Está en la página 1de 11

Applied Mathematical Modelling 34 (2010) 1710–1720

Contents lists available at ScienceDirect

Applied Mathematical Modelling


journal homepage: www.elsevier.com/locate/apm

Endoreversible radiative heat engine configuration for maximum efficiency


Lingen Chen *, Hanjiang Song, Fengrui Sun
Postgraduate School, Naval University of Engineering, Wuhan 430033, PR China

a r t i c l e i n f o a b s t r a c t

Article history: Endoreversible radiative heat engine configuration for maximum efficiency is studied in
Received 4 August 2008 this paper. The optimal configuration of a class of endoreversible heat engines with fixed
Received in revised form 5 September 2009 duration, input energy and radiative heat transfer law ðq / DðT 4 ÞÞ in the heat transfer pro-
Accepted 9 September 2009
cesses between the working fluid and the heat reservoirs is determined. The optimal cycle
Available online 17 September 2009
that maximizes the efficiency of the heat engine is obtained using optimal-control theory.
The deduced differential equations are solved by Taylor series expansion. The optimal cycle
Keywords:
has eight branches including two isothermal branches, four maximum-efficiency branches,
Maximum efficiency
Radiative heat transfer law
and two adiabatic branches. The time interval of each branch is obtained. The temperatures
Endoreversible heat engine of heat reservoirs and working fluid are also calculated. The obtained results are compared
Optimal-control theory with those obtained with the Newton’s heat transfer law and linear phenomenological heat
Optimal configuration transfer law for maximum efficiency objective, and those with radiative heat transfer law
Finite time thermodynamics for maximum power output objective.
Ó 2009 Elsevier Inc. All rights reserved.

1. Introduction

In the finite time thermodynamic analysis and optimization of energy systems, the standard problems include to
determine the objective function limits and the relations between objective functions for the given thermodynamic system,
and to determine the optimal thermodynamic process (configuration) for the given optimization objectives [1–11].
Cutowicz-Krusin et al. [12] proved that in all acceptable cycles, the Curzon–Ahlborn cycle [13] is the optimal configuration
with only First and Second Law constraints. Rubin [14,15] found the optimal configurations of endoreversible heat engines
with Newton’s heat transfer law and different constraints. The optimal configuration with fixed duration for maximum
power output is a six-branch cycle, and the optimal configuration with fixed energy input for maximum efficiency is an
eight-branch cycle [14]. The results were extended to a class of heat engines with fixed compression ratio, and an optimal
eight-branch cycle configuration of the engines for maximum power output is derived [15]. Since then, many researchers
found various optimal configurations for various systems and devices using optimal-control theory in the last three decades.
Recently, Badescu [16] studied the optimal heating paths with Newton’s and radiative heat transfer laws by taking the min-
imum entropy generation and minimum lost available work as the objectives. The obtained analytical expressions of the
optimal paths were presented in dimensionless forms. The similarities and differences between various heating strategies
under the two heat transfer laws were compared. Amelkin et al. [17,18] discussed the maximum power processes of mul-
ti-heat-reservoir heat engine with stationary temperature reservoirs, found that some reservoirs were not used in heat trans-
fer in order to achieve an optimal performance of the system, and further found that independent of the number of reservoirs
the working fluid used only two isotherms and two adiabatics. Tsirlin et al. [19] considered a stationary system that includes
a transformer of mechanical energy into heat energy or heat into mechanical energy, found that for a linear heat transfer law
and heat engine operating at maximum power the ratio of engine working body’s temperatures during contact with

* Corresponding author. Tel.: +86 27 83615046; fax: +86 27 83638709.


E-mail addresses: lgchenna@yahoo.com, lingenchen@hotmail.com (L. Chen).

0307-904X/$ - see front matter Ó 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.apm.2009.09.017
L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720 1711

reservoirs is equal to the square root of the ratio of reservoirs’ temperatures irrespective of the system’s structure and
whether the engine is internally irreversible or not. Huleihil and Andresen [20] considered optimal piston motion derived
for adiabatic process in the presence of different types of friction and with frictional energy dissipated externally as well
as internally, and found that in the externally dissipative mode optimizing power or frictional losses are equivalent, while
in the internally dissipative mode the optimal motions are different.
In general, heat transfer is not necessarily linear and also obeys other laws; heat transfer laws have the significant influ-
ences on the configuration and performance of heat engine cycles [21–36]. Song et al. [37,38] determined the optimal con-
figuration of an endoreversible heat engine for maximum efficiency objective and maximum power output objective with
linear phenomenological heat transfer law [37] and those for maximum power output with fixed duration and radiative heat
transfer law [38]. In this paper, based on Refs. [14,38], the optimal configuration of an endoreversible heat engine for max-
imum efficiency objective is obtained with fixed duration of cycle, fixed input energy and radiative heat transfer law
½q / DðT 4 Þ in the heat transfer processes between working fluid and heat reservoirs. The optimal cycle that maximizes
the efficiency of the engine is obtained using optimal-control theory. A numerical example is given. The obtained results
are compared with those obtained with the Newton’s heat transfer law for maximum efficiency objective, those with linear
phenomenological heat transfer law for maximum efficiency objective, and those with radiative heat transfer law for max-
imum power output objective.

2. Heat engine model

The following assumptions are made for the heat engine model:

(1) The engine is endoreversible.


(2) The heat conductivity between working fluid and reservoirs is q which subject to
0 6 q 6 q0 : ð1Þ
(3) The heat transfer between working fluid and reservoirs obeys radiative heat transfer law, the heat flux is

q_ ¼ qðT 4R  T 4 Þ; ð2Þ
where q is heat conductivity, T is the absolute temperature of the working fluid, T R is a constant temperature of each
heat reservoir, and
TL 6 TR 6 TH; ð3Þ
where T L and T H are the lower and upper limits of the temperature of heat reservoirs, respectively.
(4) The work done by the engine in one cycle is given by
Z s
W¼ P V_ dt; ð4Þ
0

where P and V are the pressure and volume of the working fluid, V_ is the time derivative of V, and s is cycle duration of
the engine. The working fluid is an ideal gas.

In terms of the first law of thermodynamics for an ideal gas, one has
_
C V T_ þ C V ðc  1ÞT V=V _
¼ q; ð5Þ
where C V is the isometric heat capacity of the gas, c is the ratio of isobaric heat capacity to C V , and T_ is the time deriv-
ative of T.
Substituting Eq. (2) into Eq. (5) and defining some new variables yields:
 
T_ ¼ CT þ q
^ T 4R  T 4 ; ð6Þ
b_ ¼ C; ð7Þ
b ¼ ðc  1Þ lnðV=V 0 Þ; ð8Þ
where q^ ¼ q=C V , V 0 is a constant reference volume, and C is the change rate of cylinder volume.
In terms of these variables, Eq. (4) becomes
Z s
W ¼ CV CT dt ð9Þ
0

Apparently, deriving b and C is convenient to solve the problem one confronts by using optimal-control theory. One also has
Z s    
Q 1 ¼ CV q^ T 4R  T 4 h T 4R  T 4 dt; ð10Þ
0
1712 L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720

where hðxÞ is Heaviside step function, h ¼ 1 if x > 0 and h ¼ 0 if x < 0, Q 1 is the heat be supplied to system, i.e. input energy.
The cycle efficiency is given by g ¼ W=Q 1 .
The problem now is to determine parameters q ^ ðtÞ, T R ðtÞ and CðtÞ so that the efficiency is a maximum with the fixed dura-
tion s and input energy Q 1 . It is required that C must be restricted such that
C m 6 C 6 C M ; ð11Þ
where C m and C M are arbitrary constant positive values. Then the optimal cycle of the engine for maximum efficiency can be
derived by the model.

3. Optimization procedure

To determine the optimal configuration for maximum efficiency with fixed cycle duration and input energy Q 1 is the
same as to determine the optimal configuration for maximum W  lQ 1 with fixed cycle duration, where l is an ordinary
Lagrange multiplier. The derivation of the optimal solution is complicated by the l term. One can show that
l ¼ @W max =@Q 1 , i.e., l is a measure of the sensitivity of W max with respect to small changes in the constraint, Q 1 ¼ const.
One can see that l can be both positive and negative and l vanishes for the maximum power output, as expected
[14,37,38]. To simplify the subsequent discussion, l P 0 is assumed.
The Hamiltonian function in terms of Eqs. (6), (7) and W  lQ 1 is
   
^ T 4R  T 4 h T 4R  T 4 þ w1 F 1 þ w2 F 2 ;
H ¼ CT  lq ð12Þ

where
 
^ T 4R  T 4 ;
F 1 ¼ CT þ q ð13Þ
F 2 ¼ C; ð14Þ
where T and b are state variables, and T R , q
^ and C are control variables.
Taking ðW  lQ 1 Þ=C V as the performance index, it is convenient to rewrite Eq. (12) as
h  i  
H ¼ ½ð1  W1 ÞT þ W2 C þ W1  lh T 4R  T 4 q^ T 4R  T 4 : ð15Þ

The equations of adjoint variables are given by


h  i
^ T 3 w1  lh T 4R  T 4 ;
w_ 1 ¼ @H=@T ¼ Cð1  w1 Þ þ 4q ð16Þ
w_ 2 ¼ @H=@V ¼ 0; ð17Þ
where
06q
^ 6 q0 =C V ¼ q
^0: ð18Þ

3.1. Application of maximum principle

Defining

DH ¼ H½~ u ðtÞ; ~
x ðtÞ; ~ w ðtÞ  H½~ u; ~
x ðtÞ; ~ w ðtÞ; ð19Þ
where ~
u is an admissible solution. The asterisk of symbols states the solutions are optimal. For a maximum it is required that
DH P 0, thus
   h  i   h  i  
DH ¼ 1  w1 T  þ w2 ðC   CÞ þ w1  l h T 4
R T
4
q^  T 4
R T
4
 w1  lh T 4R  T 4 q^ T 4R  T 4 P 0; ð20Þ

where 0 6 q ^ 6 q0 =C V ¼ q
^ 0 , T L 6 T R 6 T H and C m 6 C 6 C M .
Now one can consider various possible cases separately.
First, set q ^  and T R ¼ T R , then the second and the third terms of Eq. (20) vanish. To assure DH P 0 it is required
^ ¼q
that
8  
< CM ;
> if 1  w1 T  þ w2 > 0;
 
C ¼ C m ; if 1  w1 T  þ w2 < 0; ð21Þ
>
:  
undetermined; if 1  w1 T  þ w2 ¼ 0:
The last possibility corresponds to what is called the singular control problem, and for the problem it is easy to obtain that
along the singular part of the trajectory C  is constant.
L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720 1713

Next, set C ¼ C  and T R ¼ T R , then Eq. (20) becomes


h  i 
DH ¼ w1  l h T 4
R T
4
T 4
R T
4
^  q
ðq ^ Þ P 0: ð22Þ

Correspondingly, it is required that


8 h  i 
>
>
> q^ 0 ; if w1  l h T 4
R T
4
T 4
R T
4
> 0;
>
< h  i 
  4 4 4 4
q^ ¼ 0; if w1  l h T R  T

TR  T < 0; ð23Þ
>
> h  i 
>
>
: undetermined; if w  l h T 4  T 4 T 4 4
1 R R T ¼ 0:

The last possibility still corresponds to singular control problem. Because it would call for W1 ¼ l ¼ 1, further resulting in
H ¼ 0, which does not agree with the request of H > 0, one should exclude this possibility.
Finally, set C ¼ C  and q ^ ¼q^  , then Eq. (20) becomes
8   
>
<q^  w1  l T 4 4
if T R > T  ;
R  TR ;
DH ¼   ð24Þ
>
:q^  w1 T 4 4
if T R 6 T  :
R  TR ;

From Eq. (23), one can see that q ^ 0 requires W1 > l if T R > T  and W1 < 0 if T R < T  . Consider T R in the interval ½T  ; T H ,
^ ¼ q
for T R > T one can find T R ¼ T H in order that DH > 0. It then follows that DH P 0 for T R in the interval ½T L ; T   since l P 0. In
  

a similar fashion, when T R < T  , DH P 0 provides T R ¼ T L . Therefore, one has



T H ; if w1 > l ;
T R ¼ ð25Þ
TL; if w1 < 0:
And for q ^  – 0, one has T H > T > T L .
This problem could be changed into finding the optimal configuration for maximum efficiency. If l > 0; q^  ¼ 0 and
C  ¼ C, Eq. (20) can be changed into
h  i  
DH ¼  w1  l h T 4R  T 4 q^ T 4R  T 4 P 0: ð26Þ

From Eqs. (25) and (26), one can get T  > T H if w1 > l , and T  < T L if w1 < 0, so the both cases are impossible. And if
0 6 w1 6 l , T  is possible between T H and T L . From above argument, one can get that adiabatic branches are possible to take
place when l > 0. If l ¼ 0, whenever adiabatic branch vanished [38]. The problem would be more complex if l < 0, when-
ever the optimal configuration may be without adiabatic branches.

3.2. Optimal solutions

It is easy to find all possible optimal solutions now; one can obtain the optimal trajectories by solving the canonical func-
tions. All functions below are optimal. It is convenient to eliminate the asterisk of symbols.

(a) Adiabatic branches: q


^ ¼ 0, C ¼ C M or C ¼ C m

TðtÞ ¼ Tðt0 ÞeCðtt0 Þ ; bðtÞ ¼ bðt 0 Þ þ Cðt  t 0 Þ;


w1 ðtÞ ¼ 1  ½1  w1 ðt 0 ÞeCðtt0 Þ ; w2 ¼ const:; ð27Þ
H ¼ f½1  w1 ðtÞT þ w2 gC ¼ f½1  w1 ðt 0 ÞTðt 0 Þ þ w2 gC; ð28Þ
where H and w2 are constants as required and the value of C is determined by Eq. (21), t0 is initial time of branch.
(b) Maximum-efficiency branches: q ^ ¼q^ 0 , T R ¼ T H or T R ¼ T L , and C ¼ C M or C ¼ C m
 
^ 0 T 4R  T 4 ;
T_ ¼ CT þ q bðtÞ ¼ bðt 0 Þ þ Cðt  t0 Þ; ð29Þ
w_ 1 ¼ Cð1  w1 Þ þ 4q
^ 0 ½w1  l hðT 4R 4
 T ÞT ;3
w2 ðtÞ ¼ const:; ð30Þ
where the value of C is determined by Eq. (21) and the value of T R is determined by Eq. (25), t0 is initial time of branch.
Analytical solutions are seldom, thus one has to rest on numerical techniques to obtain all the solutions.
(c) Isothermal branches: q ^¼q ^ 0 , T R ¼ T H or T R ¼ T L , and ð1  w1 ÞT þ w2 ¼ 0
 
q^ 0 T 4R  T 4r
T ¼ Tr; bðtÞ ¼ bðt 0 Þ þ C r ðt  t 0 Þ; Cr ¼ ; ð31Þ
Tr
h   i h  i
T 4R þ T 4r 4lh T 4R  T 4r  1 4 1  lh T 4R  T 4r T 5r
w1 ¼ ; w2 ¼ ; ð32Þ
T 4R þ 3T 4r T 4R þ 3T 4r
1714 L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720

where T r is a constant. This is a singular case that has not been analyzed. It is easy to prove that C, T and w1 are constants by
differentiating ð1  w1 ÞT þ w2 ¼ 0 and eliminating the derivative of time using canonical functions.

The subscript r used above corresponds to R in T R , i.e. r ¼ h if T R ¼ T H and r ¼ l if T R ¼ T L . The value of T R is determined by
Eq. (25). It is shown from Eq. (32) that if T R ¼ T H , w1h > 0 implies T H > T h and if T R ¼ T L , w1l < 0 implies T L < T l . These in turn
imply that C h > 0 and C l < 0. It is easy to show that
 2 h  i
T 4R  T 4r 1  lh T 4R  T 4r
H¼q
^0 : ð33Þ
T 4R þ 3T 4r
From w2 and H are constants one has
 2  2
T 4H  T 4h ð1  lÞ T 4L  T 4l T 5h ð1  lÞ T 5l
¼ ; ¼ : ð34Þ
T 4H þ 3T 4h T 4L þ 3T 4l T 4H þ 3T 4h T 4L þ 3T 4l
When l vanishes, Eq. (34) is same as that of Ref. [38]. Substituting T h and T l into the equation of C r yields
   
q^ 0 T 4H  T 4h q^ 0 T 4L  T 4l
Ch ¼ ; Cl ¼ : ð35Þ
Th Tl
Thus there are eight different solutions of equations above, which are presented by 1 , 2 
H , 2L , 3H and 3L , respectively, where
symbol ‘‘+” refers to C ¼ C M and symbol ‘‘” to C ¼ C m and subscripts H and L correspond to the subscript of T R . If l ¼ 0, the
analytical results are the same as the optimal solutions of endoreversible heat engines for maximum power output objective
with fixed duration and radiative heat transfer law [38].
In order to determine the actual optimal trajectory, one must examine the constancy H and the continuity of the state
variables and costate variables at switchings between pairs of optimal solutions.

3.3. Switching

The surfaces in state variable phase space across which optimal-control variables change discontinuously are called
switching surface in the optimal-control theory. The switchings of the problem are summarized in Table 1.
If there be a switching between branches 1þ and 1 , Eq. (21) requires that ð1  w1 ÞT þ w2 varnishes, then Eq. (28) leads to
H ¼ 0. So there cannot be a switching between branches 1þ and 1 . Similarly, there is no switching between branches 3H and
3L because of the continuity of w1 . However, there may be switchings between T R ¼ T H and T R ¼ T L in case (b), in which C
remains constant and w1 passes through zero at the switching time. There may be switchings between C m and C M , in which
T R remains constant and ð1  w1 ÞT þ w2 vanishes. But T R and C cannot change continuously because of H – 0.
A switching between 1þ and 2þ þ
H is allowed by increasing w1 and making it across l, and a switching between 1 and 2L is
þ

allowed by decreasing w1 and making it across zero. The switching condition of each branch is shown below:

The switching condition between 3H and 2þH is ½1  w1 ðt 1 ÞTðt 1 Þ þ w2 ðt 1 Þ ¼ 0.


The switching condition between 2þ þ
H and 1 is w1 ðt 2 Þ ¼ l.
The switching condition between 1 and 2þ
þ
L is w1 ðt 2 Þ ¼ 0.
0

The switching condition between 2þ


L and 3L is ½1  w1 ðt 3 ÞTðt 3 Þ þ w2 ðt 3 Þ ¼ 0.
The switching condition between 3L and 2
L is ½1  w1 ðt 4 ÞTðt 4 Þ þ w2 ðt 4 Þ ¼ 0.
The switching condition between 2 
L and 1 is w1 ðt 5 Þ ¼ 0.
0

The switching condition between 1 and 2H is w1 ðt 5 Þ ¼ l.

3.4. Optimal controls and trajectory

The case discussed above is an autonomous system, i.e., invariant with respect to time translation, so one may choose any
point along optimal trajectory as a starting point. Here one assumes that it begins from branch 3H , i.e. T R ¼ T H and T ¼ T h , etc.

Table 1
Switchings.

1 2 3
1 a b or c a
2 b or c d d
3 a d a

a: Forbidden switchings.
b: Allowed switchings: DC ¼ 0, w1 ¼ 0.
c: Allowed switchings: DC ¼ 0, w1 ¼ l.
d: Allowed switchings: DT R ¼ 0, ð1  w1 ÞT þ w2 ¼ 0.
L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720 1715

For 0 6 t 6 t 1 , the only allowed switching is to a branch 2þ H , i.e. T R ¼ T H , C ¼ C M , etc. For t 1 6 t 6 t 2 , w1 decreases and
ð1  w1 ÞT þ w2 increases from zero, thus the only possible transition occurs at t2 when w1 ¼ 0, and one can obtain the branch
2þ þ
L . One should pay attention to that there is an adiabatic branch 1 between t 2 and t 2 , which is different to the optimal con-
0

figuration for maximum power objective [38].


From t 20 to t 3 , w1 continues to decrease and so does ð1  w1 ÞT þ w2 until it vanishes at which time another switch becomes
possible. At t3 one begins an isothermal branch 3L which lasts until the time t 4 when one switches to the branch 2 L . Along
this branch ð1  w1 ÞT þ w2 decreases from zero while w1 increases until it reaches zero at t 50 . There is also an adiabatic branch
1 between t50 and t 5 . Then one switches to branch 2 H until ð1  w1 ÞT þ w2 returns to zero at the end of the cycle t 6 ¼ s.
Consequently, one can obtain the solutions of the problem. Since w2 is constant throughout the cycle, one records it only
once.
The state and adjoint variables are given as followings:
For 0 6 t 6 t 1 :
T ¼ Th; C ¼ Ch; b ¼ C h t; q^ ¼ q^ 0 ; T R ¼ T H ; ð36Þ
T 4H þ T 4h ð4  l 1Þ 4ðl  1ÞT 5h
w1 ðtÞ ¼ ; w2 ðtÞ ; ð37Þ
3T 4h þ T 4H 3T 4h þ T 4H
where T h is determined by Eq. (34).
For t1 6 t 6 t2 :
Expanding TðtÞ with Taylor series at t1 gives
_ 1 Þðt  t1 Þ þ Oðt  t1 Þ:
TðtÞ ¼ Tðt1 Þ þ Tðt ð38Þ
Taking off high-order indefinitely-small Oðt  t1 Þ gives
_ 1 Þðt  t1 Þ:
TðtÞ  Tðt1 Þ þ Tðt ð39Þ
From the continuity of TðtÞ, one can obtain Tðt 1 Þ ¼ T h , and from
h i
_
TðtÞ ^ 0 T 4H ðtÞ  T 4 ðtÞ
¼ C M TðtÞ þ q ð40Þ

one can obtain


h  i
^ 0 T 4H  T 4h ðt  t 1 Þ:
TðtÞ  T h þ C M T h þ q ð41Þ

Expanding w1 ðtÞ with Taylor series at t1 , from the continuity of w1 ðtÞ and Eq. (16), gives

T 4H þ T 4h ð4l  1Þ ½4C M T 4h þ 4q
^ 0 T 3h ðT 4H  T 4h Þð1  lÞ
w1 ðtÞ  þ ðt  t1 Þ: ð42Þ
3T 4h þ T 4H 3T 4h þ T 4H
Combining all the analysis above yields: For t1 6 t 6 t2
h  i
TðtÞ  T h þ C M T h þ q ^ 0 T 4H  T 4h ðt  t 1 Þ; ð43Þ
h i
T 4H þ T 4h ð4l  1Þ 4C M T 4h þ 4q
^ 0 T 3h ðT 4H  T 4h Þ ð1  lÞ
w1 ðtÞ  þ ðt  t1 Þ; ð44Þ
3T 4h þ T 4H 3T 4h þ T 4H
b ¼ C M ðt  t 1 Þ þ C h t1 ; T R ¼ T H ; q ^ ¼q ^ 0; C ¼ CM : ð45Þ
From above process, the detail of each branch can be obtained. The branches of the maximum efficiency cycle are pre-
sented in detail in Appendix A. How the state variables vary can be seen in Fig. 1, which is similar to that of Ref. [14].
From the continuity of the cycle, one can obtain
Tðt6 Þ ¼ Tð0Þ; w1 ðt6 Þ ¼ w1 ð0Þ; bðt 6 Þ ¼ bð0Þ ¼ 0: ð46Þ
From the switching condition, one can obtain
w1 ðt2 Þ ¼ l; w1 ðt 20 Þ ¼ 0; w1 ðt 50 Þ ¼ 0; w1 ðt5 Þ ¼ l: ð47Þ
From the fixed cycle duration one can obtain
t1 þ ðt 4  t 3 Þ ¼ s  ðs  t 5 Þ  ðt 5  t 05 Þ  ðt 05  t 4 Þ  ðt 3  t 02 Þ  ðt 02  t 2 Þ  ðt 2  t 1 Þ: ð48Þ
From the fixed input energy, one can obtain
Z s    
Q 1 ¼ CV q^ T 4R  T 4 h T 4R  T 4 dt ¼ const: ð49Þ
0

Making discretization on Eq. (49) yields


1716 L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720

Fig. 1. State variables T (temperature of working fluid) and b (proportional to the volume of the cylinder) versus time for maximum efficiency.

(
4 ) (
4 )
Q1   T h þ Tðt 2 Þ T h þ Tðt 5 Þ
 t1 T 4H  T 4h þ ðt2  t1 Þ T 4H  þ ðt 6  t 5 Þ T 4H  : ð50Þ
CV q
^0 2 2

Combining Eqs. (34), (35), (46)–(48) and (50), one can solve out numerical solutions of all the intervals of each branch, i.e. t 1 ,
ðt 2  t 1 Þ, ðt20  t 2 Þ, ðt 3  t 20 Þ, ðt4  t 3 Þ, ðt 50  t4 Þ, ðt5  t50 Þ, and ðs  t5 Þ, as well as T h , T l , C h , C l and the Lagrange multiplier l.
Substituting them into equations of the temperatures of working fluid, one can solve out numerical solutions of Tðt 1 Þ,
Tðt2 Þ, T ðt 20 Þ, Tðt3 Þ, Tðt 4 Þ, T ðt 50 Þ, Tðt 5 Þ and Tðt6 Þ. The equations are presented in detail in Appendix B.
Maximum efficiency of the cycle is
Rs
W C V 0 CT dt
g¼ ¼ : ð51Þ
Q1 Q1

4. Numerical example

Now, a numerical example for the optimal configuration of the heat engine is provided. In the numerical calculations,
^ 0 ¼ 108 kg=ðK3 sÞ, C V ¼ 5 kJ=ðkg KÞ, and the mass of working fluid is 1 kg,
T H ¼ 1000 K, T L ¼ 400 K, C M ¼ 12:5, C m ¼ 3, q

Table 2
The parameters versus Q 1 .

t Q 1 ¼ 6450 kJ Q 1 ¼ 6500 kJ Q 1 ¼ 6550 kJ


Dt ðsÞ T ðKÞ b Dt ðsÞ T ðKÞ b Dt ðsÞ T ðKÞ b
t1 0.337818 904.0000 1.2385 0.338217 904.0000 1.2400 0.338612 904.0000 1.2414
t2 0.016587 771.6614 1.4458 0.017333 765.7131 1.4566 0.018071 759.8249 1.4673
t20 0.014300 645.3519 1.6246 0.014300 640.3773 1.6354 0.014300 635.4529 1.6461
t3 0.015200 634.0000 1.8146 0.015200 634.0000 1.8254 0.015200 634.0000 1.8361
t4 0.532400 634.0000 0.6658 0.532400 634.0000 0.6766 0.532400 634.0000 0.6873
t50 0.037018 654.0755 0.5548 0.037008 654.0699 0.5656 0.036998 654.0644 0.5763
t5 0.036935 730.7184 0.4440 0.036898 730.6304 0.4549 0.036861 730.5436 0.4657
t6 0.015130 904.0000 0 0.015793 904.0000 0 0.016452 904.0000 0
l 0.013747 0.013747 0.013747
W ðkJÞ 3139.3 3166.8 3193.8
g ¼ W=Q 1 0.4867 0.4872 0.4876

^ 0 ¼ 108 kg=ðK3 sÞ;


T H ¼ 1000 K; T L ¼ 400 K; C M ¼ 12:5; C m ¼ 3; C V ¼ 5 kJ=ðkg KÞ; q s ¼ 1 s.
L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720 1717

Fig. 2. The cycle of eight branches for maximum efficiency objective.

Fig. 3. The cycle of six branches for maximum power objective ðl ¼ 0Þ [38].

are set. Based on the method mentioned above, one can obtain all the numerical solutions. Table 2 lists the process-times of
each branch, the values of the state variables at switchings, the work output per cycle, i.e. W and the maximum efficiency, i.e.
g. Fig. 2 shows T  b diagram with Q 1 ¼ 6500 kJ. There are two solutions for this problem, but only one is reasonable, and
another includes negative solutions.
In this example, time is mostly spent on two isothermal branches, which is similar to that of the optimal configuration for
maximum efficiency objective with fixed cycle period, fixed input energy and linear phenomenological heat transfer law
[37], as well as that of the optimal configuration for maximum power output objective with fixed cycle period and radiative
heat transfer law [38] (Fig. 3 shows T  b diagram with l ¼ 0, i.e., the optimal configuration for maximum power output
objective with fixed cycle period [38]). The change of Q 1 just has a little influence on the intervals of each branch, the tem-
peratures of working fluid at the most switchings and the Lagrange multiplier l. It is shown that with the increase of Q 1 , the
work output W and the maximum efficiency g ¼ W=Q 1 increase, which is similar to that of Ref. [37], and the temperatures of
working fluid at t2 and t20 decrease. Comparing Fig. 2 with Fig. 3, one can see that when Lagrange multiplier l vanishes, the
optimal configuration changes from eight branches to six branches [38], i.e., both adiabatic branches vanish, which is similar
to that of Ref. [14], but the highest temperature and the lowest temperature hardly alter.

5. Conclusion

The optimal configuration of an endoreversible heat engine for maximum efficiency objective with fixed duration of the
cycle, fixed input energy and radiative heat transfer law has eight branches including two isothermal branches, four maxi-
mum-efficiency branches, which give maximum efficiency, and two adiabatic branches.
1718 L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720

The obtained results can be compared with those results of Ref. [14] for maximum efficiency objective with Newton’s
heat transfer law and the results of Ref. [37] for maximum efficiency objective with linear phenomenological heat transfer
law. The similarities and differences of optimal cycles among three heat transfer laws are given below: the optimal cycles
with three heat transfer laws all contain two isothermal branches, four maximum-efficiency branches and two adiabatic
branches; for three heat transfer laws, the temperatures of their isothermal branches are different, and the process paths
of four maximum-efficiency branches are different as well; the process-time is also different for different branch under three
different heat transfer laws. Since both the process-path and the process-time are different under three heat transfer laws,
the maximum efficiency for the three configurations are different.
The obtained results can also be compared with those results of Ref. [38] for maximum power output objective with radi-
ative heat transfer law. One can see that because of the introduction of the Lagrange multiplier l, the problem becomes com-
plex, the optimal configuration changes from six branches into eight branches, and the process-path and the process-time
are also different to the case of six branches. The optimal configuration is the same as the case of six branches with same
heat transfer law, if and only if l ¼ 0.
Because of the existence of differential equations, analytical solutions are seldom, one has to rest on numerical techniques
to obtain all the solutions, and Taylor series are used to simplify the differential equations. First-order item of Taylor series
was used in this paper, and the solutions would be more exact if high-order items of Taylor series can be used.
This paper made a theoretical investigation for a class of endoreversible heat engines. It would be helpful to compare the
result obtained with real world process behavior. It should be a further step for deep study in the future.

Acknowledgements

This paper is supported by the Program for New Century Excellent Talents in University of PR China (Project No.
20041006) and The Foundation for the Author of National Excellent Doctoral Dissertation of PR China (Project No.
200136). The authors wish to thank the editor of the journal and the reviewers for their careful, unbiased and constructive
suggestions, which led to this revised manuscript.

Appendix A

The state and adjoint variables are given by:


For 0 6 t 6 t 1 :

T ¼ Th; C ¼ Ch; b ¼ C h t; TR ¼ TH; q^ ¼ q^ 0 ; ðA1Þ

T 4H þ T 4h ð4l  1Þ 4ðl  1ÞT 5h


w1 ðtÞ ¼ ; w2 ðtÞ : ðA2Þ
3T 4h þ T 4H 3T 4h þ T 4H
For t1 6 t 6 t2 :
h  i
^ 0 T 4H  T 4h ðt  t 1 Þ;
TðtÞ  T h þ C M T h þ q ðA3Þ

b ¼ C M ðt  t1 Þ þ C h t1 ; TR ¼ TH; q ^¼q ^ 0; C ¼ CM ; ðA4Þ


h  i
T 4H þ T 4h ð4l  1Þ 4C M T 4h þ 4q
^ 0 T 3h T 4H  T 4h ð1  lÞ
w1 ðtÞ  þ ðt  t1 Þ: ðA5Þ
3T 4h þ T 4H 3T 4h þ T 4H
For t2 6 t 6 t20 :

TðtÞ ¼ Tðt2 ÞeC M ðtt2 Þ ; b ¼ C M ðt  t 1 Þ þ C h t 1 ; w1 ðtÞ ¼ 1  ð1  lÞeC M ðtt2 Þ ; C ¼ CM ; q^ ¼ 0; ðA6Þ


where
h  i
^ 0 T 4H  T 4h ðt2  t 1 Þ:
T ð2Þ ðt2 Þ  T h þ C M T h þ q ðA7Þ

For t20 6 t 6 t3 :
n  o
TðtÞ  Tðt2 ÞeC M ðt20 t2 Þ þ C M Tðt2 ÞeCM ðt20 t2 Þ þ q
^ 0 T 4L  T 4 ðt 2 Þe4C M ðt20 t2 Þ ðt  t 20 Þ; ðA8Þ

b ¼ C M ðt  t1 Þ þ C h t1 ; w1 ðtÞ  C M ðt  t 20 Þ; TR ¼ TL; q^ ¼ q^ 0 ; C ¼ C M : ðA9Þ


For t3 6 t 6 t4 :
L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720 1719

T ¼ Tl; b ¼ C M ðt 3  t 1 Þ þ C h t 1 þ C l ðt  t 3 Þ; ðA10Þ
T 4L  T 4l
w1 ¼ ; TR ¼ TL; q^ ¼ q^ 0 ; C ¼ C l : ðA11Þ
3T 4l þ T 4L
For t4 6 t 6 t50 :
h  i
^ 0 T 4L  T 4l ðt  t4 Þ;
TðtÞ  T l þ C m T l þ q ðA12Þ
 
T 4L  T 4l 4C m T 4l þ 4q
^ 0 T 4L  T 4l T 3l
w1 ðtÞ  þ ðt  t4 Þ; ðA13Þ
3T 4l þ T 4L 3T 4l þ T 4L

b ¼ C l ðt 4  t3 Þ þ C M ðt 3  t 1 Þ þ C h t1  C m ðt  t4 Þ; TR ¼ TL; q^ ¼ q^ 0 ; C ¼ C m : ðA14Þ
For t05 6 t 6 t5 :
0
TðtÞ ¼ T ð5 Þ ðt50 ÞeC m ðtt50 Þ ; w1 ðtÞ ¼ 1  eC m ðtt50 Þ ;
ð5Þ
b ¼ C M ðt3  t 1 Þ þ C h t 1 þ C l ðt 4  t 3 Þ  C m ðt  t4 Þ;
q^ ¼ 0; C ¼ C m ; ðA15Þ
where
h  i
^ 0 T 4L  T 4l ðt 50  t 4 Þ:
Tðt50 Þ  T l þ C m T l þ q ðA16Þ

For t5 6 t 6 t6 ¼ s:
n h io
^ 0 T 4H  T 4 ðt 50 Þe4Cm ðt5 t50 Þ ðt  t5 Þ;
TðtÞ ¼ Tðt50 ÞeCm ðt5 t50 Þ þ C m Tðt 50 ÞeC m ðt5 t50 Þ þ q ðA17Þ
b ¼ C M ðt3  t1 Þ þ C h t 1 þ C l ðt 4  t 3 Þ  C m ðt  t4 Þ; w1  l þ C m ð1  lÞðt  t 5 Þ; TR ¼ TH; q^ ¼ q^ 0 ; C ¼ C m :
ðA18Þ

Appendix B

One can obtain all the numerical solutions from the equations below:
n h io
^ 0 T 4H  T 4 ðt50 Þe4C m ðt5 t50 Þ ðt6  t 5 Þ;
T h  Tðt 50 ÞeC m ðt5 t50 Þ þ C m Tðt 50 ÞeC m ðt5 t50 Þ þ q ðB1Þ

h i  
l þ C m ð1  lÞðt6  t5 Þ  T 4H þ T 4h ð4l  1Þ = 3T 4h þ T 4H ; ðB2Þ

C M ðt 3  t 20 Þ þ C M ðt 20  t2 Þ þ C M ðt 2  t1 Þ þ C h t1 þ C l ðt 4  t 3 Þ  C m ðt6  t 5 Þ  C m ðt5  t 50 Þ  C m ðt 50  t 4 Þ ¼ 0; ðB3Þ


h  i
T 4H þ T 4h ð4l  1Þ 4C M T 4h þ 4q
^ 0 T 3h T 4H  T 4h ð1  lÞ
þ ðt 2  t 1 Þ  l; ðB4Þ
3T 4h þ T 4H 3T 4h þ T 4H

1  ð1  lÞeC M ðt20 t2 Þ ¼ 0; ðB5Þ


 
T 4L  T 4l 4C m T 4l þ 4q
^ 0 T 4L  T 4l T 3l
þ ðt50  t 4 Þ  0; ðB6Þ
3T 4l þ T 4L 3T 4l þ T 4L

1  eC m ðt5 t50 Þ ¼ l; ðB7Þ

t1 þ ðt 4  t 3 Þ ¼ s  ðs  t 5 Þ  ðt 5  t 50 Þ  ðt 50  t4 Þ  ðt3  t20 Þ  ðt 20  t2 Þ  ðt 2  t1 Þ; ðB8Þ

(

4 ) (
4 )
Q1   T h þ Tðt2 Þ T h þ Tðt 5 Þ
4 4 4 4
 t 1 T H  T h þ ðt 2  t1 Þ T H  þ ðt6  t 5 Þ T H  ; ðB9Þ
CV q
^0 2 2

 
q^ 0 T 4H  T 4h
Ch ¼ ; ðB10Þ
Th
1720 L. Chen et al. / Applied Mathematical Modelling 34 (2010) 1710–1720

 
q^ 0 T 4L  T 4l
Cl ¼ ; ðB11Þ
Tl
 2  2
T 4H  T 4h ð1  lÞ T 4L  T 4l
¼ ; ðB12Þ
T 4H þ 3T 4h T 4L þ 3T 4l

T 5h ð1  lÞ T 5l
¼ : ðB13Þ
T 4H þ 3T 4h T 4L þ 3T 4l

References

[1] B. Andresen, R.S. Berry, M.J. Ondrechen, P. Salamon, Thermodynamics for processes in finite time, Acc. Chem. Res. 17 (8) (1984) 266–271.
[2] S. Sieniutycz, J.S. Shiner, Thermodynamics of irreversible processes and its relation to chemical engineering: second law analyses and finite time
thermodynamics, J. Non-Equilib. Thermodyn. 19 (4) (1994) 303–348.
[3] M. Feidt, Thermodynamique et Optimisation Energetique des Systems et Procedes, second ed., Technique et Documentation, Lavoisier, Paris, 1996.
[4] A. Bejan, Entropy generation minimization: the new thermodynamics of finite-size devices and finite-time processes, J. Appl. Phys. 79 (3) (1996) 1191–
1218.
[5] R.S. Berry, V.A. Kazakov, S. Sieniutycz, Z. Szwast, A.M. Tsirlin, Thermodynamic Optimization of Finite Time Processes, Wiley, Chichester, 1999.
[6] L. Chen, C. Wu, F. Sun, Finite time thermodynamic optimization or entropy generation minimization of energy systems, J. Non-Equilib. Thermodyn. 24
(4) (1999) 327–359.
[7] P. Salamon, J.D. Nulton, G. Siragusa, T.R. Andresen, A. Limon, Principles of control thermodynamics, Int. J. Energy 26 (3) (2001) 307–319.
[8] K.H. Hoffman, J. Burzler, A. Fischer, M. Schaller, S. Schubert, Optimal process paths for endoreversible systems, J. Non-Equilib. Thermodyn. 28 (3) (2003)
233–268.
[9] A. Durmayaz, O.S. Sogut, B. Sahin, H. Yavuz, Optimization of thermal systems based on finite-time thermodynamics and thermoeconomics, Prog.
Energy Combust. Sci. 30 (2) (2004) 175–217.
[10] L. Chen, F. Sun, Advances in Finite Time Thermodynamics: Analysis and Optimization, Nova Science Publishers, New York, 2004.
[11] T. Yilmaz, Y. Ust, A. Erdil, Optimum operating conditions of irreversible solar driven heat engines, Renew. Energy 31 (9) (2006) 1333–1342.
[12] D. Cutowicz-Krusin, J. Procaccia, J. Ross, On the efficiency of rate process: power and efficiency of heat engines, J. Chem. Phys. 69 (9) (1978) 3898–3906.
[13] F.L. Curzon, B. Ahlborn, Efficiency of a Carnot engine at maximum power output, Am. J. Phys. 43 (1) (1975) 22–24.
[14] M.H. Rubin, Optimal configuration of a class of irreversible heat engines, Phys. Rev. A 19 (3) (1979) 1272–1276.
[15] M.H. Rubin, Optimal configuration of an irreversible heat engine with fixed compression ratio, Phys. Rev. A 22 (4) (1980) 1741–1752.
[16] V. Badescu, Optimal paths for minimizing lost available work during usual heat transfer process, J. Non-Equlib. Thermodyn. 29 (1) (2004) 53–73.
[17] S.A. Amelkin, B. Andresen, J.M. Burzler, K.H. Hoffmann, A.M. Tsirlin, Maximum power process for multi-source endoreversible heat engines, J. Phys. D:
Appl. Phys. 37 (9) (2004) 1400–1404.
[18] S.A. Amelkin, B. Andresen, J.M. Burzler, K.H. Hoffmann, A.M. Tsirlin, Thermo-mechanical systems with several heat reservoirs: maximum power
processes, J. Non-Equlib. Thermodyn. 30 (2) (2005) 67–80.
[19] A.M. Tsirlin, V. Kazakov, A.A. Ahremenkov, N.A. Alimova, Thermodynamic constraints on temperature distribution in a stationary system with heat
engine or refrigerator, J. Phys. D: Appl. Phys. 39 (19) (2006) 4269–4277.
[20] M. Huleihil, B. Andresen, Optimal piston trajectories for adiabatic process in the presence of friction, J. Appl. Phys. 100 (11) (2006) 114914.
[21] A. De Vos, Efficiency of some heat engines at maximum power conditions, Am. J. Phys. 53 (6) (1985) 570–573.
[22] C. Wu, Power optimization of a finite-time solar radiant heat engine, Int. J. Ambient Energy 10 (3) (1989) 145–150.
[23] C. Wu, Optimal power from a radiating solar-powered thermionic engine, Energy Convers. Manage. 33 (4) (1992) 279–282.
[24] M. Feidt, I. Philippi, Basis of a general approach for finite time thermodynamics applied to two heat reservoir machines, ECOS’92, 1992, pp. 21–25.
[25] S. Goktun, S. Ozkaynak, H. Yavuz, Design parameters of a radiative heat engine, Int. J. Energy 18 (6) (1993) 651–655.
[26] L. Chen, F. Sun, C. Wu, Influence of heat transfer law on the performance of a Carnot engine, Appl. Thermal Eng. 17 (3) (1997) 277–282.
[27] F. Angulo-Brown, R. Paez-Hernandez, Endoreversible thermal cycle with a nonlinear heat transfer law, J. Appl. Phys. 74 (4) (1993) 2216–2219.
[28] L. Chen, F. Sun, C. Wu, Optimal expansion of a heated working fluid with phenomenological heat transfer, Energy Convers. Manage. 39 (3/4) (1998)
149–156.
[29] L. Chen, S. Zhou, F. Sun, C. Wu, Optimal configuration and performance of heat engines with heat leak and finite heat capacity, Open Syst. Inform.
Dynam. 9 (1) (2002) 85–96.
[30] L. Chen, X. Zhu, Sun, C. Wu, Optimal configurations and performance for a generalized Carnot cycle assuming the heat transfer law Q / ðDTÞm , Appl.
Energy 78 (3) (2004) 305–313.
[31] S. Sieniutycz, P. Kuran, Nonlinear models for mechanical energy production in imperfect generators driven by thermal or solar energy, Int. J. Heat Mass
Transfer 48 (3) (2005) 719–730.
[32] L. Chen, F. Sun, C. Wu, Optimal configuration of a two-heat-reservoir heat-engine with heat leak and finite thermal capacity, Appl. Energy 83 (2) (2006)
71–81.
[33] L. Chen, X. Zhu, F. Sun, C. Wu, Effect of mixed heat resistance on the optimal configuration and performance of a heat-engine cycle, Appl. Energy 83 (6)
(2006) 537–544.
[34] S. Sieniutycz, P. Kuran, Modeling thermal behavior and work flux in finite-rate systems with radiation, Int. J. Heat Mass Transfer 49 (17–18) (2006)
3264–3283.
[35] M. Huleihil, B. Andresen, Convective heat transfer law for an endoreversible engine, J. Appl. Phys. 100 (1) (2006) 014911.
[36] M. Feidt, M. Costea, C. Petre, S. Petrescu, Optimization of direct Carnot cycle, Appl. Thermal Eng. 27 (5–6) (2007) 829–839.
[37] H. Song, L. Chen, J. Li, F. Sun, C. Wu, Optimal configuration of a class of endoreversible heat engines with linear phenomenological heat transfer law
½q / DðT 1 Þ, J. Appl. Phys. 100 (12) (2006) 124907.
[38] H. Song, L. Chen, F. Sun, Endoreversible heat engines for maximum power output with fixed duration and radiative heat-transfer law, Appl. Energy 84
(4) (2007) 374–388.

También podría gustarte