You are on page 1of 15

INTERNATIONAL JOURNAL OF ENERGY RESEARCH, VOL.

21, 113—127 (1997)

ANALYSIS OF A CONCEPT FOR INCREASING THE EFFICIENCY


OF A BRAYTON CYCLE VIA ISOTHERMAL HEAT ADDITION

J. VECCHIARELLI*, J. G. KAWALL AND J. S. WALLACE


Department of Mechanical and Industrial Engineering, University of Toronto, Toronto, Ontario, Canada M5S 1A4

SUMMARY
A thermodynamic study indicates that the hypothetical modification of gas-turbine engines to include two heat additions
rather than one may result in significant efficiency improvements of over 4% compared with conventional engines.
Specifically, the usual constant pressure heat addition would be constrained to a given temperature and then further heat
addition carried out in a manner approaching an isothermal process. Owing to the limited peak combustion temperature
of the overall heat addition process, the emissions of NO may be reduced by as much as 50%, thus offering an
x
environmental benefit as well as an efficiency advantage. This paper details the analysis of a proposed combustion
chamber in which an isothermal heat addition is approximated. The combustion chamber would consist of a converging
duct featuring discrete combustion sites positioned along the streamwise direction. A numerical analysis developed to
assess the deviation from isothermal flow in the combustion chamber shows that a reasonable approximation of such
a heat addition may be possible with two or more combustion sites. Moreover, a simplified treatment of the combustion
process implies that flame stabilization at these sites is feasible. ( 1997 by John Wiley & Sons, Ltd.
Int. J. Energy Res., vol. 21, 113—127 (1997)
(No. of Figures: 10 No. of Tables: 1 No. of Refs: 15)
KEY WORDS: gas-turbine engines; isothermal heat addition; thermal efficiency; flame stabilization

1. INTRODUCTION
Thermal power cycles provide a significant amount of the world’s power requirements. A common type of
power plant that operates on the basis of a thermal cycle is the gas-turbine engine. This type of engine is
utilized widely in applications such as aircraft propulsion, gas and oil pumping, ship propulsion, and
electricity generation. Thermal efficiency is an important performance characteristic of thermal cycles and, in
practice, has a major impact on operating costs. For instance, a 1% increase in overall efficiency in
a gas-turbine engine can translate into significant economic benefits.
In theory, the most efficient thermal cycle is the Carnot cycle, which features an isothermal heat addition at
the maximum allowable temperature. The present study examines a proposed combustion chamber (for
gas-turbine engines) in which an isothermal heat addition is approximated. The implementation of this
proposal may result in enhanced engine performance; moreover, it is postulated that the concentration of
NO and other undesirable gaseous pollutants in exhaust gases may be substantially reduced (Vecchiarelli,
x
1991). The combustion chamber in question involves flame stabilization by means similar to those employed
in afterburners and ramjets. Thermodynamically, it mimics a series of reheat processes resembling those
performed in gas-turbine engines used for large stationary applications.
Approximations of isothermal heat additions have been attempted in the past. In 1892, Rudolf Diesel
proposed a reciprocating engine that featured an isothermal combustion process (Cummins, 1989). However,

* Correspondence to: J. Vecchiarelli, Department of Mechanical and Industrial Engg., University of Toronto, Toronto, Ontario,
Canada M5S 1A4.

CCC 0363-907X/97/020113—15 Received 29 July 1993


( 1997 by John Wiley & Sons, Ltd. Revised 5 July 1995
114 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Figure 1. Brayton cycle

Figure 2. Brayton cycle with reheating

in theory the engine (which was supposed to operate on a modified Carnot cycle) had an unacceptably low
power output. Moreover, numerous problems were encountered in an experimental engine which was
developed to test the proposal. Consequently, Diesel abandoned the concept of isothermal combustion.
Currently, isothermal heat additions are approximated in some gas-turbine engines. These engines usually
operate based on the Brayton cycle, which can be represented on a temperature—entropy plot as illustrated in
Figure 1. ¹ and ¹ represent the temperature extremes which a working fluid experiences in a real
.*/ .!9
thermal cycle. Reheat processes (expansions followed by heat additions), typically two or three, are
sometimes performed in an effort to improve thermal efficiency. With reference to Figure 2, as the number of
reheats between states 3 and 4 is increased, a better approximation of an isothermal heat addition is achieved,
thereby improving efficiency. However, each reheat requires additional equipment that renders the system
more complex and costly. This practical consideration limits the extent to which an isothermal heat addition
is approached. Furthermore, due to its associated difficulties, reheating in gas-turbine engines is normally
justified only for large stationary installations (Wood, 1982).
The concept of the present proposal, in which an isothermal heat addition is attempted, is based on
(a) simple heat addition or Rayleigh flow; and
(b) simple area change or isentropic flow (Vecchiarelli, 1989).
With respect to simple heat addition, when a compressible gas with subsonic velocity flows through a friction-
less constant area duct with heat addition, the temperature of the gas in general increases along the duct. With
respect to simple area change, when a compressible gas with subsonic velocity flows through a frictionless
adiabatic duct with area change, the temperature of the gas decreases along the duct. Theoretically, based on
the nature of these two flows, simple heat addition and simple area change may be combined in such a way so
as to yield an isothermal process. The idealized isothermal process consists of a compressible gas with
subsonic velocity flowing through a frictionless converging duct while heated such that, all along the duct,
BRAYTON CYCLE EFFICIENCY INCREASE 115

Figure 3. Modified Brayton cycle

any infinitesimal decrease in temperature due to simple area change is exactly compensated by simple heat
addition. It is noted that since the temperature of the gas is constant during the heat addition, the kinetic
energy of the gas (and hence, its Mach number) must increase in order to satisfy conservation of energy.
As will be discussed later, the appropriate application of the idealized isothermal process is to gas-turbine
engines operating with air. Accordingly, the Brayton cycle was hypothetically modified by an isothermal
process. The resulting thermal power cycle is shown in Figure 3, where ¹ designates the temperature at
*40
which the idealized isothermal process occurs. In order to determine the potential merit of the modified
Brayton cycle, a thermodynamic study was conducted in which the performance of this cycle was compared
to that of the Brayton cycle. As part of the study, two constraints (arising from practical considerations) were
imposed on the modified Brayton cycle: the subsonic constraint and the temperature constraint.
The subsonic constraint restricts the flow to be subsonic throughout the modified Brayton cycle. In
particular, during the idealized isothermal process from state 3 to state 4 (Figure 3) the Mach number (M) of
the gas can increase substantially. This constraint is necessary since flow devices such as turbines cannot
operate efficiently if the flow is supersonic (unless specialized design is undertaken). Hence, the value of M at
state 4 (the turbine inlet) should not exceed unity.
The temperature constraint prevents the equipment involved in the modified Brayton cycle from attaining
a temperature in excess of the metallurgical limit. This constraint is applied specifically to the stationary
blades at the turbine inlet (state 4), since the total or stagnation temperature of the gas is highest at this point.
The stagnation temperature of the gas consists of a static component and a dynamic component. As the gas
stream flows around a stationary blade, the temperature at the stagnation point is the sum of the static
component and a fraction of the dynamic component. This fraction, termed the recovery factor (RF), depends
on the shape of the blade and is usually close to 0·85 (Morris, 1973). Therefore, for any given value of RF, the
temperature constraint restricts the temperature at the stagnation point of a stationary blade from exceeding
the metallurgical limit. This temperature is approximately 1200 K (Harman, 1981).
The results of the thermodynamic study indicate that, for certain operating conditions, the modified
Brayton cycle has significantly better performance characteristics than the Brayton cycle does (Vecchiarelli
et al. 1991). For instance, with RF"0·8 and ¹ "1104 K (hence, M "0·8), the modified Brayton cycle can
*40 4
operate with a thermal efficiency over 4% greater than that of the basic Brayton cycle while yielding an equal
amount of work. As mentioned previously, even a 1% increase in efficiency is economically advantageous.
Thus, the implementation of the idealized isothermal process in gas-turbine engines may have potential
benefits. However, it should be emphasized that achieving these benefits depends on how well the isothermal
process is approximated in practice.
In principle, the idealized isothermal process could be accomplished with any compressible gas, heated
externally along a slowly converging duct. However, this would present practical difficulties, such as the large
size and complex design of a heat exchanger to carry out the proper heat addition. Instead, it may be
convenient and practical to let air be the compressible gas and to let combustion of an injected fuel supply the
heat addition. For example, it is suggested that a reasonable approximation of the idealized isothermal
process may be achieved in the combustion chamber depicted in Figure 4. In this Figure, D is the inlet
*/
116 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Figure 4. Schematic of proposed combustion chamber

diameter of the combustion chamber, M and M are the inlet and outlet Mach numbers respectively, x is
*/ 065
the distance along the combustion chamber, x is the location of a heat addition, ¸ is the span of a heat
q q
addition, and q@ is a heat flux distribution in energy per unit mass per unit length.
Essentially, the proposed combustion chamber consists of a converging duct featuring discrete combustion
sites separated along the streamwise direction. (In Figure 4, each combustion site is represented by
a parabolic heat flux distribution). It is hypothesized that by controlling the number of heat additions as well
as the location and amount of each heat addition in the proposed combustion chamber one may obtain
a good approximation of an isothermal process. Thus, the function of the proposed combustion chamber is
to perform an approximate isothermal process. Since the velocities developed in the combustion chamber are
relatively high, the appropriate method of combustion involves flame stabilization by means similar to those
employed in afterburners and ramjets, such as bluff bodies. Therefore, the proposed combustion chamber is
best suited for gas-turbine engines using air.
It is noted that the desired isothermal temperature (&1100 K) in the proposed combustion chamber is
less than the peak temperature (&1200 K) in conventional systems. Since the rate of production of
gaseous pollutants, such as NO , increases dramatically with an increase in the maximum temperature
x
attained in the combustion chamber, it is postulated that the NO and other gaseous pollutant emissions
x
may be greatly reduced with the use of the proposed combustion chamber. For instance, data on NO
x
emissions (Munt and Danielson, 1976) taken from a conventional gas-turbine engine indicate that the
amount of NO produced may be decreased by more than 50% if the combustor inlet temperature is reduced
x
by 100 K over the range 300 K to 800 K. Assuming that the specific heats of the gases remain constant,
a 100 K reduction in the inlet temperature will result in a 100 K reduction in the peak combustor
temperature. Hence, over a wide range of peak temperatures, the amount of NO produced could be
x
decreased by a factor on the order of 50% when the maximum temperature achieved in the combustion
chamber is reduced by 100 K.

2. ANALYSIS
In theory, the idealized isothermal process has been shown to increase thermal efficiency. However, it is
necessary to assess how closely this process is approached in the proposed combustion chamber. In order to
do so, the temperature variation in the combustion chamber is needed. This is obtainable from an analysis of
the approximate isothermal process. The assumptions for such an analysis (which also accounts for the effects
of friction) are listed as follows.
(1) Steady flow.
(2) One-dimensional flow.
(3) Fully developed flow.
(4) Stream properties change continuously.
(5) Ideal gas.
BRAYTON CYCLE EFFICIENCY INCREASE 117

(6) Constant specific heats.


(7) No external heat transfer.
(8) Constant friction factor.
(9) Negligible mass additions.
(10) Negligible drag of internal bodies.
(11) Products of combustion have properties of air.
(12) Irreversibilities associated with changes in composition during combustion are ignored.

With respect to the second assumption, Ward-Smith (1980) argued that the velocity of a turbulent,
compressible flow inside a constant-area duct varies primarily in the streamwise direction. Hence, if the angle
of convergence of the combustion chamber is not too large (say, less than 30 degrees), assumption (2) is
reasonable for the approximate isothermal process. This is a key assumption and it allows for a preliminary
analysis to be carried out. As a result of assumptions (1)—(12), the theoretical efficiency of the modified
Brayton cycle may be reduced. However, the same assumptions were applied to the basic Brayton cycle; that
is, a potential increase in efficiency may still exist for the modified cycle.
Through rearrangement of the appropriate parameters tabulated by Shapiro (1953), one obtains for the
model of the flow:

1!M2 dM2 dA d¹ dx
"!2 #(1#kM2) 0#4fkM2 (1)
M2 M2 A ¹ D
1#(k!1) 0
2

where k is the ratio of specific heats, A is the cross-sectional area of the combustion chamber, ¹ is the
0
stagnation temperature, f is the friction factor (one quarter of that which appears in the Darcy—Weisbach
equation) and D is the diameter of the combustion chamber.
Equation (1) relates the changes in Mach number to the simultaneous effects of area change, heat addition
and friction. It is noted that, in the proposed combustion chamber, changes in stagnation temperature can
only occur as a result of heat addition.
From geometry,

dA dD
"2 (2)
A D
and
dD
dx"! (3)
2 tan R

Substituting equations (2) and (3) into equation (1), and rearranging yields

A B A B
1!M2 4 2f dD 1 d¹
dM2" ! ! # 1# 0 (4)

C D
M2 kM2 tan R D kM2 ¹
kM4 1#(k!1) 0
2

Direct integration of equation (4) from the inlet to the outlet of the combustion chamber is not possible
since the terms containing M in the right-hand side of this equation cannot be eliminated. Hence, this
equation must be solved numerically. The method used herein involves step-wise integration with an
incremental search between steps. The details of this method are elucidated below.
By means of step integration, the left-hand side of equation (4) is integrated from M to M "M #*M,
n n`1 n
where, *M;1. Then, the right-hand side is integrated with M"M "(M #M )/2. The result of this
!7 n n`1
procedure is

A B A B
D ¹
I "I ln n`1 #I ln 0n`1 (5)
m d D t ¹
n 0n
118 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

where

A B
1 1
!

A B
A [1#bB] A B
I "a ln #
m B[1#bA] k
4 2f
I "! !
d kC tan R
1
I "1#
t kC
a"(k#1)/2k, b"(k!1)/2, A"M2, C"M2 and B"M2 .
n !7 n`1
In equation (5), D and ¹ are functions of x. Through geometry, D is related to x by
0
D "D !2(x !x )tan R (6)
n`1 n n`1 n
From conservation of energy, ¹ is dependent on x through
0
q@
d¹ " dx (7)
0 C
1
where C is the specific heat at constant pressure.
1
With equations (6) and (7), an incremental search for the value of x is performed until equation (5) is
n`1
nearly satisfied. This involves an incremental value of x, namely *x, such that *x;*M. The solution
method progresses in this manner from M "M at x "0 to M "M .
n */ n n`1 065
Finally, with discrete values of M and ¹ computed throughout the combustion chamber, the temperature
0
profile can be determined from the definition of ¹ :
0

C D
M2
¹ "¹ 1#(k!1) (8)
0 2
where ¹ is the absolute temperature.
In order to carry out the numerical solution, one requires a mathematical description of the heat flux
distribution (i.e. q@"q@(x)) at each combustion site. Since this is not known a priori, a reasonable form of q@(x)
must be assumed. On physical grounds, one might expect the heat flux to rise from a value of zero and peak
before decreasing to a zero value again. Mathematically, such behaviour may be represented in numerous
ways; for example, by means of a Gaussian distribution. In this study, a parabolic distribution is chosen
arbitrarily. That is, each combustion site in the proposed combustion chamber is characterized by a para-
bolic heat flux distribution (this is shown schematically in Figure 4). Specifically, q@ is expressed as

C D
3 6(x!x )2
q@"q ! q , (x !0·5¸ ))x)(x #0·5¸ ) (9)
2¸ ¸3 q q q q
q q
where q is the quantity of heat addition in energy per unit mass.
It should be mentioned that any distribution can be examined.
A computer program was developed to implement the numerical analysis. From the execution of this
program, one obtains the length ¸ over which the Mach number changes from M to M due to the
*/ 065
simultaneous effects of friction, area change and heat addition. In order to verify that the program (and hence
the method of analysis) is valid, it was employed to solve special cases for which analytical solutions exist. In
particular, the program was utilized to solve isentropic flow, Fanno-line flow and isothermal flow in long
pipes. It was demonstrated that the program could accurately predict the isolated effects of friction, area
change and heat addition: the computed results were within 1% of the analytical solutions. Thus, it can be
inferred that the program is capable of accurately treating a general case of heat addition with area change
and friction, such as the approximate isothermal process. Analysis of this process yields the temperature and
Mach number variation with respect to distance along the proposed combustion chamber.
BRAYTON CYCLE EFFICIENCY INCREASE 119

The hypothetical application of the idealized isothermal process (namely, the modified Brayton cycle
shown in Figure 3) involves a temperature versus entropy plot. Therefore, assessment of the approximate
isothermal process is based on the temperature variation with entropy inside the combustion chamber. Since
the values of M and ¹ can be found for various locations in the combustion chamber, one can determine the
corresponding values of specific entropy (s) at the same locations. With this information, it is possible to
construct a plot of ¹ versus s. One measure of the deviation from isothermal flow in the proposed
combustion chamber is the weighted mean value of the temperature deviation magnitudes (E), calculated on an
entropy basis:

+ 1 (D¹ !¹ D#D¹ !¹ D) (s !s )
2 n *40 n`1 *40 n`1 n
E" n (10)
s !s
065 */
The idealized isothermal process is characterized by E"0 K. Hence, it is desirable to minimize the value of
E so that the best approximation of an isothermal process is attained. This is the criterion adopted in
optimizing the design of the combustion chamber.
Another measure of the deviation from isothermal flow is the maximum temperature difference (d ) in the
T
proposed combustion chamber. In the idealized isothermal process, d "0 K, while for a conventional
T
constant-pressure heat addition process, d "q/C . These represent the limiting values of d for the
T 1 T
approximate isothermal process carried out in the combustion chamber. Alternatively, it is convenient
to deal with the dimensionless parameter C d /q. The values of this parameter will lie between zero and
1 T
unity. Values of C d /q which are near zero in the proposed combustion chamber signify a very good
1 T
approximation of an isothermal process. In contrast, values of C d /q near unity indicate a poor
1 T
approximation.

3. RESULTS
A parameter study was conducted in which the influence of various dimensionless parameters relevant to the
proposed combustion chamber were examined. The first part of this study considers a single heat addition in
the combustion chamber. A reference set of parameter values was selected with the values of certain variables
(such as q and M ) being fixed by the operating conditions of the modified Brayton cycle, of which the
065
proposed combustion chamber is a part. As identified earlier, one desirable set of operating conditions is
represented by M "0·8 and ¹ "1104 K. Therefore, the parameter study was performed with these
065 *40
conditions.
In the parameter study, the effects of individual parameters on the calculated values of ¸, E and d were
T
examined by varying the value of each parameter about its reference value, with all other parameter values
held constant. As mentioned previously, the objective is to minimize E. Thus, the parameter study serves to
optimize the configuration of the proposed combustion chamber. It is noted that, for practical reasons, the
combustion chamber should be kept as short as possible. Therefore, ¸ should be as small as possible. The
results of the study and the range of parameter variation for the case of a single heat addition are summarized
in Table 1.
The results of the first part of the parameter study indicate that x /D (the ratio of heat-source location to
q */
inlet diameter) and ¹ /¹ (the ratio of inlet temperature to desired temperature) are the important
*/ *40
parameters in terms of controlling the deviation from isothermal flow in the proposed combustion chamber
(see Table 1). Moreover, it was found that the deviation is minimized when x +¸ and ¹ +¹ .
q */ *40
For this optimum configuration, the corresponding values of E/¹ and C d /q are 1·70% and 0·62,
*40 1 T
respectively.
From a physical standpoint, it makes sense that ¹ should be near ¹ in order to minimize the deviation
*/ *40
from isothermal flow in the proposed combustion chamber. However, the result that the single heat addition
should be near the exit of the combustion chamber requires some explanation. In general, when a compres-
sible gas with subsonic velocity is heated, a fraction of the heat received is converted to internal energy (which
120 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Table 1. Results of parameter study for single heat addition

¸/D E/¹ C d /q
*/ *40 1 T
Increase f Negligible effect Negligible effect Negligible effect
Decrease f Negligible effect Negligible effect Negligible effect
Increase ¹ /¹ No effect Increases† No effect
*/ *40
Decrease ¹ /¹ No effect Decreases† No effect
*/ *40
Increase ¸ /D No effect Minor effect Minor effect
q */
Decrease ¸ /D No effect Minor effect Minor effect
q */
Increase x /D Negligible effect Decreases Decreases
q */
Decrease x /D Negligible effect Increases Increases
q */
Increase R‡ Decreases Minor effect Minor effect
Decrease R‡ Increases Minor effect Minor effect

†For ¹ /¹ *+1, otherwise the opposite effect occurs.


*/ *40
‡With x /¸ held constant so as to isolate the effects of R alone.
q
Range of parameter variation: f"0·000—0·013; ¹ /¹ "0·98—1·02; ¸ /D "0·1—0·3;
*/ *40 q */
x /D "1·0—2·0; R"5—15 degrees.
q */

is represented by the temperature of the gas, ¹) and the remaining fraction is converted to kinetic energy
(which is represented by the Mach number of the gas, M). At relatively high Mach numbers, more of the heat
added is converted to kinetic energy than that which is converted at relatively low Mach numbers.
Accordingly, smaller temperature rises occur when heat is added at higher Mach numbers. Therefore, since
the Mach number of the flow in the combustion chamber is highest near the exit, it makes good physical
sense that heat be added at this location in order to minimize the temperature increase.
In the second part of the parameter study, two heat additions were considered. As the results from the first
part of the study imply, it is advantageous to position one of the heat additions near the exit of the
combustion chamber. For this reason, x was held equal to about ¸. Then, the effects of varying x were
q2 q1
investigated for various values of q and q subject to the constraint that the total heat added (q #q ) was
1 2 1 2
the same as for the single heat addition case. In so doing, an optimum configuration was found for the case of
two heat additions. The parameter values for this configuration are provided, along with a temper-
ature—distance plot and a temperature—entropy plot of the associated approximate isothermal process, in
Figures 5 and 6, respectively.
As shown in Figures 5 and 6, the corresponding values of E/¹ and C d /(q #q ) are 1·11% and 0·43,
*40 1 T 1 2
respectively. From a comparison of these values with those for the optimum single heat addition case, it is
concluded that the approximate isothermal process may be significantly improved by employing two heat
additions rather than one in the proposed combustion chamber.
Since the value of C d /(q #q ) is substantially less than 1, the process which occurs in the pro-
1 T 1 2
posed combustion chamber represents a good approximation of an isothermal heat addition. The implica-
tion of this value is that the maximum temperature difference anywhere in the combustion chamber is less
than half of that which would result if the same heat were added by conventional methods (i.e. at constant
pressure).
Inspection of Figure 6 reveals that the approximate isothermal process which occurs in the proposed
combustion chamber is essentially a series of reheats, similar to those illustrated in Figure 2. As discussed in
the introduction, a better approximation of an isothermal heat addition is achieved as the number of reheats
increases. Therefore, it may be assumed that the utilization of three or more heat additions in the proposed
combustion chamber would result in a further improvement over the use of two heat additions. It is also
mentioned in the introduction that reheating in gas-turbine engines is usually difficult to implement and is
only performed in large stationary power plants. However, with the combustion chamber in question, it
is anticipated that the approximate isothermal process could be established in a more practical manner. The
ramification of this is that engines featuring the proposed combustion chamber may also be employed in
relatively smaller applications, such as aircraft propulsion.
BRAYTON CYCLE EFFICIENCY INCREASE 121

Figure 5. Temperature—distance plot of the approximate isothermal process with two heat additions in the optimum configuration

Under the operating conditions selected for the parameter study (i.e. for M "0·8 and ¹ "1104 K), the
065 *40
modified Brayton cycle has a thermal efficiency which is 4·2% higher than that of the Brayton cycle, for the
same work output. This represents an ideal situation in which an exact isothermal process is carried out in
the modified Brayton cycle. However, if (in the modified Brayton cycle) the exact process is replaced by the
approximate isothermal process containing two heat additions, the thermal efficiency of the modified cycle is
estimated to be 3·5% higher than that of the Brayton cycle (for the same work output). It is noted that the
pressure drop in the combustion chamber (aside from that due to wall friction) will probably further reduce
the benefit. Such an increase in efficiency is still significant. This means that the utilization of the proposed
combustion chamber (with two heat additions) to approximate an isothermal process in the modified
Brayton cycle may not diminish the merit of this cycle. Moreover, since it has been argued that the
approximation of an isothermal process in the proposed combustion chamber improves with the number of
heat additions, the use of three or more heat additions inside the combustion chamber may allow the full
merit of the modified cycle to be approached.
122 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Figure 6. Temperature—entropy plot of the approximate isothermal process with two heat additions in the optimum configuration

4. FLAME STABILIZATION
It has been determined that a reasonable approximation of an isothermal process may be obtained in the
proposed combustion chamber with the strategic utilization of at least two discrete heat additions. For the
optimum configuration identified in the previous section, it can be shown from the analysis developed that
the velocities at the combustion sites where these heat additions occur are 206 m s~1 at combustion site 1 and
338 m s~1 at combustion site 2. Since such velocities are considered to be relatively high, the appropriate
method of combustion should be of the sort employed in ramjets and afterburners.
In these systems, combustion is normally accomplished through flame stabilization in a high-velocity
stream of premixed fuel and air. The most effective technique of flame stabilization involves the creation of
a recirculation zone within the flow. This is commonly achieved in practice by means of bluff bodies, such as
spheres and vee gutters. Bluff-body flame stabilizati on in a homogeneously premixed fuel—air mixture is
BRAYTON CYCLE EFFICIENCY INCREASE 123

Figure 7. Bluff-body flame stabilization in a homogeneously premixed fuel—air stream

depicted in Figure 7, where a region of recirculation is shown. As illustrated in this figure, flames stabilized in
homogeneous mixtures propagate at an oblique angle throughout the flow downstream of the flame
stabilizer.
One aspect of bluff-body flame stabilizers is the blowoff velocity (» ). The blowoff velocity is the approach
BO
velocity above which flame extinction occurs. This is of crucial importance in the proposed combustion
chamber; if » is less than the velocity at the location where combustion is to occur it is impossible to
BO
stabilize a flame. Thus, any attempt to approximate an isothermal heat addition would be futile. It is
therefore necessary to examine this aspect of flame stabilization in order to determine the feasibility of the
proposed combustion chamber.
Numerous theoretical and experimental studies of bluff-body flame stabilizers have been attempted in an
effort to predict blowoff conditions. Based on a simple model by Spalding (1955), Zukoski and Marble (1956)
hypothesized a critical condition for blowoff from bluff bodies. Basically, their hypothesis is that » for
BO
homogeneous mixtures depends on:
(1) the characteristic size of the bluff body (d);
(2) the laminar flame speed (or normal burning velocity) of the combustible mixture (S ); and
6
(3) the thermal diffusivity of the combustible mixture (a).
Spalding (1955) has plotted a collection of blowoff data (for homogeneous mixtures) from various
investigators. For values of Peclet number (» d/a) less than 104, the data for three-dimensional bluff-bodies
BO
are well correlated by the following expression:
» "27·44S1>5 (d/a)1@2 (11)
BO 6
It is noted that for a given combustible mixture, S is heavily dependent on the fuel-to-air ratio (F/A) and the
6
absolute temperature (¹). In fact, S is highest when the fuel-to-air ratio is nearly stoichiometric. Therefore,
6
as proven experimentally, » is a maximum in such compositions.
BO
Most applications of bluff-body flame stabilization use homogeneously premixed fuel—air mixtures. For
the proposed combustion chamber featuring two heat additions with the optimum configuration, the
required heat releases at combustion sites 1 and 2 are, respectively, 30 kJ kg~1 and 98·4 kJ kg~1 (see Figure 5
or 6). Based on fuels employed in gas-turbine engines, homogeneously premixing the necessary quantities of
fuel for these heat releases at each combustion site would result in extremely lean mixtures. In such mixtures,
it is virtually impossible to stabilize a flame at high velocities.
A more appropriate method of flame stabilization with bluff bodies involves partial mixing of fuel and air.
This is illustrated in Figure 8; with partial mixing, a flame of limited propagation is stabilized downstream
from the bluff body. Based on physical arguments, Spalding (1955) concluded that partially mixed flames
may be stabilized at much higher velocities than homogeneously premixed flames when the overall fuel-to-air
ratio is very small, as is the case with the combustion sites of the proposed combustion chamber. With partial
124 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Figure 8. Flame propagation downstream from a bluff body when the fuel is partially mixed

Figure 9. Comparison of blowoff curves for complete and partial mixing

mixtures, local regions in which the fuel-to-air ratio is nearly stoichiometric can be created. Thus, partial
mixing may indeed prove to be advantageous in the utilization of bluff bodies for flame stabilization at these
sites. Spalding (1955) further argued that the maximum blowoff velocity pertaining to a bluff body is higher
for homogeneous mixtures than it is for partially mixed flows. In summary, a comparison of the blowoff
characteristics for bluff bodies in partially and homogeneously mixed fuel—air mixtures may therefore be
represented schematically as in Figure 9.
A basis for determining the minimum (but not necessarily sufficient) bluff-body size required for flame
stabilization via partial mixing at each combustion site may now be established. The minimum size
corresponds to the bluff-body diameter required to stabilize a flame in a homogeneous stoichiometric
fuel—air mixture at the flow conditions of the particular combustion site (as elucidated earlier, the maximum
blowoff velocity in homogeneous mixtures occurs for near-stoichiometric compositions).
The minimum bluff-body diameter for a combustion site can be estimated from equation (11). In order to
determine this value, S and a for stoichiometric fuel—air mixtures at the conditions of the site are required.
6
With respect to the blowoff characteristics of flame stabilizers, the type of fuel in the combustible mixture is
BRAYTON CYCLE EFFICIENCY INCREASE 125

insignificant. This is due to the fact that most fuels possess similar values of burning velocities in air; notable
exceptions are hydrogen and acetylene, which have extraordinarily high burning velocities. Therefore, for
calculation purposes, propane is arbitrarily considered as the fuel for the proposed combustion chamber.
The value of S in centimetres per second for propane—air mixtures may be calculated within $10% from
6
the simple power law (Metghalchi and Keck, 1982):

A BAB
¹ C1 P C2
S "S (1!2·1G) (12)
6 6! ¹ P
! !
where
S "34·22!138·65(/!1·08)2 (13)
6!
C "2·18!0·8(/!1) (14)
1
C "!0·16#0·22(/!1) (15)
2
/ is the equivalence ratio ("(F/A)/(F/A) ), ¹ is the absolute temperature, ¹ "298 K, P is the absolute
450*#) !
pressure, P "1 atm and G is the diluent mass fraction (combustion products can be considered to be
!
diluents).
It is noted that equation (12) fits data measured for /"0·8—1·5, G"0—0·2, P"0·4—40 atm and
¹"350—700 K.
The minimum bluff-body diameter required for combustion site 2 of the proposed combustion chamber
with the optimum configuration is estimated to demonstrate the calculation procedure. As part of this
calculation, / is set to 1. Therefore, from equation (13), (14) and (15), one obtains S "33·3 cm s~1,
6!
C "2·18 and C "!0·16, respectively. In addition, it can be shown from the analysis developed that
1 2
¹"1082 K and P"16·1 atm at combustion site 2.
For the particular operating conditions of the proposed combustion chamber and the modified Brayton
cycle (compression ratio equal to 18·9), the burned gases from combustion site 1 and the constant-pressure
heat addition (states 2 to 3 in Figure 3) result in G"0·2 at combustion site 2. This value is determined with
the assumption of perfect combustion.
It is noted that the values of some variables lie outside the ranges for which equation (12) is applicable. In
the worst case, the value of ¹ (1082 K) is significantly greater than the limiting value (700 K) for this equation.
However, inspection of the data (Metghalchi and Keck, 1980) correlated by equation (12) suggests that S
6
increases monotonically with ¹. Hence, in order to obtain a rough estimate, equation (12) is utilized
nonetheless. Therefore, for S "33·3 cm s~1, C "2·18, C "!0·16, ¹"1082 K, P"16·1 atm, and
6! 1 2
G"0·2, equation (12) yields S "199 cm s~1.
6
The value of a for the propane—air mixture is necessary in order to make use of equation (11). This property
can be expressed as i/oC , where, i is the thermal conductivity and o is the density. It is assumed that the
1
mixture has the properties of pure air. Thus, i and C (which are temperature-dependent only) are,
1
respectively, 0·0722 W mK~1 and 1157 J kg~1 K~1 at 1082 K (O®zisik, 1985). Assuming perfect gas behaviour,
one may express o as P/R¹, where R"287 J kg~1 K~1 (the gas constant for air). Therefore, at combustion
site 2, o"(16·1)(101325)/(287)(1082)"5·2533 kg m~3. Hence, the value of a at combustion site 2 is
(0·0722)/(5·2533)(1157) m2 s~1, i.e. 0·0000119 m2 s~1 or 0·119 cm2 s~1.
Lastly, from equation (11) with S "199 cm s~1, a"0·119 cm2 s~1 and » "33 800 cm s~1, the min-
6 BO
imum three-dimensional bluff-body diameter required at combustion site 2 of the proposed combustion
chamber is estimated to be 0·2 mm. It is emphasized that this diameter is not necessarily sufficient for flame
stabilization in a partially mixed propane—air mixture at the conditions corresponding to combustion site 2.
Furthermore, this diameter may be considered as the minimum only if the local fuel—air composition in the
vicinity of the bluff body is stoichiometric. However, since this value is quite small, it is assumed that
a reasonable bluff-body size (such as a sphere 1 cm in diameter) may provide flame stabilization. Hence,
although the result of this calculation does not prove that bluff-body flame stabilization is achievable at
combustion site 2, it does not show that stability is impossible; otherwise, the diameter calculated would be
impractical, i.e. orders of magnitude larger. Experimentation is the best means of addressing this matter. It is
126 J. VECCHIARELLI, J. G. KAWALL AND J. S. WALLACE

Figure 10. Opposed-jet flame stabilization in a homogeneously premixed fuel—air stream

noted that a similar calculation indicates that flame stabilization at combustion site 1 may be attainable with
a bluff body of practical dimensions.
Another type of flame stabilizer is the opposed jet. Flame stabilization by an opposed jet in a homogene-
ously premixed fuel—air stream is depicted in Figure 10. Alternatively, the fuel can be supplied through the
opposed jet itself, i.e. in a fuel—air jet. In this case, the mainstream is void of fuel and the resulting flame is
confined to the vicinity of the stagnation region located upstream from the end of the jet tube. Based on
a bluff-body analogy, a crude method of determining the blowoff velocity from opposed jets has been derived
by Vecchiarelli (1991). Results from the utilization of this method indicate that opposed fuel—air jets may be
capable of stabilizing flames at combustion sites 1 and 2 of the proposed combustion chamber for the same
operating conditions examined previously for bluff bodies.

5. CONCLUSION
It has been established that the application of an isothermal heat addition process in gas-turbine engines may
yield significant increases of thermal efficiency. Moreover, substantial reductions of pollutant emissions are
expected. The results of a numerical analysis indicate that a reasonable approximation of such a process may
be realized in a proposed combustion chamber. This is contingent upon successful flame stabilization at each
combustion site of the combustion chamber. Calculations imply that flame stabilization is achievable with
either bluff bodies or opposed jets. Thus, it can be concluded that an approximate isothermal heat addition in
the proposed combustion chamber is feasible.

NOMENCLATURE

A "cross-sectional area of proposed combustion chamber


C "specific heat at constant pressure
1
d "bluff-body diameter
d "maximum temperature difference in proposed combustion chamber
T
D "diameter of proposed combustion chamber
D "inlet diameter of proposed combustion chamber
*/
E "weighted mean value of the temperature deviation magnitudes
f "friction factor
F/A "fuel-to-air ratio
G "diluent mass fraction
BRAYTON CYCLE EFFICIENCY INCREASE 127

k "ratio of specific heats


¸ "length of proposed combustion chamber
¸ "span of a heat addition
q
M "Mach number
M "Mach number at inlet of proposed combustion chamber
*/
M "Mach number at outlet of proposed combustion chamber
065
P "absolute pressure
q "heat addition in energy per unit mass
q@ "heat flux in energy per unit mass per unit length
R "proposed combustion chamber angle
R "gas constant for air
RF "recovery factor
s "specific entropy
s "specific entropy at inlet of proposed combustion chamber
*/
s "specific entropy at outlet of proposed combustion chamber
065
S "laminar flame speed
6
¹ "absolute temperature
¹ "temperature at inlet of proposed combustion chamber
*/
¹ "temperature of idealized isothermal process
*40
¹ "maximum allowable temperature in a thermal power cycle
.!9
¹ "minimum temperature in a thermal power cycle
.*/
¹ "stagnation temperature
0
» "blowoff velocity
BO
x "distance along proposed combustion chamber
x "location of a heat addition
q
Greek letters

a "thermal diffusivity
i "thermal conductivity
/ "equivalence ratio
o "density

Note that the subscript ‘n’ denotes an arbitrary location along the proposed combustion chamber.

REFERENCES
Cummins, C. L. (1989). Internal Fire, revised edition, Society of Automotive Engineers Inc., Warrendale, Pennsylvania, U.S.A.
Harman, C. T. R. (1981). Gas ¹urbine Engineering, MacMillan, London, U.K.
Metghalchi, M. and Keck, J. C. (1980). Combust. & Flame, 38, 143.
Metghalchi, M. and Keck, J. C. (1982). Combust. & Flame, 48, 191.
Morris, R. E. (1973). ‘Power turbine design’, Engineering Report No. 697, United Aircraft of Canada Limited, Longueuil, Quebec.
Munt, R. and Danielson, E. (1976). ‘Aircraft technology assessment, status of the gas turbine program’, U.S. Environmental Protection
Agency, Washington D.C., U.S.A.
O®zisik, M. N. (1985). Heat ¹ransfer, McGraw-Hill, New York.
Shapiro, H. A. (1953). ¹he Dynamics and ¹hermodynamics of Compressible Fluid Flow, Vol. 1, Ronald Press, New York.
Spalding, D. B. (1955). Some Fundamentals of Combustion, Butterworths Scientific, London. U.K.
Vecchiarelli, J. (1989). ‘Analysis and application of isothermal expansion’, B.A.Sc. thesis, University of Toronto, Canada.
Vecchiarelli, J. (1991). ‘A proposed combustion chamber for approximating an isothermal heat addition process’, M.A.Sc. thesis,
University of Toronto, Canada
Vecchiarelli, J., Kawall, J. G. and Wallace, J. S. (1991). ‘One-dimensional analysis of an approximate isothermal heat addition process’,
Proc. 13th Canadian Congress of Applied Mechanics, Winnipeg, p. 586.
Ward-Smith, A. J. (1980). Internal Fluid Flow, Clarendon Press, Oxford, U.K.
Wood, B. D. (1982). Applications of ¹hermodynamics, 2nd edn, Addison-Wesley, Reading, Massachusetts, U.S.A.
Zukoski, E. E. and Marble, F. E. (1956). Proc. Gas Dynamics Symp. on Aerothermochemistry, Northwestern University, Evanston,
Illinois, U.S.A., p. 205.

You might also like