You are on page 1of 23

Chapter 2 Cruise Performance

2.1 General Equations for Flight Performance


2.1.1 Equations of Motion in Wind Axes
2.1.2 Drag Polar
2.1.3 Compressibility effects
2.2 Steady Level Flight
2.2.1 Thrust and Power Required
2.2.2 Minimum Thrust
2.2.3 Minimum Power
2.3 Range and Loiter Optimization at fixed altitude
2.3.1 Specific Fuel Consumption
2.3.2 Range and Endurance definitions
2.3.3 Range Optimization at fixed altitude
2.3.4 Loiter Optimization at fixed altitude
2.4 Range Optimization with altitude and airspeed
2.4.1 Range Optimization for Propeller aircraft
2.4.2 Range Optimization for Jet aircraft
2.4.3 Flight Paths and Range
2.4.4 Longest Range at given Mach for Jet aircraft
2.5 Bibliography

Aircraft Design, 2017


2-2 CRUISE PERFORMANCE

2.1 General Equations for Flight Performance


This chapter studies the aircraft performance in level flight, that is, cruise and loiter, in
order to relate the consumed fuel with the aircraft characteristics. This analysis permits
linking the aircraft range with the aircraft mass, and it is the basis to build up the P/L-R
diagrams (aircraft payload versus aircraft range).

2.1.1 Equations of Motion in Wind Axes


The balance of aircraft accelerations involves 3 kinds of forces:
1. Aerodynamic forces
2. Propulsive forces
3. Gravity
For performance analysis it is convenient to project the equations of motion in wind axes
(Xw and Zw), as shown in figure 2-1. The longitudinal and vertical equilibrium yields:
dV
m FX T cos( T ) D W sin (2.1)
dt
d
mV FZ T sin( T ) L W cos (2.2)
dt
Where:
m = Aircraft Mass
W = Aircraft Weight = mg
T = Thrust
L = Aerodynamic Lift (by definition, lift is perpendicular to the flight direction)
D = Aerodynamic Drag (by definition, drag is parallel to the flight direction)
V = Aircraft speed
= Angle of attack
= Climb angle
T = Incidence of thrust axis with respect to the body axis

X-body axis
T
Aerodynamic Lift
V
Thrust
X-wind axis Aerodynamic

Drag Horizontal

Weight

Figure 2-1: Geometry for performance calculation

Aerodynamic lift and drag (L and D) depend on the angle of attack. In a simplified manner,
equations (2.1) and (2.2) can be understood as follows:
Angle of attack is adjusted to provide the required total lift.
Thrust is adjusted to provide the required acceleration and/or climb angle.
The required lift, climb angle and acceleration depend on the class of manoeuvre that the
aircraft has to perform: level cruise, climb, horizontal acceleration, steady turn, etc.

Aircraft Design, 2017


CRUISE PERFORMANCE 2-3

In the most of the cases, the thrust axis forms a little angle with respect to the wind axis.
That is, the angle (+T) is small enough to assume cos(+T) 1 and sin(+T) 0, letting
to simplify (2.1) and (2.2) to the following form:
T D W sin
mV (2.3)
m V L W cos (2.4)

Figure 2-2

2.1.2 Drag Polar


The aerodynamic lift and drag are usually expressed by means of the non-dimensional
coefficients CL and CD:
L QSCL (2.5)
D QSCD (2.6)
Where:
Q = Dynamic pressure = V2
S = Wing reference area (usually the full area extending to the a/c centreline)
CL and CD = Non-dimensional lift and drag coefficients
The drag at subsonic speeds is composed of two parts:
Induced drag is the drag caused by the generation of the lift. This is primarily a
function of the wing span.
Zero-lift or parasite drag is the drag that is not related to the lift. This is primarily
skin-friction drag, and as such is directly proportional to the total surface area of the
aircraft exposed (wetted) to the air.
The drag polar is a curve of the lift coefficient versus the drag coefficient, as shown in the
next two examples of figures 2-3 and 2-4.
For an uncambered wing (figure 2-3), the minimum drag (CD0) occurs when the lift is zero.
The drag polar is approximately parabolic, as defined by the following equation:
CD CD0 KCL2 (2.7)
The point at which a line from the origin is just tangent to the drag polar curve is the point
of maximum lift to drag ratio (CL/CD = L/D), as shown in figure 2-3. This ratio L/D is known
as the aerodynamic efficiency. For a parabolic drag polar without offset, as given by the
equation (2.7), the optimum lift coefficient that maximizes L/D is:
CLopt CD0 K (2.8)

Aircraft Design, 2017


2-4 CRUISE PERFORMANCE

For a cambered wing (figure 2-4), the minimum drag (CDmin) occurs at some positive lift
(CLmd). The drag polar has also a parabolic shape, but there is a vertical offset as given by
the equation (2.9):
CD CD min KCL CLmd
2
(2.9)

Uncambered wing Cambered wing


1.5 1.5

1 1

L/D max
CL

CL
L/D max

0.5 0.5

0 0
0 0.02 0.04 0.06 0.08 0.1 0 0.02 0.04 0.06 0.08 0.1
CD CD
Figure 2-3 Figure 2-4
Drag polar, uncambered wing Drag polar, cambered wing

For wings of moderate camber the offset CLmd is usually small; this implies that CD0 is
approximately equal to CDmin and equation (2.7) may be used. But even in the general
case, the equation (2.7) may be valid. It is usually found that, in a practical region of CL,
the straight line approximation of CD vs CL2 is acceptable. For instance, the figure 2-5
shows a case with CDmin = 0.025, CLmd = 0.2 and K = 0.045, comparing:
Several exact points calculated with equation (2.9).
The straight line approximation by equation (2.7), obtained with the double condition to
match the drag polar and the slope CD CL2 at the point of maximum L/D, which is in
practice a CLopt adjustment method.
CD vs CL2
3

2.5

2
Eq. (2.9)
CL2

1.5 Aprox. (2.7)


1

0.5

0
0 0.04 0.08 0.12
CD
Figure 2-5: Simplified drag polar

Aircraft Design, 2017


CRUISE PERFORMANCE 2-5

In the previous example, the obtained values are CD0=0.020 and K=0.033. Note that the
approximation provides CD0 and K values to be applied in equation (2.7) that are lower
than those CDmin and K in equation (2.9).
The figure 2-6 shows the comparison of the lift to drag ratio L/D as a function of C L,
calculated with both equations. In this example, the difference in L/D is less than 7% in
the range of CL between 0.5CLopt and 2CLopt.

L/D vs CL
25

20

15 Eq. (2.9)
L/D

Aprox. (2.7)
10

0
0 0.5 1 1.5
CL
Figure 2-6: Simplified Lift to Drag ratio

Therefore, in the use of the equation (2.7) as a simplified polar, the coefficients are
adjusted to represent the drag polar in the region of CL that is of interest for the analysis of
cruise or climb performance. In this scenario, the empirical zero-lift drag coefficient CD0
must be seen as a fictitious quantity, probably lower than the actual CD at CL=0.
The use of the simpler equation (2.7) instead of (2.9) let us to obtain easier analytical
results, without a significant loss of generality. The mathematical results of the next
sections are based on equation (2.7), providing a simplified but useful first approximation.
Nevertheless, it is feasible to get more refined results applying the same methods to the
equation (2.9).
The induced drag coefficient K is mainly related to the geometric and aerodynamic
characteristics of the wing. In the theoretical case of a 3-dimensional wing with an elliptical
lift distribution, the induced drag coefficient depends on the aspect ratio (A = b/c = b2/S,
being S the wing area, b the wing span and c the mean geometric chord) as follows:
CL2 1
Elliptical lift distribution: CDi KCL2 K (2.10)
A A
In practice, the deviation from the ideal elliptical distribution is like a reduction of the aspect
ratio. This effect is usually represented by the Oswald efficiency factor (e) as:
1
K (2.11)
Ae
Typical values of the Oswald efficiency factor are between 0.65 and 0.85 in subsonic flight,
It accounts not only for the non-elliptical form of the lift distribution, but also for the induced
drag due to fuselage, tailplanes, nacelles and interference effects.
For practical purposes, the drag coefficient CD is frequently expressed in counts, being
1 count = 104 = 0.0001 of CD.

Aircraft Design, 2017


2-6 CRUISE PERFORMANCE

2.1.3 Compressibility effects


Compressibility effects on drag are generally ignored at Mach numbers below 0.5. But for
high subsonic transport aircrafts, the effect of compressibility on CD needs to be taken into
account. An example is shown in figure 2-7.

0.8

0.6 Mach 0.5


Mach 0.6
CL

0.4 Mach 0.7


Mach 0.8
Mach 0.85
0.2

0
0 0.02 0.04 0.06 0.08
CD

Figure 2-7: Compressibility effects on drag polar (example)

In this example, between Mach 0.70 and 0.80 there is a significant increase in the drag
and at a critical Mach of approximately 0.85 a rapid rise is experienced in both the zero-lift
drag and the induced drag. The drag rise is caused by shock waves that induce boundary
layer separation.
The aerodynamic design for high subsonic aircrafts tries to achieve low compressibility
drag and a high drag critical Mach number. For initial design calculations, it is usual to
assume CD due to compressibility in the order of 5 counts (0.0005) for long range cruise
conditions and in the order of 20 counts (0.0020) for high speed cruise conditions.
Next figure 2-8 shows an example of evolution of CDmin, CLmd and K with Mach.

0.035 0.7
0.030 0.6
0.025 0.5
CDmin
K, CLmd
CDmin

0.020 0.4
0.015 0.3
K
0.010 0.2
0.005 0.1
CLmd
0.000 0
0.5 0.6 0.7 0.8 0.9
Mach

Figure 2-8: Subsonic evolution of CDmin, CLmd and K with Mach (example)

Aircraft Design, 2017


CRUISE PERFORMANCE 2-7

2.2 Steady Level Flight

2.2.1 Thrust and Power Required


If the aircraft is flying in no-accelerated level flight, the climb angle is null and the sum of
the forces must equal zero. This leads to the following equations, just stating that, in level
flight, thrust equals drag and lift equals weight.
T D QSCD (2.12)
L W QSCL (2.13)

Lift

Thrust Drag

Weight

Figure 2-9: Forces in steady level flight

Equations (2.12) and (2.13) imply that the thrust to weight ratio (T/W) in level flight must be
the inverse of the aerodynamic efficiency (L/D) as derived in (2.14). Therefore the
condition for minimum thrust at a given weight is also the condition for maximum L/D.
T D CD
(2.14)
W L CL
Power required is force times velocity, which in steady level flight equals the drag times
the speed. The power required to weight ratio is:
P DV C
V D (2.15)
W L CL
So the condition for minimum power required at a given weight occurs at minimum (DV)/L
or maximum (L/D)/V. It is achieved at a speed lower than the speed of maximum L/D, as
shown in figure 2-10.

L/D Max L/D

Max (L/D)/V

Min. Min.
Power Thrust
V
Figure 2-10: Conditions for minimum thrust and minimum power

Aircraft Design, 2017


2-8 CRUISE PERFORMANCE

2.2.2 Minimum Thrust


Using the equation (2.7), the thrust to weight ratio for level flight is:

T C D C D0
KCL (2.16)
W CL CL
At a given weight, to find the condition at which thrust is minimum and L/D is maximum,
the derivative with respect to CL is set to zero. The results are:

CD0
CL min T CLopt (2.17)
K
1/ 2 1/ 4
2W 2W K
Vmin T (2.18)
SCL min T S D0
C
If the optimal lift coefficient is substituted back into equation (2.7) the drag coefficient is
twice the zero-lift drag coefficient:

C D0
CD min T CD0 K 2CD0 (2.19)
K
Dmin T CD min T 2CD0
2 KCD0 (2.20)
W CL min T CD0 K 1/ 2

L L 1
(2.21)
D min T D max 2 KCD0
The optimal lift coefficient for minimum drag and thrust is reached by achieving the optimal
dynamic pressure. This can be obtained by varying either speed or altitude:
At a given altitude, equation (2.18) indicates that VminT decreases during the cruise
because of the reduction of weight due to the fuel burned.
On the other hand, maintaining true airspeed and minimum thrust during the cruise
would require to decrease the density () at the same rate as W. It means to gain
altitude during the cruise, following a cruise-climb flight path (figure 2-11).

Cruise-climb

Climb Descent

Figure 2-11: Cruise-climb

2.2.3 Minimum Power


Substituting the equation (2.7) into (2.15), the ratio of power required to weight is:

P C
V D
2 W C D0

SCL CL

KCL
2W

CD0 CL3 / 2 KC1L/ 2 (2.22)
W CL S

Aircraft Design, 2017


CRUISE PERFORMANCE 2-9

At a given weight and altitude, the condition at which power required is minimum can be
obtained setting the derivative with respect to CL to zero. The results are:

3CD0
CL min P (2.23)
K
1/ 2 1/ 4
2W 2W K
Vmin P (2.24)
SCL min P S 3CD0
3CD0
CD min P CD0 K 4CD0 (2.25)
K
Dmin P CD min P 4CD0 4
KCD0 (2.26)
W CL min P 3CD0 K 1/ 2
3
At a given weight and altitude, comparing the equations (2.23) to (2.26) with the equations
(2.17) to (2.21) it is seen that:
Vmin P Vmin T 31/ 4 . The velocity for minimum power required is approximately 0.76
times the velocity for minimum thrust.
CLminP is 1.73 times CLminT.
The drag coefficient for minimum power required (CDminP = 4CD0) is twice the drag
coefficient for minimum thrust (CDminT = 2CD0); however the actual drag for minimum
power required is less than twice the drag for minimum thrust, since the aircraft flies at
a slower speed. In fact:
Dmin P Dmin T 2 3 1.155
L Dmin P L Dmax
1 1.155 0.866 ; the aerodynamic efficiency when flying
at minimum power is 87% of maximum L/D.

Aircraft Design, 2017


2-10 CRUISE PERFORMANCE

2.3 Range and Loiter Optimization at fixed altitude

2.3.1 Specific Fuel Consumption


The specific fuel consumption (SFC) represents the fuel flow per unit of thrust. It is usually
denoted by C and provides the rate of change of fuel weight (not fuel mass), having units
of 1/time, normally 1/h. That is, C gives the ratio weight/time per unit of force:
dW / dt CT (2.27)
For propeller-powered aircraft, the specific fuel consumption is normally given per unit of
power delivered, that is Cpower. The associated equation is:
dW / dt Cpower PD (2.28)

The relation between Cpower and C is easily derived by means of the propeller efficiency p,
which is defined as the ratio of thrust power output to power input:
TV
p (2.29)
PD
pPD TV
T or PD (2.30)
V p
In this way, the equivalent SFC for a propeller engine is:
dW / dt Cpower PD V
C Cpower (2.31)
T T p

For cruise optimization purposes, some simplified models of SFC are frequently assumed
at constant altitude:
For subsonic jet aircrafts, C is considered approximately independent of the speed.
For propeller aircrafts, Cpower is approximately independent of the speed, so equivalent
C is approximately proportional to the speed.
C C

V V

Jet aircrafts Propeller aircrafts


Figure 2-12: Usual assumptions for C

More real variations of SFC with Mach and altitude are illustrated in next figures 2-13 and
2-14, based on general trends and expressions provided by Howe [2R6]:
For turboprop engines, as illustrated in figure 2-13, there is some slight fall of Cpower
with Mach, and small altitude effect.
For turbofan engines, the bypass ratio R has an important effect. Figure 2-14 shows
typical curves of SFC versus Mach and altitude, for a pure turbojet (R=0), a low bypass
turbofan (R=2) and a high bypass turbofan (R=10).

Aircraft Design, 2017


CRUISE PERFORMANCE 2-11

Typical Turboprop SFC per power (all altitudes) Typical Turboprop SFC per thrust
3 0.8 S/L
20000 ft
40000 ft
Cpower (N/h/kW)

0.6
2

C (1/h)
0.4

1
0.2

0 0
0 0.2 0.4 0.6 0.8 0 0.2 0.4 0.6 0.8
Mach Mach

Figure 2-13: Typical Turboprop SFC, after Howe [2R6]

Typical Turbojet SFC (1/h) Typical Turbofan SFC (1/h)


1.4 1
S/L R=2, S/L
1.2 20000 ft 0.8 R=2, 20000 ft
R=2, 40000 ft
40000 ft
1 R=10, S/L
R=10, 20000 ft
0.8 0.6 R=10, 40000 ft

0.6 0.4
0.4
0.2
0.2
0 0
0.2 0.4 0.6 0.8 1 0.2 0.4 0.6 0.8 1
Mach Mach

Figure 2-14: Typical Turbojet and Turbofan SFC, after Howe [2R6]

In general, for jet aircrafts the SFC increases slightly with Mach instead of being constant,
while for propeller aircrafts Cpower decreases slightly with Mach instead of being constant.
At a given altitude, it is possible to adopt a common formulation of SFC per unit of thrust,
valid for both jet and propeller aircraft, in the way indicated by the equation (2.32), where:
Cref is the SFC at a reference speed Vref.
kC is an empirical constant that depends on Cref, Vref and n.

V
n
C Cref k C V n
Vref
(2.32)
C
k C ref
n
Vref
For jet aircraft, the exponent n is in the vicinity of 0, being n 0. Particularly, the simplified
model that was shown in figure 2-12 has n = 0 and kC = C.
For propeller aircraft, the exponent n is in the vicinity of 1, with n 1. Particularly, the
simplified model that was shown in figure 2-12 has n = 1 and kC = Cpower / p.
In general, the model represented by equation (2.32) is usable with exponents 0 n 1;
being the jet case close to the lower limit and the propeller case close to the upper limit.

Aircraft Design, 2017


2-12 CRUISE PERFORMANCE

2.3.2 Range and Endurance definitions


The aircraft endurance (E) is the amount of time the aircraft can remain in the air, and the
aircraft range (R) is the distance the aircraft can travel. Both concepts are related to:
The amount of fuel carried by the aircraft
The rate at which the fuel is burned, which in turns is equal to the required thrust
multiplied by the specific fuel consumption. During the flight, the aircraft weight
drops as fuel is consumed, changing continuously the drag and the required thrust.
E and R can be expressed in integral form, assessing their instantaneous derivatives with
respect to the burned weight of fuel.
t2 W2 dt W2 1
E dt dW dW (2.33)
t1 W1 dW W1 dW dt

t2 W2 V
R Vdt dW (2.34)
t1 W1 dW dt
The specific endurance or instantaneous endurance is the derivative of the time flown
with respect to the weight of fuel burned. It can be expressed in terms of the specific fuel
consumption and the required thrust using the inverse of equation (2.27) and can be
related to the aerodynamic efficiency using the equation (2.14) (note that dE/dW < 0):
dE 1 1 1 L 1
(2.35a)
dW dW / dt CT CD W
dE 1 L 1
(2.35b)
dW C D W
The specific range or instantaneous range is the derivative of the distance travelled with
respect to the weight of fuel burned (note that dR/dW < 0):
dR V V V L 1
(2.36a)
dW dW / dt CT CD W
dR V L 1
(2.36b)
dW C D W
The variables involved in the equations (2.35) and (2.36) are the aerodynamic efficiency
L/D, the flight speed V and the specific fuel consumption C. If we assume that they are
approximately constant, the endurance and the range can be integrated as follows:
W2 1 L dW 1 L W1
E W 1

CD W
ln
C D W2
(2.37)

W2 V L dW V L W1
R W 1

CD W
ln
C D W2
(2.38)

Usually, the expressions (2.37) and (2.38) are not valid for the complete flight, but can be
applied with little loss of accuracy by breaking the cruise into several shorter mission
segments and using the appropriate L/D values as aircraft weight drops. Then W 1 and W 2
denote respectively the weights at the beginning and the end of each segment.
V L
The equation (2.38) is known as the Breguet range equation and the product is
C D
named the range parameter (PR), being a measure of the cruising performance.

Aircraft Design, 2017


CRUISE PERFORMANCE 2-13

Using the generic SFC model C k C V n provided by the equation (2.32), it is seen that, at
a given weight and altitude:
With n=0 (which is approximately representative of jet aircrafts):
o The maximum specific endurance |dE/dW| is reached at maximum L/D; it is
coincident with the condition of minimum thrust.
o The maximum specific range |dR/dW| is reached at maximum VL/D or
minimum D/(LV) and is achieved at a speed greater than the speed of
minimum thrust.
With n=1 (which is approximately representative of propeller aircrafts):
o The maximum specific endurance |dE/dW| is reached at maximum (L/D)/V or
minimum (DV)/L; it is coincident with the condition of minimum power.
o The maximum specific range |dR/dW| is reached at maximum L/D; it is
coincident with the condition of minimum thrust.
So in first approach, with the above very simplified models of SFC, the four conditions of
minimum thrust, minimum power, maximum specific endurance and maximum specific
range collapse in:
- 3 conditions of steady level flight for jet aircrafts with n=0.
- 2 conditions of steady level flight for propeller aircrafts with n=1.
L /D DV L D L D
Max. = Min. Max. = Min. Max. V = Min.
V L D L D LV
General
Min. Power Min. Thrust
(Jet & Prop)
Jet aircraft with n=0
Max. |dE/dW| Max. |dR/dW|
C constant
Prop aircraft with n=1
Max. |dE/dW| Max. |dR/dW|
Cpower constant
Next figures 2-15(a) and (b) illustrate the position of these points along the curves of L/D
and D/L versus speed, in non-dimensional form. Graphically it is interesting to note that:
The condition of minimum power or maximum (L/D)/V is at the point where a line from
the origin is tangent to the L/D curve. Conversely, the slope of the D/L curve at this
point is the opposite of the slope of a straight line from the origin.
The condition of maximum V(L/D) is at the point where a line from the origin is tangent
to the D/L curve. Conversely, the slope of the L/D curve at this point is the opposite of
the slope of a straight line from the origin.
L/D versus V D/L versus V
2
Min T
1 Min P Max VL/D
(L/D) / (L/D)max

(D/L) / (D/L)min

n=0 ME n=1
ME n=0 ME n=0 MR
n=1 n=1 MR n=0 n=1 MR
ME MR 1 Min P Max VL/D
0.5 Min T

0 0
0 0.5 1 1.5 0 0.5 1 1.5
V / V_minT V / V_minT

Figure 2-15(a): L/D & optimal conditions Figure 2-15(b): D/L & optimal conditions

Aircraft Design, 2017


2-14 CRUISE PERFORMANCE

Up to now we have not introduced any particular assumption about the drag polar. In the
next sections we will obtain particular results for the drag polar represented by the
simplified parabolic equation (2.7), in combination with the SFC model of (2.32).

2.3.3 Range Optimization at fixed altitude


The range parameter can be expanded as follows:
1n 2
PR
V L V 1n

CL

1 2W

CL n1 2 (2.39)
CD k C CD0 KCL k C S
2
CD0 KCL2
At a given weight and altitude, the conditions for best range are calculated setting the
derivative of the range parameter with respect to the lift coefficient to zero. The results are:
C D0 1 n
CLbestR (2.40)
K 3n
1/ 2 1/ 4
2W 2W K 3n
VbestR (2.41)
SCLbestR S C D0 1 n
CD0 1 n 4
CDbestR CD0 K CD0 (2.42)
K 3 n 3 n
L 1 (1 n) (3 n)
(2.43)
D bestR 4 KCD0
Referring L/D to the maximum aerodynamic efficiency from equation (2.21), we have:
L (1 n) (3 n) L
(2.44)
D bestR 2 D max
Next figure 2-16 shows the best-range-to-minimum-thrust ratios of aerodynamic efficiency
and airspeed. Note that:
For n = 1, best range and minimum thrust conditions are the same one.
For n = 0, (L/D) ratio is 0.866 and airspeed ratio is 1.32.

n=1 0.9 0.8 Best Range


1
Propeller
(L/D)bestR / (L/D)max

0.95

0.2
0.9 Jet
0.1

n=0
0.85
1 1.08 1.16 1.24 1.32
VbestR / VminT

Figure 2-16

Aircraft Design, 2017


CRUISE PERFORMANCE 2-15

2.3.4 Loiter Optimization at fixed altitude


The endurance parameter can be expanded as follows:

PE
1 L

1 CL

1 S
n/2
CL 1n / 2
(2.45)
C D k C V CD0 KCL k C 2W
n 2
CD0 KCL2
At a given weight and altitude, the conditions for best loiter are calculated setting the
derivative of the endurance parameter with respect to CL to zero. The results are:
CD0 2n
CLbestE (2.46)
K 2n
1/ 2 1/ 4
2W 2W K 2 n
VbestE (2.47)
SCLbestR S D0
C 2 n
CD0 2 n 4
CDbestE CD0 K CD0 (2.48)
K 2n 2n
L 1 (2 n) (2 n)
(2.49)
D bestE 4 KCD0
Referring L/D to the maximum aerodynamic efficiency from equation (2.21), we have:
L (2 n) (2 n) L
(2.50)
D bestE 2 D max
Next figure 2-17 shows the best-endurance-to-minimum-thrust ratios of aerodynamic
efficiency and airspeed. Note that:
For n = 0, best endurance and minimum thrust conditions are the same one.
For n = 1, (L/D) ratio is 0.866 and airspeed ratio is 0.76.

Best Endurance 0.2 0.1 n=0


1
Jet
(L/D)bestE / (L/D)max

0.95
0.8

0.9
0.9
n=1
Propeller
0.85
0.75 0.8 0.85 0.9 0.95 1
VbestE / VminT

Figure 2-17

Aircraft Design, 2017


2-16 CRUISE PERFORMANCE

2.4 Range Optimization with altitude and airspeed

2.4.1 Range Optimization for Propeller aircraft


At a given weight for a propeller aircraft, the optimization of the range with altitude can be
studied firstly with an exponent n=1 and secondly with n<1.
Insight with n=1
In first approach, setting n=1 in equations (2.32) and (2.39), the range parameter does not
depend on altitude:
V L 1 CL p CL
PR (2.51)
C D k C CD0 KCL2 Cpower CD0 KCL2
Therefore, with n=1:
The maximum range is achieved flying at the airspeed given by equation (2.41), which
in this case coincides with equation (2.18) (speed for minimum thrust), at any feasible
altitude.
This determines a curve of constant equivalent airspeed (EAS); there are multiple
combinations of true airspeed and altitude that maximize the range, as shown in figure
2-18 that contains lines of 100, 200 and 300 KEAS.
40000

30000
Altitude (ft)

20000

10000

0
0 100 200 300 400 500 600

TAS (kts)

Figure 2-18: Lines of constant EAS

Analysis with n<1


In second approach, if n is lower than 1:
Equation (2.39) shows that the range parameter increases with true airspeed (TAS),
since the exponent of V is (1n) > 0.
Equation (2.41) gives the same EAS for best range at all altitudes. And at constant
EAS, the true airspeed (TAS) raises with altitude, as shown in figure 2-18, so a high
altitude enhances the range parameter.
However, there are two concerns that limit in practice the operational altitude:
1. The available power decreases with altitude.
2. At constant EAS, Mach increases with altitude. This potentially raises the drag due
to compressibility effects and also reduces the propeller efficiency.

Aircraft Design, 2017


CRUISE PERFORMANCE 2-17

2.4.2 Range Optimization for Jet aircraft


To have an insight of the range optimization with altitude for a jet aircraft at given weight,
the simplified case with exponent n=0 is representative enough. Setting n=0 in equations
(2.32) and (2.39), the range parameter varies with as shown in equation (2.52):

V L 1 2W CL
PR (2.52)
C D C S CD0 KCL2
If we suppose that C is constant with altitude, it is clear from (2.52) that flying at higher
altitudes improves the range. The range parameter seems to grow indefinitely with altitude.
However this is false, because we have to take into account the compressibility effects and
the available thrust.
The longest range conditions are the optimum combination of altitude and airspeed where
the maximum range parameter is reached. Finding it requires sweeping in altitude and, as
altitude increases, it requires:
To check for compressibility effects that can change CD0 and K.
To check if there is sufficient thrust to fly at the predicted altitude.
Usually, the drag of a subsonic jet aircraft increases very rapidly at speeds beyond MDR
(Mach of drag rise). In general, past a critical altitude (which is determined by M bestR=MDR)
the best range conditions are attained while flying at constant Mach = M DR, at the altitude
maximizing the range parameter. This is quite complex and is assessed in section 2.4.4.
Figure 2-19 shows this optimization in sketch.

Figure 2-19

2.4.3 Flight Paths and Range


Keeping the conditions of L/D for best range during the cruise requires holding C L constant
in accordance with equation (2.40). But holding CL requires reducing the dynamic pressure
as the aircraft become lighter. It can be done in two different ways:
a) Reducing the airspeed at constant altitude.
b) Maintaining the airspeed (TAS) and reducing the air density by climbing.

Cruise-climb
The second option results in a flight path known as the cruise-climb. The range parameter
PR = (V/C)(L/D) is nearly constant, except for small variations of kC with altitude in equation
(2.32), and the range can be integrated as in equation (2.38). This is recalled in (2.53),
being Cave the average specific fuel consumption in case of slight variations with altitude:

Aircraft Design, 2017


2-18 CRUISE PERFORMANCE

W2 V L dW V L W1
R ln (2.53)
W1 CD W Cave D W2

Stair-step
However, the cruise-climb profiles are not normally permitted by the air traffic control that
wants to keep all aircraft at a constant altitude and speed. Nevertheless, on a long flight,
air traffic control may permit a stair-step path in which the aircraft climbs to a more
optimal altitude several times during the cruise.

Cruise-

Stair-step cruise

Descent
Climb

Figure 2-20

Flying at constant speed and altitude can be integrated in the range equation substituting
W by L and expressing CD as a function of CL:
L2 V dL V CL 2 dCL
R (2.54)
L1 C D C CL1 CD0 KCL2
And this integral is easily solved as follows:
C
V 1 CL 2 dCL V 1 CD 0 K
L2

R C K
arctg

CL
C CD 0 L1
1 CL2
C CD 0 K CD 0 CL1
CD 0
V 1 CL1
arctg CL 2
arctg C / K
(2.55)
C KCD0 CD0 / K D0

Fixed altitude with decreasing speed


When flying at fixed altitude and decreasing speed to maintain constant C L and L/D, there
are several methods to integrate the range equation, depending on the formulation of SFC.

Cruise at constant C
V1 L V2
Climb decreasing airspeed Descent

Figure 2-21

In the particular case of a propeller aircraft in which Cpower and p are independent of the
speed, the term V C p Cpower can be extracted from the integral of R, obtaining:

Aircraft Design, 2017


CRUISE PERFORMANCE 2-19

p L W1
R ln (2.56)
Cpower D W2
In the particular case of a jet aircraft in which C is independent of speed, V can be
substituted as a function of the square root of the weight, yielding:

R
1 L W2
C D W1
2W dW 2 2 CL
SCL W

C S CD
W1 W2 (2.57a)

In general, if C = kCVn with n1, we can apply the same method of speed substitution as a
function of the weight, producing the following result (note that equation (2.57a) is just a
particular case of (2.57b) with n=0):
(1 n ) / 2
R
2
(1 n) k C
2

CL(1 n) / 2
W1(1 n) / 2 W2(1 n) / 2 (2.57b)
S CD

Finally, if C = kCV (that is n=1) the range can be formulated as:


W2 V L dW 1 L W1
R ln (2.57c)
W1 CD W k C D W2

General
The previous results are only analytical approaches to very particular cases of aircraft
aerodynamic and SFC models and flight paths.
In general, the Breguet range equation (2.38) can be applied with little loss of accuracy by
breaking the cruise legs into several shorter mission segments and taking average values
of the range parameter (V/C)(L/D).

2.4.4 Longest Range at given Mach for Jet aircraft


As it was indicated in section 2.4.2, it is relevant to determine the best range conditions at
a given Mach = MDR, that is, finding the altitude that maximizes the range parameter.
Optimization in the troposphere
To maximize the range parameter (V/C)(L/D) in the troposphere at constant Mach it is
needed to take into account four contributions: the lift equation (2.58); the relation between
true airspeed and Mach given in (2.59); the standard atmosphere equations given in
(2.60); and a model of SFC suitable to account for variations in altitude at constant Mach.
Based on Howe [2R6], we can adopt the equation (2.61), with an exponent m 0.08; in
case of no variation with altitude this exponent is m = 0.
1 2
V SCL W (2.58)
2
V Ma M RT (2.59)

r T / T0
r 1
being p / p 0 (2.60)
r 5.2561 /
0

C k CMm (2.61)
With respect to the exponent m in equation (2.61), other simple interpretation is:

Aircraft Design, 2017


2-20 CRUISE PERFORMANCE

With a dependency C TASn in the troposphere, at constant Mach the SFC would
change like an n/2 n/(2(r1)). Being r = 5.2561, the exponent n/(2(r1)) would be
within 0 and 0.08 for n between 0 and 0.7 respectively.
With a dependency C TASn in the stratosphere, at constant Mach the SFC would
be constant, since the sound speed is constant at the stratosphere. So m = 0.
Nevertheless, the model C Mnm is more generally applicable and accurate than
the simple model C TASn; so the above estimations must be understood only as a
checking of the order of magnitude of the exponent m.
Introducing the standard atmosphere definitions from (2.60) in the equations (2.58), (2.59)
and (2.61) we attain:
2W
V2 r 1
(2.62)
0 SCL

V M RT0 Ma 0 (2.63)

C k CMm(r 1) (2.64)

The fraction V/C that appears in the range parameter can be expressed as a function of :
V Ma 0 [1 / 2 m(r 1)]
(2.65)
C k CM
Combining (2.62) and (2.63) it is possible to obtain a relation between and CL that does
not depend on V:
1/ r
1 2W
2 2
(2.66)
CL 0 SM a 0
Linking (2.66) with (2.65) we get the fraction V/C as a function of CL:
1 2m(r 1)
V Ma 0 1 2W 2r
(2.67)
C k CM CL 0 SM2a 02

Finally the range parameter can be expressed as follows:


1 2m(r 1)

Ma 0 (r 1)(1 2m) r CL
[1 (1 2m(r 1)) / 2r ]
V L 1 2W 2r
(2.68)
C D k CM 0 S CD
At given weight and Mach, the optimization of the range parameter only depends on the
last term of equation (2.68). To simplify, we will denote:
1 2m(r 1) 0.9049 if m 0
s 1 (2.69)
2r 0.9697 if m 0.08
This leads to the optimization of the following function of the lift coefficient:
CLs
f (C L ) (2.70)
CD0 KCL2
The maximum of this function is reached at CLbRMT as given by equation (2.71), where the
subscript bRMT indicates that it is the value of CL that produces the best range at fixed
Mach in the troposphere:

Aircraft Design, 2017


CRUISE PERFORMANCE 2-21

s CD0 s 2r 1 2m(r 1)
CLbRMT CLopt CLopt (2.71)
2s K 2s 2r 1 2m(r 1)
Drag coefficient and aerodynamic efficiency are given by:
2 4r
CD CD0 KCLbRMT
2
CD0 CD0 (2.72)
2s 2r 1 2m(r 1)

L CLbRMT s ( 2 s) 1 4r 2 [1 2m(r 1)]2 L


(2.73)
D CD 2 CD0K 2r D max
These results are extremely close to the condition of maximum aerodynamic efficiency:
CLbRMT / CLopt V / VminT (L/D)/(L/D)max
If m=0 0.909 1.049 0.9955
If m=0.08 0.970 1.015 0.9995
The altitude of this optimization has a pressure ratio that is obtained substituting (2.71) into
(2.66) and (2.60), giving:
2W K 2s 2W K 2r 1 2m(r 1)
r (2.74)
0M2 a 02 S C D0 s 0M2 a 02 S C D0 2r 1 2m(r 1)

Above results are valid while they are inside the limits of the troposphere: 1 0.22336.
If not, equations (2.71) to (2.74) must be disregarded and the maximum range parameter
in the troposphere is reached just at the tropopause. In that case the global maximum will
be in the stratosphere.

Optimization in the stratosphere


In the stratosphere, the temperature and the sound speed remain constant so, at fixed
Mach, the true airspeed V remains constant:
V M a M a0 TP (2.75)

Where TP indicates the temperature ratio at the tropopause (TP = 0.7519), which remains
constant at the stratosphere.
By equation (2.61), the specific fuel consumption varies like m. On the other hand, the
variation of is linked to the variation of CL by equation (2.58), which can be rewritten as:
2W 1
(2.76)
0M2a02TPS CL
Therefore the range parameter can be expressed as:
m
V L Ma 0 TP 0M2a02TPS C1Lm
(2.77)
CD k CM 2W C
D

At given weight and Mach, the optimization of the range parameter only depends on the
last term of equation (2.77). This leads to optimize the following function:
C1L m
f ( CL ) (2.78)
CD0 KCL2

Aircraft Design, 2017


2-22 CRUISE PERFORMANCE

The maximum of this function is reached at CLbRMS as given by equation (2.79), where the
subscript bRMS indicates that it is the value of CL that produces the best range at fixed
Mach in the stratosphere:
1 m CD0 1 m
CLbRMS CLopt (2.79)
1 m K 1 m
Drag coefficient and aerodynamic efficiency are given by:
2
CD CD0 KCLbRMS
2
CD0 (2.80)
1 m

L CLbRMS 1 m2 1 L
1 m2 (2.81)
D CD 2 CD0K D max
Again, these results are very close to the condition of maximum aerodynamic efficiency:
CLbRMS / CLopt V / VminT (L/D)/(L/D)max
If m=0 1 1 1
If m=0.08 1.083 0.961 0.9936
The density ratio of this optimized altitude is obtained by substitution of (2.79) into (2.76),
yielding (2.82). The results are valid in the stratosphere, with 0.29707.
2W K 1 m
(2.82)
0M2a02TPS CD0 1 m

Global optimization
Depending on the combination of aircraft parameters, Mach limitation and flight envelope,
a number of different situations can appear; for instance:
The maximum operational altitude of the aircraft is below the optimal altitude at the
troposphere given by equation (2.74). In that case the best range condition is at the
maximum operational altitude.
The global maximum is in the troposphere, given by (2.71) with (2.74) in bounds.
There is no local maximum in the troposphere since equation (2.74) is out of bounds.
The global maximum in the stratosphere, given either by equation (2.79) or by the
maximum operational altitude, if it is not possible to reach the optimum C L.
There is a local maximum at the troposphere, given by equation (2.71) with (2.74) in
bounds, and other local maximum at stratosphere given by equation (2.79). The global
maximum is determined by comparison of the respective values of range parameter.

Aircraft Design, 2017


CRUISE PERFORMANCE 2-23

2.5 Bibliography
References for chapter 2:
2R1. D. P. Raymer.
Aircraft Design: A Conceptual Approach.
AIAA (American Institute of Aeronautics and Astronautics) Education Series.
1st edition 1989 5th edition 2012.
2R2. E. Torenbeek.
Synthesis of Subsonic Airplane Design.
Delft University Press & Kluwer Academic Publishers.
1982.
2R3. L. M. Nicolai, G. E. Garichner.
Fundamentals of Aircraft and Airship Design.
Volume I Aircraft Design.
AIAA (American Institute of Aeronautics and Astronautics) Education Series.
1st edition 1975 Last edition 2010.
2R4. S. A. Brandt, R. J. Stiles, J. J. Bertin, R. Whitford.
Introduction to Aeronautics: A Design Perspective.
AIAA (American Institute of Aeronautics and Astronautics) Education Series.
1st edition 1997 2nd edition 2004.
2R5. L. R. Jenkinson, P. Simpkin, D. Rhodes.
Civil Jet Aircraft Design.
AIAA (American Institute of Aeronautics and Astronautics) Education Series.
1999.
2R6. D. Howe.
Aircraft Conceptual Design Synthesis.
Professional Engineering Publishing.
2000.
2R7. A. K. Kundu.
Aircraft Design.
Cambridge University Press.
2010.
2R8. M. Asselin.
An Introduction to Aircraft Performance.
AIAA (American Institute of Aeronautics and Astronautics) Education Series.
1997.

Aircraft Design, 2017

You might also like