12.12 Thrust Chamber Cooling

In this section, we examine the cooling requirements and challenges in liquid and solid propellant rocket thrust chambers.

12.12.1 Liquid Propellant Thrust Chambers

Liquid propellant rockets usually operate for extended periods, that is, typically for several minutes, and thus expose the combustion chamber walls to intense heating. The combustion products reach (near stoichiometric level) temperatures of ∼3000–4700 K depending on the propellant combination (Table 12.1). All modes of heat transfer, that is, convection, conduction, and radiation exist in a rocket combustor and contribute to the combustor wall heating. The products of combustion then expand in a convergent–divergent nozzle to produce thrust. Although expansion of the combustion gases in the nozzle convert their thermal energy into kinetic energy, extended sections of the nozzle downstream of the throat are still exposed to excessive gas temperatures and thus need to be cooled. Fortunately, the liquid propellant onboard offers a cooling capacity that is tapped for the so-called regenerative cooling purposes of the thrust chamber walls. The cooling capacity is due to the phase change, that is, from liquid to vapor, known as latent heat of vaporization. Figure 12.31 shows a definition sketch of the regenerative cooling scheme in a liquid propellant rocket. Note that we switched the chamber designation of temperature, pressure, and so on from using the subscript “c” (for chamber) to “g” for (hot) gas. Since we are cooling the walls of the thrust chamber, we need the subscript “c” for the coolant.

images

inline FIGURE 12.31 Definition sketch of a regeneratively cooled rocket thrust chamber [subscript “g” stands for (hot) gas]

As we discussed heat transfer in Chapter 10for turbine cooling, the steady-state heat transfer from the hot combustion gases to the coolant takes place convectively over the combustor inner wall, across a hot gas film, radiatively due to the volume of hot combustion gases to the wall, conductively through the combustor wall, and convectively over a coolant film on the coolant channel side. The net one-dimensional heat flux to the thrust chamber wall is the sum of the convection as well as the radiative heat transfer according to

(12.67)numbered Display Equation

The radiative heat transfer from the volume of hot gas to the solid wall in the combustor is governed by

where inlineg is the emissivity of the gas (equal to 1 for blackbody radiation and less than 1 for graybody radiation), σ is the Stefan-Boltzmann constant, which is 5.67 × 10− 8 W/m2K4 and Tg is the absolute gas temperature. Although the Stefan−Boltzmann constant is very small (due to 10−8 term), the temperature contribution to radiation heat transfer grows exponentially with temperature, that is, T4g. Therefore, according to Sutton and Biblarz (2001), radiation accounts for 3–40% of the combustion chamber heat flux for the gas temperature range of 1900−3900 K. This is a reminder that we may not neglect radiative heat transfer unless the gas temperature is low, namely, ∼700 K. The presence of solid particles in the rocket combustion chamber increases the emissivity inlineg in Equation 12.68. The upper value of 1, that is, blackbody radiation, may be chosen for high concentration of solid particles in the gas stream. The convective heat transfer is governed by the Newton’s law of cooling according to

where hg is the gas-side film coefficient, Taw is the adiabatic wall temperature (on the gas side), and Twg is the gas-side wall temperature. Fourier’s law of heat conduction governs the heat conduction through the wall,

(12.70)numbered Display Equation

where kw is the thermal conductivity of the wall material (a property of the wall), Twc is the wall temperature on the coolant side, and tw is the wall thickness (a design parameter). Note that the minus in Fourier heat conduction law is absorbed in the parenthesis, that is, the heat transfer is from the combustor toward the coolant side. Now, combining these sources of heat flow in a steady-state (one-dimensional) rocket thrust chamber-cooling problem, we can write

(12.71)numbered Display Equation

We may eliminate the gas and coolant sidewall temperatures, Twg and Twc, from the heat flux equation to get

Let us examine the parameters in Equation 12.72. The stagnation temperature of the gas often approximates the adiabatic wall temperature Taw as we encountered in turbine cooling section. This approximation, that is, TawTtg, is appropriate in applying to the rocket thrust chamber cooling problem as well due to the approximations that are inherent in heat transfer solutions. Coolant temperature Tc is also known since the stored liquid temperature in the insulated tank is known and is adjusted for the transmission line connecting the pump and the thrust chamber heat transfer. Approximate expressions for film coefficients hg and hc are also known from the heat transfer literature related to the long tubes/channels. The emissivity of gas inlineg in the presence of (high concentration of) solid particles may be approximated as 1, that is, treating the particle-laden gas as a perfect blackbody radiator. Again this approximation, in the context of other heat transfer approximations, is acceptable in the preliminary design phase. The thickness of the combustor wall tw and wall thermal conductivity kw are design choices.

The film coefficient on the gas side hg is empirically estimated for heat transfer in tubes and is related to Reynolds number and Prandtl number according to

where Dg is the (local) diameter of the thrust chamber, kg is the gas thermal conductivity, ρgvg is the average gas mass flow rate per unit area in the combustion chamber, μg is the gas coefficient of viscosity, and cpg is the specific heat of the gas at constant pressure. The left-hand side of Equation 12.73 is the Nusselt number; the parentheses on the RHS are Reynolds and Prandtl numbers, respectively. Since the fluid density and properties μg, kg, and cpg are functions of gas temperature, the question of “which gas temperature” in the combustion chamber should we use to estimate these parameters arise. There is no unique answer to this question; however, for engineering approximations an “average film” temperature Tf is used as an acceptable reference temperature, that is,

(12.74)numbered Display Equation

The coolant side film coefficient (for the coolant in the liquid state) is related to the Reynolds and Prandtl numbers on the coolant side following

(12.75)numbered Display Equation

where is the coolant average specific heat and Dc is the coolant passage hydraulic diameter. We may express the film coefficient on the coolant side in terms of on coolant mass flux, ρcvc, and specific heat:

This expression clearly shows that for a given coolant Reynolds number and Prandtl number, the coolant film coefficient hc increases with coolant mass flux and the coolant specific heat. We may thus tailor the cross section of the coolant passage in critical areas such as the nozzle throat to increase the local mass flux, ρcvc. Also, we may choose a liquid fuel with high specific heat. For example, hydrogen has a very high specific heat compared with other rocket fuels and is thus a coolant of choice for regeneratively cooling rocket thrust chambers. Table 12.4 from Sutton and Biblarz (2001) show some physical properties of some liquid propellants.

inline TABLE 12.4 Physical Properties of Some Liquid Propellants (from Sutton and Biblarz, 2001)

Propellant Liquid fluorine Hydrazine Liquid hydrogen Methane Monomethyl-hydrazine Nitric acida (99% pure) Nitrogen tetroxide Liquid oxygen Rocket fuel RP-1 Unsymmetrical dimethyl-hydrazine (UDMH) Water
Chemical formula F2 N2H4 H2 CH4 CH3NHNH2 HNO3 N2O4 O2 Hydrocarbon CH1.97 (CH3)2NNH2 H2O
Molecular mass 38.0 32.05 2.016 16.03 46.072 63.016 92.016 32.00 ∼175 60.10 18.02
Melting or freezing point (K) 53.54 274.69 14.0 90.5 220.7 231.6 261.95 54.4 225 216 273.15
Boiling point (K) 85.02 386.66 20.4 111.6 360.6 355.7 294.3 90.0 460–540 336 373.15
Heat of vaporization (KJ/kg) 1.66.26b 44.7b 446 510b 875 480 413b 213 246b 542 (298 K) 2253b
Specific heat 0.368 0.736 1.75b 0.835b 0.698 0.042 0.374 0.4 0.45 0.672 1.008
(kcal/ kg-K) (85 K) (293 K) (20.4 K) (293 K) (311 K) (290 K) (65 K) (298 K) (298 K) (273.15 K)
0.357 0.758 0.735 0.163 0.447 0.71
(69.3 K) (338 K) (393 K) (373 K) (360 K) (340 K)
Specific gravityc 1.636 1.005 0.071 0.424 0.8788 1.549 1.447 1.14 0.58 0.856 1.002
(66 K) (293 K) (20.4 K) (111.5 K) (293 K) (273.15 K) (293 K) (90.4 K) (422 K) (228 K) (373.15 K)
1.440 0.952 0.076 0.857 1.476 1.38 1.23 0.807 0.784 1.00
(93 K) (350 K) (14 K) (311 K) (313.15 K) (322 K) (77.6 K) (289 K) (244 K) (293.4 K)
Viscosity 0.305 0.97 0.024 0.12 0.855 1.45 0.47 0.87 0.75 4.4 0.284
(centipoise) (77.6 K) (298 K) (14.3 K) (111.6 K) (293 K) (273 K) (293 K) (53.7 K) (289 K) (220 K) (373.15 K)
0.397 0.913 0.013 0.22 0.40 0.33 0.19 0.21 0.48 1.000
(70 K) (330 K) (20.4 K) (90.5 K) (344 K) (315 K) (90.4 K) (366 K) (300 K) (277 K)
Vapor pressure 0.0087 0.0014 0.2026 0.033 0.0073 0.0027 0.01014 0.0052 0.002 0.0384 0.00689
(MPa) (100 K) (293 K) (23 K) (100 K) (300 K) (273.15 K) (293 K) (88.7 K) (344 K) (289 K) (312 K)
0.00012 0.016 0.87 0.101 0.638 0.605 0.2013 0.023 0.1093 0.03447
(66.5 K) (340 K) (30 K) (117 K) (428 K) (343 K) (328 K) (422 K) (339 K) (345 K)

aRed fuming nitric acid (RFNA) has 5–20% dissolved NO2 with an average molecular weight of about 60, and a density and vapor pressure somewhat higher than those of pure nitric acid.

bAt boiling point.

cReference for specific gravity ratio: 103 kg/m3 or 62.42 lbm/ft3.

In general, we have to integrate the steady-state heat transfer Equation 12.71 along the thrust chamber (axis). We may divide the thrust chamber into cylindrical sections of small axial length Δx where all the heat transfer parameters are assumed constant. By further assuming axisymmetric condition per cylindrical sections, we can march along the axis from the combustion chamber out toward the nozzle exit. The overall heat transfer to an axisymmetric thrust chamber is the integral of the heat flux along the wall, which is thus the sum of the segments according to

(12.76)numbered Display Equation

The overall heat transfer causes the temperature of the coolant to rise according to

(12.77)numbered Display Equation

where T1 represents the entrance temperature of the coolant into the jacket, and T2 is the bulk coolant temperature as it (leaves the cooling jacket and) enters the injector. As a design choice, we may limit the bulk coolant exit temperature to below the boiling point, that is, T2 < Tboiling-point. Table 12.5 lists the heat transfer characteristics of some liquid propellants (from Sutton and Biblarz, 2001). The pressure in the cooling jacket has a strong influence on the boiling point temperature of the fuel. For example, hydrazine boils at 387 K for a pressure of 0.101 MPa (i.e., 1 atm), whereas it boils at 588 K (an increase of 201 K) if the pressure is raised to 6.89 MPa. The cooling jacket pressure is to be provided by a turbopump (or a pressure) feed system and the higher pressure ratio will then require multistage pumps and very heavy turbopump feed system.

inline TABLE 12.5 Heat Transfer Characteristics of Some Liquid Propellants

Boiling characterics Nucleate boiling characteristics
Liquid coolant Pressure (MPa) Boiling Temp. (K) Critical Temp. (K) Critical Pressure (MPa) Temp. (K) Pressure (MPa) Velocity (m/sec) qmax (MW/m2)
Hydrazine 0.101 387 652 14.7 322.2 4.13  10 22.1
0.689 455 20 29.4
3.45  540 405.6 4.13  10 14.2
6.89  588 20 21.2
Kerosene 0.101 490 678   2.0 297.2 0.689   1   2.4
0.689 603     8.5   6.4
1.38  651 297.2 1.38  1   2.3
1.38  651     8.5   6.2
Nitrogen tetroxide 0.101 294 431 10.1 288.9 4.13  20 12.4
0.689 342 322.2   9.3
4.13  394 366.7   6.2
Unsymmetrical 0.101 336 522 6.06 300  2.07 10   4.9
    dimethyl 1.01  400 20   7.2
    hydrazine 3.45  489 300  5.52 10   4.7

Source: Sutton and Biblarz 2001. Reproduced with permission from Wiley.

Also note that the critical heat flux, that is, the maximum heat transfer rate per unit area, in the thrust chamber theoretically occurs at the throat where the flow passage area is at a minimum. Experimental evidence of the peak heat flux near the throat is well documented in the literature, for example, Witte and Harper (1962). To lower the wall temperature on the gas side, we have to increase the coolant side film coefficient hc, which according to Equation 12.76 behaves as

(12.78)numbered Display Equation

The design requirement on Twg at the throat dictates the following characteristics on the coolant passage side:

  • High mass flux ρcvc or since the liquid density is nearly constant, high coolant speed vc
  • Small diameter Dc cooling tubes.

The coolant speed at the throat may be as high as ∼15−20 m/s. The hydraulic diameter of the cooling tubes near the throat scales in millimeters. The most effective throat cooling in regeneratively cooled thrust chambers combines these two effects. Therefore, the cooling jacket is composed of small-diameter tubes with a tapered cross-sectional area at the throat with relatively high coolant speed.

In very small liquid propellant thrust chambers, where either insufficient coolant flow is available or insufficient (throat) surface area is available for cooling, ablative materials are used. Also, in cases where the motor is pulsed as in vernier rockets, passive cooling technique, such as ablative cooling, is used. An ablative material directly transitions from the solid-to-gaseous phase while absorbing latent heat of vaporization in the process. Additionally, the emergence (and the flow) of the (ablative) gas on the surface acts as a cooling layer that protects the surface.

12.12.2 Cooling of Solid Propellant Thrust Chambers

Solid propellant rockets may use several (nonregenerative) methods to protect the thrust chamber walls (or motorcase) against the heat loads in the combustor. The purpose of an inhibitor is to inhibit grain burning where the designer wishes to protect the thrust chamber walls (or as a means of controlling thrust-time or vehicle dynamic behavior). The purpose of an insulating layer, that is, the one with low thermal conductivity, next to the wall is also to protect the wall from excessive heating. In addition to these two techniques, solid propellant grain designer may place a cooler burning grain in the outer shell next to the wall to protect the motorcase. However, this is an expensive manufacturing proposition seldom used. Since ablative materials have been extensively used for thermal protection with success during the Apollo era, and graphite has emerged as a viable high temperature material, designers use ablative materials and graphite as throat inserts (or for motorcase insulation) in solid propellant rockets today with confidence. The use of graphite offers the added advantage of erosion resistance (as compared with ablative materials that decompose), thus maintaining the throat (cross-sectional flow) area nearly constant as a function of burn time. Detailed engineering design considerations in solid rocket motors, including practical thermal management approaches are presented by Heister (1995), which is suggested for further reading.

12.13 Combustor Volume and Shape

Chemical rocket combustion chambers are cylindrical in shape, characterized by a length L1, diameter D1, an area contraction ratio A1/Ath, (toward the nozzle throat), and a contraction length Lc (or a contraction angle θc). Figure 12.32 shows a definition sketch of a combustor volume. The combustor volume Vc includes the convergent section to the throat.

images

inline FIGURE 12.32 Definition sketch of a rocket combustion chamber geometry

Chamber diameter D1 is primarily governed by the vehicle size or thrust magnitude. In the flight performance section, we desired a low cross-sectional area for the vehicle (through m0/A), that is, our goal is to minimize D1, consistent with thrust production. The area contraction A1/Ath sets the gas speed in the chamber. Remember that the throat is sonic and area ratio (A/A*) establishes the upstream Mach number M1. The desire is toward large area ratio, yet limited by minimum A1. The contraction angle θc provides for a smooth flow at the throat with high discharge coefficient. The desire is to have a large angle consistent with high efficiency, which translates to about 30° or less (typically in the range of 20–30°). The chamber length L1 is the minimum length needed for a complete combustion in liquid propellant rockets. Therefore, chamber length is directly proportional to the product of mean speed of the propellant in the combustor and the propellant residence time, that is,

(12.79)numbered Display Equation

Residence time depends on the evaporation, mixing, and reaction timescales, which are a strong function of the propellant combination. Residence time is also a function of the injector plate design through the atomization scale. Therefore, chamber length is affected by the injector design and the propellant choice. Residence or stay time range between 1 and 40 ms for different propellants. It is customary to define a characteristic length scale L* based on the combustor (total) volume that includes the convergent section to the throat area, namely,

(12.80)numbered Display Equation

Huzel and Huang (1992) present historical numbers for L* for different propellants. As these numbers are not based on fundamental physics, they should be used for comparative purposes only. Table 12.6 shows the historical data from Huzel and Huang (1992).

inline TABLE 12.6 Combustor Characteristic Length L*

Propellants L*(m), low L*(m), high
Liquid fluorine/hydrazine 0.61 0.71
Liquid fluorine/gaseous H2 0.56 0.66
Liquid fluorine/liquid H2 0.64 0.76
Nitric acid/hydrazine 0.76 0.89
N2O4/hydrazine 0.60 0.89
Liquid O2/ammonia 0.76 1.02
Liquid O2/gaseous H2 0.56 0.71
Liquid O2/liquid H2 0.76 1.02
Liquid O2/RP-1 1.02 1.27
H2O2/RP-1 (including catalyst) 1.52 1.78

Source: Huzel and Huang 1992.

12.14 Rocket Nozzles

The aerodynamic principles of exhaust nozzles and their design considerations are presented in Chapter 6. The range of operation of the rocket nozzles, however, is vastly different than airbreathing engines. A rocket may operate in ambient conditions ranging from sea level to vacuum in space. Therefore, the nozzle pressure ratio (NPR) in a rocket may range from ∼50 at sea level to infinity (assuming ambient pressure at orbit is zero) in space. Consequently, chemical rockets with such large pressure ratios require very large expansion area ratios, that is, A2/Ath inline 1, in order to maximize their thrust production. Area expansion ratios of 100 or more are designed for rocket nozzles for space applications. It is a challenge to design a variable-area rocket nozzle due to mechanical complexity and weight penalty. A successful example is shown in Figure 12.33 where the RL-10B-2 rocket engine uses a two-piece extendable nozzle skirt.

images

inline FIGURE 12.33 The RL-10B-2 rocket engine with extendable nozzle skirt. Source: Reproduced by permission of United Technologies Corporation, Pratt & Whitney

Possible rocket nozzle configurations are shown in Table 12.7 (from Sutton and Biblarz, 2001). The nozzles are divided into two categories, (1) the ones with and (2) the ones without a centerbody. The second category of exhaust nozzles, without a centerbody, is the conventional conical, bell-shaped, and partial-bell configurations. All nozzles with an exit flow component that is not purely in the axial direction suffer the so-called (momentum) angularity loss. We have addressed and quantified this loss in Chapter 6. For example, the conical nozzle that offers a simple geometry for manufacturing; it introduces an angularity loss factor to the momentum thrust that is given by Equation 6.60 and is repeated here for convenience

numbered Display Equation

where α is the half-cone angle. The two-dimensional nozzles with exit flow angle α, suffered the following momentum angularity loss, as derived in Chapter 6.

numbered Display Equation

inline TABLE 12.7 Rocket Nozzle Configurations and their Altitude Performance

Source: Sutton and Biblarz 2001. Reproduced with permission from Wiley.

A conical nozzle with 15° half-cone angle creates an angularity momentum loss of only 1.7%. However, the low-altitude, that is, overexpanded, operation of conical nozzles involves shock formations (inside the nozzle) that cause flow separation, as shown in Table 12.7. The bell-shaped nozzles, either full bell or partial bell, are contoured and are more complex to manufacture but, by virtue of smaller exit angularity, the nozzle momentum thrust efficiency can be better than a conical nozzle of the same area ratio, or length. We note that the sea-level performance of bell-shaped nozzles (as shown in Table 12.7) tends to create a full exit flow with oblique shocks hanging at the lip. The category of nozzles that use a centerbody, as in a plug (or aerospike) nozzle or the expansion-deflection nozzle, tends to produce better off-design performance, both at the low and high altitudes, than the ones without centerbody. We may note the flow distribution at the nozzle exit, in Table 12.7, shows a fully attached flow, that is, without any shock-induced separations. The off-design performance of aerospike and other nozzles with a centerbody is thus superior to conical and bell-shaped nozzles, but the system penalty is paid by the centerbody needs to be cooled. More details of the flow pattern for a linear (truncated) aerospike are shown in Figure 12.34. This concept is used in the XRS-2200 aerospike linear rocket engine that uses 20 individual thrust cells or modules and two concave regeneratively (fuel) cooled (external) expansion ramps. The truncated base is porous where the gas generator flow (from the turbo pump system) fills the base and provides for a continuous and smooth expansion. The pressure distribution on the ramp shows the presence of (periodic) compression and expansion waves. The local compression (or shock) waves are followed by the reflected expansion waves from the jet boundary (or shear layer) and thus do not cause boundary layer separation. However, the presence of the compression waves on the ramp will cause an increase in the wall heat transfer; thus a more conservative fuel-cooling approach needs to be employed. The two-dimensional aspect of the linear aerospike offers better integration potential with a winged aircraft than the axisymmetric configuration. In addition, the action of (20) individual thrusters may be used for stability and control purposes of the aircraft.

images

inline FIGURE 12.34 Flow pattern and pressure distribution in a linear aerospike nozzle at low and high altitudes. Source: Sutton and Biblarz 2001. Reproduced with permission from Wiley

The radius of curvature upstream and downstream of the nozzle throat, the initial and exit angles of the bell nozzles, and the angularity loss parameter of conical and bell-shaped nozzles are presented and compared in Figure 12.35 (adapted from Huzel and Huang, 1992). These curves that relate the geometry and performance of various nozzles may be used in preliminary design trade studies. The nozzle length comparisons in Figure 12.35 are suitable for weight estimation studies.

images

inline FIGURE 12.35 Conical, bell and partial bell nozzle design, and performance comparison. Source: Adapted from Huzel and Huang 1992

12.14.1 Multiphase Flow in Rocket Nozzles

In our analysis so far, we have treated the nozzle fluid to be in the gaseous state. There are at least three sources for liquid or solid particle presence in a rocket nozzle flow. The first is the propellant itself, where solid particles are intentionally embedded or dispersed to increase the rocket performance through higher specific impulse. The second source is the combustion, which may lead to soot and particulate formation in the products of combustion. The third source is the appearance of liquid droplets, that is, condensation, in the exhaust stream due to static temperature drop in the gases. The very large area expansion ratios bring about very large temperature and pressure expansion ratios, which result in the condensation of a fraction of the exhaust stream. It is the impact of these solid or liquid fractions (present in the nozzle gases) on rocket performance that we address in this section.

The thermal and dynamic interaction of the two phases in a fluid, say solid and gas, pose complex fluid mechanics and heat transfer problems. However, in four limiting cases, we may arrive at closed-form solutions and shed light on the effects of multiphase flow on rocket performance.

These four limiting cases are:

  1. Solid particles reach the same temperature as the gas at the nozzle exit, the so-called thermal equilibrium case
  2. Solid particles are fully accelerated and reach the same speed as the gas at the nozzle exit, that is, they reach momentum equilibrium
  3. Solid particles maintain their temperatures, that is, the case of zero thermal interaction or heat transfer between the two phases
  4. Solid particles are not accelerated by the gas and thus maintain their negligible momentum at the nozzle exit.

Note that an underlying assumption in these four limiting cases is that the solid fraction present at the nozzle inlet remains unchanged, that is, frozen, in the nozzle and appear in whole at the nozzle exit. The four cases noted above address only the fate of these particles in heat and momentum transfer with the gas. To derive the fundamental equations of two-phase flow, we return to a slab concept where we apply the (one-dimensional) conservation principles to a slab of fluid. The fluid and flow parameters on the two sides of the slab are related to each other via a linear relationship using the Taylor series expansion of a continuous function. Figure 12.36 shows the definition sketch of a slab in a two-phase fluid flow with zero friction and heat transfer (i.e., reversible and adiabatic condition) at the wall.

images

inline FIGURE 12.36 Definition sketch of a slab in a two-phase flow in a nozzle

By stipulating a constant mass flow rate for the gas and the solid phases in the fluid, we get

Note that the density of the solid is the mass of the solid contained in a unit volume of the gas–solid mixture and is not to be confused by the density of the solid itself. The gas density is the mass of the gas per unit volume of the mixture, as expected. By conserving the mass flow rate of solid and gas individually, through Equations 12.81 and 12.82, we conserve the collective mass flow rate in the nozzle. The conservation of momentum applied to the slab gives

Note that the last term on the RHS of the Equation 12.83 is the pressure force exerted by the sidewall on the fluid (in the x-direction). Also the “momentum” in quotation mark is used as the shorthand for “time rate of change of momentum.” Momentum equation simplifies to

We may divide both sides of Equation 12.84 by the flow area A to get

(12.85)numbered Display Equation

The steady energy equation written for a (calorically) perfect gas and solid phases in a flow is the balance between the net flux of power (in fluid and solid phases) and the rate of the external energy exchange through heat transfer and mechanical shaft power, that is,

(12.86)numbered Display Equation

The energy equation simplifies to

(12.87)numbered Display Equation

The fraction of solid particles in the mixture of two-phase flow may be given a symbol X defined as

(12.88)numbered Display Equation

In terms of this parameter, that is, solid mass fraction X that constitutes a constant in our problem based on our frozen-phase assumption, we may write the energy equation as

We replace the solid particle density−velocity term ρsVs by

numbered Display Equation

in the momentum equation to get

If we divide Equation 12.89 by (1 − X) and substitute Equation 12.90 for ρgdVg, we get

Case 1

The solid and gas are in thermal equilibrium, that is, dTs = dTg = dT, therefore, Equation 12.91 reduces to

We may consider two scenarios for the solid particles momentum behavior in the nozzle flow. First, if the solid particles are accelerated by the gas to achieve the same speed as the gas, we get Vs = Vg, therefore, Equation 12.92 simplifies to

(12.93a)numbered Display Equation

This equation readily integrates for a perfect gas where ρg = p/RgT to the following form:

Also Equation 12.89 for Vs = Vg = V and Ts = Tg = T gives

(12.94)numbered Display Equation

which may be integrated from the chamber condition (with Vc ≈ 0) to the nozzle exit, V2

(12.95)numbered Display Equation

In terms of pressure ratio, nozzle exit velocity is

Since solid particles have attained the same speed as the gas, the ratio of specific impulse with and without solid particulate flow is the ratio of the two exhaust speeds, namely,

For a gas molecular weight of 20.1 kg/kmol, the ratio of specific heats γ = 1.26, and specific heat of the solid (aluminum) particles = 903 J/kg ·K, we calculate and graph Equation 12.97 in Figure 12.37.

images

inline FIGURE 12.37 The effect of solid particles on rocket performance (assuming Ts = Tg, Vs = Vg and constant Tc)

Figure 12.37 shows a drop in rocket performance with increasing fraction of the solid flow. The impact of cycle pressure ratio is small, as expected. The presence of solid particles in the propellant is, however, to enhance the burning rate and combustion temperature Tc, which was assumed constant in this calculation. The effect of higher burning rate and combustion temperature due to solids should, therefore, outweigh the penalty of the two-phase flow in the nozzle, as depicted in Figure 12.37.

The second scenario for the particle momentum/acceleration is to assume the solid particles are not accelerated by the gas, that is, dVs = 0 while their temperature has reached that of the gas. The temperature–pressure relation remains the same as Equation 12.93b. However, the energy equation gives the variation of the gas speed with temperature drop of the mixture, that is,

(12.98a)numbered Display Equation

The solution of the above equation is

(12.98b)numbered Display Equation

The thrust of the two streams of gas and solid, one moving at speed Vg2 and the second moving at speed Vs, is

(12.99)numbered Display Equation

Therefore, the stream-thrust-averaged speed at the nozzle exit (for a perfectly expanded nozzle) is

Since the second term on the RHS of Equation 12.100a is the product of two small numbers (remember that the solid was assumed to be unaccelerated and X is the small solid fraction), therefore, the thrust-averaged exit velocity is

(12.100b)numbered Display Equation

and the specific thrust is expressed in terms of the thrust-averaged exit velocity as

(12.101)numbered Display Equation

The ratio of specific impulse with/without solid particles that are in thermal equilibrium with the gas, but are not accelerated by the gas, is

The graph of Equation 12.102 superimposed on the rocket performance chart of Figure 12.37 is shown in Figure 12.38. Note that the specific impulse is further degraded when the solid particles are not accelerated by the gas (i.e., the case of small particle drag).

images

inline FIGURE 12.38 Variation of specific impulse with solid flow fraction in the gaseous exhaust nozzle (assuming Ts = Tg, and constant Tc)

Case 2

The solid and gas are not in thermal equilibrium, that is, when the solid temperature remains constant, dTs = 0 as a result of negligible heat transfer with the gas. In this case, Equation 12.91 reduces to

In addition to Ts = constant, either for full acceleration where Vs = Vg or when the solid particle remains unaccelerated, that is, when dVs = 0, Equation 12.103 gives

(12.104)numbered Display Equation

The energy Equation 12.89 reduces to the following form for Vs = Vg, that is, full acceleration case,

(12.105)numbered Display Equation

Therefore, the specific impulse ratio is

The energy Equation 12.89 for the case of constant Vs, that is, unaccelerated solid particles in the two-phase flow reduces to

(12.107)numbered Display Equation

which has a solution following

(12.108)numbered Display Equation

Although the gas phase seems to be unaffected by the presence of the solid particles, the thrust is affected since the two phases of the flow have different momentums at the nozzle exit. The thrust averaged exit velocity is

(12.109)numbered Display Equation

which again for negligible solid velocity Vs in the exhaust plane, the ratio of the specific impulse will be

Now, let us graph Equations 12.106 and 12.110 in Figure 12.39. The specific impulse diminishes with solid flow fraction, as in the previous cases. The nozzle (or cycle) pressure ratio has no effect on specific impulse ratio if the solid particles are not heated to the gas temperature, that is, when they are not in thermal equilibrium with the gas. The unaccelerated solid particles cause a larger reduction in specific impulse than the fully accelerated ones. The reason is that the fully accelerated solid particles contribute to momentum, whereas the unaccelerated particles with negligible exit momentum cause a reduction in thrust. Therefore, the solid particle drag causes the particles to accelerate and contribute to thrust. The case studies in this section are useful only in establishing a range for the impact that multiphase flow makes on rocket performance. In reality, multiphase flow is more complex and accurate simulation needs computational approach.

images

inline FIGURE 12.39 Rocket performance with solid particles in the exhaust stream (assumed Ts = constant)

12.14.2 Flow Expansion in Rocket Nozzles

The products of combustion do not in reality stop reacting once they enter the nozzle. Actually, the conversion of thermal to kinetic energy in the nozzle causes a cooling down effect of the gases as well as the pressure drop, which impact the reaction rates of some of the products of combustion. Therefore, chemical reaction takes place continually from the combustor through the nozzle. Sometimes the nozzle reverses some of the reactions that took place in the combustor. For example, dissociation in the combustor, which is a direct result of high combustion temperature, may reverse, through the recombination process, to some extent in the nozzle when the temperature of the mixture drops. In addition, chemical reactions between the constituents are often considered to be in equilibrium. However, that too is an assumption that leads to major simplifications in the analysis and may be questioned. Whether chemical equilibrium is reached in a mixture depends on the chemical kinetics and the reaction rates in the mixture. In a rocket expansion process where the timescales are very short (in milliseconds) due to high (convection) speeds, chemical equilibrium may be questioned legitimately in such environments. Computational approaches to complex chemical reactions are fortunately available, for example, Gordon and McBride (1996), which may be used for an accurate analysis of chemically reacting mixtures of gases in different thermal environments.

However, the choice of frozen (concentration of) species of gas mixture in the nozzle (known as the frozen chemistry model) is often made for preliminary rocket design calculations. As evidenced from the computational results of Olson (1962), shown in Figure 12.40 (or Figure 12.22 from Sutton and Biblarz, 2001), the corrected specific impulses for hydrogen–oxygen and JP4–oxygen are both indicated to be higher with chemical reactions in the nozzle. Also, the corrected experimental results in Olson show that the equilibrium model gives an overestimation of the actual performance. This important result points to the frozen chemistry model as a conservative estimation of the rocket performance. A conservative estimation is precisely what is needed in a preliminary design stage.

images

inline FIGURE 12.40 Comparison of specific impulse between experimental and computational models. Source: Olson 1962. Reproduced with permission from AIAA

12.14.3 Thrust Vectoring Nozzles

Conventional gimbals with hydraulic or electromechanical actuators may be used to swivel the entire rocket thrust chamber. The range of control authority using this approach is ∼ ±7°. Injection of a fluid, that is, either a propellant or an inert gas, in the nozzle, known as the secondary injection, causes an oblique shock formation and thus secondary injection thrust vector control (SITVC). The range of control authority using secondary injection is ∼ ±5°. The use of rocket clusters in the vehicle design opens the opportunity to differential operation of the opposite rockets in the cluster to achieve control. The use of jet deflectors, as in jet tabs or jet vanes, offers a limited-duration thrust vector control (in short burn-time rockets) since they are exposed to the hot exhaust gases. The control authority offered in this scheme is ∼ ±10°. Humble et al. (1995) present a practical engineering approach to rocket thrust vector control and is recommended for further reading.

A modern approach to (an old) nozzle design has produced linear aerospike configuration that uses many (e.g., 20) rocket engine cells.Figure 12.41 and 12.42 show aerospike nozzle thrust coefficient variation with nozzle pressure ratio; RS-2200 is a linear aerospike rocket engine on a test stand (from Stennis Space Center; courtesy of NASA). The linear aerospike nozzles have superior off-design performance over a bell nozzle as well as having a vector control capability.

images

inline FIGURE 12.41 Nozzle thrust coefficient for aerospike. Source: Huzel and Huang 1992. Reproduced with permission from AIAA

images

inline FIGURE 12.42 RS-2200 Linear aerospike rocket engine test. Source: Courtesy of NASA

12.15 High-Speed Airbreathing Engines

Ramjets offer the highest (fuel) specific impulse at flight Mach numbers above ∼4, as depicted in Figure 12.1. However, they produce no static thrust. The analysis of conventional, that is, subsonic combustion, ramjets is the same as other airbreathing engines such as turbojets. A typical flow path in a conventional ramjet engine is shown in Figure 12.43.

images

inline FIGURE 12.43 Flow parameter trends in a typical conventional ramjet

The Ts diagram of the ramjet cycle is shown in Figure 12.44. The ram compression in supersonic-hypersonic flow includes shock compression that causes an increase in gas temperature and a reduction in the Mach number. In the conventional ramjet, there is a normal shock in the inlet that transitions the flow into subsonic regime. The typical combustor Mach numbers in a conventional ramjet is ∼0.2–0.3. The combustor exit temperature Tt4 may reach stoichiometric levels of ∼2000–2500 K.

images

inline FIGURE 12.44 Conventional ramjet cycle and its ideal thermal efficiency

The equations that are used in the calculation of conventional ramjet performance are listed in Tables 12.8 and 12.9. The equations in these tables are listed sequentially; therefore, they are useful in computer-based calculations/simulation of ramjets. These are the same equations as in turbojets that we derived in Chapter 4, except we have set πc = 1 to simulate a ramjet.

inline TABLE 12.8 Summary of Ramjet Equations in Terms of Dimensional Parameters (for calorically perfect gas)

Given
  • M0, p0, T0 (or altitude), γ, and R
  • ηd or πd
  • πd, ηd, QR, and Tt4 or τλ
  • πn or ηn and p9
Calculate f, , TSFC, Is, ηth, ηp and ηo
get flight total temperature
get flight total pressure
get flight velocity
get total pressure at diffuser exit
pt4 = pt2b get burner exit total pressure
get fuel-to-air ratio in the burner
get nozzle total pressure ratio
pt9 = pt4 · πn get total pressure at nozzle exit
get nozzle exit Mach number
get nozzle exit static temperature
get speed of sound at nozzle exit
V9 = a9 · M9 get nozzle exit velocity
get specific thrust
get thrust-specific fuel consumption
Is=1/(g0 · TSFC) get (fuel)-specific impulse
get propulsive efficiency
get cycle thermal efficiency
get overall efficiency

inline TABLE 12.9 Summary of Ramjet Equations in Terms of Nondimensional Parameters (for calorically perfect gas)

Given
  • M0, p0, T0 (or altitude), γ, and R
  • ηd or πd
  • πd, ηd, QR, and Tt4 or τλ
  • πn or ηn, and p9
Calculate f, , TSFC, Is, ηth, ηp and ηo
get ram temperature ratio
get ram pressure ratio
get inlet total pressure ratio
get fuel-to-air ratio
get nozzle pressure ratio
get nozzle exit Mach number
get exhaust velocity
get specific thrust
get thrust-specific fuel consumption
Is = 1/(g0 · TSFC) get fuel specific impulse
get propulsive efficiency ηp
get thermal efficiency ηth
get overall efficiency ηo

12.15.1 Supersonic Combustion Ramjet

Conventional ramjets cannot produce static thrust. They require ram compression, that is, some forward flight speed, before these simple airbreathing engines (or “ducts with burners”) can produce thrust. It is seemingly ironic that the same engines (i.e., ramjets) cease to produce thrust at very high ram compressions! There are two aspects of ram compression that, at high speeds, are detrimental to thrust production. We studied them in Example 12.13. These are

  1. the inlet total pressure recovery that exponentially deteriorates with flight Mach number, and
  2. the rising gas temperature in the inlet that cuts back (ΔTt)burner to eventually zero.

The worst offender in total pressure recovery of supersonic/hypersonic inlets is the normal shock, which incidentally is also responsible for the rising gas temperatures in the burner. Therefore, if we could only do away with the normal shock in the inlet, we should be getting a super efficient ramjet. But the flow in a supersonic inlet without a (terminal) normal shock would still be supersonic! How do we burn fuel in an airstream that is moving supersonically? The short answer is not very efficiently! Despite the obvious challenges, supersonic combustion ramjets, or scramjets, are born out of necessity. Scramjets hold out the promise of being the most efficient airbreathing engines, that is, with the highest (fuel)-specific impulse, at flight Mach numbers above ∼6 to suborbital Mach numbers; although the upper limit in scramjet operational Mach number is still unknown.

Scramjet engines, more than any other airbreathing engine, need to be integrated with the aircraft. The need for integration stems from a long forebody that is needed at hypersonic Mach numbers to efficiently compress the air. In addition, a long forebody offers the largest capture area possible for the engine, A0. Aft integration with the aircraft allows for a large area ratio nozzle suitable for high altitude hypersonic vehicles. Figure 12.45 shows a generic scramjet engine that is integrated in a hypersonic aircraft. The inlet achieves compression both externally and internally through oblique shocks. The series of oblique shock reflections inside the inlet create a shock train, known as the isolator. The flow that emerges from the isolator is supersonic, say Mach 3.0, as it enters the combustor. Achieving efficient combustion at supersonic speeds is a challenge.

images

inline FIGURE 12.45 Typical arrangement of a generic scramjet engine on a hypersonic aircraft

Fuel injection, atomization, vaporization, mixing, and chemical reaction timescales should, by necessity, be short. If air is moving at Mach 3, where the speed of sound is ∼333 m/s, it traverses 1 m in ∼1 ms. This is just an indication of the convective or residence timescale in the combustor. Hydrogen offers these qualities. For example, hydrogen offers ∼1 /10 of the chemical reaction timescale of hydrocarbon fuels, as discussed in Chapter 7. Additionally, hydrogen is in a cryogenic state as a liquid; therefore, it offers aircraft and engine structure regenerative cooling opportunities.

By accepting high Mach numbers inside the scramjet engine, the static temperature of the gas will be lower throughout the engine and thus the burner is allowed to release heat in the combustor that is needed for propulsion. The h–s diagram of the scramjet cycle is shown in Figure 12.46.

images

inline FIGURE 12.46 Static and stagnation states of gas in a scramjet engine

12.15.1.1 Inlet Analysis

Scramjet inlets are integrated with the aircraft forebody, involving several external and internal oblique shocks. In addition, the vehicle’s angle of attack, α, impacts the inlet recovery through oblique shock waves’ angles. For a given forebody shape, we may use shock-expansion theory (discussed in Chapter 2) to calculate the waves’ orientations and the associated total pressure recovery. In the absence of the detailed aircraft forebody geometry, or in the preliminary phase, we may assume an inlet recovery or resort to standards such as AIA or MIL-E-5008B. We listed these in Chapter 6; they are repeated here for convenience.

(6.34)numbered Display Equation
(6.35)numbered Display Equation
(6.36)numbered Display Equation

12.15.1.2 Scramjet Combustor

Although the detailed design and analysis of supersonic combustion is beyond the scope of this book, we can still present global models and approaches that are amenable to closed-form solution. Supersonic branch of Rayleigh flow, that is, a frictionless, constant-area duct with heating is a possible (and the simplest) model for a scramjet combustor. The fact that the analysis was derived for a constant-area duct is not a major limitation, since we can divide a combustor with area change into a series of combustors with constant areas. Rayleigh flow has been detailed in Chapter 2and we will not review it here. However, we may offer to analyze a variable-area combustor that maintains a constant pressure. From a cycle analysis viewpoint, constant-pressure combustion is advantageous to cycle efficiency and thus we consider it here further. Figure 12.47 shows a frictionless duct with heat exchange, but with constant static pressure. We assume the gas is calorically perfect and the flow is steady and one-dimensional. The contribution of fuel mass flow rate to the gas in the duct is small; therefore, combustion is treated as heat transfer through the wall. The problem statement specifies the inlet condition and heat transfer rate (or fuel flow rate or fuel-to-air ratio) and assumes a constant static pressure in a duct with a variable area. The purpose of the analysis is to calculate the duct exit flow condition as well as the area A4.

images

inline FIGURE 12.47 A frictionless, constant-pressure combustor

We apply conservation principles to the slab of fluid, as shown. The continuity requires

The balance of x-momentum gives

Equation 12.112 signifies a constant-velocity flow, that is,

(12.113)numbered Display Equation

Therefore, the continuity Equation 12.111 demands ρA = constant, or area is inversely proportional to density.

The energy balance written for the duct gives

(12.114a)numbered Display Equation

Since with constant velocity, the kinetic energy (per unit mass) remains constant, then

(12.114b)numbered Display Equation

The RHS of Equation 12.115 is known (as an input to the problem), which then establishes the exit static temperature T4. Since static pressure is constant, we can calculate the exit density ρ4. Also, the area ratio follows the inverse of the density ratio, that is,

(12.116)numbered Display Equation

which establishes the exit flow area. The exit static temperature calculates the exit speed of sound and Mach number (since V = constant).

We may calculate a critical heat flux q* (or critical fuel-to-air ratio f*) that will choke the duct at its exit. Equation 12.115 is written as

(12.117)numbered Display Equation

Also, Mach number ratio M4/M2 is related to the ratio of speeds of sound following

(12.118)numbered Display Equation

For a choked exit, M4 = 1; therefore, the critical heat flux q* in terms of duct inlet flow conditions is expressed as

(12.118a)numbered Display Equation

In terms of fuel-to-air ratio, we may express q* as f* QR ηb to get

(12.118b)numbered Display Equation

With calculated exit Mach number, we may arrive at the total pressure at the exit, following

numbered Display Equation

An input to this analysis is the burner efficiency ηb. However, combustion efficiency in supersonic streams is not as well established or understood as the conventional low-speed burners. Mixing efficiency in supersonic shear layers, without chemical reaction, forms the foundation of scramjet combustion efficiency. A supersonic stream mixes with a lower speed stream (in our case the fuel) along a shear layer by vortex formations and supersonic wave interactions. These interactions require space along the shear layer or flow direction. Consequently, burner efficiency is a function of the combustor length and continually grows with distance along the scramjet combustor. Burrows and Kurkov (1973) may be consulted for some data and analysis related to supersonic combustion of hydrogen in vitiated air. Heiser et al. (1994) should be consulted for detailed discussion of hypersonic airbreathing propulsion.

12.15.1.3 Scramjet Nozzle

We have discussed hypersonic nozzles, to some extent, in Chapter 6. The large area ratios needed for high-speed, high-altitude flight prompted the use of aircraft aft-underbody as the expansion ramp for the scramjet nozzle (shown in Figure 12.45). For perfect expansion, we may approximate the area ratio requirements, A10/A9, by one-dimensional gas dynamic equations. Severe overexpansion at lower altitudes causes a shock to appear on the aft underbody of the vehicle. The complicating factors in the analysis of hypersonic airbreathing nozzles are similar to those in the rocket nozzles, that is, continuing chemical reaction in the nozzle, flow separation, cooling, and flow unsteadiness, among others. Simple nozzle design based on the method of characteristics is a classical approach that students in gas dynamics learn. However, developing robust, high-fidelity computational fluid dynamics codes for direct numerical simulation of viscous reacting flows is the most powerful tool that researchers and industry are undertaking.

12.16 Rocket-Based Airbreathing Propulsion

Integrating rockets with ramjets has long been the solution of overcoming ramjets lack of takeoff capability. A ram-rocket may be a configuration similar to the one shown in Figure 12.48.

images

inline FIGURE 12.48 Schematic drawing of a gas generator ram-rocket

A fuel-rich solid propellant rocket provides the takeoff thrust. The air intake is sealed off until forward speed of the aircraft can produce the needed ram compression for the ramjet combustor. The rocket motor serves as the gas generator for the ramjet, that is, combustion gases from the fuel-rich solid propellant in the rocket motor mix with the air and the mixture is combusted in the ramjet burner to produce thrust. Variations of this scheme, such as a separate ramjet fuel, may be used in a gas generator ram-rocket. There is no new theory to be presented here, at least at the preliminary level, that is, we have developed the tools in this book to analyze both components individually and combined. A ram-rocket configuration where the rocket provides takeoff thrust is shown in Figure 12.49. The ramjet fuel is injected in the airstream for sustained thrust, as shown.

images

inline FIGURE 12.49 A ram-rocket configuration

A rocket placed in a duct will draw air in, and through mixing of the cold and hot gases will enhance its thrust level. This is the ejector principle. Since the propulsion is based on the rocket and is assisted by the secondary air mixing through an inlet, this may be referred to as an airbreathing rocket. The duct that follows the rocket nozzle is then the mixer-ejector, which may also serve as the combustor of a conventional or scramjet engine for combined operation. Now, we have clearly passed the airbreathing and rocket propulsion boundaries and entered into the realm of rocket-based combined cycle (RBCC) propulsion (Figure 12.50). This is the future of hypersonic flight from takeoff to orbit. It clearly is an exciting time to be a propulsion engineer/researcher.

images

inline FIGURE 12.50 Concept in RBCC propulsion

Since the writing of the first edition, we have witnessed continued development and a breakthrough in airbreathing rocket technology called the Synergetic Air Breathing Rocket Engine (SABRE) from Reaction Engines Ltd in the United Kingdom The breakthrough is in the thermal management of air in the airbreathing phase of the combined cycle that starts at takeoff and ends in Mach 5.5+ and at 26 km altitude. A closed-loop cryogenic helium cycle precools the inlet air to −150°C in an ultra-light weight heat exchanger. Liquid hydrogen is used as the fuel in the common-core rocket thrust chamber, which is preheated prior to injection in the rocket combustion chamber by cooling down the helium in the helium cycle. The pure rocket mode uses onboard liquid oxygen in a LOX/LH2 rocket thrust chamber. Figure 12.51 shows the components in the SABRE (courtesy of Reaction Engines Ltd). Figure 12.52 shows the closed-loop heat exchanger systems in the SABRE, which include helium for the air intake system and hydrogen for the helium loop heat exchanger and injection into the rocket thrust chamber. Figure 12.53 shows the aerospace plane SKYLON that is designed around the SABRE for reusable SSTO capability. A comparison of propulsion concepts for SSTO reusable launchers is presented by Varvill and Bond (2003); the is recommended for further reading.

images

inline FIGURE 12.51 Critical components in SABRE engine. Source: Reproduced by permission of Reaction Engines

images

inline FIGURE 12.52 SABRE cycle showing its closed-loop helium precooler and the engine system

images

inline FIGURE 12.53 SKYLON in orbit with payload bay doors open. Source: Reproduced by permission of Reaction Engines

12.17 Summary

High-speed atmospheric flight benefits from airbreathing engines. The oxygen in the air is the infinite oxidizer supply that combustion engines need. Requirements of takeoff to earth orbit necessitate a multitude of airbreathing and rocket engines. Interestingly, the synergy of the rocket-based combined cycle brings about the promise of efficient, cost-effective, and reliable hypersonic propulsion.

The unifying figure of merit in propulsion systems is the specific impulse. Although chemical rockets’ specific impulse is in the range of 200–500 s and airbreathing engines an order of magnitude higher (i.e., ∼2000–10, 000 s), chemical rockets produce thrust that is independent of vehicle speed. The airbreathing engines lose their advantage as flight Mach number increases into the hypersonic regime. Chemical rockets come with liquid, solid, hybrid, or gaseous propellants. Liquid propellant rockets require a feed system that involves either a pressurized feed system with regulators and valves or a gas generator-driven turbopump system. These systems are heavy and add to the propulsion mass fraction and cost. In contrast, solid propellant rockets have no feed system and are, thus, lightweight and low cost and often are used as strap-on boosters. The controllability of thrust in a liquid rocket engine is the clear advantage that it has over a solid rocket motor. The ejector concept of airbreathing rockets combined with the dual-mode subsonic/supersonic combustion ramjets or the airbreathing rockets in a dual-cycle engine with precooler (as in SABRE) form the foundation of promising new engines that is suitable for continuous operation from runway to orbit.

Finally, a new era in commercial launch services, space explorations and space tourism has begun. The success of private sector in this new arena will promise the 21st century to bring space closer to mankind.

References

  1. Burrows, M.C. and Kurkov, A.P., “Analytical and Experimental Study of Supersonic Combustion of Hydrogen in a Vitiated Airstream, ” NASA TMX-2628, 1973.
  2. “Explosive Hazard Classification Procedures, ” DOD, U.S. Army Technical Bulletin TB 700-2, 1989.
  3. Forward, R.L., “Advanced Propulsion Systems, ” Chapter 11 in Space Propulsion Analysis and Design, Humble, R.W., Henry, G.N., and Larson, W.J. (Eds), McGraw-Hill, New York, 1995.
  4. Gordon, S. and McBride, S., “Computer Program for Calculation of Complex Chemical Equilibrium Compositions and Applications, Vol. 1: Analysis, ” October 1994 and “Vol. 2: User Manual and Program Description, ” NASA Reference Publication 1311, 1996.
  5. Heiser, H.W., Pratt, D.T., Daley, D.H., and Mehta, U.B., Hypersonic Airbreathing Propulsion, AIAA, Inc., Reston, VA, 1994.
  6. Heister, S., “Solid Rocket Motors, ” Chapter 6 in Space Propulsion Analysis and Design, Humble, R.W., Henry, G.N., and Larson, W.J. (Eds), McGraw-Hill, New York, 1995.
  7. Hill, P.G. and Peterson, C.R., Mechanics and Thermodynamics of Propulsion, 2nd edition, Addison-Wesley, Reading, MA, 1992.
  8. Humble, R.W., Lewis, D., Bissell, W., and Sakheim, R., “Liquid Rocket Propulsion Systems, ” Chapter 5 in Space Propulsion Analysis and Design, Humble, R.W., Henry, G.N., and Larson, W.J. (Eds), McGraw-Hill, New York, 1995.
  9. Huzel, D.K. and Huang, D.H., Design of Liquid Pro-pellant Rocket Engines, AIAA, Inc., Reston, VA, 1992.
  10. Isakowitz, S.J., International Reference Guide to Space Launch Systems, 2nd edition, AIAA, Inc., Reston, VA, 1995.
  11. Kerrebrock, J.L., Gas Turbines and Aircraft Engines, 2nd edition, MIT Press, Cambridge, MA, 1992.
  12. Kubota, N., “Survey of Rocket Propellants and their Combustion Characteristics, ” Chapter 1in Fundamentals of Solid Propellant Combustion, Kuo, K.K. and Summerfield, M. (Eds), AIAA Progress Series in Astronautics and Aeronautics, Vol. 90, AIAA Inc., Reston, VA, 1984.
  13. NASA SP-8089, “Liquid Rocket Engine Injectors, ” March 1976.
  14. NASASP-8120, “Liquid Rocket Engine Nozzles, ” July 1976.
  15. Olson, W.T., “Recombination and Condensation Processes in High Area Ratio Nozzles, ” Journal of American Rocket Society, Vol. 32, No. 5, May 1962, pp. 672–680.
  16. Price, E.W., “Experimental Observations of Combustion Instability, ” Chapter 13 in Fundamentals of Solid Propellant Combustion, Kuo, K.K. and Summerfield, M. (Eds), AIAA Progress Series in Astronautics and Aeronautics, Vol. 90, AIAA Inc., Reston, VA, 1984.
  17. Razdan, M.K. and Kuo, K.K., “Erosive Burning of Solid Propellants, ” Chapter 10, in Fundamentals of Solid Propellant Combustion. Kuo, K.K. and Summerfield, M. (Eds), AIAA Progress Series in Astronautics and Aeronautics, Vol. 90, AIAA Inc., Reston, VA, 1984.
  18. Sutton, G.P. and Biblarz, O., Rocket Propulsion Elements, 7th edition, John Wiley & Sons, Inc., New York, 2001.
  19. Witte, A.B. and Harper, E.Y., “Experimental Investigation and Empirical Correlation of Local Heat Transfer Rates in Rocket-Engine Thrust Chambers, ” Technical Report Number 32-244 Jet Propulsion Laboratory, California Institute of Technology, Pasadena, CA, March 1962.
  20. Reaction Engines Ltd website: www.reactionengines.co.uk (last accessed 7 December 2013)
  21. SpaceX Corporation website is www.spacex.com (last accessed 7 December 2013)
  22. Varvill, R. and Bond, A., “A Comparison of Propulsion Concepts for SSTO Reusable Launchers, ” Journal of British Interplanetary Society, Vol. 56, pp. 108–117, 2003.

Problems

  • 12.1 A booster stage produces 190, 400 lb of thrust at sea level and 242, 500 lb at altitude. Assuming the momentum thrust remains nearly constant with altitude, estimate the nozzle exit diameter D2 for the booster.
  • 12.2 RD-170 is a Russian launch vehicle booster. It has a nozzle throat diameter of 235.5 mm and exit diameter of 1430 mm. Propellant flow rate is 2393 kg/s and the nozzle exit static pressure is 7300 Pa. RD-170 develops a vacuum-specific impulse of 337 s. Calculate
    1. nozzle area ratio, A2/Ath
    2. effective exhaust speed c in m/s
    3. vacuum thrust in MN
    4. pressure thrust in vacuum in kN
  • 12.3 A rocket engine has a propellant mass flow rate of 1000 kg/s and an effective exhaust speed of c = 3500 m/s. Calculate
    1. rocket thrust F in kN
    2. specific impulse Is in seconds
  • 12.4 A rocket engine has a chamber pressure of pc = 1000 psia and the throat area is Ath = 1.5 ft2. Assuming that the nozzle is perfectly expanded with the gas ratio of specific heats γ = 1.2 and the ambient pressure of p0 = 14.7 psia, calculate
    1. optimum thrust coefficient CF, opt
    2. thrust F in lbf
    3. nozzle exit Mach number M2
    4. nozzle area expansion ratio A2/Ath
  • 12.5 Consider a fuel-rich combustion of the hydrocarbon fuel known as RP-1 and oxygen in a liquid propellant rocket according to
    numbered Display Equation

    Calculate

    1. number of moles of oxygen nO2
    2. the mixture ratio r, i.e., the oxidizer-to-fuel ratio
    3. the molecular weight of the mixture of gases in the products of combustion
  • 12.6 The average specific impulse of a chemical rocket is 360 s. The rocket is in a zero gravity vacuum flight. Calculate and graph vehicle terminal speed ΔV for the propellant fraction ζ, that ranges between 0.80 and 0.95.
  • 12.7 A rocket is vertically launched and operates for 60 s and has a mass ratio of 0.05. The (mean) rocket-specific impulse is 375 s. Assuming the average gravitational acceleration over the burn period is 9.70 m/s2, calculate the terminal velocity of the rocket with and without gravitational effects. Neglect the effect of aerodynamic drag in both cases.
  • 12.8 A rocket has a mass ratio of MR = 0.10 and a mean specific impulse of 365 s. The flight trajectory is described by a constant dynamic pressure of q0 = 50 kPa. The mean drag coefficient is approximated to be 0.25, the vehicle initial mass is m0 = 100, 000 kg, and the vehicle (maximum) frontal cross-sectional area Af is 5 m2. For a burn time of 100 s, calculate the rocket terminal speed while neglecting gravitational effect.
  • 12.9 In comparing the flight performance of a single-stage with a two-stage rocket, let us consider the two rockets have the same initial mass m0, the same payload mass mL, and the same overall structural mass ms. The structural mass fraction inline, which is defined as the ratio of stage structural mass to the initial stage mass, is also assumed to be the same for the single-stage rocket and each of the two stages of the two-stage rocket. For the effective exhaust speed of 3500 m/s be constant for the single-stage and each stage of the two-stage rocket, calculate the terminal velocity for the two rockets in zero gravity and vacuum flight, for
    numbered Display Equation
  • 12.10 A liquid propellant rocket uses a hydrocarbon fuel and oxygen as propellant. The heat of reaction for the combustion is QR = 18.7 MJ/kg. The specific impulse is 335 s and the flight speed is 2500 m/s. Neglecting the propellant kinetic power at the injector plate, calculate
    1. effective exhaust speed c in m/s
    2. propulsive efficiency ηp
    3. overall efficiency ηo
  • 12.11 An injector plate uses an unlike impingement design. The fuel and oxidizer orifice discharge coefficients are Cdf = 0.80 and Cdo = 0.75. The static pressure drop across the injector plate for both oxidizer and fuel jets is the same, Δpf = Δpo = 180 kPa. The fuel and oxidizer densities are ρf = 325 kg/m3 and ρo = 1200 kg/m3 and the oxidizer-fuel mass ratio is r = 3.0. Calculate
    1. oxidizer-to-fuel orifice area ratio Ao/Af
    2. oxidizer and fuel jet velocities vo and vf
  • 12.12 A solid rocket motor has a design chamber pressure of 10 MPa, an end-burning grain with n = 0.4 and r = 3 cm/s at the design chamber pressure and design grain temperature of 15°C. The temperature sensitivity of the burning rate is σp = 0.002/°C, and chamber pressure sensitivity to initial grain temperature is πK = 0.005/°C. The nominal effective burn time for the rocket is 120 s, i.e., at design conditions. Calculate
    1. the new chamber pressure and burning rate when the initial grain temperature is 75°C
    2. the corresponding reduction in burn time Δtb in seconds
  • 12.13 Extract the erosive burning parameter k from the data of Figure 12.30 for the two solid propellants shown.
  • 12.14 A regeneratively cooled rocket thrust chamber has its maximum heat flux of 15 MW/m2 near its throat. The hot gas stagnation temperature is 3000 K and the local gas Mach number is assumed to be ∼1.0. The gas mean molecular weight is MW = 23 kg/kmol and the ratio of specific heats is γ = 1.24.

    Calculate

    1. gas static temperature Tg in K
    2. gas speed near the throat in m/s
    3. gas-side film coefficient hg for Twg ∼ 1000 K
  • 12.15 A rocket combustion chamber is designed for a chamber pressure of pc = 50 MPa. The combustion gas has a ratio of specific heats γ = 1.25. If this rocket is to operate between sea level and 200, 000 ft altitude, calculate the range of area ratios in the nozzle that will lead to perfect expansion at all altitudes. Assume isentropic flow in the nozzle.
  • 12.16The propellant flow rate in a chemical nozzle is 10, 000 kg/s, the nozzle exhaust speed is 2200 m/s, and the nozzle exit pressure is p2 = 0.01 atm. Assuming the nozzle exit diameter is D2 = 2 m, calculate
    1. the pressure thrust (in MN) at sea level
    2. the effective exhaust speed c (in m/s) at sea level
  • 12.17 A solid propellant rocket motor uses a composite propellant with 16% aluminum. The same propellant with 18% aluminum enhances the combustion temperature by 5.7%. Assuming in both cases that the solid particles are fully accelerated (i.e., Vs = Vg) in the nozzle but the solid temperature remains constant (i.e., Ts = constant), calculate the ratio of specific impulse in the two cases. Aluminum specific heat is cs = 903 J/kg · K, and the specific heat at constant pressure for the gas is cpg = 2006 J/kg · K.
  • 12.18 The coefficient of linear thermal expansion for a solid propellant grain is 1.5 × 10− 4/°C.Calculate the change of length ΔL for a 1 m long propellant grain that experiences a temperature change from − 30°C to + 70°C.
  • 12.19 A ramjet has a maximum temperature Tt4 = 2500 K. The inlet total pressure recovery πd varies with flight Mach number according to
    numbered Display Equation

    The ramjet burns hydrogen fuel with QR = 120, 000 kJ/kg and combustor efficiency and total pressure ratio are ηb = 0.99 and πb = 0.95, respectively. The nozzle is perfectly expanded and has a total pressure ratio πn = 0.90. Assuming a calorically perfect gas with γ = 1.4 and R = 287 J/kg · K, use a spreadsheet to calculate the ramjet (fuel)-specific impulse, propulsive, thermal, and overall efficiencies over a range of flight Mach number starting at M0 = 3 up until ramjet ceases to produce any thrust. Altitude pressure and temperature are 15 kPa and 250 K, respectively.

  • 12.20 Consider a scramjet in a Mach 6 flight. The fuel for this engine is hydrogen with QR = 120, 000 kJ/kg. The inlet uses multiple oblique shocks with a total pressure recovery following MIL-E-5008B standards for M0 > 5, i.e.,
    numbered Display Equation

    The combustor entrance Mach number is M2 = 2.6. Use fric-tionless, constant-pressure heating, i.e., Cf = 0 and p4 = p2, to simulate the combustor with combustor exit Mach number M4 = 1.0. All component parameters and gas constants are shown in the schematic drawing below.

    images

    inline FIGURE P12.20

    Calculate

    1. Inlet static temperature ratio T2/T0
    2. combustor exit temperature T4 in K
    3. combustor static pressure ratio p4 /p2
    4. fuel-to-air ratio f
    5. nozzle exit Mach number M10
    6. nondimensional ram drag Dram/p0A1 (note that A0 = A1)
    7. nondimensional gross thrust Fg0A1
    8. fuel-specific impulse Is in seconds
    9. combustor area ratio A4/A2
    10. nozzle area ratio A10/A4
    11. thermal efficiency
    12. propulsive efficiency
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset
18.117.75.236