6.28 Overexpanded Nozzle Flow—Shock Losses

There is an aspect of nozzle performance that is hampered by the presence of shocks inside the nozzle. This behavior, which is represented by a severe case of overexpansion, is not readily apparent from Equation 6.66. An overexpanded nozzle has an area expansion ratio A9/A8 in excess of that needed for a perfect expansion. This results in an exit static pressure, which falls below the ambient pressure p9/p0 < 1. The need for an abrupt adjustment of the static pressure in the emerging supersonic flow causes a shock formation on the nozzle lip. The mild cases of overexpansion are resolved through oblique shock waves on the nozzle lip. A greater mismatch between the static pressures of the jet and ambient calls for a stronger shock, which in essence increases the wave angle with respect to stream. Eventually, a normal shock is formed at the nozzle lip. In case the normal shock at the nozzle lip is found incapable of satisfying static pressure continuity across the jet slipstream, the normal shock is then brought inside the nozzle. At this time, the exit flow is subsonic and a potential boundary layer separation needs to be investigated. Let us first examine the flow environment near the exit of an overexpanded nozzle in supersonic flight. Figure 6.62 shows the external flow around a nozzle boattail and the jet geometry dominated by wave formations. The shear layer separating the inner and outer flowfields is labeled as the jet slipstream. Due to flow curvature near the nozzle exit, local static pressure is higher than ambient pressure, that is, local ambient static pressure near plane 9 is not p0, which we had assumed all along in our analysis.

images

FIGURE 6.62 A cartoon of the inner and outer flow of an overexpanded nozzle in supersonic flight shows a wavy jet slipstream with a varying static pressure inside the jet

Note that we intentionally omitted a drawing of the wall boundary layers on the primary nozzle and divergent flaps to avoid further crowding of Figure 6.62. This figure was produced, with simplifications, to develop an appreciation of the flow complexities in supersonic exhaust flow. Now, let us graphically examine higher levels of overexpansion in the above nozzle. With larger area ratios, assuming an attached flow, the static pressure at the exit plane drops, demanding a normal shock at the lip. Due to the presence of the wall boundary layer and a curved free shear layer (i.e., the jet slipstream), a lambda shock is formed at the lip, which is graphically depicted in Figure 6.63.

images

FIGURE 6.63 Overexpanded flow from a CD-nozzle shows a lambda-shock formation at the lip

Again for simplicity, the inner boundary layer and the vortex sheet emanating from the junction of lambda shock are not graphed in Figure 6.63. Here, we clearly see a curved shock formation, which is predominantly normal to the core flow and dominated by a lambda-shock geometry near the wall and exit shear layer. In the case of a more severe overexpansion, the shock will be drawn inside the nozzle and a subsonic jet will emerge at the exit. Figure 6.64 examines this kind of overexpansion.

images

FIGURE 6.64 Curved shock inside the nozzle creates a potential for boundary layer separation and produces a subsonic exit flow

From Figure 6.64, we note that a separated boundary layer could alter the picture of exit flow, in particular with the flow direction as well as the effective exit flow area. How much of this complexity can we capture by simple one-dimensional models and correction factors? For example, in our gross thrust coefficient approach to account for viscous losses and nonaxial exit flow, are there any provisions for curved shocks or separated flows, such as demonstrated by some overexpanded nozzle flows? The short answer is no. The long answer starts with a maybe! A plausible approach calls for identifying and bracketing off zones of invalidity from our performance charts that we derived based on simple 1D loss models and assumptions. This approach relies, in part, on the experience base of the manufacturing company and other available experimental data in the literature. For example, in separated nozzle flows, the physical meaning of the velocity coefficient CV and the angularity loss factor CA are completely lost, based on our definitions and simplifying assumptions. Consequently, the nozzle gross thrust coefficient Cfg, as described by Equation 6.69, becomes at best inaccurate and at worst meaningless. Here, we need to exert judgment when dealing with overexpanded nozzle flows that include a shock inside the nozzle. Let us reproduce some of the criteria found in the literature for a passage shock that causes flow separation. In Kerrebrock (1992), a reference is made to Zukoski (1990) who suggests a separation point criterion in an overexpanded supersonic nozzle according to

(6.70)numbered Display Equation

where ps is the static pressure upstream of the shock, p0 is the ambient static pressure, and Ms is the gas Mach number upstream of the shock. Over a wide operating range of the nozzle, that is, NPR ranging between 2 and 50, the separation pressure ratio, Kerrebrock notes that ps/p0 ranges between 0.4 and 0.625. Mattingly (1996) recommends a ps/p0 ≈ 0.37, following Summerfield, Foster, and Swan (1954) and Sutton (1992) who recommend a rough separation criterion of ps/p0 ≤ 0.4. In summary, several simple or more sophisticated models for shock separation criterion are introduced, which roughly identify the zones of invalidity of our simple nozzle theory. Figure 6.65 shows the ideal gross thrust coefficient with nozzle area ratio as a function of nozzle pressure ratio. The maximum Cfg = 1 for p9/p0 = 1.0.

images

FIGURE 6.65 An ideal nozzle, i.e., pt = constant, Tt = constant, CD8 = CV = CA = 1, gross thrust coefficient with Zukoski separation criterion in overexpanded mode

Note that the nozzle overexpanded mode of operation is more prevalent for a rocket engine than an airbreathing aircraft engine. A rocket engine that operates between the ambient pressures of sea level, ∼14.7 psia (1 atm.), and near vacuum conditions of space, requires a large nozzle area expansion ratio (A9/A8) to optimize its performance. A nozzle, which is optimized for high altitude, then operates in an overexpanded mode in lower altitudes. Consequently, the thrust loss and shock interactions inside the nozzle are serious realities of rocket nozzle design. However, the airbreathing engine of an aircraft invariably operates with a smaller area expansion ratio, hence it experiences less severe over-expansion problems. In the hypersonic regime, the airbreathing and a rocket nozzle attain geometric and operational similarities and thus face the same challenges for off-design operations.

Now, we are ready to address the nozzle throat size and scheduling with afterburner operation. Also, we start the discussion on the nozzle exit area scheduling within the flight envelope and its impact on external drag penalties due to boattail and base flowfield.

6.29 Nozzle Area Scheduling, A8 and A9/A8

The exhaust nozzle throat of a modern aircraft engine operates in a choked mode over a wide operating range within the flight envelope. Consequently, the nozzle throat serves as the low-pressure turbine’s backpressure controller or equivalently as the engine aft throttle station. With an afterburner in operation, the exhaust gas temperature increases significantly with a subsequent drop of density. To accommodate a lower density gas, a choked nozzle needs to open up its throat to pass almost the same mass flow rate as with no afterburner operation. The one-dimensional continuity equation for a choked nozzle throat may be written as

We note from Equation 6.71 that the term that undergoes the most dramatic change with the afterburner operation is Tt8 due to heat release in the combustion process. The other terms in the equation undergo smaller changes. For example, the mass flow rate increases by ∼2–4%, the gas constant R hardly changes as the molecular weight of the new mixture has not changed much by the addition of the fuel; the total pressure drops with the afterburner, perhaps by 2–5%; the specific heat ratio γ8 changes with temperature and may go from ∼1.33 to ∼1.30 or 1.25. In its simplest form, the nozzle throat scheduling with afterburner for a turbojet engine may be written as

(6.72)numbered Display Equation

If the aim is to be more exact with the nozzle throat area scheduling, then one must do the bookkeeping of all the terms in the continuity equation more carefully, namely,

where the primed terms are the afterburner-on quantities.

6.30 Nozzle Exit Area Scheduling, A9/A8

Unlike the choked nozzle throat that required scheduling with an afterburner operation (i.e., a single parameter control), the nozzle exit area scheduling involves a multiparameter optimization study. The reason for this added complexity is rather obvious. Exit area variation of the nozzle impacts external aerodynamics of the nacelle boattail and base flowfields, and hence external drag. Therefore, if we gain an increment of gross thrust due to a better nozzle expansion of say p9p0, and as a result incur an increment of drag rise due to nacelle boattail and base flow changes, then the nozzle area ratio A9/A8 needs to optimize the installed performance, namely,

(6.74)numbered Display Equation
images

FIGURE 6.66 Ideal nozzle throat area variation for an afterburning turbojet engine

images

FIGURE 6.67 The impact of flow losses and gas property variation with afterburner combustion on a turbojet nozzle throat area [for γ = 1.33,  γ′ = 1.30,  pt8/pt8 = 0.96,  f = 0.02,  fab = 0.04]

Here, we are again reminded that the internal optimization of an airbreathing engine inlet and exhaust system is meaningless in isolation. Rather, the engine-installed performance should be optimized at each flight phase within the flight envelope of the aircraft. Traditionally, the nozzle exit area of a variable geometry nozzle assumed only two positions. One closed and the other open, very similar to the nozzle throat with/without afterburner. But, the limitation of only two positions for the nozzle exit area seems unnecessary today especially in light of advanced propulsion control system of modern aircraft. Here, the goal of nozzle area scheduling control is to achieve a continuous improvement of installed performance with altitude, speed, aircraft attitude, and acceleration rather than that achieved by a bang–bang (a two-position, open-closed) controller.

The contributors to the aftbody aerodynamics on an engine installation are the contributions from the boattail and base areas of the nacelle. The geometry of nozzle boattail and base areas and their impact on aftbody drag are shown in Figure 6.68 (Bowers and Tamplin, 1985). We note that the base pressure force contribution is in the thrust direction and when integrated with the boattail drag contribution, they yield the “total” pressure drag penalty, as shown in Figure 6.68. Also, a graphic definition of base area is helpful and is shown in Figure 6.69 (from Whitford, 1987).

images

FIGURE 6.68 Pressure drag contribution of the boattail and base as a function of geometry of a propulsion nacelle installation.

Source: Bowers and Tamplin 1985. Reproduced with permission from AIAA

images

FIGURE 6.69 Definition sketch of candidate nozzle/aftbody base areas. Source: Adapted from Whitford 1987

An installed nozzle performance optimization that accounts for the internal as well as external losses is reported by Younghans and Dusa (1993) and reproduced here in Figure 6.70. First, note that nozzle area ratio A9 /A8 that is needed for a perfect expansion does not give the best-installed performance. Second, note that each loss parameter has a different area ratio for its optimization. For example, flow angularity loss vanishes for A9/A8 = 1.0. The expansion loss vanishes at full expansion (p9 = p0). The friction penalty holds in a somewhat steady manner across the nozzle area ratios. Finally, the internal performance curve is shifted down by the external drag penalties. This leads to a shift in the area ratio for optimized installed performance.

images

FIGURE 6.70 Internal/external factors contributing to an optimized nozzle performance. Source: Younghans and Dusa 1993

6.31 Nozzle Cooling

The operational temperatures of a modern, multitask exhaust system, with an afterburner, have required an increasing level of active nozzle cooling. In a turbofan engine, the fan stream could provide the required nozzle cooling through an engine bypass system. The supersonic inlet/engine flow match often requires an inlet excess air bypass/bleed system to be integrated into the propulsion system design. The excess air has traditionally been used for engine bay and nozzle cooling purposes. Figure 6.71 shows a schematic drawing of a flow network between an inlet and an exhaust system in a turbofan engine (from Younghans and Dusa, 1993). A typical engine airflow variation with flight Mach number and its comparison with the inlet supply are shown for a Mach 5 aircraft by Dusa (1988).

images

FIGURE 6.71 Schematic drawing of an exhaust system cooling air sources in a turbofan engine. Source: Younghans and Dusa 1993

The primary goal of nozzle cooling is to operate the exhaust system components, that is, the liner, primary and secondary nozzle walls/ramps, seals, and the actuators at a temperature that allows for reliability of operation and good material life expectancy. The material operating temperature should be well below the melting point of the material. A very useful graph from Wazyniak (2000) is reproduced here in Figure 6.72; it shows the design temperature levels of jet engine materials. It is useful to examine the design versus the melting point temperature of different materials in addition to material density, as shown in Figure 6.72. The typical uses of the materials listed are: nickel-based alloys in turbines, titanium in compressors and fan blades, aluminum for aircraft and nacelle skin, intermetallics and nonmetals in the high-temperature environment of afterburners and exhaust systems. The acceptable operating temperature of metals (e.g., stainless steel) extends to ∼1700°F (or ∼927°C), whereas ceramics may operate up to ∼2800°F (or ∼1538°C). Carbon-carbon offers an extended operating temperature of up to ∼3400°F (or ∼1870°C). There are also refractory metals, such as tungsten and molybdenum that offer higher operating temperatures than stainless steel (of up to ∼2880°F). However, refractory metals have exhibited manufacturing-related problems as well as operational problems with embrittlement and oxidation.

images

FIGURE 6.72 Melting point/design temperature characteristics of various jet engine materials. Source: Wazyniak 2000

Now, let us examine the flight total temperature, or ram temperature, as it increases with flight Mach number. Assuming an adiabatic flow of a calorically perfect gas, the inlet air total temperature follows

(6.75)numbered Display Equation

Now assuming the supersonic/hypersonic flight occurs in the stratosphere (between 50 and 150 kft) where the static temperature is ∼ − 60°F = constant, then the air ram temperature with flight Mach number varies parabolically according to

(6.76a)numbered Display Equation
(6.76b)numbered Display Equation

A graph of the inlet air ram temperature in stratosphere (for constant γ) is shown in Figure 6.73.

images

FIGURE 6.73 Stagnation temperature variation with flight Mach number in stratosphere (50 kft < alt. < 150 kft)

We note that the ratio of specific heats γ decreases with temperature, and hence the ram temperature calculation based on cold γ of 1.4 leads to an overestimation of the temperature. To put this in perspective, for a Mach 6 flight in stratosphere, Tt0 calculated based on γ = 1.4 is 2820°F, where as γ = 1.3 yields Tt0 of 2100°F. There is a ∼720°F discrepancy between the two temperatures; although significant, it does not change the arguments about the special cooling needs of high-speed flight. Now, let us accept the goal of ∼3000°F operating temperature of say carbon–carbon, the inlet air becomes too hot to be used as a coolant for flight Mach numbers M0 = 6+. Therefore, the coolant needs to be precooled in a heat exchanger, which adds weight and complexity to the engine. The other alternative is to use a cryogenic fuel that acts as a heat sink for the propulsor walls, before its injection in the combustor and reaction with air or another oxidizer, as in a rocket engine. To use a cryogenic fuel as a coolant is a common and tried approach in rocket propulsion. The method is called regenerative cooling. From the Saturn-V Apollo launch vehicle (all three stages) to the Space Shuttle Main Engine, nozzles and the thrust chambers are regeneratively cooled using fuels and sometimes oxidizers. Cryogenic fuels of interest are liquid hydrogen, H2, and methane, CH4. Both of these fuels are available in abundance in nature in various liquid and gaseous forms. The heat sink capacity of these fuels is in part quantified by their latent heat of vaporization, namely, 446 kJ/kg for liquid hydrogen and 510 kJ/kg for liquid methane.

6.32 Thrust Reverser and Thrust Vectoring

Thrust reversers are used to decelerate the aircraft upon landing (for a shorter runway), to reduce wear on the landing gear brakes upon touchdown and landing roll, and to enhance maneuverability of a fighter aircraft in flight. The goals of thrust vectoring nozzles are to enable an aircraft to takeoff and land either vertically or use a very short runway (<500 ft) as well as to achieve supermaneuverability in flight. Thrust vectoring is offered as a means of reducing the size or ultimately eliminating the tail (i.e., tailless aircraft), as well as to provide the aircraft with stall recovery. Additional goals of thrust vectoring are the low-speed flight control where the conventional aerodynamic control surfaces become ineffective and the high angle of attack (the so-called high-alpha) capabilities. The latter two goals fall under the supermaneuverability category. To provide some degree of reverse thrust, we need to divert the exhaust jet partially in the flight direction. This is commonly achieved through a block and turn mechanism designed into exhaust systems of aircraft today. Two such concepts have been the clamshell and cascade thrust reversers. The clamshell thrust reverser is often used for nonafterburning engines to reverse the turbine exhaust gases, and the cascade type is typically used as a fan exit flow reverser in turbofan engines. Schematic drawings of the two types of thrust reverser are shown in Figure 6.74 (from a Pratt & Whitney Aircraft, 1980).

images

FIGURE 6.74 The use of clamshell and cascade thrust reversers in a turbofan engine.

Source: Reproduced by permission of United Technologies Corporation, Pratt & Whitney

For an elemental mass flow rate that attains an exit velocity from a reverser of Ve magnitude, which exits at an angle of θ with respect to the flight direction, a reverse thrust equal to Vecos θ is created. This needs to be integrated over the exit of the turning vane cascade to evaluate the reverse thrust. A schematic diagram of a fan reverser is shown in Figure 6.76.

(6.78)numbered Display Equation
images images

FIGURE 6.75 The thrust reverser on a separate- and mixed-flow turbofan engine with the contributions of fan and core. Source: Dietrich and Kuchar 2000

images

FIGURE 6.76 A fan reverser demonstrating leakage and cascade vane angle U

The actual force exerted by the fluid on the blocker and the turning vanes, that is, the reverse thrust force exerted on the aircraft, is established by a momentum balance of the fluid across the blocker and turning vanes, such as

The terms of Equation 6.79 represent momentum and pressure–area terms, similar to the engine net thrust equation. The two positive terms on the RHS of Equation 6.79 dominate the bulk of reverse thrust force generation. The two negative contributions are due to leakage and pressure imbalance in the fan duct, p13, and the ambient pressure p0. We may estimate these terms to get a feel for the reverser effectiveness. For example, the turning vane cascade can be assumed to operate with an inlet total pressure of the fan, that is, pt13, and an exit pressure of the ambient air, p0. We may also assume an adiabatic efficiency of the expansion process in the turning vanes, similar to a nozzle. The vane angle may be assumed to be ∼45°, and the leakage flow may be ∼5% of the incoming flow. To account for the boundary layer blockage in the turning vanes, we may use a discharge coefficient CD similar to CD8. We may consider the air speed through the blocker is reduced by a factor, for example, V13 ∼ 1/2V13 (note that a mistake in here will not be very large as the leakage mass flow rate is a small fraction of the incoming flow). For the fan exit speed V13, we may estimate that based on a fan exit Mach number of ∼0.5. The static pressure p13 may be estimated by the fan total pressure ratio and the M13 of ∼0.5.

The British Harrier, AV-8B, is the first operational fighter aircraft to achieve vertical takeoff through thrust vectoring. The Harrier’s four rotating nozzles turn the engine flow to a near vertical direction to achieve hover, takeoff, and landing. A three-bearing swivel nozzle in Figure 6.77 shows the mechanism for thrust vectoring in an exhaust system (from Dusa, Speir, and Rowe, 1983). The nozzle is of convergent design, hence most suitable for low nozzle pressure ratio (i.e., NPR < 6) operation of a subsonic V/STOL aircraft.

images

FIGURE 6.77 Three-bearing swivel nozzle.

Source: Dusa, Speir, and Rowe 1983. Reproduced with permission from AIAA

For a supersonic application, a single-expansion ramp nozzle (SERN) offers jet deflection capability through turning of the upper/lower flaps or engaging a separate deflector mechanism. The augmented-deflector exhaust nozzle (Figure 6.78) is capable of 110° thrust vectoring angle. The deflecting flap SERN nozzle is shown in Figure 6.79 (from Dusa, Speir, and Rowe, 1983). The two-dimensional geometry of these nozzles allows for low-drag, body-blended exhaust system integration.

images

FIGURE 6.78 Single expansion ramp configuration in an augmented-deflector exhaust nozzle design (ADEN) for V/STOL applications.

Source: Dusa, Speir, and Rowe 1983. Reproduced with permission from AIAA

images

FIGURE 6.79 Single expansion ramp exhaust nozzle configuration with deflecting flaps suitable for STOL application.

Source: Dusa, Speir, and Rowe 1983. Reproduced with permission from AIAA

Here, we note that the lower flap also serves as a reverser blocker in the STOL/SERN nozzle configuration, which is shown in Figure 6.79. To get a feel for the cooling requirements of an advanced thrust-vectoring nozzle with an afterburner, we look into the experimental results of ADEN tests (at flight weight, self-cooled with the fan air on a thrust stand at NASA-Lewis). To achieve a surface temperature below the hot streak design limit of 2700°F, a ∼5.7% cooling air was required at maximum afterburner operation of Tt8 = 3028°F and the maximum nozzle pressure ratio of 15 (from Dusa and Wooten, 1984). Now, let us examine the following twin configuration (Figure 6.80).

images

FIGURE 6.80 Schematic drawing of a multifunctional 2D-CD nozzle. Source: Younghans and Dusa 1993

Another concept for vector thrust–thrust reverser nozzle is shown as a multiflap 2D-CD nozzle in Figure 6.80 (from Younghans and Dusa, 1993). The twin configuration shown uses segmented, that is, split, divergent flaps, which are used to achieve pitch and roll control. The yaw control requires the use of thrust reverser, which causes a reduction in axial thrust. Note that the convergent flaps serve as the blocker in the reverse thrust mode.

6.33 Hypersonic Nozzle

In this section, we explore the challenges of a hypersonic nozzle design. There are several design conflicts in the hypersonic exhaust system that stem from the range of flight speeds, that is, from takeoff to Mach > 5 and the range of altitude, namely, from sea level takeoff to upper atmosphere, or 100+ kft. The combination of altitude and Mach number constitutes the flight envelope of an aircraft, and in Figure 6.81 we examine such a flight envelope (from Dusa, 1988). The shaded area in the corner of Figure 6.81 represents the flight envelope of conventional commercial and military aircraft. We examine the representative nozzle pressure ratio for airbreathing hypersonic engines in Example 6.12.

images

FIGURE 6.81 Flight envelope of hypersonic vehicles with altitude in kft. Source: Dusa 1988. Courtesy of NASA

The nozzle area ratio as a function of nozzle pressure ratio is shown in Figure 6.83, using a log-linear scale. We note that a hypersonic airbreathing propulsor will require a very large nozzle area ratio, whose integration into the vehicle is the challenge number one.

images

FIGURE 6.82 Exponential growth of stagnation pressure with Mach number

images

FIGURE 6.83 Nozzle exit-to-throat area ratio for a perfectly expanded nozzle, assuming a perfect gas with the ratio of specific heats γ a constant

To reduce the burden of excessive weight, a hypersonic nozzle will perform an external expansion of the gas, in a geometrically similar manner as a plug nozzle. However, instead of a plug, the nozzle external expansion surface may be the lower surface of the fuselage aftbody. A schematic drawing of such a nozzle is shown in Figure 6.84 where the expansion waves in the exhaust plume are shown as dashed lines and the compression waves as solid lines. Remember from supersonic aerodynamics that the wave reflection from a free boundary, as in the exhaust plume mixing layer, is in an unlike manner and the wave reflection from a solid boundary is in a like manner. Hence, a shock is reflected as an expansion wave from a free boundary and reflects as a shock from a solid surface. A similar statement may be made regarding an expansion wave reflections.

images

FIGURE 6.84 Schematic drawing of a generic hypersonic vehicle powered by an airbreathing engine shows a highly integrated inlet and exhaust system with the airframe

6.34 Exhaust Mixer and Gross Thrust Gain in a Mixed-Flow Turbofan Engine

A mixed-flow turbofan engine offers the potential of gross thrust gain and a reduction in fuel consumption as compared with an optimized separate flow turbofan engine. In this section, we examine the extent of gross thrust improvements and the nondimensional parameters that impact the performance gains.

To demonstrate the gain in the gross thrust realized by mixing a cold and a hot stream, we first treat the problem by assuming an ideal mixer. Expanding the cold, hot, or mixed gases in a nozzle to an exit pressure of p9, we shall realize a conversion of the thermal energy into the kinetic energy in the exhaust nozzle. The thermodynamics of the expansion process for the three streams is shown in Figure 6.85.

images

FIGURE 6.85 An ideal expansion of a cold, hot, and a mixed stream to the same backpressure p9

The gross thrust for a separate exhaust turbofan engine, assuming the nozzles are perfectly expanded is

We may write the isentropic pressure ratio for the temperature ratios in Equation 6.81 to get

(6.82)numbered Display Equation

The mixed exhaust nozzle shall have a gross thrust of

(6.83)numbered Display Equation

We note that the pressure terms in the brackets are identical hence they cancel when we express the ratio of gross thrusts. Assuming the same gas properties in the three streams, namely, we get

(6.84)numbered Display Equation
(6.85)numbered Display Equation

Therefore the rise of the gross thrust as a result of mixing a cold and a hot stream may be expressed in terms of the nondimensional parameters of the respective temperature ratio and the mass flow ratio, namely, the bypass ratio α. This is expressed in the following equation.

6.35 Noise

High-performance gas turbine engines are very noisy machines, by design! Energy transfer in turbomachinery can only take place through unsteady means, as presented in Chapter 8. Also, to maximize energy transfer, modern engines use high-speed wheels of rotating and stationary or counterrotating blade rows. These always involve supersonic tip speeds, which create shocks and their associated discrete tone noise. Since the energy transfer is unsteady, the blade wakes are unsteady with periodic vortex shedding structure. The unsteady wakes of a blade row are then chopped by the next row of blades, as they are in relative motion. In addition, there are ample opportunities for resonance to occur in blade row-blade row interaction. Due to blade elasticity and aerodynamic loads, the blades vibrate in various modes, for example, first bending, second bending, first torsional mode, coupled bending and torsional mode, and higher order modes. The unsteady mechanism of energy release in a combustor creates combustion noise. Finally, the pressure fluctuation in the exhaust jet due to turbulent structure and mixing is the source of broadband jet noise.

images

FIGURE 6.86 Gain in gross thrust due to an ideal mixer

Modern fan blades that are thin and incorporate blade sweep, and the inlet fan and core exit ducts that are treated with an acoustic liner, are successful solutions that alleviate the turbine noise problem. For example, the acoustic duct liners, which are Helmholtz resonators embedded in the wall, attenuate the dominant frequency in the transonic fans’ and LPTs’ noise spectra, namely the discrete tone noise at the blade passing frequency. In addition, proper selection of blade numbers in successive blade rows and blade row spacing are the solutions that alleviate, by and large, the intrablade row resonance problem.

6.35.1 Jet Noise

In a classical paper, Lighthill (1952) showed that the acoustic power in a subsonic turbulent jet is proportional to the eight power of the jet speed, Vj, that is,

(6.87)numbered Display Equation

Therefore, to reduce the broadband noise from an aircraft exhaust, the jet speed needs to be reduced. Interestingly, propulsive efficiency is also improved with reduced exhaust speed, as we discovered in the cycle analysis (Chapter 4). The development of high and ultra-high bypass ratio turbofan engines accomplish both goals, that is, higher propulsive efficiency and reduced jet noise, simultaneously. Also, the concept of Variable Cycle Engine (VCE) promises the capability of higher bypass ratio at takeoff and landing and reduced bypass ratio at cruise condition, which is advantageous in the reduction of the airport’s community noise. Since the kinetic power in a jet is proportional to Vj3, an acoustic efficiency parameter may be defined as the ratio of the radiated acoustic to the jet kinetic power (for derivation see Kerrebrock, 1992), which is then proportional to the fifth power of the jet Mach number, Mj0, which in this derivation is defined as the ratio of Vj and the ambient speed of sound, a0, that is,

(6.88)numbered Display Equation

This efficiency parameter may be used to compare different noise suppression concepts. The noise suppressors of the 1950s that were developed originally for turbojets added significant weight to the engine and caused significant loss of gross thrust. The turbofan engines of the 1960s and 1970s alleviated the jet noise problem based on their reduced jet speed and their blade and duct treatments also reduced the turbine noise; but still more stringent noise regulations were on the horizon. These limitations are set by the Federal Aviation Regulation – Part 36 Stage 3 and the new FAA Stage 4 Aircraft Noise Standards (that is used for the certification of turbofan-powered transport aircraft). One thing is for certain and that is the growth in commercial aviation and its impact on society will certainly demand yet more stringent regulatory aircraft noise standards in the future. Beyond a reduction in jet speed, as in high bypass ratio turbofan engines, are there any other means of reducing jet noise without paying a severe penalty on gross thrust? The jet noise reduction depends on many parameters but predominantly on the nozzle pressure ratio (NPR) and the jet Mach number. In this section, the most promising development in transonic jet noise reduction is briefly presented.

For the past two decades, research and development by NASA’ Glenn Research Center and the aircraft engine and airframe industry on jet noise reduction has led to the development of a new winning technology: the Chevron Nozzle. An excellent review article by Zaman et al. (2011), which chronicles the development of mixing devices, mainly tabs and chevrons, is suggested for further reading.

6.35.2 Chevron Nozzle

Chevrons are triangular serrations at the nozzle exit plane that have the potential of “calming” turbulent jets (Zaman et al., 2011). The calming effect stems from a redistribution of azimuthal vorticity at the nozzle exit that attains certain streamwise components. For optimal redistribution of vorticity, the chevron tips have to slightly penetrate the jet. Excessive penetration gives rise to thrust loss and an insignificant dipping into the jet does not produce the desired sound pressure level (SPL) reduction. Figure 6.87 shows a baseline nozzle that is used to compare the aeroacoustic and propulsive performance of various chevron configurations on the core and/or the fan nozzles in a separate flow turbofan engine.

images

FIGURE 6.87 Baseline nozzle (3BB) and examples of Chevron and tab configurations (3C12B, 3I12B and 3T24C24, respectively) in static test. Source: Saiyed, Mikkelsen, and Bridges, 2000. Courtesy of NASA

The most promising of the chevron configurations in static test were then flown to demonstrate takeoff, landing and the flyover noise reduction effectiveness. Naturally, on the propulsion side thrust loss coefficient at cruise conditions is the measure of viability of any noise reduction device.

The benefits in noise reduction of chevron-equipped aircraft in flyovers, that is, community noise, is first measured (using a logarithmic scale known as the Bel or the more commonly used unit of one tenth of a Bel, i.e., the dB) in the frequency range where human ear is sensitive (i.e., between 50 and 10 kHz) and subsequently weighted towards the frequencies where the human ear is most sensitive, that is, 2–5 kHz. There are other corrections that contribute to the calculation of what is known as the Effective Perceived Noise Level (EPNL) in dB; this is the basis of FAR-36 certification limits. Kerrebrock (1992) offers a clear methodology to EPNL calculations and is suggested for further reading on engine noise. There is ample data that support a reduction in turbulent kinetic energy in a jet where streamwise vorticity is injected at the nozzle exit, as produced by chevrons. The significance of the chevron technology is in its light weight and negligible impact (i.e., loss) on nozzle gross thrust, which is measured at 0.25%.

At the outset, the goal of jet noise reduction was to achieve 2.5 EPNL (dB) in flyover, that is, community, noise reduction while limiting the thrust penalty in cruise to less than 0.5%. As evidenced in Table 6.1, where the noise benefit and cruise thrust loss data from Zaman et al. (2011) are shown, there are several chevron configurations that achieve the desired performance.

TABLE 6.1 Noise benefit and cruise thrust loss data

Noise benefit Loss in thrust
Configuration (EPN dB) coefficient at cruise (%)
3C12B 1.36 0.55
3I12B 2.18 0.32
3I12C24 2.71 0.06
3T24B 2.37 0.99
3T48B 2.09 0.77
3T24C24 0.43
3T48C24 0.51
3A12B 0.34
3A12C24 0.49

The success of the chevron nozzle has made it into a standard component on several new aircraft, such as the Boeing 787.

6.36 Nozzle-Turbine (Structural) Integration

The turbine exhaust case, support struts, and a tail cone flange are shown in Figure 6.88 (from Pratt & Whitney, 1980). The turbine exhaust cone houses the turbine bearings and provides for a smooth flow to the nozzle (or afterburner).

images

FIGURE 6.88 Turbine exhaust’s case, struts, and a tail cone flange.

Source: Reproduced by permission of United Technologies Corporation, Pratt & Whitney

images

FIGURE 6.89 Drawing of exhaust cone, jet pipe, and a convergent nozzle.

Source: Reproduced with permission from Rolls-Royce plc

The exhaust cone, jet pipe, and a convergent nozzle are shown schematically in Figure 6.89 (from Rolls-Royce, 2005). In this view of the exhaust system, the turbine rear support struts are also shown.

6.37 Summary of Exhaust Systems

An aircraft exhaust system is a multifunctional component whose primary mission is to convert fluid thermal energy into directed forces in any desired direction (x, y, z) efficiently. These forces are usually in vector thrust mode, including hover, as well as thrust reversing directions. The system weight (and complexity) and cooling requirements are the optimizing parameters for the design and selection of exhaust systems. Advanced nozzle cooling requirements are in the ∼5–10% range. Airbreathing engines with afterburner require a variable geometry nozzle to open the throat with afterburner operation. The nozzle throat scheduling with the afterburner throttle setting is to decouple the afterburner/exhaust system from the gas generator. The nozzle exit area scheduling has to include the net effect of internal performance gains as well as the external boattail drag penalties. The internal performance is optimized when the nozzle is perfectly expanded. The build up of boundary layer in the convergent or primary nozzle is accounted for by a flow (discharge) coefficient at the throat, CD8. The effect of total pressure loss on the exhaust velocity in the divergent section of the nozzle is accounted for by a velocity coefficient CV. The effect of nonaxial exhaust flow appears as a reduction in axial force produced by the nozzle. The appropriate accounting of this loss is made using the flow angularity loss coefficient CA. Lobed mixers in a mixed-stream turbofan engines improve gross thrust by a few percentage points. Off-design operation of supersonic nozzles in overexpanded mode causes shock waves to be formed and a corresponding loss of thrust. Plug nozzles have a superior overexpanded nozzle performance over the convergent–divergent nozzles, but require additional cooling and thus system complexity and weight. In hypersonic flight, the nozzle pressure ratio reaches into several hundred level, somewhat similar to a rocket engine. The required area expansion ratio for efficient expansion of the high-pressure gas demands integration with vehicle. To achieve reduced signature, an exhaust system could employ high-aspect ratio rectangular nozzles with offset to shield the turbine as well as enhanced mixing to reduce the jet temperature. The exit plane of the nozzle placed on the upper surface of the aircraft fuselage or wing provides shielding from ground radar.

Further readings on exhaust systems and installation (references 4, 7, 8, 24 and 26) are recommended.

References

  1. Abbott, J.M., Anderson, B.H., and Rice, E.J., “Inlets, Ducts and Nozzles, ” Aeropropulsion ’87, NASA CP-3049, 1987.
  2. Abernethy, R.B. and Roberts, J.H., In-Flight Thrust Determination and Uncertainty, SAE Special Publication, SP-674, 1986.
  3. Albers, J.A, et al., “Aerodynamic Analysis of Several High Throat Mach Number Inlets for the Quiet, Clean, Short-Haul Experimental Engine, ” NASA TMX-3183, 1975.
  4. Aulehla, F. and Lotter, K., “Nozzle/Airframe Interference and Integration, ” in AGARD Lecture Series, LS-53, Paper 4, 1972.
  5. Anderson, B.H., “Three-Dimensional Viscous Design Methodology of Supersonic Inlet Systems for Advanced Technology Aircraft, ” Numerical Methods for Engine-Airframe Integration, Eds. Murthy, S.N.B. and Paynter, G.C., Progress in Astronautics & Aeronautics, Vol. 102, AIAA, Washington, DC, 1986, pp. 431–480.
  6. Anderson, J.D., Jr., Fundamentals of Aerodynamics, 4th edition, McGraw-Hill, New York, 2005.
  7. Anon., Symposium on “Developments in Aircraft Pro-pulsion, ” Haarlem, the Netherlands, 1987.
  8. Berman, K. and Crimp, F.W., “Performance of Plug-Type Rocket Exhaust Nozzles, ” ARS Journal, Vol. 31, No. 1, January 1961.
  9. Bowers, D.L. and Tamplin, G., “Throttle-Dependent Forces, ” Chapter 5in Thrust and Drag: Its Prediction and Verification, Eds Covert, E.E. et. al., Progress in Astronautics and Aeronautics, Vol. 98, AIAA, New York, 1985.
  10. Cochran, D.L. and Kline, S.J., “The Use of Short Flat Vanes for Producing Efficient Wide-Angle Two-Dimensional Subsonic Diffusers, ” NACA TN 4309, September 1958.
  11. Dietrich, D.A. and Kuchar, A.P., “High Bypass Turbo-fan Nacelles for Subsonic Transports, ” UTSI Course notes, April 2000.
  12. Dusa, D.J., “Nozzles and Turbo-Ramjet Propulsion for Mach 0 to 6 Aircraft, ” Short Course Notes in Hypersonic Propulsion, NASA-Lewis, 1988.
  13. Dusa, D.J., Speir, D.W., and Rowe, R.K., “Advanced Technology Exhaust Nozzle Development, ” AIAA Paper No. 83-1286, 1983.
  14. Dusa, D.J. and Wooten, W.H., “Single Expansion Ramp Nozzle Development Status, ” AIAA Paper No. 84-2455, 1984.
  15. Farokhi, S., Sheu, W.L., and WU, C., “On the Design of Optimum-Length Transition Ducts with Offset: A Computational Study, ” Computers and Experiments in Fluid Flow, Eds Carlomagno, G.M. and Brebbia, C.A., Springer Verlag, Berlin 1989, pp. 215–228.
  16. Hancock, J.P. and Hinson, B.L., “Inlet Development for the L-500, ” AIAA Paper No. 69-448, June 1969.
  17. Hesse, W.J. and Mumford, N.V.S., Jr., Jet Propulsion for Aerospace Applications, 2nd edition, Pitman Publishing Corp., New York, 1964.
  18. Holmes, B.J., “Progress in Natural Laminar Flow Research, ” Paper Presented at AIAA/NASA General Aviation Technology Conference, Wichita, Kansas, July 1984.
  19. Johnson, C.L., “Some Development Aspects of the YF-12A Interceptor Aircraft, ” AIAA Paper No. 69-757, July 1969.
  20. Kantrowitz, A. and Donaldson, C. du P., “Preliminary Investigation of Supersonic Diffusers, ” NACA ACR L5020, 1945.
  21. Kerrebrock, J.L., Aircraft Engines and Gas Turbines, 2nd edition, MIT Press, Cambridge, Mass, 1992.
  22. Kimzey, W.F., Wehofer, S., and Covert, E.E., “Gas Turbine Engine Performance Determination, ” Thrust and Drag: Its Prediction and Verification, Ed. Covert, E.E., AIAA progress in Astronautics & Aeronautics, Vol. 98, AIAA, Washington, DC, 1985, pp. 47–119.
  23. Kline, S.J., Abbott, D.E., and Fox, R.W., “Optimum Design of Straight-Walled Diffusers, ” Journal of Basic Engineering, 81, Series D, No. 3, September 1959, pp. 321–331.
  24. Kozlowski, H. and Larkin, M., “Energy Efficient Engine Exhaust Mixer Model Technology Report, ” NASA CR-165459, June 1981.
  25. Kline, S.J., “On the Nature of Stall, ” Journal of Basic Engineering, 81, series D, No. 3, September 1959, pp. 305–320.
  26. Kuchar, A.P., “Variable Convergent–Divergent Exhaust Nozzle Aerodynamics, ” Chapter in Aircraft Propulsion System Technology and Design, Ed.Oates, G.C., AIAA Education Series, AIAA Inc., Washington, DC, 1989.
  27. Kuechemann, D. and Webber, J., Aerodynamics of Propulsion, McGraw-Hill, New York, 1953.
  28. Mattingly, J.D., Elements of Gas Turbine Propulsion, McGraw-Hill, New York, 1996.
  29. Mount, J.S. and Millman, V., “Development of an Active Laminar Flow Nacelle, ” AIAA Paper 85-1116, 1985.
  30. MIL-E-5008B, “Military Specifications-Engines, Aircraft, Turbojet, Model Specifications for, ” January 1959.
  31. Oates, G.C., Aerothermodynamics of Gas Turbine and Rocket Propulsion, AIAA, Washington, DC, 1988, pp. 198–199.
  32. Oswatitsch, K., “Pressure Recovery for Missiles with Reaction Propulsion at High Supersonic Speeds (The Efficiency of Shock Diffusers), ” NACA TM 1140, 1947.
  33. Paul, D.L. and Younghans, J.L., “Inlets & Inlet Engine Integration, ” Chapter13 of The Aerothermodynamics of Aircraft Gas Turbine Engines, AFAPL TR-78-52, Ed. Oates, G. C., 1978.
  34. Potonides, H., et al., “Design and Experimental Studies of a Type “A” V/STOL Inlet, ” AIAA Paper No. 78-956, 1978.
  35. Powel, A.G., Welge, H.R., and Trefny, C.J., “Low-Speed Aerodynamic Test of an Axisymmetric Supersonic Inlet with Variable Cowl Slot, ” AIAA Paper 85-1210, 1985.
  36. Pratt & Whitney Aircraft, The Aircraft Gas Turbine Engine and Its Operation, PWA Oper. Instr. 200, United Technologies Corporation, East Hartford, CT, June 1980.
  37. Rodean, H.C., “Supersonic Inlet Analysis and Design, ” Ling-Temco-Vought Report No. 11174, January 1958.
  38. Rolls-Royce, The Jet Engine, Rolls-Royce plc, Derby, England, 2005.
  39. Schweikhard, W.G. and Montoya, E.J., “Research Instrumentation Requirements for Flight/Wind Tunnel Tests of the YF-12 Propulsion System and Related Flight Experience, ” in Instrumentation for Airbreathing Propulsion, AIAA Series, Progress in Astronautics and Aeronautics, Vol. 34, Eds Fuhs, A.E. and Kingery, M., MIT Press, Cambridge, Mass. 1974.
  40. Seddon, J. and Goldsmith, E.L., Intake Aerodynan-mics, AIAA, Washington, DC, 1985.
  41. Sovran, G. and Klomp, E.D., “Experimentally Determined Optimum Geometries for Rectilinear Diffusers for Rectangular, Conical or Annular Cross Section, ” Fluid Mechanics of Internal Flow, Ed. Sovran, G., Elsevier Publishing Co., Amsterdam, 1963, pp. 270–319.
  42. Strack, W.C., “Propulsion Challenges and Opportunities for High-Speed Transport Aircraft, ” in Aeropropulsion ’87, NASA CP-10003, 1987.
  43. Summerfield, M., Foster, C.R., and Swan, W.C., “Flow Separation in Over-expanded Supersonic Exhaust Nozzles, ” Jet Propulsion, Vol. 24, September–October 1954, pp. 319–321.
  44. Sutton, G.P., Rocket Propulsion Elements, 6th edition, John Wiley & Sons, Inc., New York, 1992.
  45. Wazelt, F., Ed., “Suitable Averaging Techniques in Non-Uniform Internal Flows, ” AGARD Advisory Report 182, June 1983.
  46. Wazyniak, J., “Mission Adaptation & Technology Trends in Aero Propulsion, ” Short Course Notes in Aero-Propulsion at UTSI, 2000.
  47. Whitcomb, R.T. and Clark, L.R., An Airfoil Shape for Efficient Flight at Supercritical Mach Numbers, NASA TMX-1109, July 1965.
  48. Whitford, R., Design for Air Combat, Jane’s Publishing, London, UK, 1987.
  49. Younghans, J.L. and Dusa, D.J., “Inlets and Exhaust Systems for Multi-Mission Applications (Aero-Design and Installed Performance), ” Short Course Notes, The University of Kansas, 1993.
  50. Brausch, J.F., Bangalore, A.J., Barter, J.W., and Hoff, G.E., “Chevron exhaust nozzle for gas turbine engine, ” US Patent 6360528 B1, 2002.
  51. Bridges, J.E., Wernet, M., and Brown, C.A., “Control of jet noise through mixing enhancement”, Noise-Con, 23–25 June 2003, Cleveland, OH (NASA TM 2003- 212335).
  52. Lighthill, M.J., “On Sound Generated Aerodynamically, I: General Theory, ” Proceedings of the Royal Society, Series A, Vol. 211, 1952, pp. 564–587.
  53. Saiyed, N., Mikkelsen, K.L., and Bridges, J., “Acoustics and Thrust of Separate-Flow Exhaust Nozzles With Mixing Devices for High-Bypass-Ratio Engines, ” NASA TM 2000-209948, 2000.
  54. Zaman, K.B., Bridges, J.E., and Huff, D.L., “Evolution from ‘tabs’ to ‘chevron technology’ – a review, ” International Journal of Aeroacoustics, Vol. 10, No. 5 & 6, 2011, pp. 685–710.
images

Problems

  • 6.1 In a real inlet, the total pressure loss is 10% of the flight dynamic pressure, i.e., pt0pt2 = 0.1 q0 at a flight Mach number of M0 = 0.85. Calculate
    1. inlet total pressure recovery πd
    2. inlet adiabatic efficiency ηd.
  • 6.2 A subsonic inlet in cruise condition is shown. Calculate
    1. total pressure recovery πd
    2. area ratio A2/A0
    3. static pressure ratio p2/p0
    4. adiabatic efficiency ηd
    5. entropy rise Δs/R.
    images

    FIGURE P6.2

  • 6.3 Consider a subsonic inlet at a flight cruise Mach number of 0.8. The captured streamtube undergoes a prediffusion external to the inlet lip, with an area ratio A0/A1 = 0.92, as shown. Calculate
    1. Cp (i.e., the pressure coefficient) at the stagnation point
    2. inlet lip Mach number M1
    3. lip contraction ratio A1/Ath for a throat Mach number Mth = 0.75 (assume pt, th/pt1 = 1)
    4. the diffuser area ratio A2/Ath if M2 = 0.5 and pt2/pt, th = 0.98
    5. the nondimensional inlet additive drag Dadd/p0A1.
    images

    FIGURE P6.3

  • 6.4 The captured streamtube for a subsonic inlet experiences external diffusion, where flight Mach number of M0 = 0.85 decelerates to M1 of 0.65 at the inlet lip. Calculate the inlet additive drag nondimensionalized by flight static pressure p0 and inlet area A1.
  • 6.5 An aircraft flies at an altitude where p0 is 0.1915 atm and the flight total pressure pt0 is 1.498 atm. The engine face total pressure is measured to be, pt2 = 1.348 atm. For this inlet calculate
    1. τr, ram temperature ratio
    2. , nondimensional entropy rise in the diffuser
    3. ηd, inlet adiabatic efficiency
  • 6.6 A subsonic diffuser has an area ratio of A2/A1 = 1.3. The inlet Mach number to the diffuser is M1 = 0.72 and the total pressure loss in the diffuser is characterized by Δpt = 0.1 q1. Assuming the flow in the diffuser is adiabatic, calculate
    1. the diffuser exit Mach number M2 and
    2. diffuser static pressure recovery CPR
  • 6.7 A subsonic inlet is cruising at M0 = 0.85 and the capture area ratio A0/A1 is 0.90 (as shown). For an altitude pressure of p0 = 25 kPa, the temperature of T0 = –25°C, and inlet area A1 = 3 m2, calculate
    1. the inlet Mach number M1
    2. the inlet additive drag Dadd (N)
    3. inlet mass flow rate (kg/s)
    4. the inlet ram drag Dram (kN)
    5. engine face area A2, if M2 = 0.5 (assuming πd = 0.99)
    images

    FIGURE P6.7

  • 6.8 The Mach number at the compressor face is M2 = 0.65 at takeoff (M0 ≅ 0.2). Assuming the inlet suffers a 2% total pressure loss at takeoff, calculate the capture-to-engine face area ratio A0/A2.
    images

    FIGURE P6.8

  • 6.9 A subsonic aircraft cruises at M0 = 0.85 and its inlet operates with a capture ratio of A0/A1 = 0.7. First, calculatethe lip Mach number M1. Second, assuming an engine becomes inoperative and the inlet lip Mach number drops to 0.3 (the so-called engine wind-milling condition), calculate the additive drag Dadd for an inlet area of A1 = 4 m2 and the ambient static pressure of p0 = 16.6 kPa.
  • 6.10 A subsonic inlet is flying at Mach 0.8, with an inlet capture area ratio of A0/A1 = 0.7. The inlet lip contraction ratio A1/Ath is 1.15. Calculate the 1D throat Mach number and comment on the potential shock formation, near the convex surface, at the throat.
  • 6.11 A subsonic aircraft flies at M0 = 0.85 with an inlet mass flow ratio (MFR) of 0.90. Calculate the critical pressure coefficient Cp, crit on the nacelle. Also calculate the maximum cowl (frontal) area ratio AM/A1 if this inlet is to experience an average surface pressure coefficient corresponding to the critical value, i.e., .
  • 6.12 An inlet creates a circumferential distortion at the engine face, as shown. The hub-to-tip radius ratio is rh/rt = 0.5. The spoiled sector has a 15% mass flow deficit (per unit area) as compared with a uniform flow. Assuming the static density, temperature, and pressure are uniform at the engine face and the mass flow deficit in the spoiled sector is caused by a velocity deficit (as shown), use Bernoulli equation to estimate the total pressure deficit in the spoiled sector, i.e., [pt − p]spoiled/[ptp] uniform. Also calculate
    images

    FIGURE P6.12

  • 6.13 A subsonic inlet has a capture area ratio of A0/A1 = 0.8. Assuming the flight Mach number is 0.80 and the inlet area ratio is A2/A1 = 1.25, calculate
    1. Mach number at the inlet lip M1
    2. diffuser exit Mach number if pt2/pt1 is 0.95
    3. inlet static pressure ratios p1/p0 and p2/p1 for pt2/pt1 = 0.95
  • 6.14 A normal shock inlet operates in the subcritical mode, with the shock in standoff position, as shown. The bow shock is normal to the flow at the inlet centerline and weakens into an oblique shock and eventually a Mach wave away from the inlet.

    Apply conservations of mass and momentum to the control volume a-b-c-d-e-f-a, to approximate inlet cowl external drag force coefficient, in terms of the flight and inlet parameters that are shown, for example V0, M0, p0, p1, V1, A1, AM, and Aspillage.

    images

    FIGURE P6.14

  • 6.15 Consider a variable-geometry, convergent–divergent isentropic inlet that is designed for MD = 4.0. To swallow the starting shock, i.e., to start the inlet, the throat needs to be opened. Calculate the percentage of throat area opening needed to start this inlet, (ΔAth/Ath) × 100.
  • 6.16 An external compression inlet is in a Mach-2 flow. The shocks are positioned according to the figure shown. Calculate
    1. shock total pressure ratio πs
    2. A2/Ath

    You may neglect the frictional losses in the subsonic diffuser.

    images

    FIGURE P6.16

  • 6.17 A normal-shock inlet is operating in a supercritical mode, as shown. Flight Mach number is M0 = 1:6. The inlet capture area ratio A0 = A1 = 0:90 and the diffuser area ratio A2 = A1 = 1:2. Calculate
    1. M1
    2. inlet total pressure recovery πd, i.e., pt2/pt0
    images

    FIGURE P6.17

  • 6.18 An isentropic convergent–divergent supersonic inlet is designed for MD = 1.6. Calculate the inlet’s
    1. area contraction ratio A1/Ath
    2. subsonic Mach number where the throat first chokes
    3. percent spillage at M0 = 0.7
    4. percent spillage at M0 = 1.6 (in the unstarted mode)
    5. overspeed Mach number to start this inlet, Moverspeed
    6. throat Mach number after the inlet was started, with still Moverspeed as the flight Mach number
    images

    FIGURE P6.18

  • 6.19 Consider an isentropic fixed-geometry C–D inlet, which is designed for MD = 1.75. The inlet flies at an altitude where ambient (static) pressure is 20 kPa. Calculate
    1. Overspeed Mach number that will start this inlet
    2. The flight dynamic pressure corresponding to the altitude and Moverspeed
  • 6.20 An isentropic, convergent–divergent supersonic inlet is designed for MD = 3.0. Assuming that the throat area is adjustable, calculate the percentage of the design throat area that needs to be opened to swallow the starting shock, i.e.,
  • 6.21 A supersonic C–D inlet is designed for a flight Mach number of M0 = 3.5. This inlet starts by opening its throat (from Ath to Ath). Neglecting wall frictional losses, calculate
    1. percentage throat opening required
    2. throat Mach number after it starts (with the throat open at Ath
  • 6.22 A Kantrowitz–Donaldson inlet is designed for Mach 2.0. Calculate
    1. the required contraction area ratio A1/Ath
    2. the inlet total pressure recovery with the best backpressure
  • 6.23 A Kantrowitz–Donaldson inlet is designed for MD = 1.7. Calculate
    1. the inlet contraction ratio A1/Ath
    2. the throat Mach number after the inlet self started
    3. the total pressure recovery with the best backpressure.
    images

    FIGURE P6.23

  • 6.24 Calculate the contraction ratio A1/Ath and the maximum total pressure recovery of a self-starting C–D inlet designed for MD = 3.2.
  • 6.25 A normal-shock inlet is operating in a supercritical mode, with the shock inside the inlet. If the flight Mach number is M0 = 1.6 and the shock is located at As/At = 1.2, calculate
    1. Mach number ahead of the shock wave, Mx
    2. percentage total pressure gain if the inlet were to operate in the critical mode
  • 6.26 A variable geometry isentropic supersonic inlet is designed for MD = 1.6. Calculate
    1. percentage flow spillage at M0 = 0.8
    2. percentage flow spillage at MD = 1.6 before the inlet is started
    3. percentage throat area increase needed to start this inlet
    4. throat Mach number after the shock is swallowed, Mth
    5. inlet total pressure recovery with the best backpressure (with open throat)
  • 6.27 An isentropic, fixed-geometry inlet, is designed for MD = 1.5. If this inlet is to be started by overspeeding, calculate the necessary Mach number for overspeed.
  • 6.28 A fixed-geometry, convergent–divergent, internal-compression inlet is designed for MD = 2.0 and a self-starting capability. Calculate
    1. A1/Ath
    2. Mth
    3. inlet total pressure recovery for the “best” backpressure
  • 6.29 A normal-shock inlet is flying at a Mach number of 1.8. However, due to a nonoptimum backpressure, the normal shock is inside the duct where As/Ai = 1.15. Calculate the percentage loss in the total pressure recovery due to this back pressure.
  • 6.30 A variable-geometry supersonic convergent–divergent inlet is to be designed for an isentropic operation (in the started mode) at MD = 2.6. Calculate
    1. the inlet design contraction ratio A1/Ath
    2. the percentage opening of the throat (AthAth)/Ath needed to start the inlet
    3. the throat Mach number in the open position, Mth
  • 6.31 A variable-geometry, internal-compression, C–D inlet is designed for M = 2.0. Calculate percentage opening of the throat required to swallow the starting shock.
  • 6.32 A 2D fixed geometry convergent–divergent diffuser is shown. Mach number at the entrance is M1 = 3.0 and the static pressure is pt = 10 kPa. A shock occurs at an area ratio As/Ath = 1.25 downstream of the throat, as shown. Calculate
    1. Mx
    2. Py
    3. Pty
    4. A2/A1 if the exit Mach number is 0.5, i.e., M2 = 0.5

    You may assume the throat is choked and neglect wall friction.

  • 6.33 A normal-shock inlet operates in a Mach 1.86 stream with ambient static pressure of p0 = 30 kPa. Neglecting total pressure loss in the subsonic diffuser, calculate
    1. inlet total pressure recovery with the shock at the lip, i.e., the best backpressure
    2. inlet total pressure recovery when the shock is inside the duct at Ax/At = 1.2, i.e., the supercritical mode
    3. inlet total pressure recovery in subcritical mode with 10% spillage, i.e., Aslage/A1 = 0.1
    4. flight dynamic pressure q0
  • 6.34 Consider an external compression inlet with two ramps operating in a Mach-2.5 stream of air. Calculate the total pressure recovery of the inlet shock system for the case of the best backpressure for the two ramp angles of 8° and 12°, respectively, and compare it to the normal-shock inlet at the same Mach number and with the best backpressure.
    images

    FIGURE P6.34

  • 6.35 A supersonic flow is to be decelerated over two oblique shocks and one normal shock, similar to Problem 6.34 (i.e., M0 = 2.5). The Mach number downstream of the normal shock is 0.7. Assuming the maximum total pressure recovery is obtained when the first two oblique shocks are of equal strength, calculate the necessary ramp angles.
  • 6.36 A supersonic nozzle operates in an underexpanded mode with an area ratio A9/Ath = 2. We know that the exhaust plume turns outward by 15°, as shown. Assuming the flow inside the nozzle is isentropic and γ = 1.4, calculate
    1. exit Mach number M9 and
    2. the nozzle pressure ratio pt, noz/p0 (note that p10 = p0)
    3. the Mach number after the expansion waves M10
    4. the NPR if the nozzle was perfectly expanded
    5. the nozzle pressure ratio if a normal shock appears at the exit
    images

    FIGURE P6.36

  • 6.37 A convergent–divergent nozzle discharges to ambient air and has an exit-to-throat area ratio of Ae/Ath = 2.4. The total pressure at the entrance to the nozzle is pt = 100 kPa. A normal shock occurs at the nozzle exit, as shown. Calculate
    1. the ambient pressure p0 (note that p0 = py)
    2. the exhaust temperature Ty if the temperature at the throat is known Tth = 350°C
    3. mass flow rate through the nozzle if the throat area is Ath = 0.25 m2

    Assume the medium is air with γ = 1.4, cp = 1.004 kJ/ kg · K.

    images

    FIGURE P6.37

  • 6.38 A supersonic nozzle has an exit-to-throat area ratio of Ae/Ath = 3.5 and an upstream stagnation pressure of pt = 19 × 105 Pa, which remains constant with altitude. Calculate
    1. the altitude at which the nozzle is perfectly expanded
    2. the altitude at which the exhaust plume is turned 10° outward
    3. the altitude at which the exhaust flow is turned 4° inward

    Assume: γ = 1.4 and cp = 1, 004 J/kg · K.

    images

    FIGURE P6.38

  • 6.39 A convergent–divergent nozzle operates at a pressure altitude of pamb = 1 kPa. The nozzle total pressure and temperature are pt = 101 kPa and Tt = 2000 K, respectively. Calculate
    1. nozzle area expansion ratio A9/Ath, for perfect expansion
    2. nozzle exit Mach number M9 for perfect expansion
    3. nozzle exit velocity V9

    Assume isentropic flow of air (with γ = 1.4 and R = 287 J/kg · K).

  • 6.40 Consider an overexpanded nozzle as shown. Assuming the oblique shock at the exit makes a 40° angle with respect to the exit flow and the nozzle area ratio is 2.4, i.e., A9/Ath = 2.4, calculate
    1. M9
    2. jet turning angle θ
    3. NPR, i.e., pt/pamb
    images

    FIGURE P6.40

  • 6.41 A convergent–divergent nozzle has an area ratio A9/Ath = 6.79 and a stagnation pressure of pt = 38.13 atm. First, calculate the pressure altitude (i.e., p0) for which the nozzle is perfectly expanded. If this nozzle operated at a higher altitude, i.e., discharge to a lower backpressure atmosphere, it will operate as an underexpanded nozzle with an attendant expansion fan at the exit lip. Now, if the backpressure, i.e., p0 = 1/2p0–perfect–expansion, calculate
    1. Mach number after the (first) expansion fan, M10
    2. the jet turning angle θ
    3. the tail wave angle of the first expansion fan with respect to the local flow
    images

    FIGURE P6.41

  • 6.42 The exit flow from a two-dimensional C–D nozzle is shown to turn inward an angle of 15°.

    Calculate

    1. exit Mach number, M9
    2. wave angle at the nozzle exit lip
    3. nozzle total pressure pt (kPa)
    images

    FIGURE P6.42

  • 6.43 A convergent nozzle experiences πcn of 0.98, the gas ratio of specific heats γ = 1.30, and the gas constant is R = 291 J/kg · K. First, calculate the minimum nozzle pressure ratio that will choke the expanding nozzle, i.e., NPRcrit. This nozzle operates, however, at a higher NPR than the critical, namely, NPR = 4. 2 and with an inlet stagnation temperature of Tt7 = 939 K. Assuming this nozzle operates in p0 = 100 kPa ambient static pressure, calculate
    1. the exit static pressure and temperature p9 and T9, respectively
    2. the actual exit velocity V9 in m/s
    3. nozzle adiabatic efficiency ηn
    4. the ideal exit velocity V9s in m/s
    5. percentage gross thrust gain, had we used a convergent–divergent nozzle with perfect expansion
    6. nozzle discharge coefficient CD8
    7. draw a qualitative wave pattern in the exhaust plume
    images

    FIGURE P6.43

  • 6.44 A convergent–divergent nozzle has a conical exhaust shape with the half-cone angle of α = 25°. Calculate the divergence loss CA for this nozzle due to nonaxial exhaust flow. Assuming the same (half) divergence angle of 25°, but in a 2D rectangular nozzle, calculate the flow angularity loss and compare it to the conical case.
  • 6.45 The bypass ratio in a turbofan engine is α = 3.5. We intend to mix the cold fan and the hot core flows in a mixer to enhance the engine gross thrust. Assuming the hot core temperature of the gas is Th = 3 Tc, calculate
    1. the mixed-out temperature of the gas Tm/ Tc
    2. the percentage increase in ideal gross thrust as a result of mixing the cold and hot streams in the mixer

    You may assume a reversible adiabatic mixer flow with γh = γc = 1.4.

  • 6.46 Consider a fixed-geometry supersonic nozzle with the following inlet and geometrical parameters: Tt7 = 2500 K, pt7 = 233.3 kPa, , A9/A8 = 7.45, γ = 1.4, and cp = 1004 J/kg · K. Assuming isentropic flow in the nozzle, calculate
    1. ambient pressure p0, if this nozzle is perfectly expanded
    2. ambient pressure if a normal shock appears at the nozzle exit
    3. nozzle throat and exit areas A8 and A9, respectively, in m2
    4. the nozzle gross thrust with a NS at the exit (part b)
    5. the nozzle gross thrust in the severely overexpanded case of the N.S. being inside the nozzle at As/A8 = 4.23. Assume the flow remains attached after the shock
    6. in reality the boundary layer in part e separates and reduces the exit flow area by the amount of “blockage”.

      Assuming the effective exit area is A9e = As calculate the gross thrust for this more realistic scenario.

    7. compare the gross thrust with/without separation. Why did separation help?
    images

    FIGURE P6.46

  • 6.47 A convergent nozzle discharges into an open duct, as shown (i.e., the ejector concept). Through flow entrainment in the jet, air is drawn in from the upstream open end of the duct. The two flows mix in the duct and produce a mixed-out exit flow at the ambient static pressure p0. Neglecting wall friction in the duct flow, calculate:
    1. the jet exit area Aj (in cm2) and the ejector duct area A in cm2
    2. the airflow drawn in from the duct inlet, i.e., the secondary flow rate
    3. the exit mixed-out gas velocity V2
    4. the percentage increase in mass flow rate
    5. the percentage decrease in gas static temperature with the ejector
    6. gross thrust with and without the ejector

    Hint: Apply conservation of mass, momentum, and energy to the control volume between 1 and 2. Flow between 0 and 1 is assumed to be isentropic.

    images

    FIGURE P6.47

  • 6.48 The thermodynamic states of gas in an exhaust nozzle are shown on the Ts diagram.

    Calculate

    1. exit Mach number M9
    2. nozzle adiabatic efficiency ηn

    Assume γ = 1:30.

    images

    FIGURE P6.48

  • 6.49 A convergent–divergent nozzle has an area ratio of A9/Ath = 5.9. Assuming the flow is adiabatic and frictionless with γ = 1.40, calculate
    1. the nozzle pressure ratio pt/p0 for perfect expansion
    2. the nozzle pressure ratio pt/p0 if there is an oblique shock with β = 30° at the exit lip, as shown
    3. the nozzle pressure ratio pt/p0 if there is an expansion wave at the lip that turns the flow 15° outward, as shown.
    images

    FIGURE P6.49

  • 6.50 A scramjet flies at Mach 6 with an inlet total pressure recovery of 50%. Assuming the combustor experiences a total pressure loss of 42% (from its inlet condition), calculate the NPR, assuming γ = 1 .30 and is constant.
  • 6.51 Calculate the ratio of gross thrust produced by a convergent–divergent nozzle to the gross thrust produced by a convergent nozzle when they both operate with the same nozzle pressure ratio of NPR = 10 and γ = 1 .30.
  • 6.52 An external compression inlet operates in supercritical mode, as shown. The normal shock is at Ax/Ath = 1.2.

    Calculate:

    1. Throat Mach number, Mth
    2. Normal shock Mach number, Mx
    3. Total pressure ratio across the shock system

    (Assume losses are only due to shocks)

    images

    FIGURE P6.52

  • 6.53 A convergent–divergent exhaust system has an exit-to-throat area ratio, A9/A8 = 3.0.

    Assuming the flow in the nozzle is isentropic, calculate:

    1. nozzle exit Mach number, M9
    2. nozzle exit pressure, p9 (kPa)
    3. the ratio of momentum thrust to pressure thrust at the nozzle exit, for p0 = 5 kPa
    images

    FIGURE P6.53

  • 6.54 Consider two supersonic inlets, first is a self-starting (K–D) inlet, and the second is a NS inlet. Both inlets are designed for MD = 1.7. Calculate:
    1. the inlet contraction ratio, A1/Ath, for the self–starting (K–D) inlet
    2. throat Mach number in the K–D inlet after the inlet started
    3. the total pressure recovery of K–D inlet with best back pressure
    4. % gain in total pressure recovery in K–D versus NS inlet (both at best back pressure)
    images

    FIGURE P6.54

  • 6.55 A normal-shock inlet operates in air at Mach 1.5 in subcritical mode, as shown. The Mach Number at the cowl lip is M1 = 0.62.

    Calculate

    1. inlet spillage mass fraction,
    2. inlet total pressure recovery, πd

    (Neglect viscous losses in the subsonic diffuser)

    images

    FIGURE P6.55

  • 6.56 A subsonic diffuser has an area ratio, A2/A1 = 2.02 and operates in air with γc = 1.4 and R = 287 J/kg.K. Its inlet condition is measured to be: pt1 = 100 kPa, Tt1 = 288 K and M1 = 0.5.

    Assuming the flow in the diffuser is inviscid, calculate

    1. inlet dynamic pressure, q1, in kPa
    2. exit static pressure, p2 in kPa
    3. diffuser static pressure recovery, CPR
    images

    FIGURE P6.56

  • 6.57 There is a subsonic inlet in cruise condition (as shown). We know the flight condition is: M0 = 0.8, p0 = 25 kPa, T0 = 245 K.

    The inlet mass flow rate is and the inlet capture ratio is A0/A1 = 0.837.

    For γ = 1.4, and R = 287 J/kg.K, calculate

    1. impulse at station 0, I0, in kN
    2. impulse at the inlet lip, I1, in kN
    3. additive drag force in kN
    4. ram drag in kN
    images

    FIGURE P6.57

  • 6.58 A subsonic diffuser has an inlet Mach number of M1 = 0.4. The inlet pressure and temperature are p1 = 100 kPa and T1 = 15°C, respectively. The diffuser static pressure recovery is CPR = 0.75 and the exit Mach number is M2 = 0.1.

    For γ = 1.4, and R = 287 J/kg.K, calculate

    1. inlet dynamic pressure, q1, in kPa
    2. exit static pressure, p2, in kPa
    3. exit total pressure, pt2, in kPa
    4. diffuser area ratio A2/A1
    images

    FIGURE P6.58

  • 6.59 A subsonic inlet is in M0 = 0.86 cruise with external diffusion characterized by the inlet mass flow ratio parameter, MFR=0.89, which is the same as the inlet capture area ratio, A0/A1. Assume an isentropic flow with γ = 1.4 and cp = 1004 J/kg.K, to calculate
    1. the Mach number at the inlet lip, M1
    2. the external static pressure ratio, p1/p0, i.e., as the result of external diffusion
    3. the external velocity ratio, V1/V0, i.e., as the result of external diffusion
  • 6.60 A supersonic inlet creates two oblique shocks and a normal shock, as shown. The normal component of Mach number to these shocks is 1.3, for each of the three shocks.

    For γ = 1.4, calculate

    1. the total pressure recovery of the shock system
    2. Mach number at the entrance to the subsonic diffuser, M3
    images

    FIGURE P6.60

  • 6.61 An isentropic C–D inlet is designed for Mach 4.0. Calculate the percentage opening of the throat area that is needed to start the inlet. Assume γ = 1.4.
  • 6.62 A normal shock inlet operates in subcritical mode, as shown. Calculate the percentage spillage for this inlet if M1 = 0.60. Assume γ = 1.4 (Note: percentage spillage = 100 * (A1A0)/A1)
    images

    FIGURE P6.62

  • 6.63 A convergent nozzle operates at a nozzle pressure ratio of 3.0. Assuming γ = 1.3 and πcn = 0.98,

    Calculate

    1. exit Mach number, M8
    2. ratio of exit to ambient static pressure, p8/p0
    images

    FIGURE P6.63

  • 6.64 A convergent-divergent nozzle has an area ratio of A9/A8 = 4.2. The nozzle pressure ratio is NPR = 20 with γ = 1.3. Assuming that the flow in the CD nozzle is isentropic, calculate
    1. exit Mach number, M9
    2. ratio of exit static to inlet total pressure, p9/pt7
    3. what kind of wave will form at the nozzle exit? Why?
    4. what is the static pressure ratio across the wave?
    images

    FIGURE P6.64

  • 6.65 A convergent nozzle is shown. The ambient static pressure is p0 = 30 kPa. Assuming the flow in the nozzle is isentropic with γ = 1.3 and R ≈ 287 J/kg.K, calculate
    1. nozzle exit Mach number, M8
    2. exit static pressure, p8 in kPa
    3. exit velocity, V8, in m/s
    images

    FIGURE P6.65

  • 6.66 A subsonic inlet has an average throat Mach number of and an average total pressure of 75 kPa. The engine face area is 125% of the throat area, i.e., A2 = 1.25 Ath. For the engine face Mach number of M2 = 0.5, calculate
    1. engine face total pressure, pt2, in kPa
    2. static pressure at the throat, pth, in kPa
    3. dynamic pressure at the throat, qth, in kPa
    images

    FIGURE P6.66

  • 6.67 A convergent nozzle has an inlet total pressure, pt7 = 150 kPa and inlet total temperature of Tt7 = 660 K with γ7 = 1.33 and Cp7 = 1, 156 J/kg.K. The ambient pressure is p0 = 100 kPa. Assuming that the convergent nozzle total pressure ratio is πcn = 0.98, calculate
    1. nozzle exit Mach number, M8
    2. nozzle exit velocity, V8, in m/s
  • 6.68 A subsonic inlet is in flight, as shown. Calculate
    1. the Mach number at the inlet lip, M1
    2. the Mach number at the engine face, M2
    images

    FIGURE P6.68

  • 6.69 A convergent–divergent nozzle has a throat discharge coefficient of CD8 = 0.98. The velocity coefficient in the divergent section is CV = 0.96.

    Assuming the ratio of specific heats in the nozzle is γ = 1.30, and nozzle is of conical shape with the semi-vertex angle of α = 22.5°, calculate

    1. gross thrust coefficient, Cfg
    2. percentage increase in gross thrust if the nozzle were bell-shaped, i.e., exhaust velocity was axial

    Note that the nozzle is perfectly expanded.

    images

    FIGURE P6.69

  • 6.70 An external-compression supersonic inlet creates oblique and normal shocks in flight, as shown. Calculate the inlet total pressure recovery due to shocks.
    images

    FIGURE P6.70

  • 6.71 A convergent–divergent nozzle operates at NPR of 14.86 in an altitude where the ambient pressure is p0 = 30 kPa. Assuming the flow in the nozzle is isentropic with γ = 1.3, calculate
    1. nozzle area ratio, A9/A8, if the nozzle is perfectly expanded 5
    2. exit Mach number, M9
    images

    FIGURE P6.71

  • 6.72 An external-compression supersonic inlet has a single ramp and it operates in critical mode, as shown. The flight speed is V0 = 750 m/s where the ambient speed of sound is a0 = 300 m/s and the ambient (static) pressure is 20 kPa. Assuming that the air flow rate captured by the inlet is , calculate
    1. ram drag in kN
    2. inlet total pressure recovery (due to shocks)
    images

    FIGURE P6.72

  • 6.73 Consider a subsonic inlet in cruise condition as shown. Assuming the mass flow rate (of air) is 100 kg/s (as shown), calculate:
    1. the ram drag, Dr, in kN
    2. freestream capture area, A0 in m2
    3. the static pressure at the inlet lip, p1 in kPa
    4. the static temperature at the inlet lip, T1 in K
    5. the inlet capture area, A1 in m2
    6. the additive drag, Dad, in N
    images

    FIGURE P6.73

  • 6.74 An overexpanded supersonic convergent–divergent nozzle has an inlet total pressure of pt7 = 213 kPa and an area ratio of A9/A8 = 3.02. The ratio of specific heats is γ = 1.3 and cp = 1, 250 J/kg.K. Neglecting losses in the nozzle and assuming ambient static pressure is p0 = 25.2 kPa, calculate
    1. nozzle exit Mach number M9
    2. nozzle exit (static) pressure, p9
    3. the oblique shock wave angle, β in degrees
    images

    FIGURE P6.74

  • 6.75 A subsonic pitot inlet is cruising at M0 = 0.84 at an altitude where p0 = 30 kPa and T0 = 240 K. The inlet capture ratio is A0/A1 = 0.88. The inlet lip area contraction ratio is A1/Ath = 1.05. The area ratio between the throat and the engine face is Ath/A2 = 0.85. Assuming the flow is reversible and adiabatic inside the inlet, and γ = 1.4, calculate
    1. Mach number at the inlet lip, M1
    2. Mach number at the throat, Mth
    3. Mach number at the engine face, M2
    4. overall static pressure recovery coefficient, CPR between the engine face and flight, i.e.,
    images

    FIGURE P6.75

  • 6.76 A convergent–divergent nozzle operates with a normal shock at its exit plane, as shown. The nozzle total Pressure at its inlet is pt7 = 200 kPa and the nozzle area ratio is A9/A8 = 3.0. Assuming γ = 1.3, and The flow in the nozzle is isentropic, calculate
    1. the exit Mach number, M9, before the NS
    2. the ambient pressure, p0, in kPa
    images

    FIGURE P6.76

  • 6.77 A Kantrowitz–Donaldson inlet is designed to self-start at MD = 1.5. Assume the medium is air with γ = 1.4.

    Calculate the

    1. inlet contraction ratio, A1/Ath
    2. Mach number at the inlet throat, Mth at the design Mach number when the inlet is started
    3. best inlet total pressure recovery at MD = 1.5.
    images

    FIGURE P6.77

  • 6.78 A subsonic inlet has a throat area, Ath = 1 m2, with the average axial Mach number at the throat of Mth = 0.7. The corresponding flight condition is: M0 = 0.86, p0 = 30 kPa, T0 = –50°C, γ = 1.4, R = 287 J/kg.K. Calculate
    1. the captured stream area, A0, in m2
    2. the mass flow rate, , in kg/s

    You may treat the flow to the inlet throat to be isentropic.

  • 6.79 A supersonic inlet uses isentropic pre-compression for a total of 17° turn, as shown.

    Calculate

    1. Mach number upstream of the normal shock, M1
    2. inlet total pressure recovery, pt2/pt0
    3. static pressure on the cowl lip after the cowl shock, p3 in kPa (assume lip is very thin)
    images

    FIGURE P6.79

  • 6.80 Calculate nondimensional additive drag, Dadd/p0A1, for a subsonic inlet at cruise condition, M0 = 0.8, A0/A1 = 0.835.
  • 6.81 A subsonic inlet has a contraction area ratio of 1.112, i.e., AHL/Ath = 1.112. The (mass) average throat Mach number is 0.74. Neglecting the frictional losses (on total pressure) between the highlight and the throat, calculate the Mach number at the highlight.
  • 6.82 Calculate and compare the nondimensional additive drag, Dadd/p0A1, on a subsonic inlet with a cruise Mach number of M0 = 0.86 that is designed with either of the two inlet capture ratios:
    • Case I: A0/A1 = 0.82
    • Case II: A0/A1 = 0.76
    images

    FIGURE P6.82

  • 6.83 A 2D external-compression inlet is shown in supersonic flight. The external cowl is composed of three flat panels, as shown. The Cartesian coordinates of the panels are:
    • x1 = 0.0 m, y1 = 0.00 m
    • x2 = 1.0 m, y2 = 0.25 m
    • x3 = 2.0 m, y3 = 0.35 m
    • x4 = 3.0 m, y4 = 0.35 m

    Assuming the panels’ static pressure is:

    • Panel #1: p ≈ 50 kPa
    • Panel #2: p ≈ 30 kPa
    • Panel #3: p ≈ 25 kPa

    Calculate

    1. cowl wave drag (per unit width of the inlet) in kN/m
    2. can you estimate M0? (Hint: static pressure doubled on panel #1 from free stream, and the flow turning angle to the first panel is also known.)
    images

    FIGURE P6.83

  • 6.84 A three-ramp external compression inlet is shown. The three oblique shocks and the terminal, i.e., the normal shock, are all of equal strength. Assuming that normal component of Mach number to each shock is 1.3, calculate
    1. the total pressure recovery of the shock system
    2. the inlet total pressure recovery
    3. estimate the flight Mach number, M0, if this inlet represents “the optimum” four-shock external compression inlet.
    images

    FIGURE P6.84

  • 6.85 It is desirable to design a two-ramp external compression inlet for a design-point Mach number of M0 = 2.125 that creates three shocks of equal strength. We have calculated the normal components of Mach number in this design to be 1.3, as shown. Validate that the two successive ramp angles are 10.8° and 11.8° respectively. Also, calculate the (shock) total pressure recovery for this inlet at the design point.

    [Note that the figure is not-to-scale]

    images

    FIGURE P6.85

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
3.128.206.8