Content uploaded by E. N. Slyuta
Author content
All content in this area was uploaded by E. N. Slyuta on Mar 03, 2023
Content may be subject to copyright.
Helium-3 in the Lunar Soil
E. N. Slyuta and O. I. Turchinskaya
1 Introduction
Ionizing irradiation by the solar wind and galactic cosmic
radiation is one of the main factors of cosmic weathering and
accumulation of cosmogenic isotopes in the lunar soil. The
study of the gross isotope composition of the solar wind has
a long history and practically began with the launching of
the first interplanetary automatic stations at the dawn of the
space age. For the first time, the ion component of solar wind
streams outside the Earth's magnetosphere was detected and
measured during the Luna-2 mission in September 1959.
These measurements finally confirmed the hydrodynamic
theory of the solar wind proposed in 1958 by the American
physicist Eugene Parker [1], which was immediately criti-
cized by physicists.
The solar wind is formed during solar corona gasdynamic
expansion into space and is a plasma flow of approximately
the same density, consisting mainly (about 95%) of electrons
and hydrogen ions (protons) [1,2]. In addition to protons,
helium ions are present in the solar wind, whose ratio to
hydrogen ions ranges from 0.037 to 0.056 [3,4]. Atoms of
Ne, Ar, Kr, and Xe and oxygen, silicon, sulfur, iron ions, etc.
are also present in small amounts. With the distance from the
Sun the solar wind velocity increases and its flux density
decreases. At the Earth's orbit, its speed is about 500 km s
−1
.
The temperature of protons (or electrons) is about 100,000 K
and concentration 10–20 particles per cubic centimeter. The
solar wind characteristics on the Earth's orbit vary depending
on the solar activity level. The velocity of any ions is the
same, and the ion energy depends on the nucleus mass in
accordance with the hydrodynamic theory of the solar wind
propagation [1,5]. Accordingly, at the same ion velocity, the
kinetic energy of helium nuclei is 4 times greater than proton
energy. The normal proton energy on the Earth's orbit varies
from 0.3 to 3 keV, reaching 100 keV and more during solar
flares.
Solar Wind Composition (SWC) experiments to study the
gross isotopic composition of the solar wind directly on the
lunar surface have been conducted since 1969 in almost all
manned Apollo missions except Apollo 17. A thin aluminum
foil measuring 30 140 cm (4000 cm
2
area), 15 lm thick,
and weighing 127 g for direct measurement of captured solar
wind elements was unfolded on a L-shaped support at some
angle to the lunar surface and perpendicular to the solar wind
flow. The exposure time of each subsequent experiment
increased from 77 min at Apollo 11 to 45 h 05 min at
Apollo 16 [6–10]. After completing the experiment, the foil
was rolled up, placed in a container, and transported to Earth
for further study under laboratory conditions. According to
SWC experiments, the average ratio of helium isotopes
4
He/
3
He in the solar wind near the lunar surface was
2350 ±120 [11].
The study of the bulk composition of helium isotopes in
the solar wind continued in 1978–1982 in the International
Sun Earth Explorer 3 (ISEE-3) mission (the Ion Composition
Instrument (ICI)) [12,13] and in 1991–1996 in the Ulysses
mission using the Solar Wind Ion Composition Spectrometer
(SWICS) [14–16], and in the Solar and Heliospheric
Observatory (SOHO) and Advanced Composition Explorer
(ACE) missions [17,18]. Analysis of the data obtained
showed that the ratio of helium isotope ions
4
He/
3
He in the
slow and fast solar wind fluxes is different and varies from
2450 to 3030 [15], while the average ratio of helium isotopes
in the outer convective zone of the Sun (in the corona and
chromosphere) is 2632 [19]. From 2001 to 2004, the Genesis
spacecraft exposed solar wind traps made of various ultra-
pure materials at the Lagrangian point L1 at a distance of 1.5
million km from Earth for 852.83 Earth days, which were
then delivered to Earth [20,21]. According to these mea-
surements, the average ratio of helium isotopes in the bulk
composition of the solar wind is 2155 [22].
E. N. Slyuta (&)O. I. Turchinskaya
Vernadsky Institute of Geochemistry and Analytical Chemistry,
Russian Academy of Sciences, 19 Kosygina St., Moscow, 119991,
Russia
e-mail: slyuta@geokhi.ru
©The Author(s), under exclusive license to Springer Nature Switzerland AG 2023
V. P. Kolotov and N. S. Bezaeva (eds.), Advances in Geochemistry, Analytical Chemistry, and Planetary Sciences,
https://doi.org/10.1007/978-3-031-09883-3_21
375
The purpose of studying the gross isotope composition of
solar wind isotopes was to estimate the average solar ele-
ment composition of matter, because due to its mass the Sun
represents 99.8% of the Solar system composition, and to
check the theory of proton-proton cycle of thermonuclear
fusion reaction in the bowels of our star on the basis of
helium isotope ratio. The study of isotopic composition of
implanted noble gases in the lunar soil besides the above
mentioned tasks allows to estimate the change of solar wind
composition since the formation of lunar rocks and regolith
and on the basis of Ar and Xe isotopic composition to
estimate the age of lunar regolith exposure on the surface
and trace the geochemical history of its formation [23]. Since
the 1986 publication of [24], helium-3 isotope has also been
considered as a possible promising source for thermonuclear
energy based on the “pure”thermonuclear reaction
D+
3
He = p (14.68 meV) +
4
He (3.67 meV), where p is a
proton [25–27]. On Earth, helium-3 is observed in trace
amounts, i.e., on the verge of its practical absence. Rare in
nature, the gas is mainly formed because of the radioactive
decay of tritium produced in nuclear reactors and during the
storage of nuclear and thermonuclear weapons. This gas is
collected in quantities of several tens of kilograms per year
and is commercially available for use in low temperature
physics and in neutron detectors used for both scientific and
security purposes [28,29].
In contrast to the Earth, helium isotopes and other noble
gases and cosmogenic isotopes have accumulated in parti-
cles and minerals of the lunar regolith throughout the geo-
logical history of the Moon because of constant irradiation
by the solar wind. Direct measurements of concentrations of
helium isotopes and other noble gases from the solar wind
trapped by regolith particles have been performed for all
regolith samples taken from different regions of the Moon by
Soviet lunar stations Luna-16, Luna-20, Luna-24, and
manned Apollo missions. The study of regolith samples
from the Apollo 11 expedition has revealed almost all basic
features of helium isotope concentration and distribution in
particles and minerals of lunar regolith, which were con-
firmed by the study of samples from other stations and
expeditions. It was found that the content and isotope ratio
of implanted helium in the lunar soil depend on the size of
the regolith particles: the smaller the size, the greater the
concentration and the smaller the
4
He/
3
He isotope ratio [30,
31]. The concentration also increases with increasing age of
exposure and the maturity regolith, i.e., it depends on the
ionizing irradiation dose. However, the greatest change in
the concentration of implanted helium is observed depend-
ing on the chemical and mineral composition of the regolith
particles and can vary in a very wide range, up to 3 orders of
magnitude and more [32]. The maximum content of
helium-3 is observed in ilmenite (FeTiO
3
), which is the main
ore mineral of titanium in mare lunar basalts. Selective
enrichment by solar wind ions of regolith depending on the
mineral composition leads to heterogeneous regional distri-
bution of helium isotope concentration and other noble gases
in lunar soil [33–35].
To estimate the concentration, distribution and predicted
helium-3 stocks in the lunar soil on the Near and Far sides of
the Moon, different models based on remote optical sensing
of distribution of chloroform elements TiO
2
and FeO with
correlation of helium-3 content in regolith, as well as the
degree of soil maturity and solar wind flux density at dif-
ferent latitudes [35,36], and model regolith thickness
[34,37] were used. The purpose of this work is studying the
concentration and distribution of helium isotopes in the lunar
soil based on direct measurements of helium content in
different minerals, in the bulk composition of regolith in
areas of distribution of mare low-titanium and high-titanium
basalts and Highlands’s anorthosites, as well as estimating
predicted helium reserves in different geological complexes
with different bulk composition of regolith and underlying
rocks, different helium content and different regolith
thickness.
2 Helium Implantation and Trapping
in Minerals of the Lunar Regolith
Lunar soil consists of bedrock fragments and such particles
as glass, breccias, agglutinates, which were formed during
meteoritic bombardment. The main rock-forming minerals
are anorthite, bitovnite (basic plagioclase), olivine and
pyroxenes. The main ore mineral in mare basalt is ilmenite.
The regolith contains a significant amount of glass, mainly
of impact origin. The regolith composition is mainly defined
by underlying rocks composition, with the addition of
impact craters ejects from other areas, and meteoritic com-
ponent (*1 to 2 wt%). The regolith depletion in volatile
components due to impact melting and evaporative differ-
entiation is observed [38]. Constant influence of solar wind
and cosmic radiation enriches particles and minerals of
regolith with hydrogen, helium isotopes, rare gases, cos-
mogenic isotopes, and probably, promotes formation of
reduced forms Fe, Ti, Si and other elements in surface layers
of mineral particles. According to experimental data,
metallic iron Fe
0
may form without participation of reducing
agent (implanted hydrogen ions of solar wind), as well as
bypassing process of iron condensation from impact-formed
steam. Formation of chains of nanophase iron spherules in
glasses and on surfaces of particles and minerals can occur
thermo-reductive way during shock wave passage in shock
melt because of meteorite and micrometeorite bombardment
[39].
The general structure of the noble gas content in the lunar
soil quite corresponds to the predicted cosmic prevalence of
376 E. N. Slyuta and O. I. Turchinskaya
these gases. The observed saturated
4
He concentration on the
surface of ilmenite grains equal to 10
–3
cm
34
He cm
−2
(3 10
16
atoms
4
He cm
−2
) corresponds to the integral solar
wind flux of 7 10
17
atoms cm
−2
[30]. Such a dose of
irradiation corresponds to an exposure time of about
300 years at the present intensity of the solar wind flux. At
such radiation levels, heavy saturation effects are expected,
which are likely to affect the relative content of captured
solar wind ions. For example, the maximum experimental
dose of saturation with protons on the quartz surface is
observed at 5 10
16
atom cm
−2
[40].
Lunar soil has a complex layered structure and its own
stratigraphic and geochemical history, which can be traced
back to the formation of underlying rocks (Fig. 1). An
average lunar soil formation rate is approximately 1.5 mm
per 1 Ma [41]. Layers are a stratified sequence of unstable
along strike deposits of ejections from the surrounding
impact craters, spaced from the first meters to several hun-
dred kilometers, depending on size of these craters. There-
fore, a layer of more mature (older) regolith can partially
overlap a layer of immature young regolith. Hyperbolic
distribution of lunar soil density with depth and excessive
compaction and over compaction of regolith not deep from
the surface is explained by micrometeorite and fine meteorite
bombardment, which loosens the surface layer and, on the
contrary, compacts underlying regolith layers [42]. Com-
plete mixing occurs only in the uppermost layer with a
thickness of about 1 mm, until a new ejection covers this
layer and removes it from the exposure zone by solar wind
and exogenous recycling. During this time, the 0.1 mm thick
layer will turn over more than 2000 times, the 1 mm thick
layer will turn over 250 times, and the 1 cm thick layer will
turn over once in 10
7
years [43].
Each layer can differ in bulk chemical and mineral
composition, particle size, regolith maturity, and exposure
age. At the same time in the layer itself there may be par-
ticles and minerals of lunar rocks not only of different
composition, but also age of formation. The solar wind
exposure age on the lunar soil determined by maturity index
is about 200 Ma [44]. Sometimes the lunar soil exposure age
exceeds this value twice and more. The soil exposure age
buried in 62 cm depth in the Taurus-Litre Valley (Apollo
17) is estimated to be 550 ±60 Ma [45]. Thus, the irradi-
ation dose of minerals and regolith particles by protons and
helium significantly exceeds the saturation dose.
Depending on the mechanism of retention in lunar soil of
trapped noble gases from solar wind two main types of
helium are distinguished: weakly bound, saturating pore
space of lunar regolith, and implanted in particles and
minerals of lunar soil. Due to the high diffusion and evap-
oration rate of weakly bound helium, which reaches 10
2
s
[34], most of the helium escapes from the lunar regolith.
Nevertheless, the content of the remaining helium exceeds
the solubility of helium in the lunar regolith by 15–20 orders
of magnitude under the conditions of rarefied lunar atmo-
sphere [46], which indicates a certain ability of the lunar soil
to retain implanted volatile components.
The content of weakly bound helium in the lunar soil is
determined by diffusion, the rate of which is determined by
the temperature and different densities of the solar wind flux
at different latitudes [34]. The accumulation of weakly
bound helium in the lunar regolith largely depends on the
surface temperature and does not depend on regolith com-
position. The content of this helium type substantially
increases at high lunar latitudes characterized by low diurnal
temperatures and the minimum amplitude of the diurnal
temperature variations. It is assumed that the content of
weakly bound isotope
3
He in the Polar Regions of the Moon
can reach 40–50 ppb [34,36]. This type of helium, as
compared to the implanted component, is characterized by a
more uniform distribution, the highest availability, and,
apparently, the highest concentration in the Polar Regions.
Weakly bound helium is easily released under thermal and
mechanical effect on the lunar soil and requires the devel-
opment of special methods for in situ research. The weakly
bound type of gases is one of the least studied in the lunar
regolith [47].
In contrast to weakly bound gases, implanted gases are
stable under mechanical and temperature effects on regolith
within several hundred degrees and can be sampled and
delivered to the Earth together with lunar soil without special
losses and studied in laboratory conditions. This property
makes this type of gas the most studied type. Direct mea-
surements of concentrations of noble gases trapped by
regolith particles from the solar wind composition were
performed for all regolith samples delivered by Soviet
automatic lunar stations and operated Apollo expeditions.
The study of
3
He concentration in size fractions of regolith
showed a clear dependence of concentration on the size of
fraction—the smaller the size of regolith particles, the
greater the concentration of helium-3. The observed distinct
inverse proportional correlation of gas content depending on
regolith grain size (Table 1) is approximated by the
expression
c1dn;
where cis a gas concentration, dis a particle size, and an
exponent nis estimated in the range from 0.58 to 0.65 for all
noble gases [30]. Lunar regolith consists of 50% of
particles < 50 µm in size [48], which contain up to 80%
of all trapped helium [49]. Thus 90% of the trapped helium
in lunar regolith is contained in fraction of size < 100 l[31].
For ilmenite grains also observed a clear dependence of
the content of trapped gases on the grain size. The exponent
Helium-3 in the Lunar Soil 377
Fig. 1 Schematic section of the structure of the loose layer of lunar
regolith with a borehole in the center. Parallel shading shows the
layered structure of the regolith. The underlying rocks are shown by
cross-hatching. Above is a photo panorama from the Apollo 17 landing
site (Credit NASA)
378 E. N. Slyuta and O. I. Turchinskaya
(n) for grains of ilmenite mineral (FeTiO
3
) ranges from 1.15
to 0.82. This dependence of a gas content on the grain size
can be explained by concentration of gases in the surface
layer of individual grains, i.e., the smaller the particle size,
the greater the particle surface relative to the soil density.
The dependence degree of gas content on grain size for
ilmenite is higher (Table 2) than for bulk regolith compo-
sition and is approximately constant per unit particle surface.
This means that saturation time of regolith with noble gases
does not depend on size of regolith particles and is deter-
mined only by the exposure time. Helium content in frac-
tions of approximately the same size of ilmenite in
comparison with a similar fraction of regolith, where ilme-
nite in its natural percentage is also present, is almost 4 times
higher (Tables 1and 2). Separate particles of regolith, for
example, glass spherules, which possess an amorphous
structure, in relation to bulk composition of regolith are
characterized by even greater intensity of diffusion losses of
helium and contain the smallest amount of trapped helium in
comparison with other rock-forming minerals [30].
Using the method of successive etching of the surface of
ilmenite particles, it was found that practically all trapped
helium and other noble gases are concentrated in the surface
layer 200 nm thick (Table 3). The content of the isotope
3
He
formed as a result of fission of radioactive element nuclei in
mineral particles of regolith was determined in heavily
etched samples of ilmenite grains, in which the surface layer
containing the implanted isotope
3
He was removed. The
amount of helium-3 formed by nuclear reactions was esti-
mated to be 0.7 ±0.2 ppb, which is generally in good
agreement with the regolith age estimated from xenon decay
[30]. Taking into account this value, the corrected ratio of
implanted
4
He/
3
He isotopes in ilmenite is within 2590–2840
and averages 2720 ±90. For regolith bulk composition for
different size fractions the isotopes ratio is not homogeneous.
For coarse fractions the measured isotope ratio
4
He/
3
He
exceeds 2500, and with correction this value increases about
10% and agrees with the value characteristic of ilmenite
[30]. The isotope ratio for thin regolith fraction is less than
2400. The isotope ratio correction for this fraction is less
than 2%. The difference can be explained either by the effect
of grain size or by a systematic difference in mineral com-
position of different fractions. The addition of the
4
He iso-
tope because of decay of U and Th over 4.5 billion years is
negligibly small compared to the implanted helium.
The implanted helium content is determined by compo-
sition and structure of minerals in lunar regolith and varies in
a very wide range up to three orders of magnitude or more.
Crystal lattice radiation-induced defects trap and retain
helium ions much efficiently than ones in the amorphous
structure retain [50]. The binding energy of a trapped helium
atom in vacancy traps of the crystal lattice is determined by
the occupied defect type and reaches more of 1 eV [51,52].
The theoretical calculations using the SRIM2010 pro-
gram [53] showed that the maximum ionization energy
losses of hydrogen ions (dE/dx) during interaction with
quartz should occur at 80–90 keV energy (Fig. 2)[54].
When hydrogen ions with energies less than 80 keV and
more than 90 keV are irradiated, the ionization losses in the
interaction of hydrogen ions with quartz decrease. This is
explained by the fact that at low velocities of ionizing
radiation particles, when vu, where uis the mean
velocity of electron in atom, the overcharge effect becomes
significant, which consists in the fact that the passing particle
captures (and sometimes, on the contrary, loses) electrons
and becomes neutral [55]. Overcharging leads to a reduction
of ionization losses. At high energies, the effect of electric
polarization of the medium under the action of the particle
Table 1 Distribution of helium
isotopes by size fractions of
regolith from the Apollo 11
landing site [30]
# Fraction, µm Content
3
He, ppb Content
4
He, ppm
1 130 4.6 11.5
2 90 5.6 14.5
3 42 6.8 17.8
4 15 15.8 40.6
5 3.7 37.2 89.3
6 2.0 64.2 153.4
7 1.4 95.0 221.4
Table 2 Content of helium
isotopes in different fractions of
ilmenite from the Apollo 11
landing site [30]
# Fraction, µm Content
3
He, ppb Content
4
He, ppm
1 105 7.5 19.4
2 65 15.3 37.9
3 41 22.2 59.3
4 22 45.6 120.4
Helium-3 in the Lunar Soil 379
field is effective. Polarization leads to sharp reduction or, as
they say, to screening of the particle's field and, conse-
quently, to reduction of losses.
At low irradiation energies and doses at which hydrogen
ions saturation of quartz samples surface occurs, rapid
destruction of quartz crystal lattice and formation of amor-
phous film on the crystal surface [54]. At energies close to
the maximum ionization loss energy of hydrogen ions in
quartz about 90 keV, as X-ray diffraction analysis shows,
inter-nodal defects develop to a depth of 1 µm, which lead to
the increase of the quartz lattice parameter Δd/d [54]. In this
case, the amorphous layer is not formed, but the micro-
crystalline structure of the disturbed layer is preserved.
Unlike quartz, in ilmenite at low irradiation energies, the
microcrystalline structure of the disturbed layer is preserved,
in which radiation defects of the crystal lattice continue to
develop [56], which increases the adsorption capacity and
reduces the degree of diffusion losses of trapped ions.
Amorphization of the crystal structure of ilmenite begins at
energies of about 120 keV [57], which is approximately an
order of magnitude higher than that of quartz and alumi-
nosilicates [58]. Unlike rock-forming minerals of dielectrics,
the main ore mineral of lunar basalts ilmenite is a semi-
conductor with a hole conductivity [59]. It is supposed, that
presence of conductivity at interaction with charged particles
of solar wind allows keeping crystal structure to a mineral.
As in radiation-induced amorphous films of silicates and
aluminosilicates [60], the chemical composition of the dis-
turbed microcrystalline layer of ilmenite also changes. At the
boundary with the unchanged phase of the host mineral,
compared with its composition, iron depletion is observed,
while in the upper part of the disturbed layer, on the con-
trary, enrichment with reduced metallic iron [57]. Titanium
in the disturbed layer also changes its valence from Ti
+4
to
Ti
+3
and Ti
+2
. When the ions destroy the bonds of the
ilmenite crystal lattice, a 2–threefold decrease in the O/Fe
and O/Ti ratios is also observed, indicating a significant loss
of oxygen.
Under influence of ionizing irradiation in amorphized
layer of silicates, aluminosilicates and oxides, apparently,
radiation-induced fractionation of isotopes Fe and, probably,
isotopes Ti and other elements can occur. During irradiation
of quartz samples with energy of 20 and 110 keV by protons
with implanted
54
Fe
+
and
56
Fe
+
isotopes it has been founded,
that not only general shift of Fe isotope concentration in
depth on the quantity about 20 nm occurs, but at enough
high energy of hydrogen ions (65–200 keV) radiation-
induced fractionation of Fe isotopes is observed [61]. The
value of the radiation-induced fractionation of Fe isotopes
(5 nm) is approximately equal to the observed value of the
kinetic mass fractionation (5.8 nm), but with the opposite
sign.
The metallic Fe
0
content in the amorphous mineral layer
is determined by the signal intensity of ferromagnetic reso-
nance (I
s
)[62]. The ratio of the signal intensity I
s
to the total
FeO content in regolith by weight % characterizes the
regolith maturity degree. The parameter (TiO
2
I
s
/FeO) is
used to describe the total dependence of the
3
He content on
the ilmenite content and the regolith maturity (Fig. 3)
3He ppbðÞ¼0:2043 TiO2Is=FeOðÞ
0:645;
The formation of an amorphous layer on the mineral surface
as a result of ionizing irradiation, apparently, leads to
decrease in the efficiency of traps and, accordingly, to a
decrease in the helium content. Implantation depth of solar
wind ions is determined by the ion mass, their energy and
the target composition (Fig. 4). If hydrogen ions implanta-
tion with energy of 20 keV into quartz does not exceed
250 nm, the implantation depth of hydrogen ions with
Table 3 Distribution of implanted helium in the surface layers of ilmenite grain [30]
# Fraction, µm Layer thickness, µm Content
3
He, ppb Content
4
He, ppm
141 –27.0 69.4
2 41 0.16 19.9 52.7
3 41 0.19 10.8 26.4
4 41 0.35 3.3 6.7
Fig. 2 Ionization energy loss of hydrogen ions (dE/dx) related to
interaction with quartz as a function of irradiation energy calculated
using the SRIM2010 program [53]
380 E. N. Slyuta and O. I. Turchinskaya
energy of 90 keV increases three times and reaches 850 nm
[32]. The similar distribution of hydrogen on depth
depending on irradiation energy is observed in crystals of
silicon, and the value of average run of hydrogen ions with
energy about 100 keV also makes about 1 l[40].
Calculations with the program SUSPRE [63] show that
with the increase of energy from 1 to 3 keV the depth of
proton implantation in anorthite (Ca(Al
2
Si
2
O
8
) increases
three times regardless of the flux density, whereas the degree
and depth of amorphization much more depends on the flux
density [58]. If at a flux density of 1 10
15
atom cm
−2
, the
implantation depth outpaces the development and depth of
amorphization, then at a flux density of about 1.6 10
17
atom cm
−2
the implantation is actually localized inside an
already well-developed amorphous layer (Fig. 5). The same
tendencies are observed in the case of irradiation by He
+
ions
with the only difference that the thickness of the amorphous
layer at high flux density coincides with the maximum depth
of He
+
implantation. At the same time the implantation
depth and thickness of the amorphous layer at helium irra-
diation is approximately two times greater than at proton
irradiation (Fig. 6).
At increase of energy from 1 to 3 keV the depth of proton
implantation in quartz (SiO
2
) as well as in anorthite increases
almost three times regardless of flux density. The extent and
depth of amorphization of quartz surface under the action of
ionizing irradiation depends mainly not on energy, but on
flux density. At flux densities of 1 10
16
atom cm
−2
and
less, the implantation depth greatly exceeds the development
and depth of amorphization, regardless of the proton energy.
At doses, exceeding 1 10
16
atom cm
−2
the implantation is
actually localized inside the forming amorphous layer
Fig. 3 Content of He depending on the parameter (TiO
2
I
s
/FeO)
[28]
Fig. 4 Dependence of the implantation depth of protons and He
+
ions
in quartz and anorthite on the ions’energy [32]: 1—implantation of
protons in quartz; 2—implantation of protons in anorthite; 3—
implantation of He nuclei in quartz; 4—implantation of He nuclei in
anorthite
Fig. 5 Protons’implantation depth and amorphization degree (disor-
der) in anorthite depending on protons’energy and a dose for 1 keV
D = 0.96 10
17
atom cm
−2
, for 2 keV D = 1.3 10
17
atom cm
−2
,
for 3 keV D = 1.6 10
17
atom cm
−2
:adistribution of the implanted
protons’concentration; bamorphization degree (disorder) of an
anorthite lattice
Helium-3 in the Lunar Soil 381
regardless of the proton energy (Fig. 7). The complete
(100%) amorphization of quartz surface is developed at the
doses exceeding 8 10
16
atom cm
−2
, and at flux density of
1.6 10
17
atom cm
−2
, the thickness of amorphous layer on
the quartz surface achieves 100 nm (Fig. 7). When irradiated
with He
+
ions, the depth of quartz structure destruction and
amorphization, regardless of energy and dose, always coin-
cides with the maximum ion implantation depth (Fig. 8).
Intensity of amorphization is also mainly defined by a dose
of irradiation (Fig. 8).
Thus, at dose increase up to integral value of flux
(7 10
17
atom cm
−2
) observed in lunar regolith, all
chemical and physical processes of space (geochemical)
weathering of anorthite and quartz particles in lunar regolith,
apparently, will occur inside the formed radiation-induced
amorphous film on crystal surface, not reaching the crystal
structure of mineral. In contrast to protons, the maximum
concentration of He
+
ions regardless of dose in both anor-
thite and quartz will accumulate in the transition zone
between the amorphous and crystalline phases of the
mineral.
The lowest content of noble gases in impact and volcanic
glasses is observed. Helium-3 in green glass spheres in lunar
soil samples of Apollo-15 is in the range of 0.02–0.3 ppb
(average for all size fractions about 0.2 ppb), and Helium-4
in the range 7–605 ppb (average for all size fractions 100–
200 ppb) [64–66]. It is about 100 times less, than in the bulk
regolith composition.
Similar small helium content in basic plagioclase are
observed, where
3
He content varies in the range 0.02 to
0.2 ppb, and
4
He from 33 to 500 ppb [66,67]. As noted
above, for all crystalline minerals of lunar regolith, including
plagioclase, there is an inverse proportional dependence of
helium concentration on particle size—the smaller the size,
the higher the concentration [30,66,68]. In contrast, a direct
proportional dependence, although not as pronounced, is
observed for green glasses: the smaller the particle size, the
smaller the helium concentration [64,66]. Probably, this
dependence is explained by the inclusions of mafic minerals
in larger glasses. Helium content in magnetic glasses with
Fig. 6 Implantation depth of He
+
and amorphization degree (disorder)
in anorthite depending on ions energy and a dose for 4 keV
D=210
16
atom cm
−2
, for 8 keV D = 2.8 10
16
atom cm
−2
, for
12 keV D = 3.5 10
16
atom cm
−2
:adistribution of the implanted
helium concentration; ban amorphization degree (disorder) of an
anorthite lattice
Fig. 7 Protons’implantation depth and amorphization degree (disor-
der) in quartz at constant energy of 3 keV and a doses D = 1 10
16
atom cm
−2
,D= 8 10
16
atom cm
−2
, and D = 1.6 10
17
atom cm
−2
:
adistribution of the implanted protons’concentration; ban amor-
phization degree (disorder) of a quartz lattice
382 E. N. Slyuta and O. I. Turchinskaya
large number of such inclusions significantly increases
compared to green glasses [67].
The helium highest content in ilmenite are observed. In
grains of ilmenite sized 10–14 µm, the
3
He content can
reach 130 ppb, and
4
He—360 ppm [68]. Helium isotopes
content in bulk regolith composition of the same fraction in
lunar soil samples of Apollo 12 is 12 ppb of
3
He and
30 ppm of
4
He.
3
He content in bulk regolith composition
without separation into size fractions is almost two times less
and is 7.1 ppb and
4
He is 16.5 ppm.
The
3
He content in olivine with a grain size of 25–42 lm
is 2.9 ppb and
4
He is 6.24 ppm [65]. In the mixed fraction of
olivine and pyroxene grains with a grain size of 25–42 lm,
the helium isotope content is similar: 2.9 ppb and 6.2 ppm,
respectively [66]. However, the pure pyroxene fraction with
approximately similar grain size of 30–48 µm contains
implanted helium two times more (5.7 ppb
3
He and
13.4 ppm
4
He) than olivine [67].
Thus, helium content in glasses and plagioclases of the
lunar soil is approximately 100 times less than in bulk
regolith composition, but helium content in ilmenite, on the
contrary, is approximately 10–12 times more than in bulk
regolith composition of similar fraction and about 20 times
more in comparison with bulk regolith composition without
separation into fractions. Olivine and pyroxene are charac-
terized by the medium helium content compared to glass and
ilmenite. Obviously, under equal irradiation conditions, in
which the ilmenite grains were together with other minerals
of the regolith, the different helium content is due to the
different mineral composition and structure of the target, i.e.
due to the variety and number of defects-traps.
3 Helium Content in Regolith of Lunar Maria
and Highlands
Mare rocks are represented mainly by lava rocks of the basalt
type, which fill the depressions of lunar mares. Pyroclastic
material is also found. In general, mare rocks compose about
1% of the lunar crust. The predominant mare rocks of the
Moon are mare basalts. In bulk chemical composition marine
basalts correspond to rocks of gabbro-basalts group and are
usually defined as ilmenite, olivine, cristobalite basalts (gab-
bro), etc. [69,70]. The petrographic specificity of mare basalts
is in the smaller grain size (hundreds of microns) and in the
practical absence of volcanic glass. The major rock-forming
minerals are clinopyroxenes and plagioclase (An50-95),
sometimes olivine (Fo50-75); the major ore mineral is ilmenite
(FeTiO
3
). The ilmenite content varies from 0.1–1wt%in
low-titanium basalts [71,72]to10–20 wt% in high-titanium
basalts [73]. Formation of mare basalts which cover about
20% of the Moon's surface [74] is connected with the pro-
cesses of partial mantle melting, and the difference in mineral
composition of basalts, apparently, with different depth of
melt formation, and/or lateral changes in composition of lower
crustal layers and upper mantle [41,75].
Lunar pyroclastic material is a rare type of lunar marine
formations [76,77]. Pyroclastic material is represented by
green and orange glasses—mainly in the form of glass
globules and their fragments, which have no direct equiva-
lents in chemical composition among crystalline rocks.
Green glasses have a primitive mafic composition and are
considered to be the least differentiated lunar matter. The
particle surface of green and orange glasses is strongly
enriched with Zn, Pb, F and other volatile components,
which is associated with the processes of condensation of
volcanic (fumarolic) products. The glasses have an amor-
phous structure and are characterized by the lowest content
of noble gases trapped from the solar wind.
Ultrabasic rocks, which, as well as pyroclastic material,
are extremely rare, occur as xenoliths in mare basalts. The
rocks are represented by cataclased and recrystallized
dunites, and less frequently by peridotites. They consist
mainly of olivine (Fo9-12) with an admixture of pyroxenes,
Fig. 8 Implantation depth of He
+
and amorphization degree (disorder)
in quartz depending on dose of D = 1 10
15
atom cm
−2
,
D=110
16
atom cm
−2
and D = 3.33 10
16
atom cm
−2
at constant
energy of 12 keV: adistribution of concentration of the implanted
helium; ban amorphization degree (disorder) of a quartz lattice
Helium-3 in the Lunar Soil 383
plagioclase, metallic iron, and troilite. Formation of these
rocks is associated both with the early stage of global dif-
ferentiation of the Moon (age of mantle dunites is close to
the age of the Moon), and with later stages of formation of
the lunar crust [69,70].
Highlands’s rocks are represented by feldspathic rocks,
which are subdivided according to their chemical and min-
eralogical composition into two main groups: lunar anor-
thosite rocks and lunar non-maria basalts [69,70].
Structurally, these rocks are predominantly breccias with
blast and cataclase matrix, granulites, and cataclasites. They
are the oldest lunar rocks whose age—up to 4.5 billion years
—is close to the time of planetary body formation. Minimum
age of brecciation (metamorphism) is *3.9 billion years.
Formation and brecciation of Highland rocks occurred dur-
ing intense meteorite bombardment. Highlands’s rocks are
the main type of lunar crust rocks.
Among Highlands’s rocks the most widespread are rocks
of anorthosite-norite-troctolite (ANT) series, gabbro-norites
and gabbro-anorthosite. The latter are the main type of rocks
of the lunar Highlands. The rocks are characterized by high
content of Al (Al
2
O
3
> 24 wt%), Ca, Mg and some rare
elements. The main minerals are basic plagioclase (mainly
anorthite), olivine (Fo70-95), orthopyroxenes (enstatite and
bronzite), and clinopyroxenes (diopside, augite, pigeonite).
The content of ilmenite in the anorthosite rocks is usually
less than 0.1–0.2 wt% [78]. It is assumed that the rocks of
the ANT series are the product of gravitational differentia-
tion, i.e., plagioclase surfacing or precipitation of ferro-
magnesian minerals in the primary mafic melt [75].
Lunar non-mare basalts are a sparsely distributed group
of Highland rocks of the Moon with Al
2
O
3
content of 15–24
wt%. An important feature of non-mare basalts is the
increased content in some of them of the so-called
KREEP-component characterized by enrichment of K,
REE (rare-earth elements) and P (English abbreviation
K-REE-P), as well as Zr, Ba, U, Th and some other litho-
phile elements. According to this parameter, basalts with low
content of KREEP elements and KREEP-basalts are distin-
guished. In the American special literature, the “non-marine
basalt”is a medium-potassium Fra Mauro basalt. The pet-
rographic specificity of all non-mare basalts is that most of
them are breccias, which consist of clasts with a pronounced
magmatic structure with grain sizes up to hundreds of
microns. The formation of non-mare basalts is associated
with the processes of partial melting of lunar crust rocks at
relatively low pressures.
Acidic rocks of the Moon are the most exotic type of
lunar matter. Lunar acidic rocks occur as individual small
fragments in regoliths and breccias, and as cementing mas-
ses of breccias. The main minerals are pyroxene, tridymite,
cristobalite, and K-Ba-feldspar; secondary minerals are
medium plagioclase, apatite, vitlockite, and zircon. The
biggest and the most famous granite specimen is breccia
12013, delivered from the Apollo-12 landing site [79].
Formation of acidic rocks of the Moon is associated with the
process of residual melt liquation during fractional crystal-
lization of basaltic magma.
Regolith samples delivered from 9 different regions of the
Moon by Soviet lunar automated stations and American
Apollo manned expeditions have made it possible to mea-
sure the content in the bulk composition of regolith of dif-
ferent rocks—from low-titanium to high-titanium mare
basalts and in Highland anorthosite.
The observed dependence of the helium isotope content
on the size and composition of the regolith particles imposes
significant restrictions on the use of measured values to
estimate predicted reserves of this isotope in the lunar soil.
First of all, it concerns representativeness of these estimates
to reflection of total particle size in regolith and total (bulk)
composition of regolith. The content estimates obtained for
discrete size fractions of the regolith as well as for separate
components of the regolith distinguished by mineral,
chemical, or other characteristics are unsuitable for use. In
contrast, the most objective and optimal estimates are those
obtained for representative samples of regolith with a full
range of particle sizes (< 1 mm) without partitioning into
fractions and components, i.e., for bulk samples. If data are
available for several bulk samples of regolith, the weighted
average value of
3
He content is found for the mass of the
investigated sample, which is taken as the basic value for the
specified landing area.
The average weighted contents of helium in bulk samples
of regolith with particle size fraction < 1 mm from the
Apollo 11 landing site can be obtained only for two values
(16.5 and 14.6 ppb) out of four, for which the mass of the
studied sample is indicated (Table 4). The weighted average
content of the isotope
3
He, equal to 15.1 ppb, and the iso-
tope
4
He, equal to 41.1 ppm, is taken as a value character-
izing the average isotope content in the regolith in the
Apollo 11 station area and in the Mare Tranquillitatis. The
weighted average value for the data in Table 1is 20.4 ppb,
but also cannot be used for reserves assessment because it
characterizes only a few discrete size fractions of the
regolith.
The
3
He content for a representative sample of regolith
from the Apollo 12 landing area in the Oceanus Procellarum
is 7.1 ppb and
4
He is 16.4 ppm [68]. The absence of data on
the mass of the investigated regolith samples from the
Apollo-14, and -15 landing sites does not allow us to obtain
an average weighted helium content for each of the landing
sites (Table 5), but the small scatter of values allows us to
use the average value [82]. An average value of
3
He equal to
5.7 ppb is taken for Apollo-14 station and 4.4 ppb for
Apollo-15 station. The average
4
He isotope content is taken
to be 15.1 ppm and 11.1 ppm, respectively.
384 E. N. Slyuta and O. I. Turchinskaya
Average concentration of
3
He and
4
He in regolith in the
landing area of manned expedition Apollo 16 (Table 6)is
1.4 ppb and 3.9 ppm accordingly.
The content of helium isotopes in bulk samples of rego-
lith at depth to 3 m according to the data of core drilling in
the area of Apollo 17 expedition landing in the southwest
part of the Mare Serenitatis (Table 7) is characterized by
relatively uniform distribution with small variations on
depth. This is also quite consistent with the relatively uni-
form distribution of the absolute age of the regolith expo-
sure, which averages over 62 cm depth to 550 ±60 Ma
[45]. The average content of
3
He and
4
He isotopes in this
area is 8.0 ppb and 20.8 ppm, respectively.
At the Soviet Luna-16 station in the eastern part of the
Mare Fecunditatis, sampling was carried out by means of
core drilling to a depth of 35 cm. Data on helium content are
known from several sources with different degrees of rep-
resentativeness of the samples depending on the fraction size
of the samples studied (Table 8). The most representative
sample of regolith studied is sample L16-19 [84], which is
the least enriched and most fully reflects the size range of
lunar regolith particles among the values presented in
Table 8. Thus, the content of
3
He equal to 7.9 ppb and
4
He
equal to 20.1 ppm is taken as the most representative.
Core drilling to a depth of 30 cm was carried out at the
landing site of the Luna-20 station in the Highland area. The
data obtained [88] were evaluated for the thin regolith
fraction with particle size less than 83 µm, the larger regolith
fraction was withdrawn (Table 9). The 2002-1 regolith
sample was taken from a depth of 5–10 cm from the lunar
Table 4 Content of helium
isotopes in regolith samples from
the Mare Tranquillitatis at the
Apollo 11 landing site
# Sample Mass, mg Content
3
He, ppb Content
4
He, ppm
4
He/
3
He References
1 10,084.40 3.507 16.5 44.0 2665 [80]
2 10,084.40 10.827 14.6 40.2 2750 [80]
3 10,084.18 –14.1 36.0 2550 [31]
4 10,084.59 –24.3 51.8 2130 [81]
Table 5 The content of helium
isotopes in the regolith at the
Apollo 14 and Apollo 15 landing
sites [82]
# Spacecraft Sample Content
3
He, ppb Content
4
He, ppm
1 Apollo 14 14,003 5.3 15.4
14,163 5.7 14.3
14,259 6.2 15.5
2 Apollo 15 15,601 4.6 11.0
15,091 4.2 11.2
Table 6 Content of helium
isotopes in the regolith in the
Highland region at the Apollo 16
landing site [83]
# Sample Content
3
He, ppb Content
4
He, ppm
1 63,321.7 1.6 4.5
2 63,341.4 1.5 4.2
3 63,501.22 1.4 3.9
4 67,941.15 1.0 3.0
Table 7 Content of helium
isotopes in regolith in the Apollo
17 landing site [45]
# Depth, cm Content
3
He, ppb Content
4
He, ppm Exposure soil age, Ma
10–5 11.2 27.7 360
2 26 3.4 8.9 330
3 62 7.0 19.1 560
4 93 6.6 16.8 610
5 133 7.1 18.2 490
6 173 8.2 21.6 530
7 213 9.6 26.1 610
8 253 8.5 22.6 550
9 290 9.8 26.0 490
Helium-3 in the Lunar Soil 385
surface and the 2004-1 sample from a depth of 15–20 cm
from the surface. The data obtained [89] were evaluated for
the size fraction of regolith < 400 µm. Taking into account
the dependence of the helium concentration on the size
fraction of regolith particles, this sample is less enriched and,
accordingly, more representative. Therefore, the
3
He content
equal to 3.1 ppb and
4
He content equal to 8.2 ppm are taken
as characteristic for the Highland area between the Mare
Fecunditatis and the Mare Crisium at the landing site of the
Luna-20 station.
The helium content for the offshore area in the southern
part of the Mare Crisium was obtained by studying the
regolith samples delivered by the Luna-24 station. Sampling
at the station-landing site was carried out using core drilling
to a depth of 225 cm. Helium content data were obtained for
regolith fractions of size 375 lm[89] and 250 lm
[90]. Table 10 shows the weighted average values for each
regolith sample for a particular depth interval, since the
regolith samples were divided into three or four size groups
during the study. The interval-weighted average
3
He content
at the landing site of Luna-24 station at a depth of 218 cm is
3.4 ppb and
4
He content is 7.2 ppm. It should be noted that
the data for station Luna-24 are the most optimal from the
point of view of estimation of
3
He reserves as they not only
reflect almost all basic size range of regolith particles
without exclusion, but also allow to estimate average
weighted content at a depth up to 218 cm. It should be noted
that the content of implanted helium isotopes in the regolith
of low-titanium basalts in the Mare Crisium gradually
decreases with depth [92].
In general, the average ratio of
4
He/
3
He isotopes in the
regolith of different bulk composition is about 2500. Based
on the analysis of the available data on the content of helium
in the lunar regolith in the landing areas of Soviet automatic
stations and American manned expeditions, the most repre-
sentative values have been determined, which are recom-
mended for use at estimation the predicted reserves of
helium in the lunar regolith of the corresponding composi-
tion (Table 11).
Thus, the loose regolith layer in the area of distribution of
high-titanium basalts is characterized by the highest con-
centration of implanted helium isotopes and, accordingly,
other noble gases. The concentration of implanted helium
isotopes in the bulk composition of regolith decreases with
Table 8 Content of helium
isotopes in the regolith at the
Luna-16 landing site
# Sample Fraction, µm Content
3
He, ppb Content
4
He, ppm References
1 L16-19 1–400 7.9 ±0.35 20.1 [84]
2 7-1a 83 12.4 33.1 [85]
3 G-7 125 14.3 33.6 [86]
4 G-49 40–150 4.2 14.1 [87]
Table 9 The content of helium
in the regolith at the Luna-20
landing site
# Sample Content
3
He, ppb Content
4
He, ppm References
1 L2010 3.1 8.2 [89]
2 2002–1 4.0 ±0.36 10.0 [88]
3 2004–1 5.89 ±0.45 15.0 [88]
Table 10 The content of helium
in the regolith at the Luna-24
landing site
# Sample Depth, cm Content
3
He, ppb Content
4
He, ppm References
1 24,077.9 77–85 4.9 12.1 [91]
2 24,092.4 92–95 4.4 10.9 [90]
3 24,109.13 109–121 3.3 7.4 [91]
4 24,118.4 118–121 4.3 10.4 [90]
5 24,143.4 143–146 4.1 10.5 [90]
6 24,149.15 149–163 3.1 5.9 [91]
7 24,174.10 174–183 3.1 7.3 [91]
8 24,182.15 182–196 2.2 4.3 [91]
9 24,184.4 184–187 2.3 5.7 [90]
10 24,210.9 210–218 2.6 6.0 [91]
386 E. N. Slyuta and O. I. Turchinskaya
decrease of ilmenite content in basalts from high-titanium to
low-titanium. The minimum concentration of helium is
observed in Highlands’s anorthosite rocks characterized by
the minimum content of ilmenite and a lower content of
mafic rock-forming minerals olivine and pyroxenes.
4 Lunar Soil Thickness
One of the first estimates of the thickness of the loose
regolith layer according to data from automatic stations was
made at the landing site of Luna-9 station in the South-
eastern part of the Oceanus Procellarum. Based on the
analysis of the depths of the largest craters (about 1 m)
observed on the panoramic images, it was concluded that the
thickness of the regolith at the landing site exceeded 0.2 m
[93]. Attempts to directly measure the thickness of the
regolith were made in the process of sampling the regolith
by means of core drilling and core tubing at the landing sites
of the Luna-16, Luna-20, Luna-24, and Apollo manned
expeditions. The maximum drilling depth (225 cm) by the
automatic drill rig LB-9 was achieved at the landing site of
Luna-24 in the Mare Crisium [94], and the maximum dril-
ling depth (305 cm) by the hand-held perforator was
achieved at the Apollo-17 landing site in the Mare Sereni-
tatis [95]. At the Apollo-15, and -16 stations, core drilling
was performed to a depth of 236 cm and 224.3 cm,
respectively [95]. Regolith sampling with core pipes was
carried out to a depth of 70 cm at all Apollo stations [96]. In
none of the 9 sites of direct sampling of the regolith at the
landing sites, the base of the loose regolith layer was
reached.
The thickness of the regolith was also evaluated by the
size of the craters, which had uncovered the underlying
rocks at the base of the regolith. The evidence of opening of
the underlying rocks was the presence of a placer of rocks on
the rim and in the area of the ejection of such a crater. This
method of estimation was used in the analysis of the regolith
thickness according to Lunokhod-1 and Lunokhod-2 data. In
the western part of the Mare Imbrium, in the Lunokhod-1
study area, the regolith thickness was estimated to be within
1–5m [97]. Regolith thickness in the eastern part of the
Mare Serenitatis in the Lemonier Crater (Lunokhod-2) was
estimated within of 2–3m[98].
Studies of the granulometric composition of lunar rego-
lith have shown a close relationship between the parameters
of particle size distributions, regolith thickness and regolith
maturity indices, which characterize the degree of processing
of lunar regolith [99]. The standard deviation of particle
sizes has the highest correlation dependence among the
parameters of particle size distributions with regolith thick-
ness. The found dependence is defined by expression
rcm ¼1:53 þ0:102h;
where his the regional regolith thickness. The regolith
thickness estimates thus obtained at the landing sites of
Luna-16 (Eastern part of the Mare Fecunditatis) and Luna-20
(Highland area to North of the Mare Fecunditatis) were 5.3
and 11.6 m, respectively [99].
The statistical analysis of impact crater morphology and
crater population characteristics has also been successfully
applied to estimate the regolith thickness [100]. The former
is based on the dependence of crater shape on the regolith
thickness, the latter on the determination of the characteristic
crater size from which the deviation of the equilibrium crater
distribution is observed. The size of these craters at the
boundary of the equilibrium and non-equilibrium crater
populations is directly proportional to the age of the surface
and, correspondingly, the regolith thickness. Regolith
Table 11 The average content of
helium isotopes in the regolith at
the landing sites of the Soviet
automatic stations and the
American Apollo expeditions
# Spacecraft Landing site Regolith composition Content
3
He,
ppb
Content
4
He,
ppm
1 Apollo 11 Mare
Tranquillitatis
High-titanium basalts 15.1 41.1
2 Apollo 12 Oceanus
Procellarum
Basalts with medium ilmenite
content
7.1 16.4
3 Apollo 14 Fra Mauro Low-titanium basalts 5.7 15.1
4 Apollo 15 Mare Imbrium Low-titanium basalts 4.4 11.1
5 Apollo 16 Highland Highland anorthosites 1.4 3.9
6 Apollo 17 Mare Serenitatis Basalts with medium ilmenite
content
8.0 20.8
7 Luna 16 Mare
Fecunditatis
Basalts with medium ilmenite
content
7.9 20.1
8 Luna 20 Highland Highland anorthosites 3.1 8.2
9 Luna 24 Mare Crisium Low-titanium basalts 3.2 7.2
Helium-3 in the Lunar Soil 387
thickness data were obtained for 12 sites ranging from 336 to
3179 km
2
[100]. According to these estimates, the lowest
thickness (3.3 m) is observed in the Southern part of the
Oceanus Procellarum in the area of the Flamsteed crater
(landing area of the Servier 1 spacecraft). Moderate regolith
thickness (4.6 m) is observed in western and southern parts
of the Mare Tranquillitatis (Landing Region of Apollo 11),
in the southern part of the Sinus Medii (Landing Region of
SC Servier 6), in the northern part of the Mare Cognitum
(Landing Region of SC Servier 3 and manned expedition
Apollo 12), and in the central part of the Oceanus Procel-
larum (Landing Region of SC Luna-13). The increased for
mare areas regolith thickness (7.5 m) is observed in the
northern part of the Sinus Medii and in the central part of the
Mare Imbrium. The regolith thickness at the Hipparchus
crater bottom in the Highland area reaches 16 m [100].
The regolith thickness at the landing sites of the Apollo
manned expeditions was also determined by instrumental
methods using electrical and seismic surveys. As a result of
seismic experiments it was found that the lunar regolith layer
is characterized by very low velocities of elastic longitudinal
waves (80–120 m s
−1
). The landing areas of the Apollo 11, -
12 and -15 expeditions are distinguished by relatively young
surface age and moderate thickness of the regolith. The
regolith thickness in the south-western part of the Mare
Tranquillitatis in the area of the spacecraft Apollo 11 landing
according to the data of seismic experiment is estimated at
4.4 m, in the northern part of the Mare Cognitum in the
landing area of the spacecrafts Servier 3 and Apollo 12 it is
estimated at 3.7 m, and in the foothills of the Montes
Apenninus in the Hadley Rille area in the Apollo 15 landing
area it is estimated at 4.4 m [101]. The thickness of the
low-velocity layer of lunar regolith (104 m s
−1
) in the Fra
Mauro area in the area of the Apollo 14 landing, according
to instrumental data, was 8.5 m, and in the Highland area in
the area of the Apollo 16 landing, 12.2 m [102]. The regolith
thickness within the mare area in the area of the Apollo 17
station landing (intermountain plain Taurus-Littrow to the
east of the Mare Serenitatis) according to less detailed
regional seismic model is estimated in the range of 7–12 m
[102]. The regolith thickness in the area of Apollo 17 station
according to the assessment [103], also made on the basis of
seismic experiment, does not exceed 8.5 m. A step change
of the dielectric constant typical of the transition from loose
regolith to denser underlying rocks according to electrical
prospecting data in the same area was observed at a depth of
7±1.0 m [104].
A model map of the regolith thickness distribution for the
Moon Near Side based on the correlation of the depolarized
and polarized components of the radio response at a wave-
length of 70 cm with the optical albedo (0.65 lm) and the
color index (0.65/0.42 lm) with a resolution of about 3 km
in the center of the hemisphere was obtained [105], which is
in satisfactory agreement with data obtained by other
methods (Table 12). The highest thicknesses of regolith
(more than 13 m) in Highland areas are observed in the
southeastern part of the Moon Near Side in the Highland
region. In the southwest part of the Near Side in Highland
region are observed moderate thicknesses of regolith—7to
11 m. This area represents the framing of the Mare Orientale
and corresponds to the age of formation of the Mare Ori-
entale basin. The northeastern part of the Moon Near Side is
represented by the Highland region with regolith thickness
from 9 to 13 m.
The model distribution of regolith thickness has a
bimodal character with maximums of about 5 and 10 m,
which corresponds to the average values of regolith thick-
ness for mare and Highlands [105]. The regolith thickness in
mare areas varies from 3 to 11 m. A small regolith thickness
(about 3 m) in some small areas of the Mare Serenitatis and
the Mare Nubium is observed. The maximum regolith
thickness in mare areas (9–10 m) in the areas adjacent to
some large craters (e.g., Aristillus Crater) is observed.
Among the mares, the Mare Humorum is characterized by an
average minimum regolith thickness of about 4.0 m, and the
Mare Nectaris is characterized by the deepest, about 9.6 m
(Table 13).
The dispersion of regolith thickness at Highland areas is
much wider—from 1 m in the highlands to 18 m and more
in the plateau areas with the age of underlying rocks of the
Donectarian system (>3.92 billion years). The minimum
values at the bottoms of impact craters (craters Tycho,
Theophilus, Langren, Copernicus, etc.) and near Schickard
Crater are observed. The highest thickness of regolith (more
than 13 m) in the southeastern part of the Moon Near Side at
Highland areas formed at the Donectarian period (>3.92
billion years) [108] is observed. The Donectarian age
Highland rocks are distributed at the southern, southeastern,
and central part of the Moon Far Side and occupy about half
of the hemisphere. Moderate regolith thicknesses of 7–11 m
at the southwestern part of the Near Side in the Highland
region [107] are observed. This area represents the framing
of the Mare Orientale and corresponds to the age of the
formation of the Mare Orientale basin in the early Imbrian
epoch. Approximately equal in area part of this framing is
traced on the Far Side of the Moon. The northeastern part of
the Near hemisphere of the Moon is represented by Highland
of the Nectarian period with a regolith thickness from 9 to
13 m. Highlands of the Nectarian age at the Far Side
occupies about one third of the area in the northeastern and
western parts of the hemisphere.
There is a direct dependence of the regolith thickness on
the surface age, which is explained by the time of exposure
of the rocks’surface from the moment of their formation -
the greater the surface age, the greater the regolith thickness.
This dependence agrees rather well with the
388 E. N. Slyuta and O. I. Turchinskaya
Table 12 The regolith thickness
at the landing sites of unmanned
spacecrafts and Apollo
expeditions
Spacecraft Regolith thickness [105], m Regolith thickness, m References
Luna-9 4.8 > 0.2 [93]
Luna-13 3.6 4.6 [100]
Luna-16 4.0 4.0 [106]
5.3 [99]
Luna-17 5.1 1.0–5.0 [97]
4.0 [106]
Luna-20 9.2 0.4 [106]
11.6 [99]
Luna-21 3.9 2.0 [106]
2.0–3.0 [98]
Luna-24 3.0 > 2.0 [94]
Surveyor 1 3.5 3.3 [100]
Surveyor 3 5,3 4.6 [100]
Surveyor 5 4.6 –
Surveyor 6 –4.6 [100]
Apollo 11 4.7 4.6 [100]
4.4 [101]
Apollo 12 5.3 4.6 [100]
3.7 [101]
Apollo 14 8.1 8.5 [102]
Apollo 15 6.0 4.4 [101]
Apollo 16 10.1 12.2 [102]
Apollo 17 7.0 7–12 [102]
8.5 [103]
7.0 ±1.0 [104]
Table 13 Average regolith
thickness in lunar Maria [107]Mare Regolith thickness,
m
Standard deviation (r
h
), m
Mare Serenitatis 4.1 0.8
Mare Tranquillitatis 4.1 1.1
Mare Crisium 4.6 1.6
Mare Fecunditatis 5.9 1.4
Mare Frigoris 7.4 2.1
Mare Imbrium 6.2 2.0
Mare Humorum 4.0 0.9
Mare Nectaris 9.6 2.3
Mare Vaporum 5.4 1.1
Mare Nubium 5.7 1.7
Mare Cognitum 4.9 0.7
Sinus Roris 7.5 1.9
Sinus Iridium 4.6 0.6
Sinus Aestuum 6.3 1.4
Oceanus Procellarum 4.8 1.6
Helium-3 in the Lunar Soil 389
chronostratigraphic subdivisions of lunar formations [101,
109], i.e. with the age of rocks of mare and Highland
regions. Thus, based on the geological map [108] it is pos-
sible to extrapolate these data for the Far Side of the Moon,
which can also be used to estimate the predicted helium
reserves on the Moon.
5 Predicted Reserves of Helium in the Lunar
Soil
Selective enrichment by cosmogenic isotopes of regolith
depending on its mineral composition, which is determined
mainly by underlying rocks, leads to non-uniform regional
distribution of concentration of helium isotopes and other
implanted gases in lunar soil. The highest concentration of
helium isotopes, as shown above, is observed in
high-titanium mare basalts, in which the content of the main
ore mineral ilmenite reaches 20 wt%. The areas of spreading
of high contents of TiO
2
oxides (5–10%) according to the
spectral remote sensing data actually reflect the content of
ilmenite in regolith and, correspondingly, the distribution of
high-titanium mare basalts. The spectral survey method is
based on correlation of the content of the main chromophore
elements Fe and Ti for albedo with albedo and color indices
in the visible and near-IR spectral ranges [110–112]. The
percentage content of TiO
2
according to spectral survey data
varies in the range from 0.01 in Highland areas to 10% and
more in mare areas (Fig. 9).
High-titanium mare basalts are distributed in the equa-
torial region in the Mare Tranquillitatis, the Oceanus Pro-
cellarum, the Mare Humorium, the Mare Nubium, the Mare
Vaporum, and the Sinus Aestuum, and in subordinate
amounts in the Mare Imbrium, the Mare Serenitatis, and the
Mare Fecunditatis (Fig. 9). Five main zones of TiO
2
con-
centration distribution in regolith based on the titanium
oxide distribution map were identified and delineated: V—
0.01 to 1 wt%, IV—1.0 to 3.0 wt%, III—3.0 to 5.0 wt%, II
—5.0 to 8.0 wt% and I—8.0 to 10.0 wt% (Fig. 10). Each
zone corresponds to a particular composition of lunar
regolith with different ilmenite content, including Highland
anorthosites and a series of mare rocks from low-titanium to
high-titanium basalts.
The selected zones depending on regolith mineral com-
position, as it was considered above, are characterized by
different concentration of implanted helium isotopes and
different thickness of regolith (Table 14). Thus, depending
on the concentration we can distinguish five different cate-
gories of implanted helium deposits on the Moon and esti-
mate the area of their distribution. For the estimation of
Fig. 9 Map of the TiO
2
(wt%) distribution in the lunar soil on the Moon according to the spectral survey data of the Clementine spacecraft [111]
390 E. N. Slyuta and O. I. Turchinskaya
predicted reserves of implanted helium, the average density
of the lunar soil at a depth down to 10 m is taken to be
2.0 g cm
2
[48].
The measured content of
3
He and
4
He isotopes in
high-titanium basalts of category I deposits is 15.1 ppb and
41.1 ppm, respectively (Table 14). Deposits with the maxi-
mum concentration of helium are distributed extremely
heterogeneously and are located mainly in the Mare Tran-
quillitatis and the Oceanus Procellarum, and in subordinate
quantities in the Mare Fecunditatis, in the Mare Humorum,
in the Mare Nubium and in the northern part of the Mare
Imbrium (Fig. 10). These are mare flat areas with a mini-
mum average thickness of the loose regolith layer of about
4.1 m, typical of the Mare Tranquillitatis (Table 13).
Deposits of this category on the Moon Far Side are
practically absent. The total area of the richest helium
deposits is about 733,000 km
2
, the probable reserves of
3
He
are about 9.1 10
4
tons and
4
He are about 2.5 10
8
tons
(Table 15).
The measured helium-3 contents in basalts with an
average ilmenite content in category II deposits at the
Apollo-12, Apollo 17, and Luna-16 landing sites are close
and vary from 7.1 to 8.0 ppb (Table 11). To estimate the
probable reserves in mare basalts of this type, the content of
3
He and
4
He is taken to be 8.0 ppb and 20.8 ppm respec-
tively, since measurements at the landing site of the Apollo
17 expedition reflect the average content of helium isotopes
in the regolith to a depth of 3 m and are the most repre-
sentative (Table 14). Reserves of this category are wide-
spread in mare plains in the Oceanus Procellarum, in the
Fig. 10 Zonal map of TiO
2
concentration in lunar soil
Table 14 Characteristics of the
identified Categories of deposits
of implanted helium in the lunar
soil
Category TiO
2
,wt
%
Near Side,
km
2
Far Side,
km
2
3
He,
ppb
4
He,
ppm
Soil thickness
m
I 8.0–10.0 732,477 0 15.1 41.1 4.1
II 5.0–8.0 1,571,842 9264 8.0 20.80 4.8
III 3.0–5.0 1,648,869 10,283 4.4 11.1 4.1
IV 1.0–3.0 2,432,985 92,484 3.1 8.2 9.2
V 0.01–1.0 12,613,827 18,887,969 1.4 3.9 10.1
Helium-3 in the Lunar Soil 391
Mare Imbrium, in the Mare Fecunditatis, in the Mare
Vaporum, in the Sinus Aestuum and in the Mare Nubium
(Fig. 10). The minimum average regolith thickness in this
category, which is characteristic of the Oceanus Procellarum,
is about 4.8 m (Table 13). Compared to category I, helium
reserves are distributed much more evenly and occupy an
area twice as large (Table 14), but this type of deposits is
twice as poor in helium content. Therefore, the total probable
reserves in the second category are only a quarter higher
than the reserves of the first category of deposits (Table 15).
Category III deposits cover the areas of low- titanium
mare basalt distribution, in which the minimum measured
content of
3
He and
4
He isotopes at the landing sites of the
Apollo 14 and 15 expeditions is 4.4 ppb and 11.1 ppm,
respectively (Tables 5,11 and 14). The minimum thickness
of the regolith in this category is observed in the Mare
Serenitatis and is 4.1 m (Table 13). Relatively small and
poor category III reserves, as a rule, outline the richer cat-
egory II deposits and are quite widespread in the Mare
Serenitatis and in the eastern part of the Mare Imbrium
(Fig. 10). In terms of area, this category exceeds category II
(Table 14), but in terms of helium reserves, it is significantly
inferior to deposits of a higher category (Table 15).
The fourth category includes low-titanium mare basalts
adjacent to the mareshore and mainly coastal complexes of
Highland rocks, characterized by an increased content of
TiO
2
(1.0–3.0 wt%) and, accordingly, ilmenite as compared
to anorthosites of the Highlands (0.01–1 wt%). Typical for
this category of deposits is the content of helium-3 at the
landing site of the Luna-20 spacecraft in the region of
coastal Highland rocks, which is 3.1 ppb (Tables 9,11 and
14), and at the landing site of the Luna-24 spacecraft in the
region coastal low- titanium basalts, which is 3.2 ppb
(Tables 10 and 11). The horizontal transfer of the
high-velocity component of ejections explains the enrich-
ment of coastal Highland rocks with a component of mare
basalts from impact craters with the formation of extensive
coastal mixing zones [113,114]. This process is reciprocal,
in the same way, coastal Highland rocks dilute mare rocks
with anorthosite component and reduce the concentration of
ilmenite in mare basalts, which, together with coastal
low-titanium basalts, form IV category of deposits and
outline almost all mare areas on the Moon Near and Far
Side. The minimum average regolith thickness is twice the
regolith thickness in the mare areas and in the area of the
Luna-20 landing, according to [104], is 9.2 m (Table 12).
The distribution area of the IV category is one third larger
than the III category area on the Moon Near side and almost
ten times more on the Far Side (Table 14). The total
probable reserves of helium in the regolith of this category
are similar to those of category II, which is characterized by
a much smaller area and regolith thickness (Tables 14
and 15).
The last V category with the lowest ilmenite content and
the lowest background concentration of helium isotopes
(1.4 ppb
3
He and 3.9 ppm
4
He) (Tables 11 and 14) covers
Highland anorthosite rocks, which occupy more than 80% of
the lunar surface and are characterized by an average rego-
lith thickness of about 10.1 m (Table 12)[105]. This cate-
gory is actually characterized by background minimum
possible content of the implanted helium-3 in the lunar soil,
and has no promising industrial significance. The content of
implanted helium isotopes in the regolith of Highland
anorthosites is the poorest, but given that the Highlands
occupy almost the entire surface of the Moon and are
characterized by the largest thickness of the loose regolith
layer, the total predicted reserves of helium isotopes in the
regolith of Highland anorthosites significantly exceed the
total reserves of higher categories of deposits (Tables 14 and
15). However, on the Moon Near Side in the equatorial
region, where the richest deposits are found, the total
reserves of helium in the regolith of Highland anorthosites
are even less than the total reserves of higher categories of
deposits.
Thus, in a loose regolith layer on the Moon, about
1.3 10
6
tons of helium-3 and about 3.5 10
9
tons of
helium-4 can be accumulated, which are extremely unevenly
distributed on the lunar surface. Of greatest practical interest
are the reserves of implanted helium isotopes in the richest
deposits of I and II categories, in which, over a relatively
small area, the total reserves of
3
He are about 2.1 10
5
tons, and
4
He—about 5.6 10
8
tons.
Table 15 Predicted reserves of
implanted helium isotopes in the
lunar soil
Category
deposits
Near side Far side Total, t
3
He, t
4
He, t
3
He, t
4
He, т
3
He, t
4
He, т
I 90,695 246,859,399 –– 90,695 246,859,399
II 120,717 313,865,411 711 1,849,836 121,428 315,715,247
III 59,491 150,080,056 371 935,959 59,862 151,016,015
IV 138,777 367,088,777 5275 13,953,986 144,052 381,042,763
V 356,719 993,717,291 534,152 1,487,994,198 890,871 2,481,711,489
Total, t 766,399 2,071,610,934 540,509 1,504,733,979 1,306,908 3,576,344,913
392 E. N. Slyuta and O. I. Turchinskaya
6 Summary
Depending on the mechanism of trapping and accumulation
of trapped noble gases in lunar regolith, weakly bound
helium, which saturates the pore space of lunar regolith, and
helium implanted into particles and minerals of lunar rego-
lith are released from solar wind. The accumulation of
weakly bound helium in the lunar regolith largely depends
on surface temperature and solar wind flux density
depending on geographic longitude and latitude, and prac-
tically does not depend on regolith composition. The content
of this type of helium increases essentially at high lunar
latitudes characterized by low diurnal temperature and
minimal amplitude of diurnal temperature variations; it is
extremely unstable and is easily released during temperature
and mechanical influences on the lunar soil, and requires
development of special methods of in situ investigation
[115]. It is assumed that the reserves of weakly bound
helium in the polar regions of the Moon may significantly
exceed the reserves of implanted helium.
Capture and accumulation of implanted helium in the
lunar soil depends on chemical and mineral composition of
regolith particles, on age of exposure and regolith maturity
degree, on particle sizes, and can vary in a very wide range
up to few orders of magnitude. The helium and noble gases
lowest content are characteristic of glasses of impact and
volcanic origin and the main rock-forming mineral of lunar
Highland rocks—anorthite. The highest concentration of
implanted helium is characteristic for the main ore mineral of
mare basalts—ilmenite. Amorphization of the crystalline
structure of ilmenite under solar wind irradiation begins at
energies of ions an order of magnitude higher than those of
silicates and aluminosilicates, and crystal lattice
radiation-induced defects trap and retain helium ions much
stronger than amorphous structure retain. As the content of
ilmenite in basalts decreases from high-titanium to
low-titanium, the concentration of implanted helium isotopes
in the bulk composition of regolith decreases correspond-
ingly. Minimal concentration of helium is observed in
Highland anorthosite rocks characterized by minimal content
of ilmenite and lower content of mafic rock-forming min-
erals olivine and pyroxenes with average helium content as
compared to basalts. Selective enrichment of regolith with
solar helium depending on its mineral composition, which is
determined mainly by underlying rocks, leads to
non-uniform regional distribution of helium isotope con-
centration and other implanted gases in lunar regolith. The
highest concentration of helium isotopes is observed in
high-titanium mare basalts, in which the content of the main
ore mineral ilmenite reaches 20 wt%.
Depending on mineral composition of regolith and con-
centration of implanted isotopes five categories of deposits
are distinguished. The first three categories cover areas of
distribution of high-titanium and low-titanium basalts and
are located in the territory of lunar mares in the Moon
equatorial region. Deposits with maximum content belong to
the richest and highest I category. Most reserves of
implanted helium in this category are concentrated in the two
largest fields located in the Mare Tranquillitatis and in the
Oceanus Procellarum. The predicted reserves of helium in
Category II deposits occupy an area twice as large, are
distributed much more evenly and are located in almost all
mare regions of the Moon. Approximately the same distri-
bution is typical for the deposits of the third category. The
fourth category of deposits unites low-titanium mare basalts
and adjacent complexes of Highland rocks in the coastal
zone where the mutual enrichment of Highland and mare
components occurs due to horizontal transfer of emissions
from impact craters with formation of extensive coastal
mixing zones. The fifth category, the poorest by helium
content in lunar soil but the largest by reserves unites all
lunar Highlands that occupy more than 80% of lunar surface.
In general, in the loose regolith layer on the Moon, the
total predicted reserves of helium-3 are estimated as
1.3 10
6
tons, and helium-4 as 3.5 10
9
tons, that are
characterized by an extremely inhomogeneous distribution.
The reserves of implanted helium isotopes in the richest
deposits of categories I and II, which are of the greatest
practical interest, are concentrated on a relatively small area
and are estimated at about 2.1 10
5
tons of
3
He and about
5.6 10
8
tons of
4
He.
Acknowledgements The article is dedicated to the memory of Aca-
demician E. M. Galimov, a scientist and teacher, who initiated these
studies at the Vernadsky Institute. Chapter 5.“Predicted reserves of
helium in the lunar soil”was supported by a grant fromthe Russian Science
Foundation No. 21-17-00120, https://rscf.ru/project/21-17-00120/.
References
1. Parker, E.N.: Dynamics of the interplanetary gas and magnetic
fields. Astrophys. J. 128, 664–676 (1958)
2. Parker, E.N.: The Solar Wind. McGraw-Hill, New York (1961)
3. Hundhausen, A.J.: Coronal Expansion and Solar Wind. Springer,
New York (1972)
4. Bame, S.J., Feldman, W.C., Gosling, J.T., Young, D.T., Zwickl,
R.D.: What magnetospheric workers should know about solar
wind composition. In: Johnson, R.G. (ed.) Energetic Ion Com-
position in the Earth’s Magnetosphere, pp. 73–98. Terra Scien-
tific, Tokyo/Dordrecht (1983)
5. Kovalenko, V.A.: Solar Wind. Nauka, Moscow (1983)
6. Geiss, J., Eberhardt, P., Signer, P., Buehler, F., Meister, J.: The
solar wind composition experiment. In: Apollo 11. Preliminary
Science Report, pp. 183–187. NASA. Washington, D.C. (1969)
7. Geiss, J., Eberhardt, P., Signer, P., Buehler, F., Meister, J.: The
solar wind composition experiment. In: Apollo 12. Preliminary
Science Report, pp. 99–102. NASA. Washington, D.C. (1970)
Helium-3 in the Lunar Soil 393
8. Geiss, J., Buehler F., Cerutti, H., Eberhardt, P., Meister, J.:
Solar-wind composition experiment. In: Apollo 14. Preliminary
Science Report, pp. 221–226. NASA. Washington, D.C. (1971)
9. Geiss, J., Buehler, F., Cerutti, H., Eberhardt, P.: Solar-wind
composition experiment. In: Apollo 15. Preliminary Science
Report, pp. 15-1–15-7. NASA. Washington, D.C. (1972)
10. Geiss, J., Buehler, F., Cerutti, H., Eberhardt, P., Filleux, Ch.:
Solar wind composition experiment. In: Apollo 16. Preliminary
Science Report, pp. 14-1–14-10. NASA. Washington, D.C.
(1972)
11. Geiss, J., Buehler, F., Cerutti, H., Eberhardt, P., Filleux, C.H.,
Meister, J., Signer, P.: The Apollo SWC experiment: results,
conclusions, consequences. Space Sci. Rev. 110, 307–335 (2004)
12. Ogilvie, K.W., Coplan, M.A., Bochsler, P., Geiss, J.: Abundance
ratios of
4
He++/
3
He++ in the solar wind. J. Geophys. Res. 85,
6021–6024 (1980)
13. Coplan, M.A., Ogilvie, K.W., Bochsler, P., Geiss, J.: Interpre-
tation of
3
He abundance variations in the solar wind. Solar Phys.
93, 415–434 (1984)
14. Bodmer, R., Bochsler, P., Geiss, J., von Steiger, R., Gloeckler,
G.: Solar windhelium isotopic composition from
SWICS/ULYSSES. Space Sci. Rev. 72,61–64 (1995)
15. Gloeckler, G., Geiss, J.: Measurement of the abundance of
Helium-3 in the Sun and in the local interstellar cloud with
SWICS on ULYSSES. Space Sci. Rev. 84(1/2), 275–284 (1998)
16. von Steiger, R., Schwadron, N.A., Fisk, L.A., Geiss, J.,
Gloeckler, G., Hefti, S., Wilken, B., Wimmer-Schweingruber,
R.F., Zurbuchen, T.H.: Composition of quasi-stationary solar
wind flows from Ulysses/solar wind ion composition spectrom-
eter. J. Geophys. Res. 105, 27217–27238 (2000)
17. Giammanco, C., Wurz, P., Karrer, R.: Minor ion abundances in
the slow solar wind. Astrophys. J. 681, 1703–1707 (2008)
18. Wimmer-Schweingruber, R.F., Bochsler, P., Wurz, P.: Isotopes
in the solar wind: new results from ACE, SOHO and WIND. In:
Solar Wind Nine, AIP Conference Proceedings, vol. 471,
pp. 147–152. Nantucket, MA, USA (1999)
19. Geiss, J., Gloeckler, G.: Abundances of deuterium and
HELIUM-3 in the protosolar cloud. Space Sci. Rev. 84(1/2),
239–250 (1998)
20. Burnett, D.S., Barraclough, B.L., Bennett, R., Neugebauer, M.,
Oldham, L.P., Sasaki, C.N., Sevilla, D., Smith, N., Stansbery, E.,
Sweetnam, D., Wiens, R.C.: The genesis discovery mission: return
of solar matter to earth. Space Sci. Rev. 105, 509–534 (2003)
21. Reisenfeld, D.B., Burnett, D.S., Becker, R.H., Grimberg, A.,
Heber, V.S., Hohenberg, C.M., Jurewicz, A.J.G., Meshik, A.P.,
Pepin, R.O., Raines, J.M., Wieler, R., Wiens, R.C., Zurbuchen,
T.H.: Elemental abundances of the bulk solar wind: analyses
from genesis and ACE. Space Sci. Rev. 130,79–86 (2007)
22. Heber, V.S., Wieler, R., Baur, H., Olinger, C., Friedmann, T.A.,
Burnett, D.S.: Noble gas composition of the solar wind as
collected by the genesis mission. Geochim. Cosmochim. Acta 73
(24), 7414–7432 (2009)
23. Eberhardt, P., Eugster, O., Geiss, J., Graf, H., Grogler, N.,
Guggisberg, S., Jungck, M., Maurer, P., Morgeli, M., Stettler, A.:
Solar wind and cosmic radiation history of Taurus-Littrow
Regolith. In: Proceedings Lunar Science Conference 5th,
pp. 197–199. Houston, USA (1974)
24. Wittenberg, L.J., Santarius, J.F., Kulcinski, G.L.: Lunar source of
He-3 for commercial fusion power. Fusion Technol. 10(2), 167–
178 (1986)
25. Kulcinski, G.L., Cameron, E.N., Santarius, J.F., Sviatoslavsky, I.
N., Wittenberg, L.J., Schmitt, H.H.: Fusion energy from the
Moon for the twenty-first century. In: Proceedings of the 2nd
Conference on Lunar Bases and Space Activities of the 21st
Century in Houston, pp 459–474. NASA Conference. Publication
3166 (1992)
26. Murali, A.V., Jordan, J.L.: Helium-3 inventory of lunar samples:
a potential future energy resources for mankind? In: The 24th
Lunar and Planetary Science Conference, pp. 1023–1024.
Houston, USA (1993)
27. Galimov, E.M.: Project Moon–Helium-3. Nauka v Rossii 6,15–
23 (2006)
28. Taylor, L.A., Kulcinski, G.L.: Helium-3 on the Moon for fusion
energy: the Persian Gulf of the 21st century. Solar Syst. Res. 33
(5), 338–345 (1999)
29. Cho, A.: Helium-3 shortage could put freeze on low-temperature
research. Science 326, 778–779 (2009)
30. Eberhardt, P., Geiss, J., Graf, H., Grogler, N., Krahenbuhl, U.,
Schwaller, H., Schwarzmuller, J., Stettler, A.: Trapped solar wind
noble gases, exposure age and K/Ar-age in Apollo 11 lunar fine
material. In: Proceedings Apollo 11 Lunar Science Conference,
pp. 1037–1070. Houston, USA (1970)
31. Hintenberger, H., Weber, H.W., Voshage, H., Wanke, H.,
Begemann, F., Vilscek, E., Wlotzka, F.: Rare gases, hydrogen,
and nitrogen: concentrations and isotopic composition in lunar
material. Science 167(3918), 543–545 (1970)
32. Slyuta, E.N., Yakovlev, O.I., Voropaev, S.A., Dubrovskii, A.V.:
He implantation and concentrations in minerals and lunar regolith
particles. Geochem. Int. 51(12), 959–967 (2013)
33. Cameron, E.N.: Helium mining on the moon: site selection and
evaluation. In: Proceedings of the 2nd Conference on Lunar
Bases and Space Activities of the 21st Century in Houston,
pp. 189–197. NASA Conference Publication 3166 (1992)
34. Shkuratov, Yu.G., Starukhina, L.V., Kaidash, V.G., Bondarenko,
N.V.:
3
He distribution over the lunar visible hemisphere. Solar
Syst. Res. 33(5), 409–420 (1999)
35. Johnson, J.R., Swindle, T.D., Lucey, P.G.: Estimated solar
wind-implanted helium-3 distribution on the Moon. Geophys.
Res. Let. 26(3), 385–388 (1999)
36. Cocks, F.H.:
3
He in permanently shadowed lunar polar surfaces.
Icarus 206, 778–779 (2010)
37. Fa, W., Jin, Y.: Quantitative estimation of helium-3 spatial
distribution in the lunar regolith layer. Icarus 190,15–23 (2007)
38. Sorokin, E.G., Yakovlev, O.I., Slyuta, E.N., Gerasimov, M.V.,
Zaitsev, M.A., Shcherbakov, V.D., Ryazantsev, K.M.,
Krasheninnikov, S.P.: Experimental modeling of a micromete-
orite impact on the Moon. Geochem. Int. 58(2), 113–127 (2020)
39. Sorokin, E.G., Yakovlev, O.I., Slyuta, E.N., Gerasimov, M.V.,
Zaitsev, M.A., Shcherbakov, V.D., Ryazantsev, K.M.,
Krasheninnikov, S.P.: Experimental model of the formation of
nanophase metallic iron in the lunar regolith. Dokl. Earth Sci. 492
(2), 431–433 (2020)
40. Mokrushin, A.D., Agafonov, Yu.A., Zinenko, V.I., Pustovit, A.
N.: Distribution of implanted hydrogen in a-SiO
2
amorphous
silica dioxide. Poverkhnost. Rentgen., Sinkhrotron. Neitron.
Issled. 12,5–8 (2004)
41. French, B.M.: The Moon Book. Penguin Books, London (1977)
42. Carrier, W.D. III, Mitchell, J.K., Mahmood, A.: The relative
density of lunar soil. In: Proceedings 4th Lunar Science
Conference, pp. 2403–2411. Houston, USA (1973)
43. Herz, F., Morrison, D.A., Gault, D.E., Oberbeck, V.R., Quaid, V.
L., Vedder, J.F., Brownlee, D.E., Hartung, J.B.: The microme-
teoroide complex and evolution of the lunar regolith. In:
Pomeroy, J.H., Hubbard, N.J. (eds.) The Soviet-American
conference on cosmochemistry of the Moon and planets,
pp. 605–636. NASA, Washington D.C. (1977)
44. Morris, R.V.: Surface and near surface exposure ages of lunar
soils: exposure ages based on the fine grained metal and
394 E. N. Slyuta and O. I. Turchinskaya
cosmogenic
21
Ne. In: Proceedings 8th Lunar Planet Science
Conference, pp. 685–687. Houston, USA (1977)
45. Eberhardt, P., Eugster, O., Geiss, J., Graf, H., Grogler, N.,
Guggisberg, S., Jungck, M., Maurer, P., Morgeli, M., Stettler, A.:
Solar wind and cosmic radiation history of Taurus-Littrow
Regolith. In: Proceedings 5th Lunar Science Conference,
pp. 197–199. Houston, USA (1974)
46. Haskin, L., Warren, P.: Lunar chemistry. In: Heiken, G.H.,
Vaniman, D., French, B.M. (eds.) Lunar Sourcebook, pp. 357–
474. Cambridge University Press, New York (1991)
47. Slyuta, E.N.: Basic types of Moon’s resources and their mining
and processing. Gornyi zhurnal 4,13–18 (2017)
48. Slyuta, E.N.: Physical and mechanical properties of the lunar soil
(A review). Sol. Sys. Res. 48(5), 330–353 (2014)
49. Taylor, L.: Helium-3 on the Moon: model assumptions and
abundances. In: 2nd Wisconsin Symposium on Helium-3 and
Fusion Power, pp. 49–56. University of Wisconsin-Madison
(1993)
50. Scherzer, B.: Development of surface topography due to gas
implantation. In: Behrisch, R. (ed.) Sputtering by Particle
Bombardment, vol. 2, pp. 271–355. Springer, Berlin-Heidelberg
(1983)
51. Harris-Kuhlman, K.R.: Trapping and duffusion of Helium in
lunar materials. Ph.D. thesis. Fusion Technol. Inst., Wisconsin
(1998)
52. Futagami, T., Ozima, M., Nagai, S., Aoki, Y.: Experiments on
thermal release of implanted noble gases from minerals and their
implications for noble gases in lunar soil grains. Geochim.
Cosmochim. Acta 57(13), 3177–3194 (1993)
53. Ziegler, J.F., Biersack, J.P., Littmark, U.: The Stopping and
Range of Ions in Solids. Pergamon Press, New York (2009)
54. Slyuta, E.N., Petrov, V.S., Yakovlev, O.I., Voropaev, S.A.,
Monakhov, I.S., Prokof’eva, T.V.: Application of thermodesorp-
tion mass spectrometry for studying proton water formation in the
lunar regolith. Geochem. Int. 55(1), 27–37 (2017)
55. Shirokov, Yu.M., Yudin, I.P.: Nuclear Physics. Nauka, Moscow
(1980)
56. Bernatowicz, T.J., Nichols, R.H. Jr., Hohenberg, C.M.: Origin of
amorphous rims on lunar soil grains. In: 25th Lunar Planet
Science Conference, pp. 105–106. Houston, USA (1994)
57. Christoffersen, R., Keller, L.P., Dukes, C., Rahman, Z., Baragi-
ola, R.: Experimental investigation of space radiation processing
in lunar soil ilmenite: combining perspectives from surface
science and transmission electron microscopy. In: 41th Lunar
Planet Science Conference, ID 1532. Houston, USA (2010)
58. Slyuta, E.N., Shilobreeva, S.N., Kashkarov, L.L., Kalinina, G.V.,
Voropaev, S.A.: Amorphization depth of anorthite and quartz in
dependence on H
+
and He
+
ion energy and irradiation dose. In:
42nd Lunar and Planet Science Conference ID 1127. Houston,
USA (2011)
59. Sominskiy, M.S.: Semiconductors. Nauka, Leningrad (1967)
60. Keller, L.P., McKay, D.S.: The nature and origin of rims on lunar
soil grains. Geochim. Cosmochim. Acta 61(11), 2331–2341
(1997)
61. Slyuta, E.N., Shilobreeva, S.N., Voropaev, S.A,. Kashkarov, L.
L., Zinenko, V.I., Saraykin, V.V.: Preliminary experimental data
on irradiation-induced fractionation of isotopes
54
Fe and
56
Fe. In:
42th Lunar Planet Science Conference, ID 1195. Houston, USA
(2011)
62. Morris, R.V.: The surface exposure (maturity) of lunar soils:
some concepts and Is/FeO compilation. In: Proceedings Lunar
9th Planet Science Conference, pp. 2287–2297. Houston, USA
(1978)
63. Webb, R.P.: SUSPRE, V 1.4. Guilford, UK (1987)
64. Lakatos, S., Heymann, D.: Green spherules from Apollo 15:
inferences about their origin from inert gas measurements. In: The
Moon 7, pp. 132–148. Lunar Planet. Inst., Houston, USA (1973)
65. Frick, U., Baur, H., Funk, H., Phinney, D., Schäfer, C., Schultz,
L., Signer, P.: Diffusion properties of light noble gases in lunar
fines. In: Proceedings Lunar Science Conference 4th, vol. 2,
pp. 1987–2002. Houston, USA (1973)
66. Frick, U., Baur, H., Ducati, H., Funk, H., Phinney, D., Signer, P.:
On the origin of helium, neon, and argon isotopes in sieved
mineral separates from an Apollo 15 soil. In: Proceedings
Lunar Science Conference 6th, pp. 2097–2129. Houston, USA
(1975)
67. Baur, H., Frick, U., Funk, H., Schultz, L., Signer, P.: Thermal
release of helium, neon, and argon from lunar fines and minerals.
In: Proceedings Lunar Science Conference 3
rd
, vol. 2, pp. 1947–
1966. Houston, USA (1972)
68. Eberhardt, P., Geiss, J., Graft, H., Groegler, N., Mendia, M.D.,
Moergeli, M., Schwaller, H., Guggisberg, S., Stettler, A.,
Krähenbühl, U., von Gunten, H.R.: Trapped solar wind noble
gases in Apollo 12 lunar fines 12001 and Apollo 11 breccia
10046. In: Proceedings Lunar Planet Science Conference 3
rd
, vol.
2, pp. 1821–1856. Houston, USA (1972)
69. BVSP (Basaltic Volcanism on the Terrestrial Planets). Pergamon
Press, New York (1981)
70. Bogatikov, O.A. (ed.): Igneous Rocks. Part 3. Basic rocks.
Nauka, Moscow (1985)
71. Tarasov, L.S., Nazarov, M.A., Shevaleevsky, I.D., Kudryashova,
A.F., Gaverdovskaya, A.S., Korina, M.I.: Mineralogy and
petrography of lunar Rockw from Mare Crisium (preliminary
data). In: Proceedings 8th Lunar Science Conference, vol. 3,
pp. 3333–3356. Houston, USA (1977)
72. Barsukov, V.L., Dmitriev, L.V., Tarasov, L.S., Kolesov, G.M.,
Shevaleevsky, I.D., Ramendik, G.I., Garanin, A.V.: Geochemical
and petrochemical features of regolith and rocks from the Mare
Crisium. In: Barsukov, V.L. (ed.) Lunar Soil from the Mare
Crisium, pp. 158–165. Nauka, Moscow (1980)
73. Cast, P.W., Phinney, W.C., Hubbard, N.J., et al.: Preliminary
examination of lunar samples. Apollo 17 preliminary science
report, pp. 7-1–7-45. NASA, Washington D.C. (1973)
74. Head, J.W., Wilson, L.: Lunar mare volcanism: stratigraphy,
emption conditions, and the evolution of secondary crusts.
Geochem. Cosmochim. Acta. 56, 2155–2175 (1992)
75. Moriarty, III D.P., Watkins, R.N., Valencia, S.N., Kendall, J.D.,
Evans, A.J., Dygert, N., Petro, N.E.: Evidence for a stratified
upper mantle preserved within the South Pole—Aitken Basin. J.
G.R. Planets 121(e2020JE006589), 1–27 (2020)
76. Gaddis, L.R., Pieters, C.M., Hawke, B.R.: Remote sensing of
lunar pyroclastic mantling deposits. Icarus 61, 461–489 (1985)
77. Gaddis, L.R., Staid, M.I., Tyburczy, J.A., Hawke, B.R., Petro, N.
E.: Compositional analyses of lunar pyroclastic deposits. Icarus
161(2), 262–280 (2003)
78. Vinogradov, A.P.: Preliminary data on the lunar soil delivered by
the automatic station “Luna-20”. In: Barsukov, V.L., Surkov, Yu.
A. (eds.) Soil from the Highland Region of the Moon, pp. 7–17.
Nauka, Moscow (1979) (in Russian)
79. The Lunar Sample Preliminary Examination Team. Preliminary
examination of lunar samples. In: Apollo 12. Preliminary Science
Report, pp. 189–216. NASA. Washington, D.C. (1970)
80. Heymann, D., Yaniv, A., Adams, J.A.S., Fryer, G.E.: Inert gases
in lunar samples. Science 167(3918), 555–558 (1970)
81. Reynolds, J.H., Hohenberg, C.M., Lewis, R.S., Davis, P.K.,
Kaiser, W.A.: Isotopic analysis of rare gases from stepwise
heating of lunar fines and rocks. Science 167(3918), 545–548
(1970)
Helium-3 in the Lunar Soil 395
82. Heymann, D., Yaniv, A.: Inert gases from Apollo 12, 14, and 15
fines. Geochim. Cosmochim. Acta. Suppl. 4(2), 1857–1863
(1972)
83. Eberhardt, P., Eugster, O., Geiss, J., Groegler, N., Guggisberg, S.,
Moergeli, M.: Noble gases in the Apollo 16 special soils from the
East-West split and the permanently shadowed area. In: Pro-
ceedings Lunar Science Conference 7
th
, pp. 563–585. Houston,
USA (1976)
84. Eugster, O., Grögler, N., Mendia, M.D., Eberhardt, P., Geiss, J.:
Trapped solar wind noble gases and exposure age of Luna 16
lunar fines. Geochim. Cosmochim. Acta 37(9), 1991–2003
(1973)
85. Vinogradov, A.P., Zadorozhny, I.K.: Noble gases in the regolith
from the Mare Fecunditatis. In: Vinogradov, A.P. (ed.) Lunar Soil
from the Sea of Abundance, pp. 379–386. Nauka, Moscow
(1974)
86. Kaiser, W.A.: Rare gas studies in Luna-16-G-7 fines by stepwise
heating technique. A low fission solar wind Xe. Earth Planet. Sci.
Lett. 13, 387–399 (1972)
87. Heymann, D., Yaniv, A., Ladatos, S.: Inert gases in twelve
particles and one “dust”sample from Luna 16. Earth Planet. Sci.
Lett. 13, 400–406 (1972)
88. Vinogradov, A.P., Zadorozhny, I.K.: Inert gases in regolith and
rock fragments delivered by the automatic station “Luna-20.”In:
Barsukov, V.L., Surkov, Yu.A. (eds.) Soil from the Highland
Region of the Moon, pp. 547–556. Nauka, Moscow (1979)
89. Eugster, O., Eberhardt, P., Geiss, J., Grogler, N., Jungck, M.,
Morgeli, M.: Solar wind trapped and cosmic-ray produced noble
gases in Luna 20 soil. In: Proceedings Lunar Science Conference
6th, pp. 1989–2007. Houston, USA (1975)
90. Zadorozhny, I.K., Ivanov, A.V.: Content and isotopic composi-
tion of inert gases in the base samples of the Luna-24 regolith. In:
Barsukov, V.L., (ed.) Lunar Soil from the Mare Crisium,
pp. 289–299. Nauka, Moscow (1980)
91. Bogard, D.D., Hirsch, W.C.: Noble gases in Luna 24 core soils.
In: Mare Crisium: The View from Luna 24, pp. 105–116.
Pergamon Press, Inc., New York (1978)
92. Slyuta, E.N., Abdrahimov, A.M., Galimov, E.M.: Does Helium-3
abundance decrease in dependence on depth at Mare Crisium? In:
39th Lunar Planet Science Conference ID 1054. Houston, USA
(2008)
93. Gault, D.E., Quaide, W.L., Oberbeck, V.R., Moore, H.J.: Lunar 9
photographs: evidence for a frag-mental surface layer. Science
153, 985–988 (1966)
94. Florensky, K.P., Pronin, A.A., Bazilevsky, A.T.: Geology of the
Luna-24 landing site. In: Vinogradov, A.P. (ed.) Lunar Soil from
the Mare Fecunditatis, pp. 7–18. Nauka, Moscow (1980)
95. Carrier, W.D. III.: Apollo drill core depth relationships. The
Moon 10, pp. 183–194. Lunar Planet. Inst., Houston, USA (1974)
96. Heiken, G.H., Vaniman, D.T., French, B.M. (eds.): Lunar
Sourcebook. Cambridge University Press, New York (1991)
97. Florensky, K.P., Bazilevsky, A.T., Pronin, A.A., Popova, Z.V.:
Preliminary results of a geomorphological study of panoramas.
In: Mobile Laboratory on the moon. Lunokhod-1, pp. 96–115.
Nauka, Moscow (1971)
98. Florensky, K.P., Bazilevsky, A.T.: The processes of transforma-
tion of the lunar surface in the Lemonnier region based on the
results of a detailed study on the “Lunokhod-2”. In: Tectonics
and structural geology, pp. 205–235. Nauka, Moscow (1976)
99. Stakheev, Yu.I., Lavrukhina, A.K.: Granulometric characteristics
of the lunar regolith and the thickness of the regolith layer at the
landing sites of the Luna-16 and -20 spacecrafts. In: Barsukov, V.
L., Surkov, Yu.A. (eds.) Soil from the Highland Region of the
Moon, pp. 74–76. Nauka, Moscow (1979)
100. Oberbeck, V.R., Quaide, W.L.: Genetic implication of lunar
regolith thickness variations. Icarus 9(3), 446–465 (1968)
101. Nakamura, Y., Dorman, J., Duennebier, F., et al.: Shallow lunar
structure determined from the passive seismic experiment. Moon
13(1), 3 (1975)
102. Cooper, M.R., Kovach, R.L., Watkins, J.S.: Lunar near-surface
structure. Rev. Geophys. Space Phys. 12, 291–308 (1974)
103. Duennebier, F.K., Watkins, J., Kovach, R.: Results from the lunar
surface profiling experiment. In: Lunar and Planetary Science
Conference, vol. 5, p. 183. Houston, USA (1974)
104. Strangway, D.W., Pearce, G.W., Olhoeft, G.R.: Magnetic and
dielectric properties of lunar samples. In: Pomeroy, J.H., Hubbard
(eds.) The Soviet-American Conference on Cosmochemistry of
the Moon and Planets, pp. 417–432. N.J. NASA, Washington D.
C. (1977)
105. Bondarenko, N.V., Shkuratov, Yu.G.: A map regolith layer
thickness for the visible lunar hemisphere from radar and optical
data. Solar Sys. Res. 32, 264–271 (1998)
106. Bazilevsky, A.T.: Estimation of the thickness and degree of
reworking of the lunar regolith by the prevalence of craters.
Cosmich. Issled. 12(4), 606–609 (1974)
107. Shkuratov, Yu.G., Bondarenko, N.V.: Regolith layer thickness
mapping of the Moon by radar and optical data. Icarus 149, 329–
338 (2001)
108. Wilhelms, D.E.: The geologic history of the Moon. USGS
Professional Paper 1342, Washington (1987)
109. Boyce, J.M., Johnson, D.A.: Ages of flow units in the far eastern
Maria and implications of basin-filling history. In: Proceedings
Lunar Planet Science Conference 9th, pp. 3275–3284. Houston,
USA. (1978)
110. Blewett, D.T., Lucey, P G., Hawke, B.R., Jolliff, B.L.: Clemen-
tine images of the lunar sample-return stations: refinement of FeO
and TiO
2
mapping techniques. J. Geophys. Res. 102(E7), 16319–
16326 (1997)
111. Lucey, P.G., Blewett, D.T., Hawke, B.R.: Mapping the FeO and
TiO
2
content of the lunar surface with multispectral imagery.
J. Geophys. Res. 103(E2), 3679–3699 (1998)
112. Lucey, P.G., Blewett, D.T., Taylor, G.J., Hawke, B.R.: Imaging
of lunar surface maturity. J. Geophys. Res. 105(E8), 20377–
20387 (2000)
113. Wood, J.A., Dickey, J.S., Marvin, U.B., Powell, B.N.: Lunar
anorthosites and a geophysical model of the Moon. In: Proceed-
ings of the Apollo 11 Lunar Science Conference, vol. 1, pp. 965–
988. Houston, USA (1970)
114. Florensky, C.P., Ivanov, A.V., Basilevsky, A.T.: The role of
exogenous factors in the formation of the lunar surface. In:
Pomeroy, J.H., Hubbard, N.J. (eds.) The Soviet-American
Conference on Cosmochemistry of the Moon and Planets,
pp. 571–584. NASA, Washington D.C. (1977)
115. Marov, M.Ya., Slyuta, E.N.: Early steps toward the Lunar base
deployment: some prospects. Acta Astronautica 181,28–39
(2021)
396 E. N. Slyuta and O. I. Turchinskaya