Vous êtes sur la page 1sur 18

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25 - 28 July 2010, Nashville, TN

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA 2010-6846
AIAA-2010-6846

Numerical Simulation of Graphite Nozzle Erosion with Parametric Analysis


Ragini Acharya1 and Kenneth K. Kuo2 Dept. of Mechanical and Nuclear Engineering, The Pennsylvania State Univ., University Park, PA 16802

Nozzle throat erosion is a major problem for solid rocket motors since it causes the degradation in the propulsive performance of solid rocket motors. The AP/HTPB composite propellants used in the rocket motors generate high concentrations of oxidizing species such as H2O, OH, and CO2 in the combustion products at temperatures ranging from 2,700 to 3,200 K for non-metalized propellants. Earlier, the authors utilized a comprehensive numerical program called graphite nozzle erosion minimization code for prediction of graphite nozzle throat erosion rates as a function of pressure and propellant composition. From these studies, it was established that various parameters affect the nozzle thermochemical erosion rate including oxidizing species concentrations, flame temperature, and operating pressure. In addition, the thermal properties of graphite could also affect the nozzle throat erosion rate since these are directly related to the surface temperature at the nozzle throat. In order to assess the relative importance of these parameters in terms of their impact on the nozzle throat erosion rate, a parametric analysis was performed in this study. Each of these parameters was systematically varied while keeping all the remaining parameters constant. Based upon this research, it is concluded that flame temperature can affect the thermochemical erosion rate most, followed by chamber pressure and major oxidizing species concentrations. The mechanisms associated with the influence of these parameters are explored and described. A comparison of predicted results with the available experimental data shows match within 20%. The parametric analysis performed in this research provides an in-depth understanding of the thermochemical erosion process and the controlling steps in the nozzle erosion phenomena.

Nomenclature Symbols As,j C Cp C* Ea,s,j D f hc H k M Mw me N Ns Pr pc


1

Description Pre-exponential factor in the Arrhenius reaction rate expression for jth reaction Thermal capacity Specific heat at constant pressure Characteristic velocity of a rocket Activation energy in the Arrhenius reaction rate expression for jth reaction Nozzle diameter Friction factor Convective heat-transfer coefficient N 1 2 2 Total enthalpy of the mixture Yi hi + 2 ( u + v )
i =1

Specific reaction rate constant Mach number Molecular weight Diffusion-limited mass erosion rate = c re,d Number of gas-phase oxidizing species at the solid-gas interface Number of major gas-phase species in the combustion products Prandtl number Average combustor pressure

Post Doctoral Research Associate, Dept. of Mechanical & Nuclear Engineering, 134 Research Building East, Member AIAA. 2 Distinguished Professor, Dept. of Mechanical & Nuclear Engineering, 140 Research Building East, Fellow AIAA. 1
Copyright 2010 by Ragini Acharya and Kenneth K. Kuo. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

pj q r re re,d
re,k
re j

Partial pressure of jth gas-phase species (H2O or OH or CO2) Heat flux Radial co-ordinate Net erosion rate Diffusion-limited erosion rate Kinetic-limited erosion rate Erosion rate of graphite nozzle due to jth gas-phase species only Outer radius of graphite nozzle Reynolds number Nusselt number Time Temperature Average combustor temperature Gas velocity component in axial direction Gas velocity component in radial direction Mole fraction of a gas-phase species Mass fraction of a gas-phase species Axial co-ordinate Thermal diffusivity Temperature exponent in the Arrhenius reaction rate expression Ratio of specific heats Density Viscosity Thermal conductivity Exponent in the viscosity equation =0.6 Volumetric rate of reaction

Ro Re Nu t T Tc u v X Y x

Subscripts: c Graphite cL Centerline d Diffusion D Diameter g Gas phase i ith species index j jth species index s Surface rad,net Radiation with accounting for irradiation from surface t,cL Total and centerline t Turbulent eff Effective value th Nozzle throat location Diacriticals: ~ Mass weighted average (Favre averaged) quantity I. Introduction

The ballistic performance of solid rocket motors experiences degradation due to the erosion of the exposed internal surface of the rocket nozzle, particularly the nozzle throat region. In order to achieve high propulsive thrust, it is important to minimize the nozzle throat erosion. In addition to experimental studies on this subject, it is also imperative to establish an ability to predict the nozzle throat erosion rates as a function of various operating parameters. Graphite is the most common material used for manufacturing rocket nozzles due to its excellent shock resistance, high vaporization temperature, and much lower density than metals. Several types of graphite have been used for solid rocket nozzles, including high density (1.92 g/cc) G-90 graphite, ATJ bulk graphite, pyrolytic graphite, and carbon/carbon composites. However, graphite and carbon/carbon composites have poor erosion

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

resistance, especially at high gas temperatures and long burning durations. The combustion products of the AP/HTPB based solid propellant contain oxidizing species including H2O, CO2, and OH at temperatures from 2,700 to 3,200 K. Extensive experimental research has been performed on the graphite nozzle throat erosion process during past years with several different rocket motor firings, including BATES (ballistic test and evaluation system) motor by Geisler,1 MERM (material evaluation rocket motor) motor by Klager,2 and HIPPO (high-pressure motor) by Dirling et al.3 These experimental data were obtained for rocket motor operating pressures up to 7 MPa (1,000 psia). More recently, Evans et al.4,5 have performed extensive nozzle erosion rate measurements using real-time X-Ray radiography technique when the rocket motor is operating under broad range of pressures (7-35 MPa). In order to establish a correlation between the measured throat erosion rates and operating conditions, several analytical studies have been conducted by Delaney et al.6, McDonald and Hedman7, Gowariker8, Mayberry et al.9, and Borie et al.10 Keswani et al11 also performed comprehensive modeling and numerical simulation of graphite erosion rate at pressures up to 7 MPa with the reaction kinetics of Libby-Blake12 and Golovina.13 Even though the physical and chemical processes of graphite nozzle erosion are extremely complicated, these studies have treated the graphite and carbon/carbon nozzle throat erosion as a thermochemical process. This means that the graphite nozzle throat erosion rate depends upon the diffusion rate of these oxidizing species near the surface region and the heterogeneous reaction rates with the oxidizing species (such as H2O, OH, CO2, O, and O2). More recently, numerical simulation of graphite nozzle throat erosion process was performed as a function of pressure and propellant composition (with and without aluminum addition in the AP/HTPBS based solid propellant) (Acharya and Kuo,14, 15). In these studies, the effect of motor operating pressure, propellant composition, nozzle surface roughness was examined. From these studies, it was found that contrary to popular belief; at lower pressures (~16 MPa), the heterogeneous kinetic rates seem to have a very pronounced effect on the net erosion rates. However, at higher pressures, the nozzle throat erosion process was found to depend more on the rate of diffusion of oxidizing species from the core region to nozzle surface. For metallized solid propellants with higher flame temperatures, the nozzle surface temperature was calculated to be much higher; thereby, the thermo-chemical erosion of nozzle was governed by the diffusion process. Recently, the effect of various heterogeneous kinetic schemes on the net erosion rate of graphite rocket nozzle was also studied (Acharya and Kuo,16). Several other research groups have also performed numerical simulations of graphite nozzle erosion process (Thakre and Yang17, and Bianchi et al.18). In this current study, the authors have continued development of the comprehensive predictive code called graphite nozzle erosion minimization (GNEM) for prediction of graphite nozzle throat erosion rates. From the continued work on nozzle erosion modeling, several key parameters have been established that affect the nozzle throat erosion rates. These parameters can be divided into three major groups; i.e., solid propellant, nozzle material, and nozzle geometry. The solid propellant composition affects the properties of combustion products flowing through the nozzle. The properties include the gas temperature and species composition (which affects the mixture properties including mass diffusivity, viscosity, etc.), affecting the heat and mass transfer to the nozzle material. The nozzle material properties include the thermal conductivity and specific heat, which affect the heating rate of nozzle, thereby affecting the nozzle surface temperature and surface reaction rates of the nozzle throat. In this work, the focus is on the solid propellant, nozzle material, and operating pressure. The nozzle geometry includes the radius of curvature, entrance length, and throat diameter. The nozzle geometry may affect the turbulent boundary layer near the nozzle surface, which changes the heat and mass transfer rates to the nozzle surface, thereby affecting the nozzle throat erosion rate. This work focuses on the effect of small variations in each of these parameters and the combined effect of change in more than one parameter to determine the dependency of nozzle erosion rate on dimensionless groups of these parameters. A comparison of predicted results with available experimental data (including the available highpressure data) is also shown.

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

II.

Approach

A. Graphite Nozzle Erosion Minimization Code The GNEM code couples the solution of turbulent boundary layer in the nozzle with the transient onedimensional heat conduction equation in the graphite nozzle, the heterogeneous surface reactions at the nozzle surface, and the blowing effect due to these surface reactions. The turbulent boundary layer consists of two regions: 1) a viscous sub-layer adjacent to the wall where momentum and heat transfer can be accounted for by viscous shear and molecular conduction; 2) a fully turbulent region, which comprises of the outer region of the turbulent boundary layer. It is common to apply a two-layer model to resolve the flow and heat transfer in turbulent boundary layer by using a law of the wall for the outer region and a linear equation for the inner viscous sub-layer. This approach leads to a discontinuity for the region between these two layers. An improvement in this two-layer approach can be obtained by using the van Driest hypothesis19, which provides a continuous law of the wall that can be applied to the whole turbulent boundary-layer region. The van Driest formula was further modified to incorporate the effect of surface roughness by using the method suggested by Cebeci and Chang20. The GNEM code implements the modified van Driest formula for the turbulent boundary layer in the nozzle. The flow in the nozzle is considered two-dimensional (cylindrical coordinates), compressible, viscous, and quasi-steady. The flow is assumed to be quasi-steady due to the relatively slow regression rate of the nozzle material. The instantaneous governing equations are Favre-averaged to arrive at mean governing equations. A second-order k- turbulence model has been used to achieve closure of the turbulent flow problem. The Favre-averaged values for axial velocity, pressure, and static temperature at the edge of the turbulent boundary layer (in the core region) were obtained by solving governing equations for two-dimensional, compressible, viscous, and quasi-steady flow. Due to relatively high temperatures, the gas-phase species were assumed to be in chemical equilibrium with each other. The heat conduction process in the nozzle material is considered unsteady and one-dimensional in the radial direction normal to the nozzle inner surface. The nozzle material properties are evaluated as functions of local temperature. The detailed formulation of the GNEM model and code description is given in Acharya and Kuo.14 The system of coupled, non-linear partial differential equations was solved using a numerical technique proposed by Patankar and Spalding.21 Several researchers have used this technique successfully. Patankar and Spalding introduced a transformation of coordinates in which the grid points always fit the boundary-layer region even while the thickness of the region is changing. B. Calculation of Nozzle Surface Temperature The nozzle surface temperature is an unknown parameter in the GNEM code. The governing equation for the Favre-averaged enthalpy of the gas-phase is given by Eq.(1). In order to solve this equation, two boundary conditions are required.

2 1 H H H u 2 + v = m rm + eff x r r r Pr eff r Pr eff r

(1)

The gas-phase enthalpy at centerline was determined by using the following equation:

TcL =

1 2 M cL 1 + 2

Tt ,cL

(2)

The condensed-phase temperature was obtained by solving a one-dimensional transient heat conduction equation with variable thermal properties shown by Eq.(3). The one-dimensional equation is adequate for graphite since the transverse thermal conductivity of graphite is significantly lower than the thermal conductivity in radial direction, thereby reducing the heat conduction in the transverse direction negligible in comparison to that in radial direction.

1 T ( cCcTc ) = m c r m c + re ( cCcTc ) t r r r r

(3)

The above equation also requires two boundary conditions. The first boundary condition was obtained by applying the heat-flux balance condition at the inner radius of nozzle material. The external surface of the graphite nozzle was insulated due to presence of silica-phenolic layer. Therefore, a zero heat flux condition at the external radius of nozzle was used as the second boundary condition. These two boundary equations are shown by Eqs.(4)-(5)

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

T r

" + qrad ,net + hii = c


g i =1

Tc r

(4)
r = Ri

Tc r

=0
r = Ro

(5)

Again, the flux-balance equation requires knowledge of nozzle surface temperature. In order to obtain this information, an additional equation was employed for determination of convective heat-transfer coefficient. For fully developed turbulent internal flows, the von Krmn equation given by Eq.(6) for calculation of Nusselt number was used. ( f 8 ) Re D PrL Nu D = (6) 1 + 5 f 8 ( PrL 1) + log (1 + [5 6 ] ( PrL 1) ) In the above equation, the fluid properties can be calculated at the centerline gas temperature, which is a known value. The friction factor f was calculated by using the following equation:

0.316 Re 0.25 D 0.20 f = 0.184 Re D 0.13 0.184 Re D

1 10 4 < Re D < 5 104 3 104 < Re D < 1 106 1 10 < Re D


6

(7)

The Reynolds number was calculated by using the axial velocity obtained from the solution of the momentum equation at the centerline.

Re D =

uD

(8)

The product gas mixture consists of a large number of species; however, the highest 8 species were included in the calculations of mass-averaged physical properties. These gaseous species and their mass fractions in the product gas mixture of non-metallized propellant at 6.45 MPa chamber pressure are shown in Table 1. Table 1: Mass fractions of major species in the product gas-mixture of non-metalized propellant Species H2O HCl CO2 CO Cl OH N2 H2 Mass Fraction 0.2953 0.2558 0.2233 0.0946 0.0110 0.0050 0.1053 0.0031 The thermal and physical properties of each species including specific heat, dynamic viscosity and thermal conductivity were treated as temperature-dependent properties. Specific heat of the gas-mixture was calculated by mass-averaging while the dynamic viscosity and thermal conductivity were calculated by using Wilkes method (Mathur and Saxena22). The thermal diffusivity and Prandtl number of the multi-species gas mixture was determined by using these mixture properties. The convective heat-transfer coefficient was calculated from the Nusselt number by using Eq.(9).

hc = Nu D

(9)

Once the solution of momentum equation was obtained, the convective heat-transfer coefficient was obtained by using Eqs.(6)-(9). This information was utilized in the condensed-phase heat equation to determine the nozzle surface temperature at an axial location. Once the surface temperature is known, the thermal properties of gaseous mixture at nozzle surface were determined and the enthalpy of the gas-phase at the inner nozzle surface was calculated. This information was used as a boundary condition for the solution of gas-phase enthalpy equation. C. Heterogeneous Reaction Kinetics As mentioned earlier, the major species in the combustion products of non-metallized propellant are H2O, HCl, CO2, CO, OH, H2, and N2. Among these species, only H2O, CO2, and OH are found in significant amount for oxidizing reaction considerations. As reviewed in our earlier work (Acharya and Kuo14), HCl does not react with graphite despite its relatively large concentration in the combustion products of the solid propellant considered in this work. In addition, CO, H2, and N2 are not oxidizers; therefore, the reactions between graphite and these gaseous

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

species need not be considered. The remaining gaseous species (H2O, CO2, and OH) are known oxidizers of graphite and the reaction kinetics for oxidation of carbon by these species has been known for some years. The available chemical kinetics parameters were obtained by considering reactions of solid carbon with only single gaseous species and not in a multi-species environment like the one present in the rocket nozzle. Therefore, a semi-global type reaction scheme was employed for the heterogeneous reactions between condensed-phase carbon and gaseous oxidizing species. The term semi-global implies that while multiple reactions are considered for oxidation of graphite, each reaction is considered independent and does not affect the reactions between graphite and other gaseous species. In this work, we have considered a semi-global reaction kinetic schemes for heterogeneous reactions between graphite and gaseous oxidizers. The detailed description of these reaction schemes and corresponding kinetic parameters is shown in Table 2. The reaction rate constant kj is assumed to follow the Arrhenius form given by Eq. (10).

Ea , s , j k j = As , jT j exp RuTs

(10)

where, As,j is the pre-exponential factor in the Arrhenius reaction rate expression for jth reaction, Ea,s,j is the activation energy in the Arrhenius reaction rate expression for jth reaction, j is the temperature exponent in the Arrhenius reaction rate expression for jth reaction, and Ru is universal gas constant. The rate of reaction for graphite oxidation by jth gaseous species is given by the following rate equation:

re,k j = k j p j j c

(11)

Table 2: Detailed description of reaction schemes for graphite oxidation and kinetic parameters Kinetic Parameters Kinetic Temperatur Included Ea,s Reaction Rate Scheme e Range Reactions [kcal/ As mol] MultiOxidizing Species (MOS) (1984) [p~1 atm] 1,575-1,685 800-2,500 800-2,500 C(s)+OH CO+H3 C(s)+H2O CO+H2 C(s)+CO2 2CO -0.5 0.0 0.0
361 kg K0.5 m-2 s-1 atm-1 4.80105 kg m-2 s-1 atm-0.5 9.00103 kg m-2 s-1 atm-0.5

0.0 68.8 68.1

R3 = k3 pOH

0.5 R1 = k1 pH2O
0.5 R2 = k2 pCO2

The reaction scheme consisting of three reactions is called a multiple oxidizing species (MOS) reaction mechanism. Bradley et al.23 employed this semi-global heterogeneous reaction mechanism for oxidation of graphite particles and identified free radical attack as the principle means of graphite oxidation. The reaction kinetics parameters were determined by various groups (Neoh et al.,24 Bonnetain & Hoynant,25) for oxidation of graphite particles by H2O, CO2, or OH and those are also listed in Table 2. Libby and Blake studied the combustion of single carbon particle in an oxidizing ambient gas (H2O or CO2) and concluded that the heterogeneous reactions between graphite and either of these two oxidizing species are first-order reactions with equal kinetic parameters. They established that the activation energy for either of these two reactions as 41.9 kcal/mol. Golovina also calculated the kinetic parameters for graphite oxidation by water vapor and carbon dioxide. He determined that both of these reactions are first-order reactions with activation energy as 40 kcal/mol. Neither group considered graphite oxidation by OH radical. It is also understood that at the time of their study, Libby and Blake did not have sufficient experimental evidence and assumed that the two rate constants were equal. Recently, Culbertson and Brezinsky26 performed shock-tube experiments4 on carbon-steam reaction by using carbon black and graphite as the carbon source and on the carboncarbon dioxide reaction by utilized amorphous carbon black powder. The rate constants of these two reactions were found to be equal to one another. The order of reaction in this scheme for both reactions was determined to be 0.5 with activation energy of 41 Kcal/mol. This reaction scheme is referred as Brezinsky-Culbertson scheme. The
3

For OH radicals, the reaction rate constants correspond with collision efficiencies at the graphite surface of 0.50 and 0.28, respectively. 4 To generate high-pressure and high-temperature conditions as experienced at the rocket nozzle throat. 6

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

comparison of reaction kinetics and discussion of their relative differences have been investigated in another paper published by Acharya and Kuo.16 In this paper, the emphasis on examining the influence of various parameters on chemical erosion of graphite nozzle and therefore, we have considered only one reaction kinetic scheme, i.e., MOS since these scheme predicted closer to the measured nozzle erosion rates. The nozzle geometry along with the nozzle entrance region is shown in Fig.1. The nozzle consists of two regions: 1) a nozzle entrance region with constant outer diameter, and 2) a nozzle region with a tapered outer diameter; covered by an insulating material called silicaphenolic. This nozzle was implemented for high pressure test firings and employed a nozzle boundary layer control system (NBLCS) for mitigation of nozzle throat erosion (see Evans et al.5), which is why the nozzle entrance region is much longer than the usual entrance. The parameters studied in this work and their effects on the nozzle erosion rates should not be influenced by this nozzle entrance region. Therefore, this geometry is adopted, especially since the experimental data used in the current work corresponds to this geometry.

Fig.1: Nozzle geometry and dimensions Due to implementation of a semi-global heterogeneous reaction scheme with mutually independent surface reactions, it is possible to determine the erosion of graphite nozzle surface due to each oxidizing species. The contribution of the jth species in the net erosion rate of graphite nozzle surface is given by the following equation:

re , j =

1 re ,d j

1 + 1 re ,k j

(12)

In the above equation, re,d j is the rate of mass diffusion of jth oxidizing species from core region to the nozzle surface per unit cross-sectional area and normalized by graphite density:

re,d j = Y j re ,d = Y j
re = re j
j =1 N

me,d

(13)

The net erosion rate of graphite is the sum of contributions from all oxidizing species as shown by: (14)

III. Discussion of Results The nozzle throat is actually a short flat section instead of one axial location. This geometry leads to interesting phenomena at the nozzle throat locations as shown by several figures in later part. The flow properties do not change significantly in the nozzle entrance region; however, the convergent-divergent nozzle region experience strong variation in flow field properties (axial gas velocity, gas density, Mach number, gas-temperature, and turbulent parameters) as shown in Fig. 2. The calculated results from the GNEM code are shown in following plots. The axial variation of heat-transfer coefficient in the graphite nozzle is shown in Fig. 3. Similarly, the axial variation of the calculated nozzle surface temperature at the gas-solid interface is shown in Fig. 4. The heat-transfer coefficient and surface temperature show significant dependence on axial direction. Both parameters increase up to the nozzle throat and decrease after the throat. Due to the short flat nozzle throat, the convective heat-transfer coefficient (and as a result the surface temperature) displays interesting behavior characterized by two peaks each at the beginning and the end of the flat nozzle throat.

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

(a) Gas velocity

(b) Gas density

(c) Mach number

(d) Gas temperature

(e) Turbulent dissipation (TD) rate

(f) Turbulent kinetic energy (TKE)

Fig.2: Variation of major variables in the nozzle region (from Acharya and Kuo16)

2 10 Heat Transfer Coefficient, h [W/m -K]


2

10

40 Nozzle Surface Temperature, T [K]


s

35 2200 30 25 20 1800 15 10 5 1400 0 0.05 0.1 0.15 Distance along Nozzle Axis, xd [mm] 0 Initial Nozzle Radius, r [mm]

1 10

Initial Nozzle Radius, r [mm]

6 8 10
3

4 4 10
3

0 0.14 0.15 0.16 0.17

0 0.18

Distance along Nozzle Axis, xd [mm]

Fig.3: Convective heat transfer coefficient as a function of axial distance at 6.45 MPa operating pressure and its variation due to change in stagnation temperature

Fig.4: Nozzle surface temperature as a function of axial distance at 6.45 MPa operating pressure

In the following subsections, the effect of each parameter is discussed. A. Effect of Stagnation Gas Temperature Due to the differences in the propellant composition, the flame temperature could show variations between 10%. Due to this, the stagnation gas temperature in the chamber also shows similar variation. This parameter can affect a number of quantities including the density, velocity, viscosity of gas-mixture, which in turn affect the Reynolds number based upon the diameter and the mass flow rate through the nozzle. Both of these parameters affect the rate

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

of mass transfer of oxidizing species towards the nozzle surface and the heat transfer to the nozzle surface, which results in change in the nozzle throat erosion rate. The relative change in any parameter is defined as follows:

Ref 100 Ref

(15)

The reference (or standard case) stagnation temperature is 3,020 K. The stagnation temperature is changed from 10% to +7.5% and the effect on various quantities at the nozzle throat location is shown in Table 3. The ReD,th decreases with stagnation temperature, the boundary layer thickness increases slightly. The surface temperature at the nozzle throat also increases with flame temperature and as a result the net throat erosion rate shows a four-time percentage change than the respective percentage change in stagnation temperature. Table 3: Summary of effect of stagnation gas temperature on various parameters at nozzle throat Tst,c red ,th th Tst ,c % re,th % Ts,th re,th re j ,th [ mm/s ] [K] [K] [mm] [ mm/s] H2O [ mm/s] OH CO2

ReD,th 10-6 1.03 0.973 0.919 0.895 0.870 0.849

2718 2869 3020 3096 3171 3247

0.108 0.149 0.192 0.214 0.235 0.256

0.097 0.138 0.180 0.201 0.221 0.242

0.00931 0.00911 0.00901 0.00896 0.00890 0.00886

0.0015 0.0023 0.0034 0.0040 0.0048 0.0057

0.655 0.647 0.640 0.636 0.632 0.630

-10.0 -5.0 0.0 +2.5 +5.0 +7.5

-44 -22 0 11 22 33

2180 2240 2300 2330 2360 2390

1.08 1.10 1.12 1.13 1.14 1.14

The relative change in any quantity at the nozzle throat including the net erosion rate, the species-specific erosion rate, and diameter-based Reynolds number, boundary-layer thickness, density, and convective heat-transfer coefficient as a function of relative change in stagnation temperature is shown in Fig. 5. It can be observed that nozzle throat erosion rate due to OH shows slight decrease whereas the rate due to H2O increases drastically. The surface temperature at the nozzle throat increases linearly with Tst,c but the nozzle throat erosion rate increases much more strongly than the Tst,c. As shown in Eq.(10), the reaction rate constant increases exponential with temperature. Therefore, a linear increase in surface temperature results in exponential increase in the reaction rate constant and therefore the reaction rate. Since the nozzle throat erosion rate mirrors the trend shown by the reaction rate, it implies that the erosion rate at these conditions is kinetic-controlled. This means that the kinetic rate at the surface is slower than the rate of mass transfer from the center line to the nozzle surface via mass diffusion. In order to establish a correlation between dimensionless parameters, the nozzle throat erosion rate is normalized by the mass flow rate at the throat location and multiplied by the graphite density. The average mass flux of the combustion products flowing through the nozzle throat is calculated by the following equation:

Gavg = th uth

(16)

Similarly, the boundary-layer thickness is normalized by the throat radius. These two dimensionless parameters are plotted against the Reynolds number (based upon throat diameter) and are shown in Fig.6. The dimensionless throat erosion rate decreases linearly with the ReD,th whereas the dimensionless boundary-layer thickness shows a power law dependence upon the ReD,th.

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

80 % Change in Parameter

r r

. .e, th .
e,OH, th e, H2O s,th

c, th

1.2
avg

BL,th th

/r

t, th
D, th

Re
th

0.200
avg

c e, th

/G

/G

40

BL,th th

c e, th

/r

0.194 0.8
/r =0.373Re /G
avg -0.291 D D,th

0
BL,th th

c e, th

=-0.406Re

+ 4.62

-40 0 5 10 % Change in Temperature 15

0.4 8 9 Re
D, th

0.188 X10
6

10

Fig.5: Relative change in various parameters at the nozzle throat due to stagnation temperature change at 6.45 MPa operating pressure

Fig.6: Relationship between the normalized net throat erosion rate and the Reynolds number at 6.45 MPa operating pressure

The boundary-layer thickness variations with the change in Tst,c at all axial locations in the nozzle are shown in Fig. 7. Similarly, the nozzle throat erosion rates with the change in Tst,c at all axial locations in the nozzle are shown in Fig. 8. Again, in the nozzle throat region, the distribution of erosion rate shows a peak followed by a short plateau region.
10 Boundary Layer Thickness [mm] % Change in T st,c +15 +10 0 -15 0.3 10

Net Erosion Rate, r [mm/s]

2.2

8 0.2 6
% Change in T
st,c

Initial Nozzle Radius, r [mm]

Initial Nozzle Radius, r [mm]

1.8

4 1.4 2 1.0 0.14 0.15 0.16 0.17 0 0.18

0.1

+15 +10 0 -15

0 0.14 0.15 0.16 0.17

0 0.18

Distance along Nozzle Axis, xd [mm]

Distance along Nozzle Axis, xd [mm]

Fig.7: Boundary-layer thickness variation due to stagnation temperature at 6.45 MPa operating pressure

Fig.8: Net erosion rate variation due to stagnation temperature at 6.45 MPa operating pressure

B. Effect of Chamber Pressure The operating pressure directly affects the gas density, which alters the Reynolds number and the average mass flow rate through the nozzle. Both of these parameters increase the rate of mass transfer of oxidizing species towards the nozzle surface and the heat transfer to the nozzle surface, which results in increase in the nozzle throat erosion rate. In this study, the reference (or standard case) operating pressure is 6.45MPa. In order to assess the impact of chamber pressure on the nozzle throat erosion rate, it is changed within 15%. A summary of its effect on various quantities at the nozzle throat location is shown in Table 4. The net nozzle erosion rate and boundary-layer thickness at all axial locations are shown in Figs. 9 and 10, respectively. At the nozzle throat location, the ReD,th increases with

10

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

pressure, the boundary-layer thickness decreases slightly. The surface temperature at the nozzle throat does not change much and yet the net throat erosion rate increases at the same % change as pressure.
Table 4: Summary of effect of stagnation pressure on various parameters at nozzle throat Pc hc,th red ,th re,th re j ,th [ mm/s ] re ,th % [MPa] [kW/m2-K] [ mm/s] H2O [ mm/s] OH CO2 5.48 0.164 0.153 0.0079 0.0028 0.560 12.9 -14.9 5.81 0.173 0.162 0.0083 0.0030 0.586 13.5 -9.9 6.13 0.183 0.171 0.0086 0.0032 0.614 14.1 -5.0 6.45 0.192 0.180 0.0090 0.0034 0.640 14.7 0.0 6.77 0.202 0.189 0.0095 0.0035 0.676 15.3 5.0 7.10 0.212 0.198 0.0099 0.0037 0.703 15.9 10.0 7.42 0.220 0.206 0.0103 0.0039 0.728 16.4 14.4

Pc %

Ts,th [K] 2280 2290 2300 2300 2300 2310 2310

ReD,th 106 0.782 0.829 0.873 0.919 0.966 1.01 1.06

-15 -10 -5 0 5 10 15

0.3

10

10 Boundary-Layer Thickness [mm]


% Change in P
st, c

Net Erosion Rate, r [mm/s]

8 0.2 6
% Change in P
st,c

2.2

-15 -0 +15

Initial Nozzle Radius, r [mm]

Initial Nozzle Radius, r [mm]

1.8

0.1

+15 +10 0 -15

4 1.4 2 1.0 0.14 0.15 0.16 0.17 0 0.18

C 0

0.14

0.15

0.16

0.17

0 0.18

Distance along Nozzle, xd [mm]

Distance along Nozzle, xd [mm]

Fig.9: Net erosion rate variation due to variation in average chamber pressure

Fig.10: Boundary-layer thickness variation due to variation in average chamber pressure

The relative change in any quantity including the net erosion rate, the erosion rate due to each species, and diameterbased Reynolds number, boundary-layer thickness, density, and convective heat-transfer coefficient as a function of relative change in chamber pressure is shown in Fig. 11. It can be observed that nozzle throat erosion rates due to H2O and OH show same relative change as the pressure. The surface temperature at the nozzle throat does not change much with the chamber pressure but the nozzle throat erosion rate still increases with pressure. As shown in Eq.(11), the kinetic reaction rate has a power-law dependence on the pressure. Therefore, the reaction rate increases with pressure even though the surface temperature is constant; however, the dependence of reaction rate on pressure is mostly linear. It is also interesting to observe that the convective heat-transfer coefficient increases with the pressure and yet the surface temperature does not increase. This observation sheds some light on the complex relationship between the surface temperature and throat erosion rate. The reaction rates have exponential dependence upon the temperature but the graphite oxidation reaction by H2O is an endothermic reaction. Therefore, as the reaction rate increases, it also absorbs the energy, which causes the surface temperature to decrease. This is exactly what takes place in this case. Due to increase in the chamber pressure, the boundary layer thickness decreases, the rate of oxidizing species reaching the nozzle surface increases, and the heat transfer rate to the nozzle surface also increases. But the extra heat available is balanced by the extra heat required for the heterogeneous chemical reactions. This balance results

11

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

in almost constant surface temperature. Therefore, it should be understood that the increase in nozzle throat erosion rate is caused by a different physical process than the increase due to stagnation temperature increase. Here also, the correlation between the non-dimensional nozzle throat erosion rate and ReD,th as well as between non-dimensional boundary layer thickness are shown in Fig.12. The non-dimensional throat erosion rate does not show a strong dependence upon the ReD,th whereas the non-dimensional boundary layer thickness shows a power law dependence upon the ReD,th. The pressure effect on the nozzle throat erosion rate is absorbed in the average mass flow rate.
20

. .e, th r .e,OH, th
r r
e, H2O, th s,th

20
h
c, th

1.2
avg

BL,th th

/r

t,th
D, th

% Change in Parameter

Re
th

c e, th

/G

avg

0.198

10

10
.

/G

BL,th th

c e, th

/r

0.8

0.195

0
r

c e, th

/G

avg

=-0.013*Re

D, th

+ 0.868

0.4 0.75

0.192 0.85 Re
D, th

-10 0 5 10 % Change in Pressure

-10 15

0.95 6 X10

1.05

Fig.11: Relative change in various parameters at the nozzle throat due to variation in average chamber pressure

Fig.12: Relationship between the dimensionless throat erosion rate and the Reynolds number

C. Effect of Species Concentration The nozzle throat erosion rate with the non-metalized propellant shows strong dependency upon the reaction kinetics between the gaseous oxidizing species and the graphite surface and it can be said that the rate of reactions is slower than the rate of mass diffusion to the nozzle surface. Therefore, it is very important to evaluate the effect of oxidizing species concentration on the nozzle erosion rate. As shown previously, the major oxidizing species is H2O, followed distantly by the OH radical. The effect of CO2 on the net erosion rate is negligible in comparison and therefore, the effects of variations in its mole fraction are not considered in the following section.

The species concentration could change due to propellant composition variations. This parameter directly affects the kinetic rates of reactions at the nozzle throat surface. The reference (or baseline) H2O mole fraction is 0.4298 and that of OH is 0.008. Since the species concentrations may show higher variations than temperature or pressure, they were varied within 30%. A summary of the effect of variations of H2O mole fraction on various quantities at the nozzle throat location is shown in Table 5. Since the species concentration does not affect ReD, boundary-layer thickness, and the convective heat-transfer coefficient, these quantities are not shown. As expected, reH O ,th increases
2

with H2O mole fraction. However, reOH ,th decreases slightly with increasing H2O mole fraction. The surface temperature also decreases with increasing H2O mole fraction. The relative change in any quantity at the nozzle throat including the net erosion rate, the species-specific erosion rate by both H2O and OH, and surface temperature as a function of relative change in H2O is shown in Fig. 13. It can be observed from Fig. 15 that the throat erosion rate by H2O increases with the mole fraction of H2O. However, the throat surface temperature and throat erosion rate by OH decrease with H2O mole fraction. This effect can be explained as follows. Due to increase in the reaction rate of graphite oxidation by H2O, the heat required for the heterogeneous chemical reaction increases since this reaction is endothermic. This causes the surface temperature to decrease, which in turn decreases the rate of graphite-H2O reaction. Despite of this counter effect, the overall value of reH O ,th increases since the negative effect caused by the
2

reduction of surface temperature is compensated by the increase in the H2O mole fraction. The species specific erosion rate by OH decreases since surface temperature decreases and the mole fraction of OH is constant. 12

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

Table 5: Summary of effect of H2O concentration on various parameters at nozzle throat Ts,th red ,th X H 2O re,th % X H 2O % re,th re j ,th [ mm/s ] [K] [ mm/s] H2O [ mm/s] OH CO2 0.3009 0.1737 0.160 0.0097 0.0040 0.529 -9.73 -30 2330 0.3438 0.1812 0.168 0.0094 0.0038 0.566 -5.83 -20 2320 0.3868 0.1868 0.174 0.0092 0.0036 0.603 -2.93 -10 2310 0.4083 0.1896 0.177 0.0091 0.0035 0.621 -1.47 -5 2300 0.4212 0.1915 0.179 0.0090 0.0034 0.632 -0.49 -2 2300 0.4298 0.1924 0.18 0.0090 0.0034 0.640 0.00 0 2300 0.4384 0.1933 0.181 0.0090 0.0034 0.647 0.49 2 2300 0.4513 0.1942 0.182 0.0089 0.0033 0.658 0.96 5 2300 0.4728 0.1971 0.185 0.0089 0.0032 0.677 2.44 10 2290 0.5158 0.2008 0.189 0.0087 0.0031 0.713 4.37 20 2290 0.5587 0.2056 0.194 0.0086 0.0030 0.751 6.86 30 2280
10 10

% Change in Parameter

. .e, th r .e, OH, th


r r
e, H2O, th

T h

s, th

c, th

-10 0 15 % Change in X 30
H2O

-10

Fig.13: Relative change in various parameters at the nozzle throat due to variations in mole fraction of H2O

A summary of the effect of variations of OH mole fraction on various quantities at the nozzle throat location is shown in Table 6. The relative change in the net erosion rate and species-specific erosion rates at the nozzle throat as a function of relative change in OH mole fraction are shown in Fig. 14. The value of reOH ,th increases with increasing OH mole fraction. However, the value of reH O ,th decreases only very slightly. The surface temperature at the nozzle
2

throat also does not change by any appreciable amount due to the change in OH mole fraction. This effect can be explained as follows. The heterogeneous reaction of graphite and OH is also endothermic but the mole fraction of OH is a very small quantity to significantly affect the surface temperature. A nearly constant surface temperature and constant species concentration of H2O result in nearly constant reaction rate between graphite and H2O. The overall erosion rate at the throat does not show significant variation due to change in mole fraction of OH as shown in Table 6 and Fig. 14.

13

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

X OH

0.0064 0.0072 0.0076 0.0078 0.0079 0.0080 0.0082 0.0088 0.0096

Table 6: Summary of effect of OH concentration on various parameters at nozzle throat Ts,th red ,th re,th % X OH % re,th re j ,th [ mm/s ] [K] [ mm/s] H2O [ mm/s] OH CO2 0.1906 0.1800 0.00721 0.00340 0.64 -0.93 -20% 2300 0.1915 0.1800 0.00811 0.00339 0.64 -0.46 -10% 2300 0.1920 0.1800 0.00856 0.00339 0.64 -0.23 -5% 2300 0.1922 0.1800 0.00883 0.00338 0.64 -0.09 -2% 2300 0.1923 0.1800 0.00892 0.00338 0.64 -0.05 -1% 2300 0.1924 0.1800 0.00901 0.00338 0.64 0.00 0 2300 0.1926 0.1800 0.00919 0.00338 0.64 0.09 2% 2300 0.1928 0.1800 0.00946 0.00338 0.64 0.24 10% 2300 0.1932 0.1790 0.01080 0.00336 0.64 0.40 20% 2300
20 15 % Change in Parameter 10 5 0 -5 -10 -15 -20 0 5 10 % Change in X 15
OH

20 15 10 5 0

r r
.

e, th e, OH, th

r
.

e, H2O, th

-5 -10 -15 -20 20

re, CO , th
2

Fig.14: Relative change in various parameters at the nozzle throat due to variations in mole fraction of OH

The relationship between the overall dimensionless nozzle throat erosion rate and species mole fraction is shown in Figs. 15 and 16 for H2O and OH, respectively. On these plots, the correlation between species-specific nozzle throat erosion rate and species mole fraction is also plotted. By comparing these two plots, it can be observed that H2O affects the throat erosion rate more strongly than OH. The species-specific erosion rates show much stronger dependence on the species mole fractions in accordance with their respective rate laws shown in Table 2. Even though graphite-H2O reaction has an order of 0.5, the correlation between dimensionless erosion rate by H2O shows a power law dependence with exponent of 0.31. This is due to the fact that the surface temperature is decreasing as a result of increase in H2O mole fraction, which lowers the exponent. Similarly, the graphite-OH reaction has an order of 1.0, the correlation between dimensionless erosion rate due to OH shows a power law dependence with power 0.95 due to the same reason. The correlation curve for overall dimensionless throat erosion rate shows similar exponent to that of the dimensionless throat erosion rate by H2O in Fig. 15. On the contrary, it shows a very low exponent for OH mole fraction in Fig. 16. This difference demonstrates that H2O concentration can affect the throat erosion rate significantly whereas OH does not.

14

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

. r . r

1 /G
avg

0.05
r

c e, th

c e,OH,th

/G

avg

=3.95X

0.95 OH

or cre, H O, th/Gavg

c e, H2O, th

/G

avg

0.88 0.04

0.9

/G

c e, th

/G

avg

=1.073X

H2O

avg

0.9
0.27

r .

c e, OH, th

c e, th

0.84

0.03
/G
avg

/G

c e, th

=1.01X

0.03 OH

avg

avg

/G

0.8
r

0.8 .
/G =1.036X
0.31 H 2O

c e, th

c e, H2O, th

avg

. r /G c e, th, OH avg . r /G
c e, th avg

0.02

0.7 0.3

0.7 0.4 X
H 2O

0.5

0.8 0.006

0.007

0.008 X
OH

0.009

0.01 0.01

Fig.15: Relationship between the dimensionless throat erosion rates and the mole fraction of H2O

Fig.16: Relationship between the dimensionless throat erosion rates and the mole fraction of OH

D. Effect of Thermal Properties of Graphite Another interesting parameter that may affect the nozzle throat erosion rate is the thermal properties like thermal conductivity and heat capacity of the graphite. These properties were varied by 30% and the results are shown in Table 7. These properties do not affect the quantities such as heat-transfer coefficient and boundary-layer thickness and therefore these quantities are not shown in this table. Both of these parameters affect the surface temperature at the nozzle throat, which affects the reaction rate of graphite with the oxidizing species. Table 7: Summary of effect of thermal properties of graphite on various parameters at nozzle throat c % Parameter Varied Cc% re,th % re,th % Ts,th Ts,th re,th re,th [K] [K] [ mm/s] [ mm/s] % Change 30 0.185 2290 -3.14 0.172 2280 -9.95 20 0.187 2300 -2.09 0.178 2280 -6.81 10 0.189 2300 -1.05 0.184 2290 -3.66 5 0.190 2300 -0.52 0.187 2300 -2.09 0 0.191 2300 0.00 0.191 2300 0.00 -5 0.192 2300 0.52 0.194 2310 1.57 -10 0.193 2300 1.05 0.198 2310 3.66 -20 0.196 2310 2.62 0.206 2320 7.85 -30 0.200 2310 4.71 0.215 2330 12.57

The relative change in the overall nozzle throat erosion rate as a function of relative change in these parameters is shown in Fig. 17. It can be observed that nozzle throat erosion rates decreases as either thermal conductivity or heat capacity is increased, since the increase of either of these parameters can result in decreasing surface temperature. The heat capacity affects the nozzle throat erosion rate more strongly than the thermal conductivity. The variation of overall erosion rate and surface temperature at the throat with relative change in either of these parameters is shown in Fig. 18. The change of overall erosion rate is solely due to the change in surface temperature.

15

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

0 Relative Change in Erosion Rate (%)


[K]

2400 2380 2360 2340 Themal Conductivity Heat Capacity

0.22

s,th

Overall Erosion Rate, r

Surface Temperature at Throat, T

0.21

-5

0.20 2320 2300 2280 2260 0.18

e, th

-10 Thermal Conductivity Heat Capacity -15 0 5 10 15 20 25 30 35 Relative Change in Parameter (%)

[mm/s]

0.19

-30

-25

-20
%
c

-15
or

-10
%C
c

-5

Fig.17: Relative change in overall erosion rate at the nozzle throat due to graphite properties

Fig.18: Variation of surface temperature and overall erosion rate at the nozzle throat due to graphite properties (Symbols = Surface temperature at throat, Lines = overall erosion rate)

E. Comparison with Experimental Data A comparison of numerical results with the experimental data obtained by Evans et al.5 is shown in Table 8. The parameters (stagnation temperature, species mole fraction) for these simulations were obtained via CEA calculations for a given chamber pressure. The comparison between measured and predicted erosion rates show a match within 20% error, except the measured erosion rate Pc,avg= 5.41 MPa. Table 8: Summary of effect of thermal properties of graphite on various parameters at nozzle throat Mole Fraction Pc,avg Tst,c Tth,s re,th re,th ( exp ) [K] [MPa] [K] [ mm/s] [ mm/s] H2O CO2 OH 5.41 2798 0.4298 0.1329 0.0080 0.162 0.087 2280 6.41 2798 0.4297 0.1329 0.0080 0.191 0.179 2300 6.44 2791 0.4290 0.1324 0.0084 0.192 0.162 2300 15.79 2831 0.4333 0.1352 0.0059 0.280 0.236 2310 16.28 2831 0.4333 0.1352 0.0059 0.320 0.261 2310

A summary of the effects of all of the above parameters on the percentage change in overall nozzle throat erosion rate is presented in Fig. 19. From this plot, it can be seen that the stagnation temperature is the most dominant parameter followed by pressure. Oxidizing species mole fractions and the thermal properties of graphite also affect nozzle throat erosion rates but not to the same extent as temperature and pressure. The effect of these parameters on the percentage change in nozzle throat surface temperature is shown in Fig. 20. Again, stagnation temperature affects the surface temperature most whereas all other quantities affect the surface temperature only slightly.

16

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

c H2 O OH

80

X X

c c

80
10
s,th

c H2 O OH

X X

c c

10

% Change in r

40

40

% Change in T

e,th

-40 0 5 10 % Change in Parameter

-40
15

-10 0 5 10 % Change in Parameter

-10 15

Fig.19: Summary of the effects of various parameters on the relative change in overall erosion rate at the throat

Fig.20: Summary of the effects of various parameters on the relative change in surface temperature at the throat

IV. Conclusions The objective of this work was to assess a range of parameters that affect the nozzle erosion rate by the combustion products of non-metalized propellants and to establish how these parameters affect the nozzle erosion rate. The major findings are listed below: 1. Stagnation temperature in the rocket motor has dominant effect on the nozzle throat erosion rate and its surface temperature. Chamber pressure has the second most dominant effect. 2. For non-metalized propellants, the nozzle erosion rate is more kinetically-dominated than the mass diffusion of oxidizers to the nozzle surface. Among the oxidizing species, water vapor is the most important oxidizing species, even though the OH may have higher heterogeneous reaction rate with graphite at the same temperature. The influence of OH on the overall throat erosion rate is small due to its low concentration in the combustion products. Despite of vast differences in mole fractions, OH has a greater influence on the erosion rate than CO2. 3. The correlation between the Reynolds number at the throat and dimensionless overall erosion rate show nearly linear dependency with the dimensionless erosion rate decreasing with increasing Reynolds number (when the mass erosion rate is non-dimensionalized by the mass flux at the nozzle throat). The dimensional overall erosion rate of the nozzle throat is almost linearly proportional to the Reynolds number at the throat for non-metalized propellants considered in this study. 4. The graphite properties (thermal conductivity and heat capacity) can also influence the erosion rate via the change in the surface temperature at the nozzle throat. Although, their influence is not as strong as the operating conditions. Acknowledgments This research investigation was conducted under the sponsorship of ONR as a part of multi-university research initiative (MURI) project funded under the Contract N00014-04-1-0683. The authors would like to acknowledge the interest, support, and encouragement of Dr. Clifford Bedford of ONR and the former program manager Dr. Judah Goldwasser of ONR. We would also like to acknowledge the suggestions made by Mr. Daniel O. Miller, formerly working at NAWC-China Lake for his input on the selection of the baseline Propellant S and the G90 nozzle material for this research. The experimental data was obtained through the work of Dr. Brian Evans, Mr. Drew Cortopassi who are graduate students at High Pressure Combustion Laboratory at Penn State University.

17

46th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit 25-28 July 2010, Nashville, TN

AIAA-2010-6846

References
Geisler,R.L., The Prediction of Graphite Rocket Nozzle Recession Rates, CPIA publication 342, May 1981, pp. 173-196. 2 Klager, K., The Interaction of the Efflux of the Solid Propellants with Nozzle Materials, Propellants and Explosives, Vol. 2, 1977, pp 55-63. 3 Dirling, R. B., Jr., Heightland, C. N., Loomis, W. C., Ellis, R. A., and Kearney, A. J., Private Communication, 1983. 4 Evans, B., Kuo, K., Boyd, E., and Cortopassi, A., Comparison of Nozzle Throat Erosion Behavior in a Solid-Propellant Rocket Motor and a Simulator, presented at 45th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Denver, Colorado, 2009, AIAA-2009-5421. 5 Evans, B., Kuo, K., and Cortopassi, A., Characterization of Nozzle Erosion Behavior under Rocket Motor Operating Conditions, presented at 8th International Symposium on Special Topics in Chemical Propulsion, Cape Town, South Africa, November 2-6, 2009. Accepted for publication in International Journal of Advancements in Energetic Materials and Chemical Propulsion, 2010. 6 Delaney, L. J. Eagleton, L. C., Jones, W. H., A Semi-quantitative Prediction of the Erosion of Graphite Nozzle Inserts," AIAA Journal, Vol. 2, No 8, August 1964, pp. 1428-1433. 7 McDonald, A. J., Hedman, P.O. Erosion of Graphite in Solid Propellant Combustion Gases and Effects on Heat Transfer, AIAA Journal, Vol 3, No 7, July 1965, pp. 1250-1257. 8 Gowariker, V. R., Mechanical and Chemical Contributions to the Erosion Rates of Graphite Throats in Rocket Motor Nozzles, J. of Spacecraft, Vol. 3, No. 10, 1966, pp. 1490-1494. 9 Mayberry, J. L., Kordig, J. W., Zeamer, R. L., Browning, S. C., Correlation of Graphite Nozzle Throat Erosion in SolidRocket Motors, AIAA Journal, Vol 6, No 11, 1968, pp. 2222-1257. 10 Borie, V., Brulard, J., and Lengelle, G., Aerothermochemical Analysis of Carbon-Carbon Nozzle Regression in SolidPropellant Rocket Motors, J. of Propulsion, Vol 5, No 6, Nov.-Dec 1989, pp. 665-673. 11 Keswani, S. T., Andiroglu, E., Campbell, J. D., and Kuo, K. K., "Recession behavior of graphitic nozzles in simulated rocket motors," J. of Spacecraft and Rockets, Vol. 22, No. 4, July Aug., 1985, pp. 396-397. 12 Libby, P. A., and Blake, T. R., Burning carbon particles in the presence of water vapor, Combustion and Flame, Vol.41, 1981, pp.123-147. 13 Golovina, E. C., The gasification of carbon by carbon dioxide at high temperatures and pressures, Carbon, Vol.18, 1980, pp. 197-201. 14 Acharya, R. and Kuo, K.K., Effect of Pressure & Propellant Composition on Graphite Rocket Nozzle Erosion Rate, J. of Propulsion and Power, Vol. 23, No. 6, 2007, pp.12421254. 15 Acharya, R. and Kuo, K.K., Graphite Rocket Nozzle Erosion Rate Reduction by Boundary-Layer Control Using Ablative Materials at High Pressures, International Journal of Advancements in Energetic Materials and Chemical Propulsion, Vol. 7, No. 4, 2008, pp. 402-412. 16 Acharya, R. and Kuo, K.K., Effect of Reaction Kinetic Schemes on Graphite Rocket Nozzle Erosion Rates, presented at the 8th-International Symposium in Special Topics in Chemical Propulsion, 2009, Cape-town, South Africa. In press in International Journal of Advancements in Energetic Materials and Chemical Propulsion, 2010. 17 Thakre, P., and Yang, V., Chemical Erosion of CarbonCarbon/Graphite Nozzles in Solid-Propellant Rocket Motors, Journal of Propulsion and Power, Vol.24, No. 4, 2008, pp.822-834. 18 Bianchi, D., Nasuti, F., Martelli, E., Coupled Analysis of Flow and Surface Ablation in CarbonCarbon Rocket Nozzles, Journal of Spacecraft and Rockets, Vol. 46, No. 3, 2009, pp. 492-500. 19 van Driest, E. R., On the Turbulent Flow Near a Wall, Journal of the Aeronautical Sciences, Vol. 23, 11, 1956, pp.1007. 20 Cebeci, T., and Chang, K. C., Calculations of Incompressible Rough-wall Boundary Layer Flows, AIAA Journal, Vol. 16, No. 7, pp. 730-735, 1978. 21 Patankar, S. V., and Spalding, D. B., Heat and Mass Transfer in Boundary Layers, Inter-text Books, London, 1970. 22 Mathur, S., and Saxena, S. C., Viscosity of Polar Gas Mixtures: Wilkes Method, Applied Science and Research, 15, Section A, 1965, pp. 404-410. 23 Bradley, D., Dixon-Lewis, G., Habik, S.E., and Mushi, E.M.J., The Oxidation of Graphite Powders in Flame Reaction Zone, Proceedings of 20th Symposium (International) on Combustion, Vol. 20, 1984, pp. 931-940. 24 Neoh, K.D., Howard, J.D., and Sarofim, A.F., Soot Oxidation in Flames, In D.C. Siegla and G.W. Smith, 1981, Eds., Particulate Carbon Formation during Combustion, New York, Plenum Press. 25 Bonnetain, L. and Hoynant, G., Les Carbones, (A. Pacault, Ed.) 2, 1965, Masson et Cie, Paris 270277. 26 Culbertson, B., and Brezinsky, K., (2009) High-Pressure Shock Tube Studies on Carbon Oxidation Reactions with Carbon Dioxide and Water, Energy and Fuels, 23(12) pp. 58065812.
1

18

Vous aimerez peut-être aussi