Sie sind auf Seite 1von 16

41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit AIAA 2005-3935

10 - 13 July 2005, Tucson, Arizona

AIAA-2005-3935
Joint Propulsion Conference, Tucson, Arizona 2005

Modeling of Radiation Heat Transfer in Liquid


Rocket Engines
M.H. Naraghi*
Department of Mechanical Engineering, Manhattan College, Riverdale, NY
10471

S. Dunn† and D. Coats‡


SEA Inc., 1802 North Carson Street, Suite 200 Carson City, NV 89701

This paper presents a radiation heat transfer model for liquid rocket
engines. The radiation model is conjugated to an existing conductive and
convective model for liquid rocket engines (TDK-RTE). The new TDK-RTE
model is used to analyze the effects of radiative heat transfer on thermal
characteristics of two regeneratively cooled rocket engines. One of the
engines is the Space Shuttle Main Engine (SSME), which uses liquid
hydrogen and liquid oxygen as propellant and liquid hydrogen as coolant.
The other engine has RP1 (a hydrocarbon fuel) and liquid oxygen as
propellant and liquid oxygen as coolant. It is shown that the gas radiation
has some effect on the wall temperature for the SSME and a small effect on
its coolant flow characteristics. For the RP1-LO2 engine, however, the gas
radiation significantly changes the wall temperature and coolant flow
characteristics. It is also shown that in order to maintain reasonable wall
temperatures and cryogenic coolant flow temperature and pressure, the
design of cooling channels is significantly influenced by radiation.

I. Introduction

Thermal analysis is an essential and integral part of the design of rocket engines. The need for
thermal analysis is especially important for reusable engines, where an effective and efficient
cooling system is crucial in extending the engine life. The rapid and accurate estimation of
propulsion system thermodynamic heat loads and thermal protection system effectiveness is
required if new vehicle propulsion concepts are to be evaluated in a timely and cost effective
manner. In high-pressure engines hot-gas temperatures are extremely high (approximately
7000°R). It is therefore essential to be able to estimate the wall temperature and ensure that the
material can withstand high temperatures. Furthermore, an accurate thermal model enables an
engine designer to modify the cooling channel configuration for optimum cooling at regions of
high thermal loads. It should be noted that the under-cooling of an engine could result in
catastrophic failure, while over-cooling would result in loss of engine performance. This loss of
performance may be the result of higher coolant compressor requirements, or due to decreased

*
Professor, mohammad.naraghi@manhattan.edu, Senior Member AIAA.

Vice President, stu@seainc.com, Member AIAA.

President, doug@seainc.com. Associate Fellow AIAA.
1
DISTRIBUTION A
Approved for public release, distribution is unlimited
Copyright © 2005 by the American Institute of Aeronautics and Astronautics, Inc. All rights reserved.
effective flow area at the throat (larger boundary layer displacement when the liner is temperature
is very low).

The thermal phenomena in rocket engines involve interactions among a number of processes,
including: combustion in the thrust chamber; expansion of hot-gases through the nozzle; heat
transfer from hot-gases to the nozzle wall via convection and radiation; conduction in the wall;
and convection to the cooling channel. The complexity of the thermal analysis in rocket engines is
due to: three-dimensional geometry; coolant and hot gas heat transfer coefficient dependence on
the pressure and wall temperature; unknown coolant pressure drop and properties; axial
conduction of heat within the wall; and radiative heat transfer between gases and surfaces of the
engine. A comprehensive thermal model must account for all of these items.

The most commonly used hot-gas model for chemical propulsion systems (TDK’04, Two
Dimensional Kinetics Nozzle Performance Computer Program)1 considers only convective heat
transfer to the wall in evaluating heat flux from combustion gases. In fact, most thermal models
neglect the effect of radiative heat transfer, even when the propellant is a hydrocarbon fuel, such
as RP1. The work presented by Hammad and Naraghi2 shows that radiative heat transfer can be
up to 30% of total heat transfer from combustion gases to the nozzle wall. However, they used a
one-dimensional model and did not conjugate gas radiation with wall conduction and coolant
convection. Liu and Tiwari3 studied the effects of radiative heat transfer in chemically reacting
nozzle flows. They concluded that for large nozzles, the radiative heat transfer is dominant over
the convective wall fluxes. Badinand and Fransson4 studied radiative heat transfer in film cooled
LH2/LO2 rocket engines thrust chamber with film cooling. Their results show that the radiation
effect would raise the wall temperature of the nozzle by approximately 140K for a shocked
nozzle. Recently, Wang5 used a CFD model to evaluate convective and radiative heat flux for the
Space Shuttle Main Engine (SSME), a LH2/LO2.system. His results show that radiative heat flux is
about 5% of the convective heat flux in the thrust chamber.

This paper presents a radiation heat transfer model for TDK. This radiation model is based on the
Discrete Exchange Factor (DEF) method for axisymmetric configurations given in Nunes et al.6-8.
The model incorporates all geometric complications of rocket nozzles (e.g., blockage and
shadowing effects due to the throat area). The formulation presented by Modest9 is used to
model the blockage due to the throat. The convective wall heat flux is calculated by TDK’s
Boundary Layer module, and the radiative exchange factors are calculated by TDK’s radiation
model. These are passed to RTE, which subsequently evaluates the wall and coolant
characteristics. The procedure for passing the convective heat fluxes from TDK to RTE is
presented by Naraghi et al.10. This combined TDK-RTE model is used to study the effect of
radiative heat transfer on the rocket thermal characteristics, including coolant temperature,
pressure and Mach number. Two engines were analyzed in this work: first, the Space Shuttle
Main Engine (SSME) a high-pressure chamber with LH2-LO2 (liquid hydrogen and liquid oxygen)
propellant and LH2 as coolant; and second, a high-pressure chamber with RP1-LO2 (C12H23 and
Liquid Oxygen) as propellant and LO2 as coolant.

II. Radiative Heat Transfer Model

The hot-gas species that participate in radiation are water vapor (H2O), CO2, CO and soot. The
absorption coefficients for these species are evaluated based the Planck-mean absorption
coefficients from the HITEMP11 and HITRAN12 databases. As stated by Zhang and Modest13, the
new HITEMP database is an attempt to theoretically predict a large number of hot lines for CO2
and water vapor. The results of the absorption coefficients presented by Zhang and Modest13 for
CO2 show that there is a large discrepancy between the results of HITEMP and HITRAN at
temperatures above 1000K. As discussed by Zhang and Modest13, this discrepancy is due to the
fact that the HITRAN database misses some bands at high temperatures. As a result, it under-
predicts gas absorption coefficients at high temperatures.

2
DISTRIBUTION A
Approved for public release, distribution is unlimited
The Discrete Exchange Factor (DEF) method is used to evaluate the radiative heat flux. This
radiation model is based on the configuration of a typical rocket nozzle shown in Figure 1. In the
current model, radiative exchanges between surfaces and/or volumes are expressed by four
exchange factors: 1) between two surface elements ( dss (ri , r j ) ); 2) between a surface and gas

elements ( dsg (ri , r j ) ); 3)between gas and surface elements ( dgs (ri , r j ) ); and 4)between two

gas elements ( dgg (ri , r j ) ). The equations for these four mechanisms of radiative transport are
(see Figure 1 for nomenclature):

2r j ds j ψ max cos β i cos β jτ (ri − r j )


dss (ri , r j ) =
π ψ min
∫ ri − r j
2
dψ j (1)

ψ max


2 K t j r j dr j dx j cos β iτ (ri − r j )
dsg (ri , r j ) = dψ j
π ri − r j
2

ψ min (2)

r j ds j ψ max
cos β jτ ν (ri − r j )
dgs ν (ri , r j ) =
2π ∫
ψ min ri − r j
2
dψ j
(3)
ψ max


K t j r j dr j dx j τ (ri − r j )
dgg (ri , r j ) = dψ j
2π ri − r j
2

ψ min
(4)

where symmetry with respect to the azimuth angle ψ has been incorporated; ri denotes the
location at which radiation is emitted; r j the position at which radiation is received; β the angle
between the surface normal and the vector connecting ri and r j ; K t j the extinction coefficient at
node j; and τ is the transmittance, which can be defined as:

r j

− ∫ K t (r )dr (5)
τ ( ri − r j ) = e ri

The hot-gases in the thrust chamber and nozzle form a non-homogenous medium (composition,
pressure and temperature of hot-gases change with axial position); hence the extinction
coefficient changes with position. The limits of integration in equations (1) – (4) are ψ min and
ψ max which are the minimum and maximum azimuth angles at which ring element j is seen from
a point on ring element i.

3
DISTRIBUTION A
Approved for public release, distribution is unlimited
rgi dgi

rsi dsi
θi

x
rgj dgj
θj
rsj dsj

Figure 1: Configuration of surface and gas rings within a nozzle and thrust chamber with
throat blockage (shadowing)

4
DISTRIBUTION A
Approved for public release, distribution is unlimited
The allowable range of ψ is dictated by the orientation and relative position of the ring element
pair and blockage effects by the throat. Details concerning the determination of the limiting
azimuth angles are shown below. Geometric consideration of any ring element pair depicted in
Figure 1 reveals:

2
ri − r j = ri 2 + r j2 − 2ri r j cosψ + ( x j − xi ) 2 (6)

and for surface ring elements:

ri − r j cos β i = −(ri − r j cosψ ) cosθ i − ( x j − xi ) sin θ i (7)

ri − r j cos β j = (ri cosψ − r j ) cosθ j + ( x j − xi ) sin θ j


(8)

Where θ k is the angle residing in the r-x plane, measured from the z-axis in the direction of
increasing radius, onto the backside of element k. All surface elements satisfy π / 2 ≤ θ ≤ 3π / 2 .
Combining equations (1)-(8) gives the resulting exchange factor expressions:

ψ max


2ri ri cos θ i cos θ j ds j
2
(φ i − cos ϕ )(φ j − cos ψ )τ (ri − r j )
dss (ri , r j ) = dψ j (9)
π ri − r j
4

ψ min

ψ max


− 2 K t j ri 2 cos θ i cos θ j dr j dx j (φ i − cos ϕ )τ (ri − r j )
dsg (ri , r j ) = dψ j
π ri − r j
3 (10)
ψ min

ψ max


− ri r j cos θ j ds j (φ i − cos ϕ )τ (ri − r j )
dgs (ri , r j ) = dψ j
2π ri − r j
3
(11)
ψ min

ψ max


K t j r j dr j dx j τ (ri − r j )
dgg (ri , r j ) = dψ j
2π ri − r j
2

ψ min
(12)

Where:

ri z j − z i rj zi − z j
φi = + tan θ i and φj = + tan θ j (13)
rj rj ri ri

The limiting angles ψ min and ψ max remain to be determined. The limiting azimuth angles for
surface-to-surface exchange are governed by the configuration and/or blocking surfaces.
5
DISTRIBUTION A
Approved for public release, distribution is unlimited
In many instances it is possible that the view between ring element pairs is partially obstructed by
−1
the throat. The blockage angle, cos Γ , is evaluated by projecting a line from a point on an
emitting ring element (denoted by i) around the periphery of the blocking body at an axial position
xk, such that xk is between xi and xj. The minimum azimuth angle is determined by the intersection
point between the receiving ring element (denoted by subscript j) and the shadowing produced by
the blocking body at xk. This procedure is repeated for several values of xk and can
mathematically be stated as:

 [ DG ( x k ) / 2] 2 ( x j − xi ) 2 − ri 2 ( x j − x k ) 2 − r j2 ( x k − xi ) 2 
Γ= min   (14)
 2ri r j ( z k − z i )( z j − z k ) 
  xk ∈( xi , x j )

The minimum and maximum azimuth angles can then be calculated from the following equations:

ψ min = cos −1 [min(φ i , φ j , Γ,1)] and ψ max = π


(15)

The direct exchange factors calculated with the above formulation account for the direct
exchange of radiation between surface and gas elements. To account for multiple reflections and
scattering of radiation, total exchange factors are introduced. The total exchange factor between
two elements is defined as the fraction of the radiative energy that is emitted from one element
and is absorbed by the other element via direct radiation, multiple reflections, and scatterings
from surfaces and gas. These terms are calculated using the following equations:

[ { [
DSS = I − dss + dsgω 0 Wg I − dggω 0 Wg ] dgs}ρW ] −1
s
−1

⋅ {dss + dsgω W [I − dssω W ] dgs}α


−1 (16)
0 g 0 g

[
DGS = I − dggω 0 Wg ]
−1
dgs

[ {
⋅ I − ρWs dss + dsgω 0 Wg I − dggω 0 Wg [ ]−1
dgs }] α
-1
(17)

Where: [ ] [ ]
DSS = DS i S j , DVS = DVi S j are matrices representing total exchange factors
from surface and gas axisymmetric rings to surface elements; dss = ds i s j , dsg = ds i g j , [ ] [ ]
[ ] [ ]
dgs = dg i s j , dgg = dg i g j are matrices of direct exchange factors between differential
surface/volume ring elements; Ws = ws ,i δ i , j [ ] and [ ]
Wg = wg ,i δ i , j are diagonal matrices of
numerical integration weight factors for surface/volume ring elements, and ρ = [ρδ ]
i, j and
[ ]
α = αδ i , j are diagonal matrices for reflectivities and absorptivities for surface ring elements.

Once the total exchange factors are evaluated using equations


(16) and

(17) the radiative heat flux at the n-th station is computed using the following energy balance
equation:

6
DISTRIBUTION A
Approved for public release, distribution is unlimited
∑ ∑
2 nr + m m⋅ n r
(18)
q r ,n = ws , j DS j S n E s , j + w g , j DG j S n E g , j − E s , n
j =1 j =1

where E sn and E g n are surface and gas emissive powers at station n. Note that the first term in
the right-hand-side of equation (18) is the radiative flux at the surface due to emission from other
surface elements, the second term is due to the radiative flux from gas elements, and the last
term is the radiative heat loss due to emission.

The above model is benchmarked against a number of known analytical solutions and solutions
that are available for cylindrical problems. The results reported by Nunes et al.6-8 show excellent
agreement between the results of this model and published data for cylindrical configurations.

Before performing axial marches, the radiation model uses the Discrete Exchange Factor (DEF)
method to evaluate the total exchange factors. In this model the nozzle is subdivided into a
number of volume and surface nodes as shown in Figure 2. The number of radial nodes is
designated as NCLMN, which is currently set to 5. The number of axial nodes is the same as the
number of stations used in the RTE-TDK model10. The positions of axial nodes coincide with
those of the stations. Since the exchange factors are dependent on gas and surface radiative
properties and the geometry of the nozzle, they are calculated first. Since the composition of
combustion gases varies with axial position, radiative properties of combustion gases are not
constant (the gases are non-homogenous), and are calculated subsequently.

Nodal points in axial


direction, the same
as stations
Nodal points in
radial direction

Figure 2: Position of surface nodes and gas nodes for the radiation model

A radiation module based on the formulations given by equations (1)-(18) has been developed.
Before running the radiation module, TDK must run to determine the mole fraction of radiatively
participating gases. The HITEMP and HITRAN databases are used to evaluate gas absorption
coefficients at various locations in the combustion chamber and nozzle. The radiation module
uses the gas absorption coefficients and engine contour to evaluate the total exchange factors
7
DISTRIBUTION A
Approved for public release, distribution is unlimited
between gas and surface rings. The total exchange factors calculated by the radiation module
and the convective heat transfer table generated by TDK are then passed to the RTE. The
procedure for passing the table of convective heat transfer from TDK to RTE is described by
Naraghi16. The radiative heat flux based on various wall temperatures are then evaluated in
RTE14 using equation (18).

III. Results and Discussion

In order to show the effects of radiation heat transfer on the wall temperature and coolant flow
characteristics, two cases were selected for analysis. Both cases represented typical
regeneratively cooled rocket thrust chambers and nozzles, with commonly used propellants and
coolants. The first case is for the Space Shuttle Main Engine (SSME), a high chamber pressure
design with liquid hydrogen and liquid oxygen (LH2-LO2) as propellant, and hydrogen as the
coolant. The second case is for a high chamber pressure design with RP1 (a hydrocarbon fuel)
and liquid oxygen (RP1-LO2) as propellant, and oxygen as the coolant. For both engines the
chamber and nozzle wall emissivities are 0.9.

A. LH2-LO2 Chamber (SSME)

The specifications of this engine are:


Chamber pressure 3027 psia
O/F 6.0
Contraction ratio 3.0
Expansion ratio 77.5
Throat diameter 10.3 inches
Propellant LH2-LO2
Coolant LH2
Total coolant flow rate 29.06 lb/s
Coolant inlet temperature 95°R
Coolant inlet stagnation pressure 6452 psia
Number of cooling channels 430

The radiation model in the combined TDK-RTE was used to generate results with and without
radiation. TDK provides the composition and mole fractions of species in the combustion gases.
The gas is assumed non-homogenous (absorption coefficient of the gas varies with gas
composition, temperature and pressure along the axial direction). The HITRAN and HITEMP
Planck mean absorption coefficients reported by Zhang and Modesto13 were used to evaluate the
gas absorption coefficients. For this engine H2O is the only species that has a significant thermal
radiation absorption coefficient. The mole fraction of H2O varies from 60% to 74% in the
combustion gases.

Figure 3 shows the resulting wall heat flux distribution. As shown, the effect of radiation is
negligible for the diverging section of the engine. This low radiative flux effect is due to the low
gas emission (the gas temperature and emissivity are low in the diverging section of the nozzle)
and emission from the surface to low temperature gases at the exit of the nozzle. The effect of
radiative heat flux is more pronounced in the converging section of the nozzle and the thrust
chamber where the gas temperature and gas emissivity are high. The results indicate that the
radiation heat flux can reach 10% of the overall wall heat flux. Although apparently small, this
increase in wall heat flux due to radiation can have a significant effect on the resulting wall
temperature. Figure 4 shows the wall surface temperature axial distribution with and without
radiation. As expected, the effect of radiation on the wall temperature is negligibly small for the
diverging section of the engine. This figure shows that the peak temperature occurs at the throat
area (axial position 0). There are also some local peak temperature points, which are caused by
step changes in cooling channel width or height. When radiation is neglected, these local peak
temperatures are smaller than the throat temperature. The effect of including radiative flux, as
8
DISTRIBUTION A
Approved for public release, distribution is unlimited
shown in Figure 4, results in a substantial increase in the local peak temperatures, such that one
of them becomes larger than the throat temperature. This figure also shows that the difference
between wall temperatures for no-radiation and with-radiation cases is larger downstream of the
cooling channel. This is due to the fact that additional radiative heat is continually picked up by
the coolant, which results in a steady increase in coolant temperature, and hence larger wall
temperature in the thrust chamber. Figure 5 shows the increased stagnation temperature of the
coolant due to the radiation heat flux. Using the HITEMP database, the coolant temperature is
increased by 7% at the exit of the cooling channel, as shown in Figure 5.

110
100 No Radiation
Wall Heat Flux (Btu/in s)

With Radiation, HITRAN


90
2

With Radiation, HITEMP


80
70
60
50
40
30
20
-15 -10 -5 0 5 10
Axial Position (in)

Figure 3: Effects of radiation on the wall heat flux at different axial location for the SSME

1600
Wall Surface Temperature (R)

1400

1200

1000
No Radiation
With Radiation HITRAN
800
With Radiation HITEMP

600
-15 -10 -5 0 5 10
Axial Position (in)

Figure 4: Effects of radiation on the wall temperatures of the SSME

9
DISTRIBUTION A
Approved for public release, distribution is unlimited
700

Coolant Stagnation Temperature (R) No radiation


600
With Radiation, HITRAN
With Radiation, HITEMP
500

400

300

200

100
-15 -10 -5 0 5 10
Axial Position (in)

Figure 5: Effects of radiation on the coolant stagnation temperature versus axial position
for the SSME

Coolants at the exit of the cooling circuit are often used to run a turbo-pump. Hence, coolant
pressure drop across the cooling circuit is an important parameter in turbo-pump design. Figure 6
shows the effect of thermal radiation on the coolant pressure distribution. As shown in this figure
thermal radiation increases the coolant pressure drop in the cooling channels by about 5%.
7000

6500
Coolant Stagnation Pressure (psi)

6000

With Radiation, HITEMP


With Radiation, HITRAN
5500
No Radiation

5000

4500

4000
-15 -10 -5 0 5 10
Axial Position (in)

Figure 6: Effects of radiation of the coolant stagnation pressure of the SSME

10
DISTRIBUTION A
Approved for public release, distribution is unlimited
B. RP1-LO2 Chamber

The specifications of this engine are:

Chamber pressure 2,000 psi


O/F (mixture ratio) 1.8
Contraction ratio 3.4
Expansion ratio 7.20
Throat diameter 2.6 inch
Propellant RP1-LO2 (C13H23-LO2)
Coolant LO2
Total coolant flow rate 32.893 lb/s
Coolant inlet temperature 160°R
Coolant inlet pressure 3,000 psi
Number of cooling channels 100
Throat region channel aspect ratio 2.5

Figure 7 shows the contour of this thrust chamber and nozzle. As with the previous engine, TDK
is used to determine the mole fraction of combustion gas species. The species in the combustion
gases that contribute to the absorption and emission of radiation for this engine are H2O, CO2 and
CO. While the mole fractions vary across the chamber and nozzle, the average percentages of
the mole fractions are 33% H2O, 17% CO2 and 30% CO. As in the previous case, the HITEMP
and HITRAN databases were used to determine the gas absorption coefficients based on the gas
composition, pressure, temperature and optical length. It was noted that the absorption
coefficients for HITRAN are substantially lower than that of HITEMP for this engine.

10

8
Liner radius (in)

0
-10 -8 -6 -4 -2 0 2 4
Axial location (in)

Figure 7: RP1-LO2 rocket thrust chamber and nozzle contour showing station locations

DISTRIBUTION A 11
Approved for public release; distribution is unlimited
The resulting wall heat flux for the RP1-LO2 engine is similar to the LH2-LO2 engine (see Figure
8), showing almost no difference at the diverging section of the engine between the radiation and
no-radiation cases. However, the difference between the two heat fluxes is much larger in the
thrust chamber and converging section of the engine. The resulting wall surface temperature
distributions for radiation and no-radiation are shown in Figure 9. These results demonstrate that
there is very little increase in the wall temperature due to radiation in the diverging section of the
nozzle and the throat area (positive axial position). However, for the converging section, the effect
of radiation results in a substantial increase in wall temperature (up to 30%).

An important issue in designing cooling channels is to ensure that the coolant pressure at the exit
of the cooling channel is adequate for the injector. For the first engine, the increased heat flux
due to radiation did not make a significant change in the coolant pressure. However, for the
hydrocarbon engine, as shown in Figure 10, the increase in heat flux due to radiation causes a
significant increase in coolant pressure drop across the cooling circuit (13% increase in coolant
pressure drop). Similarly, Figure 11 shows that the coolant stagnation temperature increases by
approximately 7%

Another coolant characteristic that should be watched closely is the coolant Mach number.
Generally, the coolant Mach number should be kept below 0.3 throughout the cooling passage.
Figure 12 shows that the coolant Mach number is in fact less than 0.3 throughout the engine
when the effect of radiation is neglected. However, the results with radiation show that the coolant
Mach number exceeds this critical value at the cooling channel exit. As shown in Figure 12,
including the radiation heat transfer resulted in 18% increase in the coolant exit Mach number.

50

45

40

35 No radiation
With Radiation HITEMP
QW, BTU/s in2

30

25

20

15

10

0
-10 -5 0 5
Axial location (in)

Figure 8: Effect of radiation on the wall heat flux of a LO2 cooled RP1-LO2 engine

DISTRIBUTION A 12
Approved for public release; distribution is unlimited
1400

1200

1000
TW (R)

800
No Radition
With Radition HITEMP

600

400

200
-10 -5 0 5
Axial location (in)

Figure 9: Effects of radiation on the wall temperature of a LO2 cooled RP1-LO2 engine

3000

2900
Coolant stagnation pressure, psi

2800

2700

2600
No radiation
With radiation
2500

2400

2300
-10 -5 0 5
Axial location, in

Figure 10: Effect of radiation on coolant pressure of a LO2 cooled RP1-LO2 engine

DISTRIBUTION A 13
Approved for public release; distribution is unlimited
450

400

No Radiation
With Rdiation HITEMP
Coolant temperature, R

350

300

250

200

150
-10 -5 0 5
Axial location, in

Figure 11: Effect of radiation on the coolant stagnation temperature of a LO2 cooled RP1-
LO2 engine

0.35

0.3

0.25 No radiation
With Radiation, HITEMP
Coolant Mach number

0.2

0.15

0.1

0.05

0
-10 -5 0 5
Axial location, in

Figure 12: Effect of radiation on the coolant Mach number of a LO2 cooled RP1-LO2
engine

DISTRIBUTION A 14
Approved for public release; distribution is unlimited
IV. Concluding Remarks

The effects of gas and surface radiation on the wall temperature, coolant pressure, temperature
and Mach number were studied. The results presented demonstrate that although the increase in
heat flux due to radiation is small, it can have a significant effect on the wall temperature and
coolant flow characteristics.

For a LH2/LO2 engine it is shown that the radiation has a small effect on the wall temperature of
the diverging section of the nozzle. However, the radiation results in a substantial increase in the
wall temperature of the thrust chamber and converging section of the nozzle, such that the local
peak temperature is the same order of magnitude as the throat temperature.

For the RP1/LO2 engine, radiative heat transfer resulted in a 30% increase in wall temperature.
Additionally, it significantly increased the coolant pressure drop and Mach number. Therefore,
neglecting radiation during the design phase may result in a faulty cooling system.

V. Acknowledgement
This work is supported by Edwards Air Force SBIR Phase II contract F04611-03-M-3015.

VI. References

1. Dunn, S.S., Coats, D.E., and French, J.C., “TDK’04™ Two-Dimensional Kinetics (TDK)
Nozzle Performance Computer Program”, User’s Manual, prepared by Software &
Engineering Associates, Inc., Dec 2002.

2. Hammad, K.J., and Naraghi, M.H.N., “Exchange Factor Model for Radiative Heat
Transfer Analysis in Rocket Engines,” AIAA Journal of Thermophysics and Heat Transfer,
Vol. 5, No. 3, pp. 327-334, 1991.

3. Liu, J., and Tiwari, S.N., “Radiative Heat Transfer Effects in Chemically Reacting Nozzle
Flows,” AIAA Journal of Thermophysics and Heat Transfer, Vol. 10, No. 3, 1996.

4. Badinand, T. and Fransson, T.H., “Radiative Heat Transfer in Film Cooled LH/LO Rocket
Engine Thrust Chamber”, AIAA Journal of Thermophysics and Heat Transfer, Vol. 17,
No. 2, pp. 29-34, 2003.

5. Wang, Tee-See, “Multidimensional Unstructured-Grid Liquid Rocket Engine Nozzle


Performance and Heat Transfer Analysis,” AIAA paper 2004-4016 presented at the 40th
AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit July 11-14, 2004, Fort
Lauderdale, Florida.

6. Nunes, E.M., Naraghi, M.H.N., and Modi, V., “Radiative Transfer in Axisymmetric Bodies
With Non-homogenous Media,” presented at the 35th National Heat Transfer Conference,
Anaheim California, June 10-12, 2001.

7. Nunes, E.M., Modi, V., and Naraghi, M.H.N., “Radiative Transfer in Arbitrarily-Shaped
Axisymmetric Bodies with Anisotropic Scattering Media," International Journal of Heat
and Mass Transfer, Vol. 43, pp. 3275-3285, 2000.

8. Nunes, E.M., Naraghi, M.H.N., “A Model for Radiative Heat Transfer Analysis in
Arbitrarily-Shaped Axisymmetric Enclosures“, Numerical Heat transfer, Part A, Vol. 33,
pp. 495-513, 1998,.

DISTRIBUTION A 15
Approved for public release, distribution is unlimited
9. Modest, M.F., “Radiative Shape Factors Between Differential Ring Elements on
Concentric Axisymmetric Bodies,” Journal of Thermophysics and Heat Transfer, Vol. 2,
No. 1, pp. 86-88, 1988.

10. Naraghi, M.H.N., Dunn, S., and Coats, D., “A Model for Design and Analysis of
Regeneratively Cooled Rocket Engines,” AIAA paper 2005-3852, present at the Joint
Propulsion Conference, Fort Lauderdale, July 2004.

11. Rothman LS, Camy-Peyret C, Flaud J-M, Gamache RR, Goldman A, Goorvitch D,
Hawkins RL, Schroeder J, Selby JEA, Wattson RB. “HITEMP, the high-temperature
molecular spectroscopic database 2000”, available through http://www.hitran.com.

12. Rothman, L.S., Rinsland, C.P. , Goldman A., “The HITRAN Molecular Spectroscopic
Database and HAWKS (HITRAN Atmospheric Workstation) 1996 Edition,” J. Quant.
Spectrosc. Radiat. Transfer, Vol. 60, NO. 5, pp. 665-710, 1998.

13. Zhang, H., Modest, M.F., “Evaluation of the Plank-Mean Absorption Coefficient from
HITRAN and HITEMP databases,” JQSRT, Vol 73, pp. 649-653, 2002.

14. Naraghi, M.H.N., “RTE - A Computer Code for Three-Dimensional Rocket Thermal
Evaluation, “ Grant NAG 3-892 report, prepared for NASA Lewis Research Center, 1994
(see http://home.manhattan.edu/~mohammad.naraghi//rte/rte.html for the report).

DISTRIBUTION A 16
Approved for public release, distribution is unlimited

Das könnte Ihnen auch gefallen