You are on page 1of 12

AIAA SciTech Forum 10.2514/6.

2021-1405
11–15 & 19–21 January 2021, VIRTUAL EVENT
AIAA Scitech 2021 Forum

Preliminary Study on Design and Testing of a


Metallic Thermal Protection System for
Spaceplane Vehicles
Vinh Tung Le1 and Nam Seo Goo2
Konkuk University, Seoul, 05029, Republic of Korea

This study presented a development of design and testing of a metallic thermal protection
system (TPS) for spaceplane vehicles. A metallic TPS was proposed to shield the fuselage of a
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

spaceplane vehicle from the extreme heat flux for returned missions. A conceptual design of a
metallic TPS panel which consisted of a hot load-carrying structure, a low thermal
conductivity thermal insulation, support brackets, and a cool structure, was modified from
the ARMOR panel developed by NASA. The thermal load was selected from a specific location
on the X-33 spaceplane (5.5 W/cm2). A transient semi-infinite heat transfer model was used to
predict the required thickness of the thermal insulation for an allowable temperature at the
cool structure. Due to high stresses and heat shorts in bracket support structures, a thin strip
with a brazed joint was used to connect the hot structure and the cool structure. A prototype
metallic TPS panel and supporting hardware made of S304 stainless steel were fabricated for
thermomechanical testing in room conditions. A panel-to-panel gap in panel assemblies was
designed to prevent possible contact between two adjacent panels. A numerical analysis was
conducted to determine whether the designed gap would change thermal responses at the cool
structure as well as at supporting hardware structures. A non-contact thermal-mechanical
measurement method was employed to measure full-field temperatures and deformations of
the cool structure and supporting hardware structures. A comparison of temperature
responses and panel deformations was performed between numerical analysis and
experimental testing. The results indicate that the current metallic TPS panel satisfied the
design requirements in terms of temperature and permanent deformation limits. The panel-
to-panel gap affected significantly the temperature rise in the cool structure and supporting
hardware structures. The experimental method provided a good approach to testing TPS
panels at the early stage of spaceplane development.

I. Introduction
A primary function of a thermal protection system (TPS) panel is to maintain the temperature limit of the inner
structure and to prevent possible impacts from outside. In the structural design of spaceplane vehicles, the extreme
conditions during a hypersonic flight must be considered. Thus, data estimation on the durability and reuse of TPS
panels under extreme conditions is necessary. High-temperature-resistant ceramic tiles and blankets had been used in
the Space Shuttle Orbiter as TPS structures which primarily prevented heat from reaching the vehicle interiors [1].
However, these tiles were very brittle and susceptible to damage even under applying small impact loads. Therefore,
passive metallic thermal protection systems with a low thermal insulation material are an excellent candidate due to
the inherent ductility and design flexibility of metallic material [2, 3]. To deal with the design of the TPS panel, the
thermal performance and thermo-mechanical behaviors are crucial issues. There is little information available to
indicate how to improve and modify the design of the TPS panel based on the testing data. In the early 2000s, NASA
published a design development of an advanced metallic TPS panel which included thermal performance issues and
prototype hardware. They classified the thermal issues and studied the structural design for bracket supports [4]. An

1
Postdoctoral Researcher, Department of Aerospace Information Engineering; vinhtung@konkuk.ac.kr.
2
Professor, Department of Aerospace Information Engineering, AIAA Member; nsgoo@konkuk.ac.kr.

Copyright © 2021 by Vinh Tung Le; Nam Seo Goo. Published by the American Institute of Aeronautics and Astronautics, Inc., with permission.
array of four TPS panels was fabricated based on the design parameters [2, 5, 6]. AIRBUS SAFRAN developed a
metallic TPS panel using beta titanium alloys [7]. However, the testing method and thermo-mechanical behaviors of
the TPS panel as well as of the assembled TPS panels associated with the supporting hardware structure have not yet
reported. Furthermore, because of the limitation on the measurement technique, the deformation measurement of the
TPS panel was complicated in high-temperature conditions. Recently, a hypersonic-vehicle program has been running
in the Republic of Korea, that aims to develop a technology for a spaceplane to travel in an orbital spaceflight mission
of reaching the low Earth orbit. In this program, we initially developed a superalloy TPS panel based on the design of
the ARMOR TPS panel invented by NASA [8, 9]. The common advantages of the introduced metallic TPS panels are
reusability and ease of installations, inspections, and replacements. The limitations include service temperatures of
less than 1173 K, heat shorts of mechanical connection, and increased weight.
Also, integrated TPSs have become popular recently. NASA proposed a new design of an integrated TPS panel
which had both thermal protection and load-bearing capabilities intended to use in the Orion crew module [10]. Meng
et al. [11] presented a design of the integrated TPS panel using numerical analysis. They proposed promising thermal
protection concepts for future hypersonic vehicles. The fabrication of the prototype and the testing have not been
reported yet due to the complexity of their proposed structures. In lately, Wei et al. [12-15] developed integrated TPSs
which were intended to be made of high-temperature ceramics. Li et al. [16] tested the thermal performance of an
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

integrated TSP made of C/SiC composite. Lin et al. [17] fabricated a multi-layer integrated TPS with ceramic material
and measured the temperature response at the inner structure. Cao et al. [18] developed an integrated TPS panel using
the addition of a shape-stabilized phase change materials. They investigated the heat transfer characteristics through
the panel and the result revealed a reduction of about 18% of temperature at the cool structure.
The advantage of the recent integrated TPS panels is that the structural weight was low, and the temperature service
of ceramic materials was extremely high. Although much research development of integrated TPS panels has been
achieved so far, however, the engineering applications of integrated TPS panels are still restricted. The design shapes
are extremely complicated to be fabricated in large quantities. The assembly of integrated TPS panels is limited for
large areas of vehicles and the installation and removal of integrated ceramics TPS panels to the vehicle’s structure
are complex. Therefore, the use of metallic materials is still needed in the design of future TPS panels. Especially, for
assembling and connecting between structural components, metallic materials are a good candidate for manufacturing
TPS panels. A disadvantage of the metallic material is its structural weight that should be enough in thickness to
guarantee the strength for the TPS panel. By using a lightweight metallic corrugated core sandwich structure [19], the
weight of metallic TPS panels became more competitive and attractive. Another disadvantage is the significant loss
of the strength of the metallic materials at extremely high temperature, therefore, the metallic materials are only used
to design for specific locations in the spaceplane which have a temperature not being elevated extremely.
Our previous work [8] presented a design of a superalloy TPS panel which consisted of a dense Inconel superalloy
for the hot structure, thermal insulation, spacer-fastener brackets, and thin titanium alloy for the cool structure. We
thoroughly examined the thermomechanical behaviors of a superalloy TPS panel under a simulation aerodynamic
heating. We found that permanent deformations and heat shorts were crucial issues in the current design; therefore, in
this study, we develop a new version of the metallic TPS panel which mostly focused on reductions of permanent
deformations, heat shorts, and total weight. A new version of the metallic TPS panel is designed along with supporting
hardware structures for further attaching to the vehicle’s structure. The development of the testing method is
emphasized in this study. A new prototype of stainless-steel metallic TPS panels and supporting hardware structures
were manufactured for testing at room conditions. The development of a non-contact testing method for the whole
TPS structure is emphasized in this study. The design development of the current metallic TPS panel is evaluated in
terms of temperature responses at supporting hardware structures and thermal deformations of the whole TPS
structures under simulated aerodynamic heating. Numerical finite element analysis was performed to compare with
experimental results. Results obtained from the experiment and numerical model would provide valuable testing data
for evaluating the current design at the early stage of development and to toward manufacturing the next prototypes
made of advanced materials.

II. Conceptual Design and Fabrication


A specific location on the X-33 spaceplane was chosen as a basis for calculating thermal loading history, as also
discussed in our previous works [8, 19]. The new conceptual design of the metallic TPS panel was developed for this
specific design point. This chosen design point based on the simulated temperature distribution over the body of the
X-33 spaceplane [4]. This location had a higher temperature than most of the surfaces of the vehicle except the
locations near the nose and wing edges where Carbon/Carbon and Carbon/SiC ceramics were used. Palmer et al. [20]
presented the use of metallic TPS panels for most areas of the X-33 windward surface.

2
A. Thermal Load and Requirements
The thermal load used for the current design was taken from the simulation results of the X-33 spaceplane [21].
Fig. 1a shows the heat-flux history and surface temperature distribution over the windward surface of the X-33
spaceplane. The current design of the metallic TPS panel used the equivalent temperature surface at the point marked
in Fig. 1a. This point resulted in approximately the temperature history shown in Fig. 1b. The peak temperature was
at 1050 K at the 550th sec and the temperature was maintained at 1023 K in the next 1250 sec.
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 1 Thermal load applied to the metallic TPS panel [6]. a) surface temperatures of the X-33 spaceplane and
the heat flux at the location used in the current study, b) Temperature profile applied on the outer surface of
the TPS panel. Temperature unit: Kelvin
Because the TPS panel forms the shape of the vehicle, it must withstand all the environmental conditions
experienced by the vehicle. Dorsey et al. [22] listed sufficiently general requirements for TPS panels which should
not be considered complete at this early stage of development. In this current study, we concerned about the thermal
protection function of the TPS panel and its structural deformation during the thermal load to maintain an acceptable
aerodynamic surface. The temperature of the internal structure should be within an acceptable limit to maintain the
good operations of the inner materials or devices. The maximum deflection should also be within an acceptable limit
to prevent the increase in local thermal loads due to the boundary-layer transition and to minimize permanent
compactions of the thermal insulation. Localized heating increment due to panel deflections was assumed to be
negligible. In addition to the thermal and structural functions, the TPS panel must be designed so that the overall
vehicle costs should be reduced by considering the low cost in the fabrication and installation, reducing required
inspection and maintenance.
In the current design of the target spaceplane, metallic TPS panels were connected to the vehicle’s structure
through supporting hardware structures. The designated material for the inner structure of the target vehicle was
aluminum which could operate well within 523 K without losing significant strength. Thus, the design and test of
metallic TPS panels should be along with the supporting hardware structure so that the temperature limit of supporting
hardware structures is estimated properly.

B. Required Thickness of Thermal Insulation and Support Bracket Structure


The transient problem was considered to determine the size of the thermal insulation material in the TPS panel. In
this study, to simplify the problem, we only considered the TPS panel with one layer of the thermal insulation material.
The one-dimensional (1D) semi-infinite model was used to calculate analytically the optimum thickness to get the
desired temperature within a specific time. Because the 1D semi-infinite model is an idealized body that has single
plane surface and extends to infinity in all directions. The temperature change in the part of the semi-infinite body
was due to the thermal conditions applied on a single plane surface. Therefore, the 1D semi-infinite body matched the
boundary condition in the initial sizing design of the thermal insulation material. Considering a semi-infinite model
with constant thermophysical properties, no internal heat generation, uniform thermal conditions on the exposed
surface, the initial temperature was uniformly distributed. The required thickness of the thermal insulation material to
satisfy the temperature of 523 K at the back side at 1000 sec was obtained using the thermophysical properties of the
Saffil insulation material at 523 K. This initial sizing result helped to select the available forms and densities of the
insulation material in the market for further experiments and numerical simulation.

3
Before the design of the new support bracket for the TPS panel, the requirement of the support bracket was listed:
(1) a thinner structure to minimize the heat conduction, (2) reducing the stress concentration on the support bracket
due to low bending stiffness. Based on the design of the support bracket of NASA [2, 6], we developed a thin strip of
the support bracket. By using the brazing method, both straight ends of the bracket could be brazed to the outer surface
and inner frame. To simply size the support bracket, we performed a thermomechanical analysis in ABAQUSTM
software. The 3D model is shown in Fig. 2. The temperature profile applied on the outer surface was taken from the
data in Fig. 1b. The residual stress in the support brackets and peak temperature on the back surface of the inner frame
were considered to choose the size of the support bracket.
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 2 Simple model for sizing the support bracket.

C. Fabrication
A metallic TPS panel with nominal dimensions of 170 mm×170 mm×33.6 mm was fabricated by Doowon Heavy
Industrial company, Republic of Korea. At this early stage of development, the stainless-steel material was chosen to
fabricate all the structural components of the TPS panel and the supporting hardware structure. A picture of the TPS
panel is shown in Fig. 3. The TPS panel consisted of a stainless steel corrugated-core sandwich structure on the top
side (outer side), a stainless-steel base frame (inner structure), four stainless steel support brackets, Saffil insulation
material, a thin stainless-steel foil for enclosing sides. The corrugated-core sandwich structure was 6.6 mm tall and
had the corrugated core with a 0.2 mm thick wall. The outer face sheet of the corrugated-core sandwich structure had
a 10 mm extension along two edges to overlap the gaps between the adjacent TPS panels. These overlap areas were
functioned as seals to cover the gaps between the adjacent TPS panels, as shown in Fig. 3a. Note that, the gap between
the adjacent TPS panels should be existed to allow the in-plane expansion of the TPS panels.
The corrugated core sandwich structure and the 1mm thick base frame were brazed to the four support brackets
(0.5 mm). The base frame was slotted at four corners to allow for inserting the fasteners, as shown in Fig. 3b. A 26
mm thick Saffil insulation material was inserted into the TPS panel. A 0.05 mm stainless steel foil was brazed to the
surrounding sides of the TPS panel. Finally, a 0.05 mm stainless steel foil was brazed to the base frame to close out
the bottom of the TPS panel. The weight of the fabricated metallic TPS panel was 0.52 kg which was equivalent to an
area weight ratio of 18 kg/m2. Fig. 4 shows the supporting hardware structure for an array of 3×3 flat TPS panels. This
unit array of the supporting hardware structure was designed to assemble on the curved vehicle’s structure. Two slip
joints and fixed joints formed the connection between the support bars and the corner brackets. The metallic TPS panel
was connected to the supporting hardware structure at corner brackets with corner legs, as shown in Fig. 4.

4
Fig. 3 A completely fabricated TPS panel: (a) view from the top side and (b) view from the bottom side.
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 4 A Supporting hardware structure is assembled with slip joints.

III. Thermomechanical Analysis and Experimental setup

A. Finite Element Analysis


A transient numerical simulation using ABAQUSTM software was performed to consider the coupled
thermomechanical analysis of the TPS panel and the supporting hardware structure. The thermal load was the
temperature profile obtained from the experiment. The temperature of the ambient was also inputted from the
experimental data. A single TPS panel integrated with a unit of the support hardware structure was simulated, as
shown in Fig. 5. Radiation in the cavity of the corrugated core structure was considered. The side of the TPS panel
was assumed to be an adiabatic condition because the side was surrounded by the thermal insulation in the experiment.
The maximum in-plane displacement of a single TPS panel was about 1.1 mm, as shown in Fig. 6. To prevent
possible contact between two adjacent panels, the gap distance should be larger than 2.2 mm.

Fig. 5 3D model of the TPS panel integrated with the supporting hardware structure. Boundary conditions
were applied in the bottom of the corner brackets, the adiabatic assumption was assigned for the sides, and
the radiation and convection dissipations were applied at the backside.

5
Fig. 6 In-plane displacement of the TPS panel. The displacement history was used to determine the minimum
gap between TPS panels.

To evaluate the ability to withstand the aerodynamic pressure of the designed support bracket, the suggested model
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

of the metallic TPS panel will be investigated under buckling analysis. We consider the worst-case scenario that the
venting system is not working well, and the vented space is at zero pressure after landing during a certain lag time
while the outer local static is at atmospheric pressure. There will be a pressure load on the outer surface of the TPS
panel. Therefore, the maximum pressure load on the outer surface is considered equal to 1 atm or 0.1 MPa, as also
assumed by Blosser [4] and Bapanapalli et al. [10].

B. Experimental Setup
Fig. 7 shows the arrangement of the TPS panels on the supporting hardware structure. Due to the size limitation
of the radiation heater (300-mm square), only the TPS panel at the center of the array was interested. The temperature
and displacement of the center TPS panel and the center cell of the supporting hardware structure were fully measured
in the experiment. One real TPS panel was installed at the center of the array and the other eight dummy TPS panels
were installed surrounding the real one. The insulation material was filled out with the dummy TPS panels and the
panel-to-panel gaps. The front side of the whole structure was faced with a radiation heater.

Fig. 7 Assembly of the TPS panels on the supporting hardware structure. a) the view from the outer surface
of the eight dummy TPS panels and one real TPS panel, b) the back view of the testing system.
A schematic of the experimental setup is shown in Fig. 8. The non-contact digital image correlation (DIC) method
[19, 23-26] was used to measure the full-field deformation of the whole testing structure. Thermocouples were
attached to the support bars and backside of the metallic TPS panel to monitor the temperature responses. An infrared
camera was also used to measure the full-field temperature distribution of the testing structure. The temperature profile
applied to the front side of the metallic TPS panel was monitored by thermocouples and controlled by a controller, the
temperature data from the front side was used to generate a DIC-camera trigger signal as the temperature reached the

6
desired value. A series of the captured images were then processed in the DIC-based software to obtain the
deformation.
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 8 Schematic of the experiment setup.

IV. Results and Discussion


The temperature distribution of the TPS panel was shown in Fig. 9. The temperature of the upper areas was higher
than that of the lower areas because of the imperfect experimental setup and the rising of the hot air. The temperature
at the bracket joints was highest among the other areas in the backside of the TPS panel. The temperature at the support
bars increased significantly. The temperature of the support bars was plotted in Fig. 9d. The peak temperature of the
backside of the TPS panel was at the left bracket joint, about 533 K which was very localized, while the temperature
at the rest of the backside was lower than 523 K which satisfied the temperature limit at the backside. Note that the
temperature was highest at the bracket joints because the heat conduction transferred from the outer surface to the
inner structure through the support bracket and arrived at the bracket joints. The temperature at the center backside of
the metallic TPS panel was about 423 K. The temperature at the support bars and corner brackets was about 393 K
and 378 K, respectively.
The full-field temperature distribution of the supporting hardware structure showed that the temperature of the
support bars was higher than that of the other components (corner brackets) except the corner legs. The corner legs
were connected to the base frame of the TPS panel by the fasteners. Thus, the temperature rise in the corner legs was
due to the heat conduction from the base frame. Therefore, I concluded that the temperature rise in the support bars
was mainly due to the heat transfer through the insulation strips filled in the panel-to-panel gaps.

7
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 9 Temperature measurement results. a) the interesting points, b) full-field temperature distribution, c)
measured temperatures at bracket joints, d) measured temperatures at support bars.
The out-of-plane deformation of the TPS panel was shown in Figs. 10a (experiment) and b (simulation). The
deformed shape was a concave shape from the view of the backside. Specifically, the center areas deformed outward
while the corner areas deformed inward because of the temperature difference between the front surface and back
surface of the TPS panel. The temperature of the front surface was higher than that of the back surface. To consider
the deflection of the base frame, there two points in Figs. 10a and b were extracted to calculate the total deflection.
Note that, the center area of the backside of the TPS panel was the thin foil, which was used to seal up the TPS panel,
but it did not withstand loads. Therefore, the total deflection of the base frame was determined by considering the two
marked points in Fig. 10. The total deflection history was shown in Fig. 10c for the four thermal cycle tests and the
numerical simulation. The maximum deflection was about 0.8 mm and the total absolute deflection was permanently
about less than 0.04 mm at the final time of the tests. The permanent deflection was a small value that satisfied the
permanent deflection limit of less than 1% of the diagonal length.

Fig. 10 Out-of-plane deformation; a) experimental result, b) numerical simulation, c) comparison of total


deflection between experiments and simulation.
It is difficult to estimate the deformation of the hot plate (outer surface) from the experiment. Thus, we estimated
the total deflection of the corrugated core sandwich structure from the numerical simulation. Fig. 11a shows the total
deflection of the corrugated core sandwich structure which was calculated by the deformation difference between the
two marked points, center and corner points. Fig. 11b shows the maximum deflection was 2.6mm and the absolute

8
permanent deformation was 0.06 mm. This maximum deflection sastified the deflection of the sandwich structure
which could prevent the compaction of the insulation material (less than 10% of total thickness of the TPS panel).
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

Fig. 11 Out-of-plane deformation of the outer corrugated core sandwich structure: a) deformation field in the
out-of-plane direction, b) total deflection versus time.

Fig. 12 shows the deformation of the support bars in the supporting hardware structure. This out-of-plane
deformation at the final time was almost zero. There was no existing permanent deformation found in the supporting
hardware structure after the experiment. According to the deformation results, the design of two slid joints and two
fixed joints between the support bars and the corner brackets is suitable for the supporting hardware structure.

Fig. 12 Out-of-plane deformation of the support bars in the supporting hardware structure.

The TPS panel was investigated to check the possible damages after thermal cycle tests. As a result, the TPS panel
was in good condition. There were several localized areas where the foil sides were distorted permanently, as shown
in Fig. 13a. This could be explained by the high-stress concentration in these areas after the experiments, the results
of stress distribution obtained from the numerical simulation are shown in Fig. 13b.

9
Fig. 13 Checking the status of the TPS panel after the thermal cycle test. a) A picture of the TPS panel
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

showed several distorted areas at the side joints, b) Mises-stress distribution of the TPS panel after a thermal
cycle test.
We did not observe any wrinkles or deflections at the overhang areas of the central TPS panel, as shown in Fig.
13a in the manuscript. The overhang areas of the central TPS panel were still straight. As shown in Fig. 14a, the
oxidized area on the tested array of the assembled panels indicates the heated area. There would be significant thermal
gradients and distortions in the surrounding panels. Gaps or steps were formed between two overlapped outer face
sheets of two adjacent TPS panels, as shown in Figs. 14b and c. This problem was due to the distortions of the
surrounding panels. In the tests, the thermal gradient occurred because of the size mismatch between the radiation
heater and the array of assembled TPS panels and resulted in distortions of the surrounding panels. However, the
distortions of TPS panels due to the thermal gradients over the outer surface of the TPS panel might be very small in
a real flight mission, because the incident heat would be applied uniformly over a TPS panel.

Fig. 14 A tested array of assembled TPS panels: a) heated area, b) and c) steps were formed at overhang
areas.

V. Conclusion
A new robust metallic thermal protection system (TPS) panel has been redesigned, fabricated, and tested. The
weight was significantly reduced in a comparison with the previous metallic TPS panel while the strength was still
satisfied (60% weight reduction). The new TPS panel could assemble on the vehicle’s structure through the fastener.
A full investigation of the TPS panel integrated with the supporting hardware structure was successfully tested by the

10
non-contact deformation and temperature measurement methods. The permanent deformation of the current TPS panel
was a small value that satisfied the permanent limit. The supporting hardware structure also operated well resulting in
the very small permanent deformation after thermal cycle tests. The temperature at the corner brackets, where the
whole TPS structure was intended to attach to the vehicle’s structure made of aluminum, was within the temperature
limit (less than 523 K). The developed metallic TPS panel could withstand well the thermal load of 5.5 W/cm2 which
corresponded to a safety design point for most of the surfaces on the X-33 spaceplane. The experimental method
provided valuable data for the design of metallic TPS panels at this early stage of metallic TPS development.

Acknowledgments
This research was supported by Ministry of Education Science and Technology as a project of “Space Core
Technology Development program” (NRF-2018M1A3A3A02065278). The authors are grateful for financial support.

References
[1] C. National Research. Reusable Launch Vehicle: Technology Development and Test Program. Washington, DC:
The National Academies Press, 1995.
[2] M.L. Blosser, C.C. Poteet, R.R. Chen, J.T. Dorsey, I.H. Schmidt, R.K. Bird, K.E. Wurster. "Development of
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

advanced metallic-thermal-protection system prototype hardware," Journal of Spacecraft and Rockets Vol. 41,
No. 2, 2004, pp. 183-194.
doi:10.2514/1.9179
[3] C.C. Poteet, M.L. Blosser. "Improving metallic thermal protection system hypervelocity impact resistance through
numerical simulations," Journal of Spacecraft and Rockets Vol. 41, No. 2, 2004, pp. 221-231.
doi:10.2514/1.9193
[4] M.L. Blosser. "Fundamental modeling and thermal performance issues for metallic thermal protection system
concept," Journal of Spacecraft and Rockets Vol. 41, No. 2, 2004, pp. 195-206.
doi:10.2514/1.9182
[5] D.E. Myers, C.J. Martin, M.L. Blosser, Parametric weight comparison of advanced metallic, ceramic tile, and
ceramic blanket thermal protection systems, NASA/TM-2000-210289, USA, 2000.
[6] M. Blosser, R. Chen, I. Schmidt, J. Dorsey, C. Poteet, R. Bird, Advanced metallic thermal protection system
development, 40th AIAA Aerospace Sciences Meeting & Exhibit, DOI 10.2514/6.2002-504 (2002), pp. 504.
doi: 10.2514/6.2002-504
[7] W. Fischer. "Maturation of AIRBUS D&S thermal protection systems portfolio," Processing and Properties of
Advanced Ceramics and Composites VII: Ceramic Transactions. Vol. 252, 2015, pp. 265-276.
[8] V.T. Le, N.S. Goo, J.Y. Kim. "Thermomechanical behavior of superalloy thermal protection system under
aerodynamic heating," Journal of Spacecraft and Rockets Vol. 56, No. 5, 2019, pp. 1432-1448.
doi:10.2514/1.a34400
[9] V.T. Le, N.S. Ha, N.S. Goo, J.Y. Kim. "Insulation system using high-temperature fibrous insulation materials,"
Heat Transfer Engineering Vol. 40, No. 17-18, 2019, pp. 1523-1538.
doi:10.1080/01457632.2018.1474602
[10] S. Bapanapalli, O. Martinez, C. Gogu, B. Sankar, R. Haftka, M. Blosser, Analysis and design of corrugated-core
sandwich panels for thermal protection systems of space vehicles, 47th AIAA/ASME/ASCE/AHS/ASC Structures,
Structural Dynamics, and Materials Conference DOI 10.2514/6.2006-1942 (2016), pp. 1942.
doi: 10.2514/6.2006-1942
[11] S. Meng, Q. Yang, W. Xie, G. Han, S. Du. "Structure redesign of the integrated thermal protection system and
fuzzy performance evaluation," AIAA Journal Vol. 54, No. 11, 2016, pp. 3598-3607.
doi:10.2514/1.J054996
[12] K. Wei, X. Wang, X. Yang, Z. Qu, Y. Tao, D. Fang. "Heat transfer mechanism and characteristics of lightweight
high temperature ceramic cellular sandwich," Applied Thermal Engineering Vol. 154, 2019, pp. 562-572.
doi:10.1016/j.applthermaleng.2019.03.126
[13] K. Wei, Y. Peng, Z. Qu, R. He, X. Cheng. "High temperature mechanical behaviors of lightweight ceramic
corrugated core sandwich panel," Composite Structures Vol. 176, 2017, pp. 379-387.
doi:10.1016/j.compstruct.2017.05.053
[14] K. Wei, K. Wang, X. Cheng, Y. Peng, M. Li, X. Yang. "Structural and thermal analysis of integrated thermal
protection systems with C/SiC composite cellular core sandwich panels," Applied Thermal Engineering Vol. 131,
2018, pp. 209-220.
doi:10.1016/j.applthermaleng.2017.12.009

11
[15] X. Wang, K. Wei, Y. Tao, X. Yang, H. Zhou, R. He, D. Fang. "Thermal protection system integrating graded
insulation materials and multilayer ceramic matrix composite cellular sandwich panels," Composite Structures
Vol. 209, 2019, pp. 523-534.
doi:10.1016/j.compstruct.2018.11.004
[16] Y. Li, L. Zhang, R. He, Y. Ma, K. Zhang, X. Bai, B. Xu, Y. Chen. "Integrated thermal protection system based
on C/SiC composite corrugated core sandwich plane structure," Aerospace Science and Technology Vol. 91, 2019,
pp. 607-616.
doi: 10.1016/j.ast.2019.05.048
[17] Y. Xu, N. Xu, W. Zhang, J. Zhu. "A multi-layer integrated thermal protection system with C/SiC composite and
Ti alloy lattice sandwich," Composite Structures DOI 10.1016/j.compstruct.2019.111507, 2019, p. 111507.
doi:10.1016/j.compstruct.2019.111507
[18] C. Cao, R. Wang, X. Xing, W. Liu, H. Song, C. Huang. "Performance improvement of integrated thermal
protection system using shaped-stabilized composite phase change material," Applied Thermal Engineering Vol.
164, 2020, p. 114529.
doi: 10.1016/j.applthermaleng.2019.114529
[19] V.T. Le, N.S. Goo. "Thermomechanical performance of bio-inspired corrugated-core sandwich structure for a
Downloaded by Vinh Tung Le on January 12, 2021 | http://arc.aiaa.org | DOI: 10.2514/6.2021-1405

thermal protection system panel," Applied Sciences Vol. 9, No. 24, 2019, p. 5541.
doi:10.3390/app9245541
[20] G. Palmer, D. Kontinos, B. Sherman. "Surface heating effects of X-33 vehicle thermal-protection-system panel
bowing," Journal of Spacecraft and Rockets Vol. 36, No. 6, 1999, pp. 836-841.
doi:10.2514/2.3522
[21] B.R. Hollis, R.A. Thompson, S.A. Berry, T.J. Horvath, K.J. Murphy, R.J. Nowak, S.J. Alter, X-33 computational
aeroheating/aerodynamic predictions and comparisons with experimental data, NASA/TP-2003-212160, NASA
Langley Research Center, Hampton, VA 23681-2199, 2003.
[22] J.T. Dorsey, R.R. Chen, K.E. Wurster, C.C. Poteet. "Metallic thermal protection system requirements,
environments, and integrated concepts," Journal of Spacecraft and Rockets Vol. 41, No. 2, 2004, pp. 162-172.
doi:10.2514/1.9173
[23] T. Jin, N.S. Ha, V.T. Le, N.S. Goo, H.C. Jeon. "Thermal buckling measurement of a laminated composite plate
under a uniform temperature distribution using the digital image correlation method," Composite Structures Vol.
123, No. 0, 2015, pp. 420-429.
doi:10.1016/j.compstruct.2014.12.025
[24] V.T. Le, N.S. Ha, T. Jin, N.S. Goo, J.Y. Kim. "Thermal interaction of a circular plate-ring structure using digital
image correlation technique and infrared heating system," Journal of Mechanical Science and Technology Vol.
30, No. 9, 2016, pp. 4363-4372.
doi:10.1007/s12206-016-0750-0
[25] T. Zhao, V.T. Le, N.S. Goo. "Global-local deformation measurement of stress concentration structures using a
multi-digital image correlation system," Journal of Mechanical Science and Technology Vol. 34, 2020, pp. 1655–
1665.
doi:10.1007/s12206-020-0328-8
[26] N.S. Ha, V.T. Le, N.S. Goo. "Investigation of punch resistance of the Allomyrira dichtoloma beetle forewing,"
Journal of Bionic Engineering Vol. 15, No. 1, 2018, pp. 57-68.
doi:10.1007/s42235-017-0004-6

12

You might also like