Applicability of Composite Materials For Space Radi - 2021 - Life Sciences in SP

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Life Sciences in Space Research 31 (2021) 71–79

Contents lists available at ScienceDirect

Life Sciences in Space Research


journal homepage: www.elsevier.com/locate/lssr

Applicability of composite materials for space radiation shielding


of spacecraft
Masayuki Naito a, Hisashi Kitamura a, Masamune Koike a, Hiroki Kusano a, Tamon Kusumoto a,
Yukio Uchihori a, Toshiaki Endo b, Yusuke Hagiwara b, Naoki Kiyono b, Hiroaki Kodama b,
Shinobu Matsuo b, Ryo Mikoshiba b, Yasuhiro Takami b, Masahiro Yamanaka b,
Hiromichi Akiyama c, Wataru Nishimura c, Satoshi Kodaira a, *
a
Radiation Measurement Research Group, National Institute of Radiological Sciences, National Institutes for Quantum and Radiological Science and Technology, Chiba
263-8555, Japan
b
Space Systems Division, Integrated Defense & Space Systems, Mitsubishi Heavy Industries, Ltd., Aichi 455-8515, Japan
c
Manufacturing Technology Research Department, Research & Innovation Center, Mitsubishi Heavy Industries, Ltd., Aichi 455-8515, Japan

A R T I C L E I N F O A B S T R A C T

Keywords: Energetic ion beam experiments with major space radiation elements, 1H, 4He, 16O, 28Si and 56Fe, have been
Space radiation conducted to investigate the radiation shielding properties of composite materials. These materials are expected
Shielding to be used for parts and fixtures of space vehicles due to both their mechanical strength and their space radiation
Composite material
shielding capabilities. Low Z materials containing hydrogen are effective for shielding protons and heavy ions
Total charge changing cross section
due to their high stopping power and large fragmentation cross section per unit mass. The stopping power of the
composite materials used in this work is intermediate between that of aluminum and polyethylene, which are
typical structural and shielding materials used in space. The total charge-changing cross sections per unit mass,
σ UM , of the composite materials are 1.3–1.8 times larger than that of aluminum. By replacing conventional
aluminum used for spacecraft with commercially available composite (carbon fiber / polyether ether ketone), it
is expected that the shielding effect is increased by ~17%. The utilization of composite materials will help
mitigate the space radiation hazard on future deep space missions.

1. Introduction 2014). HZE composition is driven in part of the physics of stellar


nucleosynthesis and nuclei, such as He, C, N, O, Ne, Mg, Si and Fe, are
The radiation environment in space, which consists primarily of relatively abundant. Data from the Mars Science Laboratory indicate
galactic cosmic rays (GCRs) and solar energetic particles (SEPs), is that a round-trip to Mars will result in a dose equivalent of ~660 mSv or
recognized as one of the principal health hazards to humans engaged in more (Zeitlin et al., 2013). Radiation shielding is one of the measures
space exploration. Although the geomagnetic field partially shields being investigated to help reduce risks to crews due to radiation
space travelers in low-Earth orbit (LEO), where the International Space exposure.
Station (ISS) is located, crews will encounter more serious radiation In accordance with the review by Durante (2014), passive shielding
hazards on upcoming long duration missions to the Moon, NASA’s Deep is considered as a practical solution. Because it is not technically feasible
Space Gateway (Crusan et al., 2018) and Mars due to the small magnetic to completely stop HZE particles in a shielding material, one possible
field. Protons and high atomic number and high energy (HZE) particles and realistic approach is to absorb relatively low energy particles in the
from He to Fe nuclei in the GCR are known to increase radiation risks in material and to break up the HZE particles into lighter particles, whose
space (Cucinotta, 2015). Although a large fraction (~87%) of the GCR is LET and dose contribution are smaller than primary heavier nuclei.
represented by protons (Simpson, 1983; George et al., 2009), the HZE Shielding properties of various materials such as aluminum (Al) and
particles significantly contribute to the radiation dose due to their high polyethylene (PE, (C2H4)n) have been studied through ground-based
linear energy transfer (LET) and high biological effect (e.g., Cucinotta, experiments using particle accelerators (e.g., Miller et al., 2003; La

* Corresponding author.
E-mail address: kodaira.satoshi@qst.go.jp (S. Kodaira).

https://doi.org/10.1016/j.lssr.2021.08.004
Received 19 April 2021; Received in revised form 17 August 2021; Accepted 18 August 2021
Available online 30 August 2021
2214-5524/© 2021 The Committee on Space Research (COSPAR). Published by Elsevier B.V. This is an open access article under the CC BY license
(http://creativecommons.org/licenses/by/4.0/).
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Tessa et al., 2005; Guetersloh et al., 2006; Zeitlin et al., 2006; DeWitt 2.2. Radiation dosimeters
et al., 2009). Hydrogen is known to be the most effective element for
shielding protons and heavy nuclei. However, since hydrogen is not We used two types of radiation dosimeters, an optical stimulated
chemically stable and not easy to handle, hydrogen-rich materials such luminescent detector (OSLD, ϕ7.2 mm × 0.35 mmt, Landauer Inc., USA)
as PE walls and water-rich towels in packages were proposed as and a CR-39 nuclear track detector (~25 mm × 25 mm × 0.9 mmt,
shielding materials on the ISS (Shavers et al., 2004; Kodaira et al., 2014). TechnoTrak, Chiyoda Technol Corp., Japan), to evaluate shielding ef­
In our previous work, we evaluated the physical parameters, stop­ ficiency of the target materials. OSLD has high sensitivity to radiation
ping power and nuclear fragmentation cross sections of various potential with low LET (< 10 keV/μm) such as protons and He ions, while CR-39
space radiation shielding materials (Naito et al., 2020) and found that records individual ion tracks with high LET particles (> 10 keV/μm) as
composite materials have a potential for use as structural materials with etch pits (Somogyi et al., 1976). CR-39 provides the LET spectrum
relatively high shielding efficiency. containing the primary ion beam and its fragments produced in the
In this paper, we have characterized the shielding properties of some target material. The absorbed dose behind the target material was ob­
types of composite materials by means of heavy ion accelerator exper­ tained by a combination of OSLD and CR-39, which is well-known as a
iments and Monte Carlo simulations. The enhanced shielding perfor­ standard space radiation dosimetry (Doke et al., 1995; Benton et al.,
mance achieved by replacing conventional Al in an actual spacecraft 2002; Hajek et al., 2008; Tawara et al., 2011; Kodaira et al., 2014). The
with composite materials has been reported. analyses using CR-39 for the derivation of the absorbed dose were
summarized elsewhere (Kodaira et al., 2016). An OSLD unit consists of
2. Materials and methods four pieces covered with black tape for light tightness (the unit size is
about 29 mm × 8 mm). The absorbed dose recorded in the OSLDs was
2.1. Shielding materials measured by a commercial reader (microStar, Landauer Inc., USA).

The composite materials are made of resins, reinforced to enhance


2.3. Beam experiments
mechanical strength. The resin is an organic polymer such as polyether
ether ketone (PEEK), polyimide (PI) and polypropylene (PP), which is
To evaluate the shielding efficiencies of target materials, the mate­
selected based on its the chemical and thermal properties. Carbon fiber
rials were irradiated with protons and heavy ion beams representing
(CF) or silicon carbide is used for reinforcement. The selection of
major components of the GCR in the biological irradiation room of
appropriate resins and reinforcements allows the composite materials to
HIMAC (Heavy Ion Medical Accelerator in Chiba) of QST, Japan. Table 2
be more shock- and heat-resistant than resin and reinforcement alone.
lists the beam energies and calculated LET values in water. We note that
Composite materials are already used in automobiles, airplanes and
these energies are the maximum available for each ion at the HIMAC,
spacecraft because of their low density and high mechanical strength (e.
which are somewhat lower than the mean energy of the corresponding
g., Chen, 1997). Composite materials contribute not only to structural
GCR particles (~1 GeV/n). The ranges of beams in the target materials
strength in spacecraft but also to radiation shielding by their relatively
are listed in Table 3. As shown in a schematic of the experimental set-up
higher stopping power and a larger nuclear fragmentation cross section
(Fig. 1), the target materials were stacked with the two types of radiation
per unit mass compared to Al (Naito et al., 2020). We evaluated three
dosimeters. The dosimeter at 0 g/cm2 was attached in front of first target
types of target materials: reference materials (Al and PE), commercially
material. According to the Monte Carlo simulation, 228 MeV 1H particle
available composite materials and mixed composite materials as sum­
go through 10 g/cm2 Al slab with backward radiations, contributing to
marized in Table 1. The commercial materials were CF/PI (AURUM
~2% to absorbed dose. The backward radiations consist of 75% pho­
JCN3030, Mitsui Chemicals, Inc., Japan), CF/PEEK (Toray Cetex
tons, 20% neutrons, and 5% other particles. Therefore, the contribution
TC1200, Toray Advanced Composites, USA), CF/PP (TAFNEX CF/PP,
of backward primary and secondary particles would be small in the
Mitsui Chemicals, Inc., Japan) and CF/Epoxy (Mitsubishi Chemical
Corporation, Japan). The mixed composites were produced by mixing
resin (PE or PP) and reinforcement material (C or SiC powder) in the Table 2
volume ratios of 8:2 or 6:4 to imitate the elemental composition of a Energy and calculated LET in water of each beam.
composite material. For instance, PE SiC is made of 80% PE and 20% Ion Energy (MeV/n) LET in water (keV/μm)
SiC, and PP SiC20 is made of 80% PP and 20% SiC in volume. The un­ 1
H 228 0.4
certainties due to the variations of material thickness were included in 4
He 145 2.3
16
the measurement results. O 380 20.0
28
Si 440 56.9
56
Fe 410 199.5

Table 1
Target materials used in our experiments and their physical parameters.
Materiala Density (g/cm3) Size of sample (cm × cm) Resina Reinforcementa Resin fraction (wt%) ZT /AT AT
− 1/3

Al 2.7 10 × 10 —– —– —– 0.481 0.333


PE 0.94 10 × 10 —– —– —– 0.570 0.598
PP SiC20b 1.36 10 × 10 PP SiC 53 0.537 0.515
PP SiC40b 1.81 10 × 10 PP SiC 30 0.520 0.462
PP C 1.17 10 × 10 PP C 62 0.543 0.547
PE SiC 1.36 10 × 10 PE SiC 53 0.537 0.516
CF/PI 1.43 6×6 PI CF 70 0.511 0.466
CF/PEEK 1.62 5.5 × 5.5 PEEK CF 33 0.508 0.434
CF/PP 1.31 10 × 10 PP CF 33 0.523 0.502
CF/Epoxy 1.57 10 × 10 Epoxyc CF 34 0.510 0.464
*
PE, polyethylene; PI, polyimide; CF, carbon fiber; PEEK, polyether ether ketone; PP, polypropylene.
#
SiC20 and SiC40, the numbers indicate percent volume of SiC in the composite material.
+
Composition of the epoxy material is not publicly available. Therefore, ZT and AT of CF/Epoxy were estimated from its mass and density.

72
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Table. 3 The fluence histograms as a function of LET measured by CR-39 were


Calculated range of each beam in each target material. CF/PI, CF/PEEK and CF/ fitted to a gaussian function to obtain the mean LET and the number of
PP were not tested with 1H, 4He and 16O beams. primary particles. The variations of the LET values and the number of
Material Range (mm) counts as a function of material thickness were obtained from the fitting
1
H 228 4
He 145 16
O 380 28
Si 440 56
Fe 410 results. The total charge-changing cross section, σ CC , was obtained from
MeV MeV/n MeV/n MeV/n MeV/n the reduction rates of the number of counts of primary particles, dCp /dx,
Al 153.1 70.1 90.0 64.6 34.4 as:
PE 326.2 149.5 193.0 138.8 73.9 /
PP 249.4 114.5 147.3 105.9 56.2 dCp dx
SiC20 σ CC = (2)
N
PP 198.0 91.3 116.7 84.0 44.7
SiC40 σ CC generally depends on the Bradt-Peters equation (Bradt and Pe­
PP C 276.5 127.2 163.9 118.0 62.6 ters, 1950; Durante and Cucinotta, 2011):
PE SiC 248.8 113.9 146.3 105.6 56.2
( )2
CF/PI 104.6 55.7
(3)
—– —– —–
CF/ —– —– —– 92.6 49.3
σ CC = c1 A1/3 1/3
p + Ap + c2

PEEK
CF/PP —– —– —– 110.4 58.8 where Ap is atomic mass of the projectile. c1 and c2 are semi-empirical
parameters. When we consider the cross section per target unit mass,
σ UM , Eq. (2) should become:
( )2
1/3 1/3
σCC c1 Ap + AT + c2
σ UM = = (4)
AT AT
These equations can be simplified with a parameter, c, as a rough
approximation (Durante, 2014):

σ CC ∝cA2/3 (5)
− 1/3
T , σ UM ∝cAT

Therefore, ZT /AT and A−T 1/3


are useful parameters to compare SUM
and σUM among the target materials, respectively. We note that ZT and
AT have to be the mean values in the case of a chemical compound; for
instance, ZT and AT of PE are (1 × 4 + 6 × 2)/6 = 2.67 and (1.01 × 4 +
12.01 × 2)/6 = 4.68, respectively. These values of target materials are
also listed in Table 1. The composite and mixed materials have param­
eters intermediate between those of Al and PE.
Fig. 1. Schematic of the experimental setup. Target materials were stacked
with the two types of radiation dosimeters.
2.5. Numerical simulation
measurement data. We note that CF/PI, CF/PEEK and CF/PP were not
tested with 1H, 4He and 16O beams due to limitations on beam time. A We conducted Monte Carlo simulations with Geant4 ver. 10.04.02
stack with OSLDs was irradiated with each beam of a dose of 100 mGy in (Agostinelli et al., 2003; Allison et al., 2006, 2016) to reproduce our
front of the stack, while that with CR-39 detectors was irradiated with experimental results. For the simulations, ion beams with the same en­
each beam of a fluence of ~3000 cm− 2. The irradiated dose and fluence ergies in the experiments were projected onto the target materials ar­
were chosen based on availability of beam intensity and analysis pro­ ranged in configurations similar to the experiment setup. It should be
cedure of the detectors. Each beam intensity was independently moni­ noted that CF/Epoxy was excluded from the simulations since its
tored with an ionization chamber and a plastic scintillator installed in elemental composition is not publicly available. The LET and the num­
the beam line. The beam spot size was uniformly spread out to be 100 ber of primary particles were extracted and compared to the experi­
mm in diameter using a wobbler-scatterer method. The uniformity over mental data. The effective dose equivalent due to primary and secondary
the 100 mm diameter was better than 5%, leading to a 5% systematic particles (neutrons, electrons, positrons, photons, positive and negative
error in the results. Finally, the measured data with OSLD and CR-39 pions, muons and lighter charged particles) behind the target materials
were combined to obtain the total absorbed dose by assigning a of 5 g/cm2 mass thickness were also obtained from the particle track
scaling parameter equal to the ratio of the irradiated dose on a stack with information using the ICRP fluence-dose conversion coefficient (ICRP,
OSLDs and fluence on a stack with CR-39 detectors. The factor was 2013). We note that the contribution of neutral pions was not considered
applied to the dose data measured with CR-39 for each stack. because the corresponding conversion coefficient is not given by ICRP.
Particle transport was terminated when the residual range in the ma­
terial was to be less than 0.7 mm. The simulations were conducted using
2.4. Shielding parameters
two Geant4 reference physics lists of “Shielding” and “QGSP_BIC_HP”. It
was confirmed that the two models gave similar results. The only result
The stopping power of the target materials was evaluated by the
obtained by the “Shielding” list is presented.
position of the Bragg peak. The stopping power, S, is defined by the
Spatial dose distribution due to space radiation in the spacecraft was
Bethe-Bloch equation, depending on NZ2p ZT /E where N, Zp , ZT and E are
calculated using Monte Carlo simulations to assess the shielding effect of
the number density of target molecules, atomic numbers of the projectile composite materials. The Japanese spacecraft HTV (H-II Transfer
and target, and kinetic energy of the projectile, respectively. Therefore, Vehicle, “The KOUNOTORI”) pressurized logistics carrier model (Fig. 2)
the mass stopping power, SUM , is described using the atomic mass of was exposed to GCR particles ranging from protons to iron nuclei with
target, AT , as: GCR energy spectra ranging from 10 MeV/n to 100 GeV/n. The HTV
NZp2 ZT ZT module has been used for the transportation of supplies to ISS. Although
SUM ∝ ∝ (1) human transportation has not been conducted yet, an update of the HTV
EAT AT
pressurized module for manned work space is planned. The HTV model

73
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Fig. 2. The HTV module model; (a) side view and (b) top view.

is imported from CAD model into the Monte Carlo simulation. The en­ any interactions (so-called the ray-tracing technique). The shielding
ergy spectra of GCR particles were derived from the DLR model at the thickness at any region can be obtained from the track information of
solar minimum phase (Matthiä et al., 2013), assuming the worst case each particle.
radiation exposure. Al and CF/PEEK were employed as the structural
materials of the HTV module for comparison. Although the employment 3. Results and discussion
of CF/PEEK for whole structure may introduce other issues such as heat
resistance, transmission and emission, our calculation is a representative 3.1. Shielding properties of composite materials
comparison for future design studies of spacecraft from the viewpoint of
space radiation shielding. CF/PEEK was selected because of its superior The variation of absorbed dose in the target materials was obtained
shock resistance compared to the other composite materials tested in for experimental beam irradiation, as shown in Fig. 3. The dose data
this work. We note that the scale of geometric model was fixed in units of were normalized to those at 0 g/cm2 (i.e., in front of a stack). The Bragg
mm or cm since the CAD model is too complicated to implement the peaks of composite materials were located between those of PE and Al as
density factor converting to g/cm2 scale. A 2000 cm × 2000 cm planar expected from the stopping power parameters in Table 1. No large dif­
source with a cosine angular distribution was fixed at z = ±250 cm. The ferences among the composite materials were observed. The differences
cosine angular distribution on the planar surface reproduces an isotropic in the stopping power among the composite materials were trivial. We
radiation environment. 10 particles/cm2 of primary particles were note that there was a restriction on the measurement points due to the
projected for each ion. The spatial distribution of effective dose equiv­ available thickness of the materials.
alent due to the primary and secondary particles was obtained from the The LET spectra measured at different target thicknesses with CR-39
track information. The spatial distribution of mean shielding thickness were obtained for 16O, 28Si and 56Fe beam irradiations. For instance, LET
in the HTV model was also obtained using particles that do not undergo spectra for 56Fe irradiation are shown in Fig. 4. The broadness of peak at

Fig. 3. Variations of absorbed dose experimentally obtained in the target materials for irradiated beams: (a) 1H 228 MeV, (b) 4He 145 MeV/n, (c) 16O 380 MeV/n, (d)
28
Si 440 MeV/n and (e) 56Fe 410 MeV/n. Vertical axis was normalized to the dose in front of each stack (0 g/cm2).

74
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Fig. 4. The LET spectra of the 56Fe beam at in front of a stack (0 mm) and behind some of the target materials for (a) Al, (b) PE, (c) PP SiC20, (d) PP SiC40, (e) PP C,
(f) PE SiC, (g) CF/PI, (h) CF/PEEK, (i) CF/PP and (j) CF/Epoxy. Thin solid lines denote the gaussian fitting results of primary Fe ions. Some fragments with LET lower
than primary Fe ions were observed in the enlarged scale view of PE data (b).

0 mm mainly comes from the detector resolution and fluctuation of Al target are ~1200 mb (400 MeV/n, Zeitlin et al., 2011), 1550 mb (a
energy loss in air layer. The LET resolution in CR-39 has been verified by few hundred MeV/n to ~2 GeV/n, Binns et al., 1989) and 2000 mb
Kodaira et al. (2016). The peaks of Fe ion beam become broad with (1.05 GeV/n, Zeitlin et al., 1997) with errors of 2–3%, respectively. Our
increasing the target thickness due to the degradation of kinetic energy. results are consistent within ~10–20% deviations. A plausible expla­
The LET of primary Fe ions increased due to the ionization energy loss, nation of this discrepancy is systematic error of the experiments. The
while the number of counts decreased due to nuclear fragmentation in ratios of experiment to simulation for the number of primary particles
the target material. The LET and the number of counts of each Fe ion offered 10–15% differences (Figs. 5(d-f)), which were caused by the
peak fitted with a gaussian function are indicated in the subfigure. The differences in the set-up of a stack and variation of material thickness.
experimentally-measured relative variations of the LET and the number The statistical and systematic errors due to the beam irradiation and
of counts of 16O, 28Si and 56Fe peaks are represented as a function of measurement offered the uncertainty of about 10%. Around the Bragg
target material thickness (Figs. 5a-c). Notably that ion tracks of low LET peak, the differences between the experiment and simulation would
1
H and 4He ions were not recorded in CR-39 due to the detection limit become large, especially for the LET. Judging from these views, the
(Kodaira et al., 2016). The rates of LET increase and count decrease were experimental result is reasonable with the simulations, meaning that we
comparable over the experimental energy range. This balance rendered can successfully reproduced the fragmentation and transport of heavy
the flatness of the dose curve in the plateau regions in Fig. 3. The ions in the plateau region of the Bragg-curve using a Monte Carlo
stopping power depends on the reciprocal of particle kinetic energy (Eq. simulation.
(1)), while energy dependence of the fragmentation cross section is The effective dose equivalent values behind 5 g/cm2 mass thickness
relatively weak for particle energies ranging from a few hundred MeV/n of target materials were obtained by numerical calculations (Fig. 7). The
to about 2 GeV/n (e.g., Zeitlin et al., 2011). It is expected that frag­ vertical axis was normalized to the effective dose equivalent at 0 g/cm2.
mentation becomes dominant at the cosmic ray energy having a peak It should be noted that the effective dose equivalent was calculated by
around 1 GeV/n (Matthiä et al., 2013). the conversion coefficient for irradiation of a voxel phantom; i.e.,
We plot the values of σ UM and σ CC obtained from the reduction rates effective dose equivalent in the voxel phantom. Meanwhile, the exper­
of the number of primary particles by Eqs. (2) and (4) as a function of imental data in Fig. 3 comes from the absorbed dose on each dosimeter.
target material mass, AT (Fig. 6) The plots were fitted to Eq. (5) using a Although, the variation of absorbed dose in Fig. 3 was minimal at the
least square method. The fitted lines are also shown. σUM , σ CC and fitted plateau regions, the effective dose equivalent was decreased with a thin
parameter, c, are summarized in Table 4. The experimentally acquired (5 g/cm2) target material. This is because of the difference in scales of
σ UM and σ CC are proportional to A−T 1/3 and A2/3 phantom and dosimeter. The fragmentation of projectile particles re­
T , respectively. In accor­
dance with the previous studies, the σCC of 16O, 28Si and 56Fe ions in the duces effective dose equivalent for heavier ions except for protons.
Protons do not break down further and LET values increase due to the

75
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Fig. 5. (a-c) Variations of the LET and the number of counts of primary ions as a function of material thickness. Closed plots with solid lines denote LET and open
plots with dotted lines denote the counts, respectively. (d-f) Their ratios of experimental to simulation results. The horizontal dotted and solid lines denote to ±10%
and±20% variation against the ratio = 1.0, respectively.

Fig. 6. (a) σUM and (b) σCC experimentally obtained as a function of AT . Fitted lines to Eqs. (3) and (4) are also shown.

energy loss in the material. The increase of LET added to the effective fragmentation reactions increased with projectile mass: ~4.0% for 4He,
dose equivalent behind 5 g/cm2 shielding (Fig. 7(a)). The production of ~5.3% for 16O, ~9.5% for 28Si and ~24% for 56Fe in PE target. How­
secondary particles due to target fragmentation also plays an important ever, the effective dose equivalent reduction by shielding was more than
role. The secondary particles contributed to increase the effective dose compensated for the secondary yields (Fig. 7 (b-e)). In other words, the
equivalent, specifically ~1.0% for Al, ~0.6% for PE and ~0.7% for fragmentation of primary particles contributes to the dose reduction of
composite and mixed materials in average. For heavy ion irradiations, space radiation in spite of the production of secondary particles. The
the contribution of secondary particles due to both projectile and target composite materials provided >30% higher dose reductions than Al for

76
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Table. 4 3.2. HTV module simulation


σMU and σCC experimentally obtained and parameter c obtained by fitting in
Fig. 6. The spatial distribution of mean shielding thickness of the HTV
Material 16
O 380 MeV/n 28
Si 440 MeV/n 56
Fe 410 MeV/n module was calculated and is shown as a function of each axis (x and y)
σMU σCC σMU σCC σMU σCC in Figs. 8(a-b), where the ranges of y-z and z-x were fixed from − 50 cm
to +50 cm, respectively. Although any y-z and z-x ranges can be
Parameter 107 ±1 148 ± 2 224 ± 5
c selected, small area results very long calculation time to reduce statis­
Al 35 ± 5 940 ± 47 ± 5 1270 ± 80 ± 7 2150 ± tical errors. We chose these y-z and x-z ranges, where the errors along
140 128 174 with the results are statistically acceptable. The shielding thickness of Al
PE 66 ± 310 ± 98 ± 459 ±60 141 ± 660 ± 60 is larger than that of CF/PEEK by a factor of their density ratio, 2.7/1.62
10 45 13 13 = 1.67, since the calculations were made using the same model. The
PP SiC20 54 ± 390 ± 77 ± 559 ± 114 ± 835 ± shielding thickness at the central region becomes the maximum for the
10 71 17 120 22 160
axis direction (x) and minimum for the radius direction (y). This position
PP SiC40 49 ± 8 490 ± 67 ± 9 681 ± 91 98 ± 5 994 ± 48 dependence is due to the isotropic angular distribution of GCR particles.
80
The mean shielding thickness relies on the effect of inclination,
PP C 60 ± 370 ± 78 ± 474 ± 134 ± 821 ± 41
16 97 17 100 7
increasing the apparent thickness.
PE SiC 57 ± 410 ± 71 ± 514 ± 105 ± 765 ± 45
The spatial distribution of effective dose equivalent for each axis (x
11 79 15 110 6 and y) is presented in Figs. 8(c-d) in the same manner as for Figs. 8(a-b).
CF/PI —– —– 70 ± 6 693 ± 57 103 ± 1020 ± It should be noted that the vertical axes were normalized to the effective
10 96 dose equivalent in free space. Despite the fact that the shielding thick­
CF/PEEK —– —– 73 ± 890 ± 118 ± 1440 ± ness of CF/PEEK was 1.67 times smaller than that of Al, the relative
17 210 19 231 effective dose equivalent of CF/PEEK was comparable to that of Al.
CF/PP —– —– 73 575 ± 112 ± 890 ± 54 When we consider shielding material for space radiation, σUM is the
110 7
±14
dominant parameter rather than the mass stopping power due to the
CF/Epoxy 65 ± 655 ± 65 114 ± 1140 ±
—– —–
very high energy of the GCR. Depending on the projectile ions, the σUM
12 20 195
values of CF/PEEK were 35–70% larger than that of Al in our experi­
ments. In other words, the high shielding efficiency of CF/PEEK pro­
heavy ions. The dose reductions of the composite materials for heavy vided a comparable radiation environment to that in the module
ions were <30% lower than that of PE. Among the mixed and composite fabricated from Al, even though the spacecraft mass fabricated with CF/
materials, PPC provided the highest dose reduction rates (4.3% for 4He, PEEK was almost half of that with Al (1/1.67).
20% for 16O, 24% for 28Si, and 16% for 56Fe). This result is consistent A general comparison with the same shielding thickness could be
with the largest A− 1/3 values (Table 1) although the differences among conducted through the effective dose equivalent as a function of
the mixed and composite materials are not large, compared with the shielding thickness (Fig. 9). The mean values of relative dose equivalent
errors in the experimental results. inside the CF/PEEK HTV module (− 230 < x < 230 and − 150 < y < 150)

Fig. 7. Effective dose equivalent behind 5 g/cm2 mass thickness of target materials for (a) 1H 228 MeV, (b) 4He 145 MeV/n, (c) 16O 380 MeV/n, (d) 28Si 440 MeV/n
and (e) 56Fe 410 MeV/n. Vertical axis was normalized to the dose at 0 g/cm2. Bottom dark colors denote to the effective dose equivalent due to the primary particles.
Top light colors denote to the effective dose equivalent due to the secondary particles.

77
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Fig. 8. (a) x and (b) y distributions of mean shielding thickness in the HTV module derived from the ray-tracing technique. (c) x and (d) y distributions of effective
dose equivalent in the HTV module. Vertical axis was normalized to the dose equivalent in free space. The mean relative dose equivalents inside the module are
indicated at the center for Al and CF/PEEK.

higher than those of Al HTV one in the same shielding thickness (g/cm2).
Therefore, the mean shielding effect of CF/PEEK was expected to be
~17% higher than that of Al. This result is consistent with the experi­
mental results on shielding performance with slabs irradiated with
heavy ion beams. For mitigating radiation hazards, there are strong
advantages to using composite materials in structures and fixtures in
space vehicles rather of conventional Al.

4. Conclusion

We have investigated the radiation shielding properties of composite


materials compared to conventional materials of polyethylene as a
shielding material and aluminum as a structural material in spacecraft.
The composite materials had shielding efficiency intermediate between
that of polyethylene and aluminum: >30% higher shielding efficiency
than aluminum and <30% lower than polyethylene. By using a
commercially available composite, CF/PEEK, for the HTV module
structure, the effective dose equivalent due to galactic cosmic ray par­
Fig. 9. Normalized effective dose equivalents as a function of shielding thick­
ticles was found to be comparable to that with the aluminum HTV
ness of Al and CF/PEEK. module despite its small mass by a factor of their density ratio of 1.67. A
35–70% larger CF/PEEK fragmentation cross section per unit mass
provides effective radiation shielding. The use of composite materials in
were calculated to be 0.64 for x average and 0.49 for y average. These
place of aluminum in spacecraft is a promising option for mitigation of
values correspond to 13 and 31 g/cm2 shielding thicknesses, respec­
space radiation exposure.
tively. By multiplying these corresponding shielding thicknesses by the
density factor 1.67, the mean relative dose equivalents in the same
5. Funding
thickness can be evaluated as 0.55 for x and 0.41 for y. The mean values
of relative dose equivalent inside the Al HTV module were calculated to
The work by members of the National Institutes for Quantum and
be 0.62 for x average and 0.49 for y average. The dose reduction rates of
Radiological Science and Technology has been partially funded by
CF/PEEK HTV module were 1.18 times (for x) and 1.16 times (for y)

78
M. Naito et al. Life Sciences in Space Research 31 (2021) 71–79

Mitsubishi Heavy Industries, Ltd. Guetersloh, S. et al. (2006) ‘Polyethylene as a radiation shielding standard in simulated
cosmic-ray environments’, Nuclear Instruments and Methods in Physics Research
Section B: Beam Interactions With Materials and Atoms. North-Holland, 252(2), pp.
Declaration of Competing Interest 319–332. doi: 10.1016/j.nimb.2006.08.019.
Hajek, M., et al., 2008. ‘Convolution of TLD and SSNTD measurements during the
The authors declare that they have no known competing financial BRADOS-1 experiment onboard ISS (2001)’. Radiat. Measurements. Pergamon 43
(7), 1231–1236. https://doi.org/10.1016/j.radmeas.2008.04.094.
interests or personal relationships that could have appeared to influence ICRP, 2013. ‘Assessment of radiation exposure of astronauts in space’. ICRP Publication
the work reported in this paper. 123. Annuals of the 42 (4), 1–339. https://doi.org/10.1016/j.icrp.2013.05.004.
ICRP.
Kodaira, S., et al., 2014. ‘Verification of shielding effect by the water-filled materials for
References space radiation in the international space station using passive dosimeters’.
Advances in Space Res. 53 (1), 1–7. https://doi.org/10.1016/j.asr.2013.10.018.
Agostinelli, S., et al., 2003. Geant4—A simulation toolkit’, nuclear instruments and Kodaira, S., et al., 2016. ‘A performance test of a new high-surface-quality and high-
methods in physics research section a: accelerators, spectrometers. Detectors and sensitivity CR-39 plastic nuclear track detector – TechnoTrak’. Nuc. Instrum.
Associated Equipment 506 (3), 250–303. https://doi.org/10.1016/S0168-9002(03) Methods in Phys. Res. Section B: Beam Interactions with Materials and Atoms 383,
01368-8. 129–135. https://doi.org/10.1016/j.nimb.2016.07.002.
Allison, J., et al., 2006. Geant4 developments and applications’. IEEE Trans. Nucl. Sci. 53 Matthiä, D., et al., 2013. ‘A ready-to-use galactic cosmic ray model’. Advances in Space
(1), 270–278. https://doi.org/10.1109/TNS.2006.869826. Res.. Pergamon 51 (3), 329–338. https://doi.org/10.1016/j.asr.2012.09.022.
Allison, J., et al., 2016. Recent developments in Geant4’, nuclear instruments and Miller, J., et al., 2003. ‘Benchmark studies of the effectiveness of structural and internal
methods in physics research section a: accelerators, spectrometers. Detectors and materials as radiation shielding for the International Space Station’. Radiat. Res.
Associated Equipment 835, 186–225. https://doi.org/10.1016/j.nima.2016.06.125. 159, 381–390.
Benton, E., et al., 2002. Passive dosimetry aboard the mir orbital station: internal Naito, M., et al., 2020. ‘Investigation of shielding material properties for effective space
measurements’. Radiat. Measurements. Pergamon 35 (5), 439–455. https://doi.org/ radiation protection’. In: Life Sciences in Space Research, 26. Elsevier Ltd,
10.1016/S1350-4487(02)00075-6. pp. 69–76. https://doi.org/10.1016/j.lssr.2020.05.001.
Binns, W.R., et al., 1989. ‘Charge, mass, and energy changes during fragmentation of Shavers, M.R., et al., 2004. ‘Implementation of ALARA radiation protection on the ISS
relativistic nuclei’, physical review C. Am. Phys Society 39 (5), 1785–1798. https:// through polyethylene shielding augmentation of the service module crew quarters’.
doi.org/10.1103/PhysRevC.39.1785. Advances in Space Res.. Pergamon 34 (6), 1333–1337. https://doi.org/10.1016/j.
Bradt, H.L., Peters, B., 1950. ‘The heavy nuclei of the primary cosmic radiation’, physical asr.2003.10.051.
review. Am. Phys. Society 77 (1), 54–70. https://doi.org/10.1103/PhysRev.77.54. Simpson, J.A., 1983. ‘Elemental and Isotopic composition of the galactic cosmic rays’.
Chen, W.-.C., 1997. ‘Some experimental investigations in the drilling of carbon fiber- Ann. Rev. of Nucl. Particle Sci. 33 (1), 323–382. https://doi.org/10.1146/annurev.
reinforced plastic (CFRP) composite laminates’. Int.J. Machine Tools and ns.33.120183.001543. Annual Reviews 4139 El Camino Way, P.O. Box 10139, Palo
Manufacture. Pergamon 37 (8), 1097–1108. https://doi.org/10.1016/S0890-6955 Alto, CA 94303-0139.
(96)00095-8. Somogyi, G., et al., 1976. ‘Revision of the concept of registration threshold in plastic
Crusan, J.C., et al., 2018. ‘Deep space gateway concept: extending human presence into track detectors’. Nucl. Instruments and Methods. North-Holland 134 (1), 129–141.
cislunar space’. 2018 IEEE Aerospace Conference. IEEE, pp. 1–10. https://doi.org/ https://doi.org/10.1016/0029-554X(76)90133-6.
10.1109/AERO.2018.8396541. Tawara, H. et al. (2011) ‘Characteristics of Mg2SiO4:tb (TLD-MSO-S) relevant for space
Cucinotta, F.A., 2014. ‘Space radiation risks for astronauts on multiple international radiation dosimetry’, Radiation Measurements. Elsevier Ltd, 46(8), pp. 709–716.
space station missions’. PLoS ONE. Edited by P. J. Janssen. Public Libr.Sci. 9 (4), doi: 10.1016/j.radmeas.2011.05.058.
e96099. https://doi.org/10.1371/journal.pone.0096099. La Tessa, C., et al., 2005. ‘Fragmentation of 1 GeV/nucleon iron ions in thick targets
Cucinotta, F.A., 2015. ‘Review of NASA approach to space radiation risk assessments for relevant for space exploration’. Advances in Space Res.. Pergamon 35 (2), 223–229.
mars exploration’. Health Phys. 108 (2), 131–142. https://doi.org/10.1097/ https://doi.org/10.1016/j.asr.2005.02.007.
HP.0000000000000255. Zeitlin, C., et al., 1997. ‘Heavy fragment production cross sections from 1.05 GeV/
DeWitt, J.M., et al., 2009. ‘Assessment of radiation shielding materials for protection of nucleon 56Fe in C, Al, Cu, Pb, and CH2 targets’. Phys Rev C Nucl Phys 56 (1),
space crews using CR-39 plastic nuclear track detector’. Radiat. Measurements. 388–397. https://doi.org/10.1103/PhysRevC.56.388.
Pergamon 44 (9–10), 905–908. https://doi.org/10.1016/j.radmeas.2009.10.041. Zeitlin, C. et al. (2006) ‘Measurements of materials shielding properties with 1 GeV/nuc
Doke, T., et al., 1995. ‘Estimation of dose equivalent in STS-47 by a combination of TLDS 56Fe’, Nuclear Instruments and Methods in Physics Research Section B: Beam
and CR-39’. Radiat. Measurements. Pergamon 24 (1), 75–82. https://doi.org/ Interactions With Materials and Atoms. North-Holland, 252(2), pp. 308–318. doi:
10.1016/1350-4487(94)00084-E. 10.1016/j.nimb.2006.08.011.
Durante, M., 2014. ‘Space radiation protection: destination mars’. Life Sci. Space Res. Zeitlin, C., et al., 2011. ‘Fragmentation of N14, O16, Ne20, and Mg24 nuclei at 290 to
Elsevier 1, 2–9. https://doi.org/10.1016/j.lssr.2014.01.002. 1000 MeV/nucleon’. In: Physical Review C - Nuclear Physics, 83. American Physical
Durante, M., Cucinotta, F.A., 2011. ‘Physical basis of radiation protection in space Society. https://doi.org/10.1103/PhysRevC.83.034909.
travel’, reviews of modern physics. Am. Phys. Society 83 (4), 1245–1281. https:// Zeitlin, C., et al., 2013. ‘Measurements of energetic particle radiation in transit to Mars
doi.org/10.1103/RevModPhys.83.1245. on the Mars Science Laboratory’. Sci.. Am. Association for the Advancement of Sci.
George, J.S., et al., 2009. ‘Elemental composition and energy spectra of galactic cosmic 340 (6136), 1080–1084. https://doi.org/10.1126/science.1235989.
rays during solar cycle 23’. In: The Astrophysical Journal, 698. IOP Publishing,
pp. 1666–1681. https://doi.org/10.1088/0004-637X/698/2/1666.

79

You might also like