Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Defence Technology 29 (2023) 181e192

Contents lists available at ScienceDirect

Defence Technology
journal homepage: www.keaipublishing.com/en/journals/defence-technology

Measurement of alumina film induced ablation of internal insulator in


solid rocket environment
Ji-Yeul Bae*, In Sik Hwang, Yoongoo Kang
Agency for Defense Development, 160 488th Bukyuseong-street, Daejeon, South Korea

a r t i c l e i n f o a b s t r a c t

Article history: This study investigates the ablation of internal insulation induced by the deposition of an alumina film
Received 30 November 2022 with different lateral film speeds. A sub-scale test solid rocket motor (SRM) was designed in an
Received in revised form impinging jet configuration to form an alumina film on the sample and to encourage the lateral
28 February 2023
movement of the film by a high-speed wall jet. Fifteen static fire tests of the test SRM were conducted
Accepted 9 March 2023
with six different jet velocities (Vjet ¼ 100 m/s, 150 m/s, 200 m/s, 268 m/s, 330 m/s, and 450 m/s) that
Available online 16 March 2023
indirectly affected the velocity of the wall jet and the deposition rate of alumina droplets. The ablation
velocity was deduced from the difference in the sample thickness after a test using a coordinate
Keywords:
Internal insulation
measuring machine. The droplet deposition mass flux and wall jet velocity were obtained via two-phase
Solid rocket motor flow simulation with the same jet velocity and effective pressure. As a result, the characteristics of
Ablation alumina-induced ablation and the changes in ablation with jet velocities were obtained. The area within
Alumina 0.8  jet diameter was focused upon, where the ratio of ablation velocity to incoming alumina mass was
constant for each jet velocity, and showed a similarity in jet structure. When the ablation velocity was
increased from 2.05 to 9.98 mm/s with increasing jet velocity, the ratio of the ablation velocity and
alumina mass flux decreased from 1.07104 to 0.49104 m3/kg as Al2O3eC reactions became less
efficient with a reduced residence time of the film. Because the decrease in residence time by the wall jet
is more pronounced for slow reactions involved in Al2O3eC reactions, fast reactions in Al2O3eC reactions
are less affected and result in a convergence of the volumetric rate of ablation per unit mass of alumina.
© 2023 China Ordnance Society. Publishing services by Elsevier B.V. on behalf of KeAi Communications
Co. Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/
licenses/by-nc-nd/4.0/).

1. Introduction illustrated in Fig. 1. Once the insulator is exposed to the combustion


gas, the temperature of the material increases owing to convective
In a solid rocket motor (SRM), an internal insulator is placed and radiative heat transfer. When it reaches its decomposition
between the case and propellant, whose role is to transmit the temperature, the molecular structures in the virgin material
pressure load from the combustion gas to the case while main- decompose owing to endothermic chemical reactions (charring).
taining the case temperature at an acceptable level for structural Eventually, the material is converted into a porous char layer and
integrity [1]. To achieve these requirements, elastic materials such pyrolysis gases. As the heating process continues, the increasing
as nitrile butadiene rubber (NBR) or ethylene propylene diene thickness of the char layer reduces the heat transfer into the virgin
monomer (EPDM) reinforced with aramid or carbon fibers are material. In addition, pyrolysis gases cool the char layer and reduce
commonly used as internal insulators [2]. As the insulator un- the heat and mass transfer on the char surface as they flow through
dergoes a complex destruction process upon exposure to the the char layer and into the boundary layer of the combustion gas.
combustion gas, numerous studies have been conducted to un- However, the ablation of the char layer limits the heat-repelling
derstand the destruction characteristics. effect of the material. One of the main reasons for ablation is the
The general destruction process of an internal insulator is chemical reaction between the carbonaceous char and oxidizing
species such as H2O and CO2 in the combustion gas (chemical
ablation). In addition, the mechanical strength of a highly porous
* Corresponding author. char layer can be extremely low; thus, the char layer can get me-
E-mail address: jybae@add.re.kr (J.-Y. Bae). chanically removed by shear stress from the flow (mechanical
Peer review under responsibility of China Ordnance Society

https://doi.org/10.1016/j.dt.2023.03.007
2214-9147/© 2023 China Ordnance Society. Publishing services by Elsevier B.V. on behalf of KeAi Communications Co. Ltd. This is an open access article under the CC BY-NC-
ND license (http://creativecommons.org/licenses/by-nc-nd/4.0/).
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

Fig. 1. General destruction mechanism for an internal insulator.

ablation). Finally, the alumina (Al2O3) droplets included in the alumina and carbon by investigating the mass loss of a graphite
combustion gas can also cause mechanical and thermochemical sample in contact with molten alumina at elevated temperatures.
damage to the char surface when they are deposited on it (alumina- In their experiments, the consumed mass ratio between carbon and
induced ablation). Therefore, the rate of thermal decomposition alumina was approximately 0.2 with the following carbothermic
and each ablation mode of an internal insulator should be inves- reaction:
tigated under various environments before it can be applied to an
actual SRM.
In the early phase of material development, the material is Al2 O3 þ 3C / 2Al þ 3CO (1)
heated with lasers [3,4] or torches [5,6] to study its thermal It is important to note that the stoichiometric mass ratio of the
destruction characteristics and to screen candidate materials. suggested global reaction was found to be approximately 0.35,
However, the chemical/thermal characteristics of actual SRM con- instead of 0.2. They attributed this discrepancy in the mass ratio to
ditions cannot be closely replicated in these experiments, whose the removal of intermediate species, such as Al2O, along with CO
use for studies on chemical and mechanical ablations is limited. bubbles moving away from the char surface during the reaction.
Therefore, sub-scale SRMs that develop a flow parallel to the ma- Guan et al. [26] evaluated the ablation rate in the recirculation
terial surface have been utilized for further investigations [7e12]. zone of a test motor, where alumina droplets were deposited to
These test motors are typically designed to maintain a constant form a film with minimal mechanical impact. Their results showed
flow velocity on the material surface to control the ablation rate for that the ablation rate can be as high as 0.3 mm/s only through
a given test duration. The resulting destruction of the material can chemical reactions of the AleOeC system between alumina and the
be measured by comparing the thickness of the sample before and carbonaceous char layer. In the following research, they analyzed
after the test [7e11] or through in-situ measurements based on in- chemical kinetics and reaction mechanism of Al2O3eC reaction by
depth thermocouples [11,13] or real-time X-rays [12]. And results DT-TGA and summarized reaction rates for carbothermic reaction
from these tests have been widely utilized for extensive theoretical/ [29]. Finally, based on the measured chemical kinetics of Al2O3eC
numerical studies to predict a thermal response and destruction of reaction and heat flux from the slag pool [30], they predicted a
the material since 1960s [14e22]. In these research, decomposition thermal destruction and ablation of the EPDM insulator under the
of matrix usually has been modeled based on the Arrhenius thermochemical influence of the alumina pool deposited on the
equation and chemical kinetics and diffusion of the oxidizing spe- insulator surface [31]. However, thermal effect of the Al2O3eC re-
cies on the surface have been considered for thermochemical action was not considered in their research. Since the thermal effect
ablation of the char layer. Mechanical ablation of the char layer was of the reaction (1) reaches 1310.2 kJ [32], the interfacial tempera-
modeled as proportional to the shear stress from the gas flow [21] ture between alumina and char layer would be decreased as the
or a concept of fail temperature [18] was utilized. Recently, a model reaction proceed, which then reduces the reaction rate itself. This
for a formation of loose/compact char structure by the coking of the behavior can be inferred from the predicted ablation rate that
pyrolysis gas was developed [21]. initially increased to 0.48 mm/s, then decreased to the converged
For analyzing alumina-induced ablation [23e26], test motors value around 0.38 mm/s [31].
were designed to deliberately provoke the impact of alumina Although the measured ablation rate would be valid if the re-
droplets on the material. These studies mostly focused on the action occurs beneath a stationary alumina pool, the ablation rate
mechanical effect of the droplet impact, and high-speed droplets may become different when the deposited alumina film moves
were introduced to the material surface with a controlled impact inside the motor as usually observed at an aft dome and backward
angle [23e25]. Then, the relationships between the impulse of side of the cylindrical section in a motor case [33e36]. When the
droplets, char layer strength, and ablation rate were investigated alumina film moves along the insulator surface due to a shear of the
based on Wirzberger et al.’s approach [27]. However, in actual SRM gas flow or a body force (gravity or acceleration), the reaction rate
environments with a high droplet mass flux, droplets can aggregate of the Al2O3eC would be changed by both supply of high temper-
on the surface to form liquid alumina film or pool. ature alumina with flow and removal of intermediate species like
Under such circumstances, the mechanical effect of the droplets Al2O, Al2OC and Al4C3. Therefore, the relationship between the rate
is absorbed by the alumina film, and alumina-induced ablation is of alumina-induced ablation and movement of the alumina film
governed by the thermal/chemical interactions between the film should be investigated to accurately predict the behavior of the
and the surface. However, very few studies have investigated the internal insulator.
chemical reaction between the liquid alumina film and the internal In this study, a sub-scale test SRM with an impinging jet
insulator. Gubertov et al. [28] studied the reaction mechanism of configuration was designed and test fired 15 times to study the
182
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

ablation of a Kevlar-reinforced EPDM insulator induced by a designed to withstand a maximum test pressure of 6.9 MPa (1000
deposited alumina film with lateral movement. A jet with an psia). Most of the internal parts, such as the jet nozzle and blast
alumina droplet was introduced with a high mass flux for all the tubes, were made of chop-molded or tape-wrapped carbon/
tests to ensure the formation of the alumina film with high tem- phenolic materials because of the low Mach number inside the test
perature. The design velocity of the jet was varied from 100 to chamber. However, the throat inserts of the main nozzles were
450 m/s to change the lateral velocity of the alumina film move- made of copper-infiltrated tungsten to prevent the ablation of the
ment along with the velocity of the wall jet. The thickness of the nozzle, which can result in a variation in exhaust pressure. Two bag
Kevlar/EPDM sample was measured using a coordinate measuring igniters with BKNO3 powder were attached near the entrance of the
machine (CMM) before and after each test to deduce the ablation jet nozzle to evenly ignite the propellant. The igniter wires were
thickness and ablation rate. In addition, a two-phase flow simula- connected through a hole in the two nozzle closures made of ure-
tion for each static fire test was carried out to characterize the flow thane foam.
inside the test SRM and to deduce the effective conditions on the The condition of each static fire test was determined based on
sample surface, including the droplet deposition mass flux, velocity the key dimensions of the test SRM. First, the average pressure and
of the wall jet and wall shear stress. Finally, the measured ablation mass flow rate of the combustion gas during a test were deter-
rate was analyzed along with the mass of incoming alumina mined by the diameters of the propellant grain, Dgrain, and the main
droplets and the movement speed of the deposited film. nozzle, Dnoz. In addition, the ratio of the exit diameter of the jet
nozzle, Djet, to the effective exit diameter of the two main nozzles,
2. Experimental apparatus and procedure Dnoz,eff, controls the design velocity of the jet, Vjet. The incoming
droplet mass flux also changes as the jet is concentrated in a
2.1. Sub-scale test SRM smaller area when Djet decreases. Finally, the distance between the
jet and the test sample, Hjet, affects the jet geometry because it is
The sub-scale test SRM designed for this research has an related to the curvature of the flow and pressure gradient at the
impinging jet configuration, as shown in Fig. 2. An end-burning stagnation point.
propellant grain was used to ensure a uniform distribution of
alumina droplets. When the propellant is ignited, the combustion 2.2. Conditions of static fire tests
gas, including alumina droplets, is accelerated through the jet
nozzle and impinges on the surface of the test sample. Unlike other In this study, 14 static fire tests with different jet velocities and
test apparatuses [23e26], high mass flux conditions were imposed pressure were conducted to observe alumina-film-induced ablation
at the exit of the jet nozzle to encourage the formation of a liquid of Kevlar/EPDM with different droplet mass fluxes and wall jet
film on the char surface and prevent direct impact of the droplet. velocities. A single test was conducted with a copper test sample to
The jet was located close to the test sample to form a strong capture and visualize the deposited alumina film on the sample
pressure gradient at the stagnation point, which induces the for- surface. For each test, an HTPB/AP/Al propellant composed of 69%
mation of a high-speed wall jet [37]. The wall jet causes the flow of AP, 19% aluminum, 7.3% HTPB and other additive (without nitr-
the liquid film away from the stagnation point with its shear stress, amine) was applied, and Dgrain and Dnoz were determined to ach-
thereby affecting the rate of the Al2O3eC reactions. After the ieve a design pressure of approximately 2.06 MPa (300 psia) and
impingement, small droplets and the combustion gas are dis- 6.89 MPa (1000 psia). Djet was reduced from 36.0 to 17.5 mm to
charged through two main nozzles, while large droplets are increase Vjet from 100 to 450 m/s and the converging half angle of
deposited and remain on the test sample in the form of a liquid film. the jet nozzle was 10 in all cases. The Hjet/Djet value for each case
To prevent large droplets and the liquid film from entering the was held constant at 0.9 to maintain the similarity of the jet
nozzle, each nozzle was equipped with a blast tube that extended structure in all tests. For each test, the dimensions of the jet nozzle
the position of the inlet higher than the surface of the test sample, were changed according to the design jet velocity and all other
thus reducing the pressure oscillation and clogging of the nozzle components such as chamber insulation, test sample, blast tubes,
during the static fire test. and main nozzles were identically manufactured. The amount of
The casing of the SRM was made of AISI 4130 steel and was propellant for each test was adjusted according to the anticipated
thickness of alumina-induced ablation and increase in Hjet/Djet to
avoid an excessive change in the jet geometry. For a test with a
copper sample, the minimum amount of propellant and lowest Vjet
were chosen to reduce the loss of alumina film by the wall jet.
During each test, the chamber pressure was measured at two
pressure holes on the SRM casing located close to the burning
surface of the propellant. Hydraulic oil was filled between the
pressure holes and DYNISCO G831-101 pressure sensors to protect
the sensors from the high-temperature environment. The sampling
rate of measurement was 1000 Hz and the measurement accuracy
of the sensor was 0.25%. A typical history of the measured pressure
is shown in Fig. 3. The first peak originated from burnout of the
igniter, and a slight increase in pressure continued until 2.4 s due to
coning of the end-burning grain. The pressure was then decayed
back to the atmospheric pressure. To deduce a representative
condition for each test, an effective time, teff, is defined as the time
between the igniter peak and the moment when a pressure be-
comes 10% of the maximum pressure, as shown in Fig. 3. The
effective pressure, Peff, is defined as the average pressure during the
effective time.
Fig. 2. Schematic view of sub-scale test SRM with key components and dimensions. Table 1 lists the design conditions and values of Peff and teff for
183
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

theoretical approach or experimental measurement as these values


are related to a two-phase flow under thermally harsh conditions.
Thus, in this study, a numerical simulation was utilized to deduce
the conditions on the sample surface necessary to analyze the
experimentally measured ablation of the sample. Since the depo-
sition of droplet is mostly affected by the freestream flow structure
and droplet/wall interaction, complicated interactions between the
flow and surfaces of the test SRM including reactive/blowing wall
condition, gas/particle radiation were not considered. Instead,
steady state solutions of combustion gas flow coupled with alumina
droplet trajectories at average test conditions were obtained to
deduce a flow structure and average droplet deposition rates.

2.3.1. Numerical setup for two-phase flow


To predict the conditions on the sample surface, we calculated
the two-phase flow inside the test SRM for each condition. Because
of the high mass fraction and low volume fraction of alumina
droplets in the combustion gas, gas-phase conservation equations
and droplet trajectories were solved based on the Euler-Lagrange
approach using ANSYS Fluent. For the gas phase, compressible
three-dimensional RANS equation was solved using a two-equation
Fig. 3. Typical pressure curve of test SRM and definition of effective test condition.
SST k-u model. A second-order upwind scheme was used for all the
conservation equations. For the discrete phase, the velocity and
position of each droplet in the Lagrangian frame were deduced by
each test in this study. For tests #1, #2, #3, #4, #9, and #10, the
duration of the test was intentionally varied to determine the effect stepwise integration of a discretized particle force balance equation
of the exposure time of the sample. Thus, the Peff of the test with the based on a trapezoidal rule [38]. The drag from the gas phase was
same geometries and same Vjet also varied. Peff for these tests be- assumed to be the only force acting on the droplet, and the drag
comes higher with longer teff owing to a longer period with con- coefficient of the droplet was taken from the work of Gilbert et al.
stant pressure in the pressure curve. An unintentional variation in [39], shown below.
Peff was also observed in some tests, with an elongated tail-off ( 
24 Rep ; 0 < Rep  0:34
pressure caused by the conning of the end-burning propellant. In CD ¼ (2)
such cases, Peff decreased with increasing teff because the amount of 0:48 þ 28Re0:85
p ; 0:34 < Rep
propellant was fixed.
Even though there are complicated drag models that consider
the effect of the Mach number between the gas phase and droplets
2.3. Conditions on sample surface [40], the above model of the spherical particle is sufficient for this
study because most of the area inside the test SRM, including the
To determine the relationship between the characteristics of the test chamber, is in a subsonic condition. Collision, breakup and
liquid alumina film and ablation of the test sample, the deposition turbulent dispersion of the droplet were neglected as most of
rate of alumina droplets on the sample surface should be derived droplets were rapidly accelerated thorough the converging jet
first. Not all droplets in the exhaust gas would be deposited on the nozzle into the sample surface and would not be greatly affected by
sample because small alumina droplets with low inertia would be those phenomena. After calculating each droplet trajectory, the
pushed away by the pressure gradient at the stagnation point and result of momentum/energy exchanges between the gas phase and
may not reach the surface. In addition, the speed of the alumina droplets was considered as the source term in the gas-phase
film movement on the surface should be determined to correlate it equations to accomplish the global conservation of momentum
with the Al2O3eC reaction between the alumina film and the test and energy.
sample. However, it is difficult to obtain these conditions using a

Table 1
Dimensions and effective conditions for static fire tests.

Test Djet/mm Hjet/mm Vjet/(m$s1) Design test duration/s Teff/s Peff/MPa Sample

1 36.0 32.4 100 1.0 1.21 2.055 Kevlar/EPDM


2 36.0 32.4 100 3.0 3.04 2.504 Kevlar/EPDM
3 29.0 26.1 150 3.0 2.94 2.549 Kevlar/EPDM
4 29.0 26.1 150 1.0 1.14 2.373 Kevlar/EPDM
5 25.0 22.5 200 1.0 1.19 2.337 Kevlar/EPDM
6 25.0 22.5 200 1.0 1.25 2.006 Kevlar/EPDM
7 22.0 19.8 268 1.0 1.09 2.519 Kevlar/EPDM
8 22.0 19.8 268 1.0 1.24 2.102 Kevlar/EPDM
9 20.0 18.0 330 3.0 2.95 2.664 Kevlar/EPDM
10 20.0 18.0 330 1.0 1.23 2.242 Kevlar/EPDM
11 17.5 15.8 450 0.75 0.76 2.418 Kevlar/EPDM
12 17.5 15.8 450 0.75 0.71 2.884 Kevlar/EPDM
13 21.0 18.9 150 1.0 1.50 4.922 Kevlar/EPDM
14 21.0 18.9 150 1.0 1.51 5.182 Kevlar/EPDM
15 36.0 32.4 100 0.25 0.34 2.452 Copper

184
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

When a droplet reaches the sample surface, an appropriate wall median diameter and mass fraction were released from the pro-
interaction model should be applied to predict the deposition and pellant surface.
post-collision behavior of the droplet. Numerous studies have been Fig. 4 shows the computational grid and applied boundary
conducted on the droplet-wall interaction model, including the conditions for the test apparatus with a Vjet value of 200 m/s
works of Mundo et al. where the Sommerfeld number was used to (Djet ¼ 25.0 mm). Because the adoption of a symmetry condition
determine the fate of a droplet being trapped or splashed [41,42], could distort the droplet distribution at the jet axis, a full volume of
and the study by Wang et al. of a deposition probability criterion fluid inside the test SRM was modeled. An unstructured tetrahe-
based on the Weber number [43]. However, the effect of the dral/prism mesh with approximately 9 million cells was generated
deposited film on the post-collision behavior of the droplet was not for each jet velocity. The average first cell height on the sample
considered in these models. Therefore, the model used for this surface within Djet was adjusted from 3.1106 to 9.3106 m
research is the Stanton-Rutland wall film model adopted in ANSYS according to Vjet to ensure that yþ is around 1.0 in the area of the
Fluent, in which the fate of a droplet incident on a wall or a liquid interest. The spatial resolution of the mesh on the area near the
film is divided into stick, spread, splash, and rebound. Number of stagnation point within Djet was approximately 2.5104 m.
smaller drops generated by a splash was set as the default value of The inlet and outlet of the computational domain corresponded
4. For the other walls inside the test SRM, an elastic collision be- to the initial surface of the propellant and exit planes of the two
tween a droplet and wall was assumed. main nozzles. The mass flow rate was applied at the inlet for the gas
The equilibrium composition of the exhaust gas and alumina phase instead of the inlet pressure. This was done to fix the ratio of
mass fraction at Peff was calculated using the approach of Gordon the mass flow between the gas phase and droplets to the value
and McBride [44], and the properties related to the gas phase at calculated using the CEA program. However, several calculations
chamber condition are provided in Table 2. The combustion gas was with different total mass flow rates were carried out to determine
assumed as ideal gas. The properties of the droplet required for the absolute value of the mass flow that makes the simulated inlet
trajectory prediction are summarized in Table 3. The droplet pressure equal to the Peff value of each test. The deduced total mass
properties are based on the experimental results of previous flow rate was larger than the choked mass flow rate calculated
studies [45,46] and extrapolated to an exhaust temperature of using Peff with Dnoz,eff (i.e., around 20% higher mass flow rate
3442 K to consider the effect of temperature on the properties. The compared to the theoretical value at Peff to achieve same pressure).
size distribution of alumina droplets at the nozzle inlet was This discrepancy was caused by the deposition and recirculation of
deduced using an in-house program for agglomeration prediction the alumina droplet, which led to a significant loss of mass inside
that considers the packing of aluminum powder in the propellant the test SRM. For all cases, the temperatures of the gas phase and
and agglomeration with a stochastic pocket model [47,48]. As AP droplet at the inlet were assumed to be equal and set to 3442 K. The
powders with two distinct diameters were used in the propellant, initial speed of injected droplets was also assumed to be identical
alumina droplet showed a trimodal size distribution including with that of the gas phase on the propellant surface around 4.5 m/s.
smokes and agglomerates with maximum diameter around The condition at the exit of the two main nozzles was set as the
200 mm. In the deduced size distribution, the mass fraction of atmospheric pressure, as the flow was supersonic at the exit and
droplets with diameters less than 1.5 mm (smoke) was 23.4% of total not affected by the given condition.
alumina, and the mass mean diameter of the droplets was The wall temperature inside the test SRM was fixed at 2500 K for
13.67 mm. In order to apply the derived droplet distribution for the the carbon/phenolic parts and 3200 K for the EPDM test sample,
simulation, it was divided into 11 sections (i.e., Section 1 for droplet considering the thermal diffusivity of each material and previous
diameter from 0.0 to 1.0 mm and Section 2 for droplet diameter from literature on material temperature [49e51]. The Stanton-Rutland
1.0 to 2.0 mm). Then, identical droplets for each section with the wall film model was applied to the surface of the test sample and

Table 2
Exhaust gas properties for gas-phase simulation.

Temperature/K Specific heat Ratio Mole fractionof Al2O3 Molar mass/(kg$kmol1) Specific heat/(J$kg1,K) Viscosity/(Pa,s)

3442 1.122 0.0838 28.98 2035.3 9.214  105

Table 3
Droplet properties for trajectory simulation.

Density/(kg$m3) Specific heat/(J$kg1,K) Thermal conductivity/(W$m$K1) Surface tension/(N$m1)

1840.2 1888.0 7.4 0.45

Fig. 4. Domain, mesh and boundary conditions for two-phase flow simulation (Vjet ¼ 200 m/s).

185
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

jet nozzle, where the formation of the liquid film was observed in a
test SRM after a static fire test. However, the droplets colliding on
the chamber wall were assumed to be trapped on the surface to
prevent the continuous accumulation of droplets in the simulation
by a recirculating flow. An elastic collision was assumed for all other
walls, including the blast tubes and main nozzles.

2.3.2. Droplet deposition mass flux


Fig. 5 shows the distribution of two-phase flow near the test
sample for a Vjet value of 200 m/s. Both the gas phase and droplets
are accelerated through the jet nozzle and impinge on the sample
surface. The actual jet velocity at the nozzle exit is higher than the
design value, approximately 220 m/s, owing to a higher mass flow
through the jet nozzle incurred by the aforementioned loss of
alumina inside the test SRM. A large number of droplets is depos-
ited near the jet axis and form a plateau of droplet deposition mass
flux on the sample surface. Subsequently, the mass flux gradually
decreases away from the center as smaller droplets are pushed
away by the pressure gradient at the stagnation point. Droplets
with even lower inertia (diameters smaller than 5.5 mm for a jet
velocity of 200 m/s) cannot reach the sample surface because they Fig. 6. Radial distribution of droplet deposition mass flux with different jet velocity
are carried away by the gas phase. The gas-phase flow is also turned and pressure.

and pushed away by the pressure gradient and forms a radially


propagating wall jet. As the incoming jet pushes the wall jet against
distribution of qdep for each Vjet was obtained by averaging a radial
the sample, the boundary layer of the wall jet becomes very thin
strip with a width of 1 mm for each radius (i.e., the value at
and exerts a high shear stress on the deposited alumina film. At the
r ¼ 2.0 mm was average of values from 1.5 to 2.5 mm) for a
edge of the sample in the 0 section, the wall jet separates from the
smoother distribution, and shown in Fig. 6. For a low Vjet (e.g.,
sample and forms a recirculating flow that impinges on the bottom
100 m/s), droplets are spread out over a large area that extends to a
of the jet nozzle. Then, the flow circles around the test sample to-
radius of 30 mm owing to the large Djet and low momentum of the
ward the 90 section and discharges through the blast tube and
droplets. However, the jet is concentrated at the center, and qdep
main nozzle. The recirculating flow slightly reduces the speed of
increases rapidly when Vjet increases. For every Vjet, the size of the
the wall jet as it takes momentum from the wall jet while moving
plateau in the qdep distribution is larger than 0.8Djet, and this is the
around the test sample. On the contrary, the wall jet at 90 section
main area of interest for this research. For validation purposes, the
directly hits the blast tube and is split into two streams, showing a
mass of deposited alumina on the copper sample of test #15 was
different flow structure from that of 0 section. However, the effect
calculated by integrating qdep over the entire sample surface and
of this flow disturbance does not appear in the central region,
multiplying it by teff of test #15. The predicted mass of alumina was
especially in the area within Djet, owing to the high strength of the
12.4 g, which is in good agreement with the experimental result of
jet.
12.6 g gathered from the surface of the copper sample after test
For a quantitative analysis of the ablation of the test sample in
#15.
relation to alumina deposition, a local distribution of the droplet
deposition mass flux, qdep, on the sample is required for each jet
2.3.3. Wall jet speed and shear stress
velocity. However, the raw deposition mass flux showed vibratory
In this work, the velocity of the deposited alumina film on the
behavior owing to the discrete nature of the Euler-Lagrange
sample surface is one of the key parameters, as it affects the resi-
approach based on finite streams of particles. Therefore, a radial
dence time of the film and the Al2O3eC reactions. In a two-phase
flow simulation, the velocity of the liquid film on a sample sur-
face can be obtained as part of the Stanton-Rutland wall film model.
However, the film model was not calibrated for a high-temperature
alumina film in an SRM, especially considering the thermal-
echemical interactions between the film and surface. In this study,
the average wall shear stress exerted by the gas flow on a certain
area were used instead of the speed of the wall film itself to char-
acterize the film behavior. This was based on the assumption that
the radial film movement on the surface is solely driven by the
shear stress from the flow over the film.
Because the wall jet is generated by the flow turning at the
stagnation point, the shear stress of the jet is zero at the jet axis. It
then increases with a radius to the outer section owing to accel-
eration and thinning of the wall jet. Although the velocity of the
incoming jet and developing wall jet are determined by the jet
geometry and gas properties, the wall shear stress can become
different even for tests with same Vjet due to the variation of the
flow density and Reynolds number resulting from different pres-
Fig. 5. Contours of velocity and velocity vectors at 0 , 90 section of the test SRM along sure. Therefore, wall shear stress for each test were directly calcu-
with droplet deposition mass flux on the sample surface. lated from the simulation with corresponding Peff. And the average
186
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

velocity of the wall jet was taken at the surface 0.5 mm away from of 1 mm, as shown in Fig. 7(b). A hexagonal Croma 686 CMM was
the thickest point of the boundary layer as a supplementary value used with a measuring error of less than 1.3 mm in the thickness
to characterize the film movement. direction. Pre-test measurements showed that the maximum
standard deviation of thickness for the 14 samples was 0.15 mm,
and the average thickness of the samples was 40.3 mm. After the
2.4. Test sample and measurement of ablation
static fire test, thickness of the char residue was negligible, usually
less than 0.2 mm due to a fast ablation rate. Thus, char residue was
Fourteen cylindrical test samples made of Kevlar fiber-
removed from the sample surface using a steel brush before a
reinforced EPDM were identically manufactured with a diameter
measurement to prevent contamination of the measurement
of 139 mm and a height of 40 mm. The composition of the material
probe. The ablation thickness, dabl, was obtained by comparing the
is the same as that used by Kang et al. [11], which includes 30 phr. of
measured thicknesses at each point before and after the static fire
silica sphere and 8 phr. of chopped Kevlar short fiber without
test. The ablation rate Vabl was derived by dividing dabl at each point
directionality. Fire retardant was not present in the material. The
by the teff of each test. Finally, the volumetric rate of ablation per
initial material was made into a square plate by stacking rubber
unit mass of alumina, Qabl, was defined as the division of Vabl by
sheets, and the stacked plate was cured in an autoclave. The cured
qdep at each point to compare the rate of ablation between the cases
plate was then machined into a cylindrical test sample. The finished
with different Vjet and qdep. As Vabl is related to the speed of char
test sample was bonded to a stainless-steel sample holder using
consumption by the Al2O3eC reaction and qdep is related to the
Araldite AW 106 adhesive, as shown in Fig. 7(a), and fastened to the
supply of alumina mass to the surface, Qabl indicates the
aft casing of each test SRM using a bolt. For test #15, a copper test
completeness Al2O3eC reactions on the surface with a moving
sample of the same size as that of the Kevlar/EPDM test sample was
alumina film. When Qabl is multiplied by a density of char layer, it
clamped to the sample holder using four bolts. Copper was used to
indicates a mass ratio of consumed carbon and alumina and would
maintain the surface temperature of the sample to be as low as
be around 0.2 for stationary Al2O3eC reaction as reported by
possible with its high thermal mass and thermal diffusivity to so-
Gubertov et al. [28].
lidify the alumina film deposited on the sample.
The thickness of each Kevlar/EPDM sample before and after the
3. Results
static fire test was measured on every 45 line with a radial spacing
3.1. Basic characteristics of ablation

The rear view of the disassembled test SRM after test #5 is


shown in Fig. 8(a), and the aft casing of the SRM with a test sample
and two blast tubes is shown in Fig. 8(b). A trail of the separated
wall jet that circled around the test sample and discharged through
the blast tube is visible at the bottom of the jet nozzle. In particular,
the black marks on the bottom of the jet nozzle show a reattach-
ment point of the separated flow where the wall shear is low; thus,
smoke particles are not washed away. On the contrary, the alumina
film on the sample surface is washed away by the wall jet and
alumina residue is found around the sides of the test chamber. At
the 90 and 270 sections of the chamber where the blast tubes are
located, the sediment is found at the side of the blast tube instead of
the chamber because the wall jet impinges on the blast tube at
these sections. In addition, there is hardly any alumina deposited
inside the blast tubes and nozzles owing to flow turning at the
entrance of the blast tubes, as intended by the SRM design.
The ablation of the K/EPDM sample can be more clearly un-
derstood by referring to the shape of the alumina film deposited on
the copper sample. As shown in Fig. 8(c), the alumina droplets from
the incoming jet are mainly deposited at the center of the test
sample. The flow of the deposited film in the radial direction
induced by the wall jet is clearly visible in the outer region of the
sample. More importantly, the region around the jet is fully covered
with the alumina film, even in the test with the lowest Vjet and qdep,
either by direct deposition from the jet or radial flow of the film.
Because the sample surface is not exposed to the combustion gas,
the chemical and mechanical ablation of the K/EPDM sample by the
gas flow can be neglected. Thus, the ablation of the sample shown
as a pit at the center of the sample in Fig. 8(b) can be solely
attributed to alumina-induced ablation. A small amount of alumina
is visible at the center of the test sample, as most of the alumina
film is either consumed by the Al2O3eC reaction or washed away by
the wall jet. Further, almost no char layer is left on the surface
owing to fierce ablation.
Fig. 9 shows a test sample after removing the char layer and the
Fig. 7. Design of test sample and measurement locations: (a) Design of test sample and measured residual and ablation thicknesses for test #5. Before
sample holder; (b) Lines of CMM measurement sweep on sample surface. measurement by CMM, the surface of the sample was brushed to a
187
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

decomposition zone that is noticeable by the gluey melted silica in


this layer. This was done to reduce the contamination of a CMM
probe during measurement and the difference in the measured
thickness due to a char layer is negligible (less than 0.2 mm
compared to a measured ablation thickness of approximately
6.8 mm for test #5). The measured sample shape showed a flat area
at the center within a radius of approximately 6 mm. Then, the
residual thickness increases linearly with the radius to a point of
approximately 30 mm. At this point, the shear stress of the wall jet
reaches its maximum and then starts to decrease. Therefore, the
residual thickness after that point slightly decreases owing to the
increased residence time of the alumina film, along with a drop in
the wall jet shear stress.
As shown in Fig. 9(c), the maximum dabl of test #5 is 6.92 mm
near the center of the jet. The average dabl value of the flat surface
with a diameter of 12 mm was 6.81 mm and the standard deviation
of dabl in this region is only 0.06 mm. dabl linearly drops away from
this region and becomes 1.5 mm at a radius of 34 mm, and increases
slightly to 2.08 mm at a radius of 45 mm in connection with the film
movement. The dabl at the edge of the test sample is approximately
1.5 mm.
Fig. 10 shows the radial distributions of ablation rate, droplet
deposition mass flux, and volumetric rate of ablation per unit mass
of alumina for test #5. Vabl within 6 mm radius is almost flat with an
average value of 5.72 mm/s, and it is significantly higher than the
maximum ablation rate around 1.0 mm/s reported in previous
works [23,26,31]. From the measured chemical kinetics of the re-
action [31], a reactant conversion (from 0 to 1), a, can be expressed
with first-order reaction model and Arrhenius equation with pre-
exponential factor, A ¼ 5.8109 min1, and activation energy, Ea
439.5 kJ/mol, as equation below.

lnð1  aÞ ¼ kðTÞt (3)

kðTÞ ¼ AeEa ∕RT (4)

where t is time in min, R is 8.314 J/mol,K and T is absolute tem-


perature. Time required for a conversion of 90% of reactant are
calculated as 298.5 s at 2000  C and 19.5 s at 2300  C. And it cor-
responds to the ablation rate of 0.52 mm/s and 8.03 mm/s,
respectively. Therefore, the reason for the higher reaction rate is the
active Al2O3eC reaction on the surface with a continuous supply of
high-temperature alumina into the deposited film in this research.
qdep deduced from the simulation varies from 74.2 to 83.0 kg/
m2s on a plateau area within 9 mm radius and sharply drops to
nearly zero at a radius of 22 mm. As the rate of the reaction de-
creases by a faster wall jet, lower qdep and temperature drop of film
caused by the endothermic heat at the outer region, Vabl decreases
to 1.32 mm/s at a radius of 34 mm. At this point, almost no alumina
droplets are deposited from the incoming jet, and the alumina
reacting with the char layer is provided by a radial flow of the film.
As a factor considering both the deposition and reaction of
alumina, Qabl is 7.32105 m3/kg at the center of the jet, and the
deviation is only 6.6% in the region within Djet. Notably, the area
with uniform Qabl is larger than that with uniform Vabl. In this re-
gion, the incoming alumina mass and removal of intermediate
species of the reaction by the film movement are balanced, and the
rate of the Al2O3eC reaction is maintained even with a decreasing
qdep at the edge of the jet. Outside of Djet, Qabl rapidly increases to
8.42  104 m3/kg at a radius of 20 mm. However, Qabl in this region
does not have a physical meaning because most of the reacting

Fig. 8. Condition of the test SRM and test sample after static fire test: (a) Bottom side
#5 (Vjet ¼ 200 m/s); (c) Deposited alumina film on the copper test sample after test
of the jet nozzle after test #5 (Vjet ¼ 200 m/s); (b) Aft-casing and test sample after test
#15 (Vjet ¼ 100 m/s).

188
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

alumina is supplied by the radial flow of the film. In this region, the
mass flow rate of the alumina film in the radial direction should be
used for Qabl instead of qdep. However, a model of the thermal and
chemical interactions between the char layer and alumina film is
required, which is extremely complex and currently not available.
Therefore, the ablation of samples with different wall jet shear
stress was analyzed only in the region within Djet, where the
definition of Qabl with qdep is valid.

3.2. Effect of film movement

Fig. 11 shows the radial distributions of ablation rate, droplet


deposition mass flux, and volumetric rate of ablation per unit mass
of alumina with different jet velocities. In order to compare the
results with different jet sizes, each distribution is illustrated with
R/Djet, normalized radius with Djet of each test. First, Vabl at the
center of the jet changes from 2.41 to 7.57 mm/s when Vjet increases
from 100 to 330 m/s. The area with uniform Vabl extends to R/Djet
around 0.4 for Vjet ¼ 100 m/s, and reduces slightly with increasing
Vjet. Like the increase in Vabl with Vjet, qdep at the center also in-
creases rapidly, from 23.0 to 134.8 kg/m2$s, as the jet is concen-
trated on a smaller Djet. The plateaus of qdep for different Vjet values
have a similar size compared to Djet of each casedaround 0.8Djet for
all cases. The average qdep values in this area are 23.3 kg/m2$s,
78.3 kg/m2$s, and 133.4 kg/m2$s, respectively. Further, maximum
deviation of qdep within this area is 3.7% for test #2. Normalized
distributions of Vabl and qdep for different Vjet suggest that a simi-
larity of the jet structure is maintained in each test, as intended in
the design of the test SRM.
Although both Vabl and qdep increase with Vjet, Qabl that related
to a completeness of Al2O3eC reaction of incoming alumina mass
exhibits a different trend. As shown in the lowest part of Fig. 11, Qabl
at the jet axis is 1.05104 m3/kg for Vjet ¼ 100 m/s and decreases
to 0.56104 m3/kg for Vjet ¼ 330 m/s. If we assumed that the
ablation is mainly caused by the mechanical erosion from the
particle, the ablation rate divided by the incoming particle mass
flux, Qabl should increase with the impact velocity, as the higher
momentum of the particle induce more deformation and fracture of
the char layer based on the Hertzian theory of impact [52e55].
Therefore, it is likely to be caused by a reduced residence time of
the film and incompleteness of the Al2O3eC reactions, as stated in
this research. To determine the relationship between Qabl and the
residence time of the film, a region within 0.8Djet, where Qabl is
uniform for all cases, was chosen for further analysis. The values
averaged in this region are Vabl,avg, qdep,avg, Qabl,avg, twall,avg, and
Wwalljet,avg where each denotes the average ablation rate, average
droplet deposition mass flux, average volumetric rate of ablation
per unit mass of alumina, average wall shear stress and average wall
jet velocity, respectively.
Table 4 summarizes Vabl,avg, qdep,avg, Qabl,avg, twall,avg, and Wwall-
jet,avg for every test conducted with the Kevlar/EPDM sample.
Because Peff related to the total mass flow rate through the test SRM
is varied intentionally by the amount of propellant or uninten-
tionally by the longer tail-off pressure, qdep,avg also varied from test
to test, even for the tests with same Vjet. The largest variation of
qdep,avg is 21.7% between tests #1 and #2 with Vjet of 100 m/s.
Similar variation was also observed in twall,avg as it was affected by
the Reynolds number. On the contrary, Wwalljet,avg of the 0.8Djet area
barely changes when Vjet is the same because Vjet and the velocity
around the jet nozzle are mainly determined by the geometric ratio
Djet/Dnoz,eff and exhaust gas properties. As a result, two tests for

Fig. 9. Sample shape after removal of char layer and distribution of residual/ablation
Measured sample thickness on CMM sweep lines after test; (c) Measured ablation
thickness after test #5 (Vjet ¼ 200 m/s): (a) Test sample after removal of char layer; (b)
thickness on CMM sweep lines.

189
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

Vjet, although twall,avg of these tests were higher than those of test
with Vjet of 268 m/s at 2.5 MPa due to higher pressure.
By comparing the results from test #1 to test #12, the effect of
the wall jet velocity increase from 30.5 to 136.0 m/s can be
observed. It is important to note that a variation in Qabl,avg for same
Vjet is less than 5.1% when Vjet is below 200 m/s. However, the
scatter of Qabl,avg increases with the scatter of twall,avg at a higher
Vjet. Maximum variation of Qabl,avg is 43.2% between tests #9 and
#10 with Vjet of 330 m/s. The reason for this is the large change in
the jet geometry during test #9, especially Hjet/Djet, caused by the
long exposure to high qdep,avg. In fact, dabl in test #9 is larger than
20 mm and Hjet/Djet is increased from 0.9 to 1.97 for this case. This
change alters the actual value of qdep,avg and twall,avg in the test SRM
over time and cannot be predicted by the simulation. The increase
in the variation of Qabl,avg can also be explained by this phenome-
non, which is the reason why the target test duration is shorter for a
test with a higher Vjet. Except for a large scatter in test #9, Qabl,avg
decreased constantly from 1.07104 to 0.49104 m3/kg with
Fig. 10. Distribution of ablation rate, droplet deposition mass flux and volumetric rate increasing Wwalljet,avg from 30.5 to 136.0 m/s.
of ablation per unit mass of alumina for test #5 (Vjet ¼ 200 m/s). However, Qabl,avg of test #13 and #14 with higher Peff are
significantly lower than the value of test #3 and #4 with similar Vjet
and Wwalljet,avg. Instead, they are comparable with the value of test
#7 and #8 with similar twall,avg. A scatter plot of the average Qabl,avg
and twall,avg shown in Fig. 12 confirms that twall,avg is the major
factor related to the film movement and resulting change of
Al2O3eC reactions. And it also indicates that the momentum of the
incoming droplet is not the main factor of the ablation in this study.
As shown in Fig. 12, Qabl,avg decreases with increasing twall,avg and
converges to a certain value at higher twall,avg as the slope de-
creases. A polynomial curve fit shows good agreement with the
experimental results including test #13, #14 and converges to a
Qabl,avg value around 0.49104 m3/kg. And This convergence of
Qabl,avg can be explained by the intermediate reactions involved in
the Al2O3eC reaction.
Guan et al. [26] compared three reactions between Al2O3 and C,
which generate pure Al, Al2OC, and Al4C3, respectively. The reaction
that produces Al4C3 was speculated to be the fastest under SRM
conditions because it had the lowest Gibbs free energy at an
elevated temperature. From this point of view, Qabl,avg in this study
can be understood as a combined effect of fast and slow reactions
between Al2O3 and C. Because a fast reaction has a much smaller
characteristic reaction time, the reduction in residence time due to
the movement of the film would have a more pronounced effect on
slow reactions. A drop in the reaction rate only for slow reactions
would yield a decaying Qabl,avg with increasing twall,avg. However,
further research is needed to truly correlate the species, mo-
mentum and energy conservations of the film with the heteroge-
neous Al2O3eC reaction. At the current stage, deduced relationship
between Qabl,avg and twall,avg can be used to predict the ablation rate
in the area where a liquid alumina formed with incoming droplet
mass flux, qdep, is under movement by the shear stress twall,avg like
aft dome of the rocket motor and inlet section of the nozzle.

4. Conclusions

In the present study, the alumina film-induced ablation of


Kevlar/EPDM with different lateral film speeds was investigated in
Fig. 11. Distribution of ablation rate, droplet deposition mass flux and volumetric rate a test SRM with an impinging jet configuration. Unlike previous
of ablation per unit mass of alumina with different Vjet. studies, a jet with a high alumina droplet flux was located close to
the sample to ensure the formation of an alumina film on the
surface and encourage the lateral movement of the film by a strong
each Vjet were performed with different qdep,avg, twall,avg and similar
wall jet. The lateral speed of the deposited alumina film was gov-
Wwalljet,avg. For test #13 and #14, Wwalljet,avg and momentum of the
erned by the velocity of the wall jet and was indirectly changed
incoming droplet were similar to those of test #3 and #4 with same
using six different design jet velocities (Vjet ¼ 100 m/s, 150 m/s,

190
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

Table 4
Summary of test results of alumina induced ablation.

Test Vjet/ Teff/ Peff/ Vabl,avg @ 0.8Djet/ Wwalljet,avg @ 0.8Djet/ twall,avg @ 0.8Djet/ qdep,avg @ 0.8Djet/ Qabl,avg @ 0.8Djet/
(m$s1) s MPa (m$s1) (m$s1) Pa (kg$m2$s) (104m3$kg1)

1 100 1.21 2.055 2.05 30.5 39.8 19.2 1.07


2 100 3.04 2.504 2.47 31.1 46.6 23.3 1.06
3 150 2.94 2.549 4.20 48.8 96.0 53.1 0.79
4 150 1.14 2.373 4.09 47.5 90.7 49.5 0.83
5 200 1.19 2.337 5.55 65.8 122.4 78.3 0.71
6 200 1.25 2.006 4.77 64.0 108.3 67.2 0.71
7 268 1.09 2.519 6.74 84.1 138.5 116.1 0.58
8 268 1.24 2.102 6.08 83.3 119.8 96.9 0.63
9 330 2.95 2.664 7.03 102.9 176.0 158.4 0.44
10 330 1.23 2.242 7.26 102.4 153.4 133.4 0.54
11 450 0.76 2.418 9.14 133.6 206.7 171.8 0.53
12 450 0.71 2.884 9.98 136.0 238.0 204.8 0.49
13 150 1.50 4.922 6.39 49.9 143.1 103.3 0.62
14 150 1.51 5.182 6.25 50.4 149.1 108.7 0.57

Qabl,avg and twall,avg obtained from the test SRM can be directly used
for full-scale SRMs to predict the ablation of an insulator exposed to
a droplet deposition and high-speed lateral flow (i.e., at the aft-
cylinder or aft dome of a case). The alumina-induced ablation on
a submerged nozzle can also be predicted from the experimental
results if the difference in the carbonaceous char density between
different materials is properly considered.

Declaration of competing interest

The authors declare that they have no known competing


financial interests or personal relationships that could have
appeared to influence the work reported in this paper.

References

[1] Twichell SE, Keller Jr RB. Solid rocket motor internal insulation. 1976. NASA-
SP-8093.
[2] Natali Maurizio, Maria Kenny Jose, Torre Luigi. Science and technology of
polymeric ablative materials for thermal protection systems and propulsion
devices: a review. Prog Mater Sci 2016;84:192e275.
Fig. 12. Change in averaged volumetric rate of ablation per unit mass of alumina with [3] Russell Gerald W, Strobel Forrest. Modeling approach for intumescing char-
averaged shear stress on 0.8Djet area. ring heatshield materials. J Spacecraft Rockets 2006;43:739e49.
[4] Brewer William D. Graphite and ablative material response to CO2-laser,
carbon-arc, and xenon-arc radiation. Washington, D.C.: NASA; 1976.
[5] Koo J, Lin S, Kneer M, Miller M. Comparison of ablative materials in a simu-
200 m/s, 268 m/s, 330 m/s, and 450 m/s). Fifteen static fire tests
lated solid rocket exhaust environment," 32nd Structures, Structural Dy-
were conducted, including one with a copper sample, to visualize namics, and Materials Conference. 1991. p. 978.
the alumina film. A two-phase flow simulation was used to analyze [6] Beall G, Shirin Z, Harris S, Wooten M, Smith C, Bray A. Development of an
the conditions on the sample surface. ablative insulation material for ramjet applications. J Spacecraft Rockets
2004;41(6):1068e71.
The ablation rate was considerably higher than the previously [7] McComb JC, Hitner JM. In: Technique for evaluating the erosive properties of
reported value due to a continuous supply of high temperature ablative internal insulation materials. vol. 1. Johns Hopkins Univ., The 1989
alumina to the surface by the jet. The ablation rate increased with JANNAF Propulsion Meeting; 1989.
[8] Li Jiang, et al. Ablation and erosion characteristics of EPDM composites under
the jet velocity, as a smaller jet diameter resulted in an increased SRM operating conditions. Compos Appl Sci Manuf 2018;109:392e401.
droplet deposition mass flux. However, Qabl,avg that indicates frac- [9] Li Jiang, et al. Characteristics and formation mechanism of compact/porous
tion of Al2O3 reacted with char layer initially decreased with structures in char layers of EPDM insulation materials. Carbon 2018;127:
498e509.
increasing average wall shear stress (twall,avg) and converged at a [10] Sapozhnikov I, Leitner A, Natan B, Mograbi E. Investigation of the ablative
higher twall,avg. This phenomenon indicates that the consumption properties of an EPDM\kevlar insulator in a solid rocket motor. Propellants,
of alumina in Al2O3eC reactions became less efficient at higher Explosives, Pyrotechnics, vo. 2022;47(1).
[11] Kang YoonGoo, Park JongHo. A comparison with thermal reaction character-
twall,avg because of the reduced residence time of the film. The in- istic of kevlar/EPDM internal insulator by change of chamber pressure.
fluence of the residence time was found to be more pronounced for J Korean Soc Propulsion Eng 2016;20(3):71e7.
slow reactions among the several intermediate reactions involved [12] Martin Heath T, Cortopassi Andrew C, Kuo Kenneth K. Assessment of the
performance of ablative insulators under realistic solid rocket motor oper-
in the Al2O3eC reactions. In addition, the fast reactions were less
ating conditions. Int J Energ Mater Chem Propuls 2017;16(1).
affected by the residence time, resulting in the convergence of [13] McWhorter Bruce, et al. Real-time measurements of aft dome insulation
Qabl,avg with twall,avg. We expect the result from this research will erosion on space shuttle reusable solid rocket motor. 40th AIAA/ASME/SAE/
contribute to extend the knowledge of the reaction mechanism ASEE Joint Propulsion Conference and Exhibit; 2004.
[14] Rindal RA. An analysis of the coupled chemically reacting boundary layer and
between alumina and carbonaceous char layer and it will become charring ablator: Part 6: an approach for characterizing charring ablator
further useful with further analysis of species/momentum/energy response with in-depth coking reactions", NASA. Tech. Rep. NASA CR-
of the alumina film. At the moment, the relationship between 1968;1065.
[15] Moyer C, Rindal RA. An analysis of the coupled chemically reacting boundary

191
J.-Y. Bae, I.S. Hwang and Y. Kang Defence Technology 29 (2023) 181e192

layer and charring ablator. Part 2- finite difference solution for the in-depth strategic solid rocket motor," 32nd AIAA, ASME, SAE, and ASEE, joint pro-
response of charring materials considering surface chemical and energy bal- pulsion conference and exhibit. Florida: Lake Buena Vista; 1996.
ances,” NASA CR-1061. June 1968. [35] Belz RA. Radiography diagnostics for tactical missile development. Arnold Air
[16] Chen Y-K, Milos FS. Ablation and thermal analysis program for spacecraft Force Base Report; 2021.
heatshield analysis. J Spacecraft Rockets 1999;36(3):475e83. [36] JasperLal C, Sridharan P, Krishnaraj K, Srinivasan V. Investigation of slag
[17] Chen Y-K, Milos FS. Two-dimensional implicit thermal response and ablation accumulation in solid rocket motors," 49th AIAA/ASME/SAE/ASEE Joint Pro-
program for charring materials. J Spacecraft Rockets 2001;38(4):473e81. pulsion Conference. 2013. p. 3861.
[18] Russell GW, Strobel F. Modeling approach for intumescing charring heatshield [37] Zuckerman N, Lior N. Jet impingement heat transfer: physics, correlations, and
materials. J Spacecraft Rockets 2006;43(4):739e49. numerical modeling. Adv Heat Tran 2006;39:565e631.
[19] Dec John A, Braun Robert D, Bernard Laub. Ablative thermal response analysis [38] ANSYS Fluent Theory Guide, Release 17.0, ANSYS, Inc., 2016.
using the finite element method. J Thermophys Heat Tran 2012;26(2): [39] Gilbert M, Davis L, Altman D. Velocity lag of particles in linearly accelerated
201e12. combustion gases. Jet Propulsion 1955;25:26e32.
[20] Milos FS, Chen YK. Ablation, thermal response, and chemistry program for [40] Hwang CJ, Chang GC. Numerical study of gas-particle flow in a solid rocket
analysis of thermal protection systems. J Spacecraft Rockets 2013;50(1): nozzle. AIAA J 1988;26(6):682e9.
137e49. [41] Mundo CHR, Sommerfeld M, Tropea C. Droplet-wall collisions: experimental
[21] Liu Y, Yang S, He G, Li J. An overall ablation model of ethylene-propylene- studies of the deformation and breakup process. Int J Multiphas Flow
diene monomer based on porous characteristics in char layer. Adv Mech 1995;21(2):151e73.
Eng 2016;8(2):1e10. [42] Mundo CHR, Tropea C, Sommerfeld M. Numerical and experimental investi-
[22] Shilav R, Khosid S. Development of ablative thermal response modeling of gation of spray characteristics in the vicinity of a rigid wall. Exp Therm Fluid
EPDM-based thermal protection systems. Int J Energ Mater Chem Propuls Sci 1997;15(3):228e37.
2020;19(4):275e92. [43] Wang Junlong, et al. Experimental and numerical study on slag deposition in
[23] Xu Y-H, Hu C-B, Zang S-M, Chen J. Experimental research on dynamic erosion solid rocket motor. Aero Sci Technol 2022;122.
of EPDM insulation subjected to particle-laden flow. J China Ordnance [44] Gordon Sanford, McBride Bonnie J. Computer program for calculation of
2010;6(4):225e33. complex chemical equilibrium compositions and applications. Part 1: anal-
[24] Wang J, Zha B, Zhang W, Zhang Y, Su Q. A new method for studying the ysis,” No. NAS 1 1994;61:1311.
ablation/erosion properties of silicone rubber composites based on multi- [45] Hara Shigeta, Norihito Ikemiya, Kazumi Ogino. Surface tensions and densities
phase flow. J Rubber Res 2019;22(2):59e68. of molten Al2O3 and Ti2O3. Iron Steel 1990;76(12):2144e51.
[25] Guo MengFei, et al. Carbon nanotube reinforced ablative material for thermal [46] Hakamada Shinya, et al. Surface oscillation phenomena of aerodynamically
protection system with superior resistance to high-temperature dense par- levitated molten al2o3. Internat J Microgravity Sci Appl 2017;34(4).
ticle erosion. Aero Sci Technol 2020;106. [47] Song Soon Ho, Kwon Hyung Joon, Yoon Ji Sang, Kim Dae Yu, Ho Park Nam.
[26] Guan Yiwen, Jiang Li, Liu Yang. Ablation characteristics and reaction mecha- A study on Aluminum Agglomeration Model in Solid Propellants Combustion
nism of insulation materials under slag deposition condition. Acta Astronaut (ADD internal report) 2018. ADDR-416-182955.
2017;136:80e9. [48] Yoon J, Lee K, Kim D, Park N, Ko S, Yoon W. Study of aluminum agglomeration
[27] Wirzberger Hod, Yaniv Sara. Prediction of erosion in a solid rocket motor by model during solid propellant combustion. J Korean Soc Propulsion Eng
alumina particles. 41st AIAA/ASME/SAE/ASEE Joint Propulsion Conference & 2019;23(2):78e86.
Exhibit; 2005. [49] Bae JY, Kim T, Kim JH, Ham H, Cho HH. Conjugate simulation of heat transfer
[28] Gubertov AM, Mironov VV, Borisov DM, Baskakov VN. Gasdynamic and and ablation in a small rocket nozzle. J. Comput. Struct. Eng. Inst. Korea.
thermophysical processes in solid-propellant rocket engines. Moscow: 2017;30(2):119e25.
Mashinostroenie; 2004. [50] Shimada T, Sekiguchi M, Sekino N. Flow inside a solid rocket motor with
[29] Guan YW, Li J, Liu Y, Yan QL. Reaction kinetics and a physical model of the relation to nozzle inlet ablation. AIAA J 2007;45(No. 6):1324e32.
charring layer by depositing Al2O3 at ultra-high temperatures. Phys Chem [51] Bae JY, Hwang IS, Kang Y. Prediction method for thermal destruction of in-
Chem Phys 2018;20:24418e26. ternal insulator in solid rocket motor. Journal of the Korean Society of Pro-
[30] Guan Y, Li J, Liu Y, Yan N. Deposits evolution and its heat transfer charac- pulsion Engineers; Feb. 2023. Accepted article for.
teristics research in solid rocket motor. Appl Therm Eng 2021;184:116266. [52] Neilson JH, Gilchrist A. An experimental investigation into aspects of erosion
[31] Liu Yang, Gao Yong-gang, Xiao-cong Li, Dong Zhi-chao, Guan Yi-wen. Kinetic in rocket motor tail nozzles. Wear 1968;11(2):123e43.
study and model verification of reaction between alumina and charred layer. [53] Thakre Piyush, et al. Mechanical erosion of graphite nozzle in solid-propellant
Ceram Int 2023;49:375e82. rocket motor. J Propul Power 2013;29(3):593e601.
[32] Walker Matthew S. Fundamentals of several reactions for the carbothermic [54] Yang BC. A theoretical study of thermomechanical erosion of high-
reduction of alumina," Thesis. Carnegie Mellon University; 2010. temperature ablatives. Ph.D. dissertation, Pennsylvania State Univ.; 1992.
Dissertation(Ph.D. [55] Xu YH, Hu X, Yang YX, Zeng ZX, Hu CB. Dynamic simulation of insulation
[33] Chang I-Shih. Slag and thermal environment of a spinning rocket motor. material ablation process in solid propellant rocket motor. J Aero Eng
J Spacecraft Rockets 1991;28(5):599e605. 2015;28(5):04014118.
[34] Frederick RA, Nichols J, Rogerson J. Slag accumulation measurements in a

192

You might also like