction due to com-
referred to other publications for detailed discus- bustion;
sions, for example references [8] to [10], and [23]. (b) the optimal IC engine operating cycle (i.e. the
Peak combustion temperature may also be reduced engine architecture) is composed of separable
via dilution. A thermodynamic discussion of this elements of reactant preparation, combustion,
reactant preparation strategy is deferred to the and work extraction from the products.
companion paper [17].
Based on the first finding, the thermodynamic
To summarize, while the system homogeneity and
properties of the combustion system in its frozen
adiabatic assumptions are essential for the dynami-
reactant and equilibrated product states are exam-
cal system control problem of minimum exergy
ined (by considering the respective fundamental
destruction to be solved exactly in reference [18],
relation manifolds) to quantify the exergy destroyed
the thermodynamic conclusions that can be drawn
due to constant-UV combustion. Exergy destruction,
from the optimal solution remain valid when the
Xdest, is shown to be only mildly dependent on the
assumptions are omitted. Equilibrium entropy Seq
reactant mixture volume V, but a strong function of
minimization, through steering of the reactants to
its internal energy U at the point of combustion. In
the highest possible internal energy state before
particular, raising the internal energy necessarily
reaction, is shown to be a general strategy for
lowers exergy destruction because combustion is
minimizing the exergy destroyed during combus-
tion. exothermic. Therefore, the destruction term can only
be reduced by work and/or heat transfers into the
closed system, i.e. gas compression, heating, or some
6 CONCLUSIONS combination of the two.
This result is readily extensible to non-premixed
This is the first of a two-part paper in which the and non-adiabatic systems. For non-premixed sys-
strategies for maximizing expansion work output tems, exergy destruction from liquid fuel evapora-
from the IC engine are investigated. The discussion tion and mixing are an order of magnitude below
focuses on the fundamental thermodynamic re- that due to combustion. This is not the case with
quirements for the optimal engine operating cycle. respect to exergy destruction due to heat loss for
Some related implementation issues are also con- non-adiabatic systems. Nevertheless, provided that
sidered within this context, including reactant mix- heat loss can be kept low (or at least not exacer-
ing, ignition phasing, and engine heat loss. bated), then equilibrium (product) entropy Seq
Underpinning all analyses and conclusions pre- minimization, through steering of the reactants to
sented in this article is the optimal control solution the highest possible internal energy state before
for minimum exergy destruction due to combustion reaction – following what may be called the ‘extreme
in a homogeneous, adiabatic piston engine model state’ principle – remains the general strategy
[18]. In thermodynamic terms, that solution reduces for minimizing the exergy destroyed during
to a strategy of minimizing the entropy of the combustion.
constant-UV equilibrium to which the chemical The second finding above implies that the remain-
system is attracted. For many systems, this strategy ing elements of optimal IC engine architecture – i.e.
is well approximated by the constant-volume com- the reactant preparation and product expansion
bustion cycle. However, the authors’ assertion is (work extraction) processes – can be separately

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9


464 K-Y Teh, S L Miller, and C F Edwards

analysed from the standpoint of managing the internal combustion engines. SAE Paper 890832,
associated exergy transfers so as to maximize the 1989.
overall engine cycle efficiency. The exergetic and 16 Caton, J. A. The effect of burn rate parameters on the
operating attributes of a spark-ignition engine as
efficiency implications of these processes will be
determined from the second law of thermody-
discussed in the companion paper [17]. namics. In Proceedings of the Spring Technical
Conference of the ASME ICE Division, San Antonio,
Texas, April 2000, ICE-Vol. 34-2, Paper 2000-ICE-274.
REFERENCES 17 Teh, K.-Y., Miller, S. L., and Edwards, C. F.
Thermodynamic requirements for maximum inter-
1 OECD. IEA world energy statistics and balances, nal combustion engine cycle efficiency. Part 2:
Vol. 2006, Release 1, 2006 (Organisation for Eco- work extraction and reactant preparation strate-
nomic Co-operation and Development (OECD), gies. Int. J. Engine Res., 2008, 9(6), 467–481.
Paris). 18 Teh, K.-Y. and Edwards, C. F. An optimal control
2 IEA. CO2 emissions from fuel combustion, 1971– approach to minimizing entropy generation in an
2004, 2006 (International Energy Agency (IEA), adiabatic internal combustion engine. J. Dyn. Syst.,
Paris). Measmt, Control, 2008, 130(4), 041008.
3 Heywood, J. B. Internal combustion engine funda- 19 McBride, B. J., Gordon, S., and Reno, M. A.
mentals, 1988 (McGraw-Hill, New York). Coefficients for calculating thermodynamic and
4 Szargut, J., Morris, D. R., and Steward, F. R. Exergy transport properties of individual species. NASA
analysis of thermal, chemical, and metallurgical Technical Memorandum No. TM-4513, 1993.
processes, 1988 (Hemisphere, New York). 20 ASM International. ASM handbooks online, at http://
5 Caton, J. A. A review of investigations using the www.asmmaterials.info/, accessed 1 December 2006
second law of thermodynamics to study internal- (ASM International, Materials Park, Ohio).
combustion engines. SAE Paper 2000-01-1081, 2000. 21 Teh, K.-Y. Thermodynamics of efficient, simple-
6 Rakopoulos, C. D. and Giakoumis, E. G. Second- cycle combustion engines, PhD thesis, Stanford
law analyses applied to internal combustion en- University, 2007.
gines operation. Prog. Energy Combust. Sci., 2006, 22 Teh, K.-Y. and Edwards, C. F. Optimizing piston
32(1), 2–47. velocity profile for maximum work output from an
7 Caton, J. A. Results from the second-law of IC engine. In Proceedings of ASME IMECE,
thermodynamics for a spark-ignition engine using Chicago, Illinois, November 2006, Paper IM-
an engine simulation. In Proceedings of the Fall ECE2006-13622.
Technical Conference of the ASME ICE Division, 23 Borman, G. and Nishiwaki, K. Internal-combus-
Ann Arbor, Michigan, October 1999, ICE-Vol. 33-3, tion engine heat transfer. Prog. Energy Combust.
Paper 99-ICE-239. Sci., 1987, 13(1), 1–46.
8 Primus, R. J., Hoag, K. L., Flynn, P. F., and Brands,
24 Shaver, G. M., Roelle, M. J., Caton, P. A.,
M. C. An appraisal of advanced engine concepts
Kaahaina, N. B., Ravi, N., Hathout, J. P., Ahmed,
using second law analysis techniques. SAE Paper
J., Kojić, A., Park, S., Edwards, C. F., and Gerdes,
841287, 1984.
J. C. A physics-based approach to the control of
9 Flynn, P. F., Hoag, K. L., Kamel, M. M., and
homogeneous charge compression ignition en-
Primus, R. J. A new perspective on diesel engine
gines with variable valve actuation. Int. J. Engine
evaluation based on second law analysis. SAE
Res., 2005, 6(4), 361–375.
Paper 840032, 1984.
25 Smith, G. P., Golden, D. M., Frenklach, M.,
10 Alkidas, A. C. The use of availability and energy
Moriarty, N. W., Eiteneer, B., Goldenberg, M.,
balances in diesel engines. SAE Paper 890822, 1989.
Bowman, C. T., Hanson, R. K., Song, S., Gardiner,
11 Dunbar, W. R. and Lior, N. Sources of combustion
W. C., Jr, Lissianski, V. V., and Qin, Z. GRI-Mech
irreversibility. Combust. Sci. Technol., 1994, 103(1–
home page, at http://www.me.berkeley.edu/gri_
6), 41–61.
mech/, accessed 24 January 2006.
12 Caton, J. A. On the destruction of availability
(exergy) due to combustion processes – with 26 Goodwin, D. G. An open source, extensible soft-
specific application to internal-combustion en- ware suite for CVD process simulation. In Proceed-
gines. Energy, 2000, 25(11), 1097–1117. ings of Chemical Vapor Deposition XVI and
13 Chavannavar, P. S. and Caton, J. A. Destruction of EUROCVD 14, 2003, vol. 2003-08, pp. 155–162.
availability (exergy) due to combustion processes: a
parametric study. Proc. IMechE, Part A: J. Power
and Energy, 2006, 220(7), 655–668. APPENDIX 1
14 Van Gerpen, J. H. and Shapiro, H. N. Second-law
analysis of diesel engine combustion. J. Engng Gas Notation
Turbines Power, 1990, 112(1), 129–137.
15 Shapiro, H. N. and Van Gerpen, J. H. Two-zone R̄ universal gas constant,
combustion models for second law analysis of 8.314 kJ/kmol21 K21

Int. J. Engine Res. Vol. 9 JER01508 F IMechE 2008


Thermodynamic requirements for maximum IC engine cycle efficiency. Part 1 465

X exergy system, its associated fundamental thermodynamic


relation manifolds, and its variants are used to
Dr property change from reactants to illustrate the thermodynamic implications of mini-
products mum entropy generation due to combustion (Figs 1
to 5, and 7).
Figure 6 in section 3.3 shows how entropy genera-
Subscripts tion varies with the volume at which adiabatic
eq chemical equilibrium state (at con- combustion takes place for a hydrogen–air system
stant U, V) (with decreasing system mole number from reactants
f final (exhaust) state to products, SNj,P 2 SNj,R , 0), to be contrasted with
i initial state Fig. 5 for the above-mentioned propane–air system
j property of species j (with increasing mole number, SNj,P 2 SNj,R . 0). The
P product state hydrogen–air system is stoichiometric
R reactant state
0 environmental (dead) state H2 z0:5ðO2 z3:76N2 Þ

and is initially at temperature Ti 5 298 K, pressure


APPENDIX 2 Pi 5 1 atm.
Figure 7 in section 4 considers propane (denoted
Notes on the illustrative examples
here by subscript ‘1’) separately from the remaining
species listed in equation (14), i.e. stoichiometric
The thermodynamic analyses and discussions pre-
proportion of air with 50 per cent exhaust (subscript
sented in this paper are fundamental in nature, and
‘2’). The unmixed reactant manifold is thus con-
are not dependent on specific engine parameters,
structed by summing the internal energy, volume,
e.g. the choice of fuel, stoichiometry, or compression
and entropy for propane (U1, V1, S1) at fixed T1 5
ratio. For consistency, and without loss of generality,
25 uC, P1 5 1 atm, with those for the air–exhaust
most of the illustrative examples used are based on
mixture (U2, V2, S2) at varying T2, P2.
the same chemical system first considered in
reference [18]. (Parameters used in that paper, in The same approach applies when comparing
turn, approximate the single-cylinder homogeneous- entropy generated due to mixing versus combustion
charge, compression-ignition research engine para- in Fig. 8. The mixing entropy arises when separate
meters in reference [24].) components of propane and air (in stoichiometric
The system initially comprises a stoichiometric proportion, T 5 25 uC, P 5 1 atm) as well as combus-
propane–air mixture diluted with 50 per cent EGR tion products (varying residual fraction, T 5 25 uC or
(complete-combustion products only) 600 K, P 5 1 atm) are combined.
The NASA thermochemical database [19] is used
C3 H8 z5ðO2 z3:76N2 Þzð3CO2 z4H2 Oz18:8N2 Þ in all the calculations, as is the GRI-Mech 3.0 natural
ð14Þ gas combustion mechanism [25] when reaction
kinetics are involved, e.g. Figs 1 and 2. The calcula-
at temperature Ti 5 562 K and pressure Pi 5 1 atm. tions are carried out using Cantera, an open source
The initial temperature is chosen such that auto- program for chemical kinetics and thermodynamic
ignition occurs close to top centre upon adiabatic, simulations [26], in conjunction with the computing
13:1 slider-crank compression at 1800 r/min. This software MATLAB.

JER01508 F IMechE 2008 Int. J. Engine Res. Vol. 9

View publication stats

You might also like