Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

Author’s Accepted Manuscript

Water delivery to the Moon by asteroidal and


cometary impacts

V.V. Svetsov, V.V. Shuvalov

www.elsevier.com

PII: S0032-0633(15)00261-5
DOI: http://dx.doi.org/10.1016/j.pss.2015.09.011
Reference: PSS4062
To appear in: Planetary and Space Science
Received date: 13 March 2015
Revised date: 12 August 2015
Accepted date: 8 September 2015
Cite this article as: V.V. Svetsov and V.V. Shuvalov, Water delivery to the Moon
by asteroidal and cometary impacts, Planetary and Space Science,
http://dx.doi.org/10.1016/j.pss.2015.09.011
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Water delivery to the Moon by asteroidal and cometary impacts

V.V. Svetsov*, V.V. Shuvalov

Institute for Dynamics of Geospheres, Leninskiy Prospekt 38-1, Moscow 119334, Russia.

Moscow Institute of Physics and Technology, Institutskiy Per. 9, Dolgoprudny, Moscow Region,

141700, Russia.

*
Corresponding author. E-mail address: svettsov07@rambler.ru (V.V. Svetsov).

ABSTRACT

Recent spacecraft missions detected presence of hydroxyl or water over large areas on the lunar

surface. Several craters near the lunar poles have increased concentrations of hydrogen

suggesting impact delivery of water. Using a numerical model, we have carried out computer

simulations of the impacts of asteroids and comets in order to estimate the fate of water that can

be contained in the projectiles. We find that at impact velocities below ~10 km/s a significant

fraction of a stony projectile remains in the crater and is heated to temperatures below 1000 K.

At these velocities hydrated minerals contained in carbonaceous projectiles decompose only

partly. We conclude that the impacts of water-bearing carbonaceous asteroids could produce

deposits of free and chemically bound water inside some lunar craters. The relative number of

these craters may reach several percent. In contrast to asteroids, water from cometary impacts,

even at low velocities, is vaporized, and vapor plume expands and disperses over the lunar

surface.

Keywords: Moon, water, impact, carbonaceous chondrite, crater, dehydration

1
1. Introduction

The Moon has been considered to be anhydrous for a long period of lunar exploration, but

recent observations have shown that water on the lunar surface is present in the forms of

hydroxyl (OH) or water ice. Enhanced concentrations of hydrogen were first detected at the lunar

poles by the Lunar Prospector neutron spectrometer (Feldman et al., 1998). Infrared

spectroscopic measurements of the lunar surface from Cassini, Deep Impact, and Chandrayaan-1

spacecrafts have given evidence for the presence of OH or water not only at the poles but also

over the entire lunar surface (Clark, 2009; Sunshine et al., 2009; Pieters et al., 2009).

Observations by Lunar Exploration Neutron Detector onboard the Lunar Reconnaissance Orbiter

spacecraft show some specific regions near the lunar south pole with enhanced hydrogen content

(Mitrofanov et al., 2010). These regions do not necessarily correlate with permanently shadowed

areas at the bottom of polar craters and are observed in both permanently shadowed and

illuminated areas (Mitrofanov et al., 2012). The Lunar Crater Observation and Sensing Satellite

(LCROSS) impact experiment directly confirmed the presence of water ice in permanently

shadowed regolith of the Cabeus crater (Colaprete et al., 2010). Significant epithermal neutron

flux suppressions, indicative of hydrogen, were detected at Cabeus (crater diameter ~100 km,

distance ~150 km from the south pole), Haworth (diameter 35 km, distance ~90 km), and

Shoemaker (diameter 50 km, distance ~60 km) craters (Sanin et al., 2014). However, some other

craters at the same distances from the south pole give a neutron flux close to the average.

Water and/or hydroxyl has been also found in volcanic glasses (e.g., Saal et al., 2008;

Hauri et al., 2015) and in glasses formed in the lunar regolith by micrometeorite impacts (Liu et

al., 2012). Numerous studies of apatites from different lunar samples (basaltic lavas and plutonic

rocks) show a significant amount of water (OH) in apatites (hundreds to thousands of ppm) and

indicate that water may be ubiquitous within the lunar interior (e.g., McCubbin et al., 2010;

Greenwood et al., 2011; Tartèse et al., 2013; Barnes et al., 2014). It is likely that water on the

2
Moon is of the same origin as on the Earth (Saal et al., 2013; Barnes et al., 2014), it could have

been delivered to the Moon’s interior by late accretion of comets and chondritic asteroids

(Greenwood et al. 2011; Tartèse, Anand, 2013). However, many questions remain unanswered,

and further studies are necessary for correct estimates of abundances of water and other volatile

elements in the lunar mantle, crust, and regolith (McCubbin et al., 2015). Hydrogen at the

surface can be emplaced by solar wind and endogenic sources, and, as have been supposed in

many papers, could be derived from comets and water-rich asteroids impacting the Moon (see,

Anand, 2010). The presence of water in some distinct craters (Sanin et al., 2014) supports the

hypothesis of water delivery by the impacts.

Ong et al. (2010) numerically simulated vertical impacts of comets on the Moon at

velocities of 5 to 60 km/s and observed an escape of 92% to 98% of cometary water at moderate

velocities (20 to 30 km/s). Retained water can migrate over the lunar surface and be captured by

cold traps at the lunar poles. Formation of a transient atmosphere and transport of water to cold

traps after a cometary impact at 30 km/s were modeled by Stewart et al. (2011) and Prem et al.

(2015). The model of Stewart et al. (2011) shows that after the impact of a 2-km-diameter comet

at latitude 45°S about 0.1% of the initial comet mass can reach the cold traps in several months,

and ~1 mm layer of ice can be accumulated at the floors of the shadowed craters. Stewart et al.

(2010), using Ong et al’s (2010) cometary flux and size distribution and taking into account loss

mechanisms during water migrations and gardening in the cold traps, estimated that about

5.0×1010 – 1.6×1011 kg of water should be present inside the cold traps after 1 Ga; these deposits

are equivalent to an ice layer thickness of 8 – 30 mm. Despite the deposition of ice is non-

uniform (Prem et al., 2015), it seems that the deposition by the process of water migration from

numerous cometary impacts cannot explain why similar nearby craters near the south pole have

noticeably different hydrogen concentrations in the regolith and why hydrogen was detected in

sunlit areas. Various reasons, e.g., crater age and impact gardening, could potentially explain

variations in hydrogen abundance that have been observed through remote sensing, but we also

3
can suggest that high concentration of hydrogen may be caused by remnants of impactor material

remaining in a crater. At impact velocities lower than 15 km/s the bulk of cometary water is

retained by the Moon (Ong et al., 2010), but it remains unknown if any significant amount of

water (that produces hydrogen content at least about 100 ppm in the upper layer of the crater

floor) can settle in a crater after a cometary impact. (The minimum hydrogen concentration

across the entire lunar surface is 10 to 50 ppm (Mitrofanov et al., 2010)).

The impacts of carbonaceous chondrites potentially deliver more water to the Moon than

comets (Ong et al., 2010). Although some material of a stony impactor will be ejected from a

crater, approximately 40 to 60% of asteroid mass could remain in the crater at low impact

velocities (3 – 7 km/s) and most probable impact angle 45° (Bland et al., 2008). Bland et al.

found that at these velocities a sizeable fraction of the projectile remaining in these craters will

be essentially unaltered, experiencing shock pressures <10 GPa. Potter and Collins (2013)

calculated the fraction of dunite impactors that remains solid or unvaporized for various impact

velocities and impact angles. They found that at an impact velocity of 5 km/s all the impactor is

weakly shocked below 50 GPa and remains solid. However, dunite is not quite suitable for

modeling of carbonaceous chondrite. In addition, it is unclear if water in the meteoritic remnants

is vaporized or remains in the bound form. It is necessary to calculate the temperatures of the

remaining projectile material. The objectives of this work are to simulate the impacts of asteroids

and comets, estimate the masses of cometary and carbonaceous material remaining in the craters,

to clarify the fate of water contained in projectile remnants, and assess the relative number of

impact craters on the Moon with increased content of hydrogen.

2. Methods

Water can be delivered to the Moon by hydrated and hydroxylated minerals contained in

asteroids similar in composition to carbonaceous chondrites, such as CI, CM, and CR groups.

Water in hydrous chondrites can exist in molecular form and can be locked up in phyllosilicate

4
and hydroxide minerals, with water content making up approximately 10 wt.% of the meteorites

on average. Release of water from hydrous meteorites depends on the temperature to which a

projectile is heated during the impact. Studies of phyllosilicates in carbonaceous chondrites show

that hydrated minerals do not begin to decompose until they reach temperatures of about 600–

700 K (Akai, 1992). Molecular water contained in the meteorite and water from hydroxylated

minerals can release at lower temperatures from 300 to 500 K (Garenne et al., 2014), and,

contrariwise, at a temperature of 573 K hydration of olivine can go on (Ivanova et al., 2013). The

highest temperatures of complete dehydration of serpentine and saponite with formation of

water-free olivine and enstatite are from 1100 to 1200 K (Akai 1990; Ivanova et al., 2010). In

our simulations we calculated the temperatures of projectile fragments and used these

temperature levels as indicators of dehydration degree.

We used a hydrodynamic computer code SOVA (Shuvalov, 1999) for two- and three-

dimensional impact simulations. The SOVA method is a Lagrangian-Eulerian approach for

solving the equations of hydrodynamics with a constant gravitational field. The effects of dry

friction are taken into account by solving the equations similar to the Navier-Stokes equations.

The equations contain the same derivatives of velocity components with viscosity coefficients

depending on a friction coefficient, pressure and velocity field (Dienes, Walsh, 1970). The

dimensionless coefficient of friction in our simulations was equal to 0.7, which is typical for

rocks and sand. The hydrodynamic equations of motion are approximated on a rectangular grid.

Differential equations are solved in two steps – Lagrangian and Eulerian at which flows of

materials between grid cells are calculated. Each cell may have several different materials with

different densities and internal energies but with the same pressure, and a special procedure is

used which constructs boundaries between different materials in the cells. Vacuum is simulated

by a gas with a very low density of 10-30 g/cm3.

The numerical grid consisted of 300×150×300 cells along X, Y, and Z axes respectively.

(Y axis is perpendicular to the vertical plane XZ in which the impactor trajectory lies). We

5
assumed symmetry relative to the plane XZ, which allowed us to model only the half-space in Y

direction. The size of cells in 3D simulations of oblique impacts was equal to 1/80 of the

projectile’s diameter. The cell size was doubled when the shock wave approached the boundaries

of computation grids, and the cell sizes increased to the outer boundaries to avoid problems with

boundary conditions only at the latest stages of crater formation. 2D simulations of vertical

impacts in cylindrical coordinates were made both with the maximum mesh resolution 80 and

160 cells per the projectile diameter. The results differed by no more than 3–5 %.

The temperature of material is calculated using its equation of state. However, after the

impacts at speeds below 15 km/s considered in this paper, projectile fragments do not vaporize

and form a very thin layer on the crater floor, much smaller than the crater size, and typically

smaller than the resolution of a computational grid. Because of the limitations in mesh

resolution, the layer of impactor material mixes with the material of the target, and the accuracy

of temperature computations reduces. We used tracer particles to calculate the temperatures with

acceptable accuracy (~5% for a given equation of state). The tracers are passive particles that

move through the mesh with velocities equal to the velocities of the medium. 16000 tracer

particles were initially distributed in the projectile. Each particle represents a mass

corresponding to the mass of a volume containing the particle before the impact and is assigned

the pressure, temperature, and velocity of a cell in which this particle is at a given point of time.

The heating of the projectile and target occurs primarily in the shock wave. We determine the

maximum pressure of the tracers when the shock wave goes through the projectile and calculate

the temperature at this moment. Eventually the real temperature can change due to the drop of

pressure, thermal conductivity, and friction, but not substantially for the solid fraction of a

projectile at the stage of crater excavation and collapse. As artificial numerical effects can

significantly alter the temperature, we determined the temperature of projectile material by its

state after the passage of a shock wave. The tracer particles also make it possible (in addition to

6
the integration of mass over cells) to estimate position of impactor material after crater

formation, and to calculate the mass of excavated material and the mass remained in the crater.

The diameter of impacting spherical bodies in the simulations was 1 km. Asteroids of this

size hit the Moon approximately once in 10 Ma and produce craters about 20 km in diameter

(see, e.g., Ivanov, 2008). (10-km-diameter asteroids hit the Moon once in 2 Ga). Shackleton

crater at the south pole has such diameter, other craters containing high abundance of water,

Cabeus, Haworth and Shoemaker, are larger (see above). However, it is not necessary to make

simulations for various sizes of impacting asteroids and comets because the flow pattern,

pressures, projectile temperatures, etc, will be the same for different sizes of impactors due to

hydrodynamic similarity. The size of a projectile may influence only the fraction of material

remaining in the crater because the crater diameter and the time of excavation depend on the

projectile size. We varied the impact velocities of asteroids from 6 to 14 km/s, in this case the

bulk of a projectile is not vaporized (Potter, Collins, 2013), and, on the other hand,

approximately 25-30% of asteroidal impacts on the Moon occur at velocities below 12 km/sec

(Marchi et al., 2009; Yue et al, 2013). The velocities of comets were 8–10 km/s because

cometary ice can potentially remain in a crater only at low velocities, and about 25% of short

period comets probably hit the Moon at velocities 8–10 km/s (Jeffers et al., 2001).

In our numerical simulations the pressures and temperatures are determined by material

density and specific internal energy through equations of state. However, equations of state

present a problem. ANEOS equation of state (Thompson, Lauson, 1972) is widely used in

hydrodynamic simulations by many authors (Melosh, 1989), but, unfortunately, there is only a

restricted number of materials for which reliable equations of state are available. They are

dunite, granite, quartz, and some other rocks, but there are no equations of state for real targets

and impactors of planetary science interest. Specifically, we have no equations of state for

carbonaceous chondrites, and we have to approximate real rocks by materials with available

equations of state. Tillotson’s (1962) analytical equation of state is also often used, but it does

7
not contain temperature and is inaccurate at phase transitions. Note that temperature is not

included in the hydrodynamic equations, and only dependence of pressure on density and

specific internal energy is necessary for simulations of impacts, but we must calculate the

temperature of a projectile.

Experimental data on equations of state of chondrites are very scarce. Hugoniot curves

obtained in shock compression experiments with Murchison (carbonaceous chondrite of CM

group) and Bruderheim (L6 ordinary chondrite) meteorites (Anderson, Ahrens, 1998) are shown

in Fig.1. Comparison with the ANEOS shows that dunite and quartz equations of state can serve

as approximations for ordinary and carbonaceous chondrites respectively. However, we are

interested not only in dynamical patterns of post-impact flows but also in the degree of material

heating. Post-shock temperatures as functions of shock wave pressures were calculated for H-

and L-chondrites by Schmitt et al. (1994) and for Murchison CM chondrite by Tomioka et al.

(2007). These temperatures are shown in Fig.2. Dunite gives a good approximation of the

heating of ordinary chondrites (the curves for H- and L-chondrites coincide), but Murchison is

heated to higher temperatures than quartz due to its high porosity (~20%). We adopted dunite as

our template for ordinary chondrites and, for lack of a better equation of state, quartz as our

template for hydrated carbonaceous chondrites. In the latter case we should keep in mind that

quartz is heated to lower temperatures than CM carbonaceous chondrites. We used ANEOS

equation of state with input data from (Melosh, 2007). The target was approximated as quartz,

and gabbroic anorthosite with Tillotson’s equation of state (Ahrens, O’Keefe, 1977). Gabbroic

anorthosite is likely a better analog to the composition of the lunar rocks than quartz, but the

ANEOS equation of state for quartz is more accurate than Tillotson’s equation of state for

gabbroic anorthosite that does not include phase transitions. Note that the pressure-particle

velocity Hugoniot relation calculated with the ANEOS equation of state of quartz is close to the

linear shock-particle velocity relation for basalt (Melosh, 1989) founded on the results of shock-

wave experiments. As it is unclear what equation of state gives better approximation to the lunar

8
crust in the hydrodynamic simulations, the ANEOS for quartz or less accurate Tillotson’s

equation of state for gabbroic anorthosite, we used both of them. In simulations of cometary

impacts we used Tillotson’s equation of state of water (Melosh, 1989) for the projectile and the

ANEOS of quartz for the target.

Fig.1. Shock pressure-particle velocity curves from Fig.2. Calculated post-shock temperatures
experimental data for Murchison and Bruderheim as functions of post-shock pressures for H-
meteorites (Anderson, Ahrens, 1998) and for and L-chondrites (Schmitt et al., 1994) and
serpentine (Mader et al., 1980). Comparison with the for Murchison CM chondrite (Tomioka et
curves for dunite and quartz calculated using the al., 2007). Comparison with the post-shock
ANEOS equations of state. temperatures for dunite and quartz
following from the ANEOS equations of
state.
3. Results of simulations

Figure 3 shows the relative masses of projectile materials (dunite and quartz), that are

heated to temperatures below 600, 1000, and 1200 K, as functions of impact velocity for oblique

and vertical impacts. The simulations of the impacts of dunite projectiles, which are a good

approximation for ordinary chondrites, show that even at impact velocity of 14 km/s (and most

probable impact angle 45°) a half of the projectile mass remains solid during the impacts on both

quartz and gabbroic anorthosite targets. About one third of all asteroidal impacts on the Moon

occur at velocities below 14 km/s (see, e.g.., Ivanov, 2008). This confirms the conclusion of

(Yue et al, 2013) that a large quantity of projectile material can remain on the Moon in a solid

state, although being fragmented and deformed. About a half of unmelted projectile material

9
remains in craters. Our results also show that terrestrial meteorites would survive impacts on the

Moon (see Crawford et al., 2008).

10
Fig. 3. Heating of projectile material. Fractions of unmelted impactor material (indicated as
Unm) and fractions of impactor mass (dunite and quartz) heated below given temperature levels,
600, 1000, and 1200 K, during oblique (at 45° ) and vertical impacts on quartz (left column) and
gabbroic anorthosite (GA, right column) targets as functions of impact velocity. Materials of the
projectile and the target are indicated in the upper right corners of the panels.

Quartz melts more easily and is heated to higher temperatures than dunite. The projectile

materials are heated to higher temperatures in vertical impacts. The impacts on gabbroic

anorthosite lead also to higher temperatures in comparison to the impacts on quartz targets, but

only slightly. The curves for the impacts on the gabbroic anorthosite target will be close to the

curves for the quartz target if we shift them along the velocity axis by no more than 1 km/s. At a

velocity of 7 km/s and the most probable impact angle 45° about 20% of the mass of the quartz

impactor (for the impacts on quartz and gabbroic anorthosite targets) are heated below the

temperature of 600 K at which, according to (Akai, 1992), hydrated minerals begin to

decompose. According to our simulations, at the impact angle of 45° almost complete ( 92‒ 95

%) melting of a quartz projectile takes place only at an impact speed of about 13 km/s.

Impactor material is ejected from the crater, but at low impact velocities, a substantial

part of it remains on the crater floor (40 to 60% according to Bland et al., 2008). We have also

obtained large proportions (30 – 40% for oblique and 60 – 70% for vertical impacts) of projectile

material remaining within the crater, which are shown in Fig.4. Comparison of dunite and quartz

variants shows that the total impactor mass remaining in the crater only slightly depends on

projectile and target materials. At vertical impacts greater mass of impactors remains in a crater,

and, despite the impactor material is heated to higher temperatures, the remaining impactor mass

with low temperatures is not smaller than in the case of oblique impacts. Note that in small

simple craters the projectile is dispersed across a crater floor, and in complex craters much of the

projectile debris can be swept back to a central peak by a collapse flow (Yue et al, 2013).

11
Fig. 4. Projectile material in the crater. Fractions of projectile material, total, unmelted (Unm),
and heated below given temperature levels (600, 1000, and 1200 K), remaining within the crater.
Materials of the projectile and the target are indicated in the upper right corners of the panels.
The figure is similar to Fig.3.
12
Comets may also contain hydrated minerals and a substantial fraction of water ice. Water

probably dominates up to ~90% of the volatiles that outflow from the cometary nucleus within

~3–4 AU from the Sun (Combi et al., 2004). Comets may contain a water ice fraction near 50%

by mass, but some estimates for 2P/Encke and C/Hale Bopp comets give only 10% fraction of

water (see Jewitt, 2004). Nevertheless, in numerical simulations comets are typically treated as

bodies made from pure water ice (e.g., cited above Ong et al. (2010), Stewart et al. (2011), and

Prem et al. (2015)). As was mentioned above, after impacts of icy comets with typical velocities

20–50 km/s the major portion of cometary water (95 – 99.9%) escapes the Moon (Ong et al.,

2010). However, a considerable amount of short-period comets, about 25%, hit the Moon at

velocities 8–10 km/s (Jeffers et al., 2001). Bottke et al. (2002) calculated that about 6% of the

near-Earth objects come from the Jupiter-family comet region. They did not include the Halley-

type comets and the long-period comets in their model, however, the impacts of nearly isotropic

comets may be responsible for 10–30% of the craters on Earth (see references in Bottke et al,

2002). Therefore, using modeling results of Bottke et al. (2002) and the data of Jeffers et al.

(2001), we can estimate that at least 1.5% of lunar craters are produced by low-velocity comets.

We have simulated cometary impacts at 8–10 km/s, assuming that the comet consists of

water ice and that the comet consists of water ice with grains of quartz as shown in Fig.5. We

obtained that in both cases all cometary water is vaporized and only less than 1% (which is not

quite well resolved in the simulations) of water in the form of vapor and condensate mixed with

the target material remains in the crater at all impact angles. As these values are beyond the

accuracy of our simulations, we cannot show the diagrams similar to Fig.4, and as it is unclear

how our granular model approximates the real structures of cometary nuclei, we can demonstrate

only qualitative results. The difference between the impacts of an icy comet, a comet with quartz

grains, and an asteroid is shown in Fig. 5. After the impact of an asteroid at a speed of 10 km/s

its material, partially melted, settles on the crater surface. There is no silicate vapor. After the

13
impact of an icy comet all water is vaporized and forms a plume which, in 0.6 s after the impact,

has various densities and expands into vacuum. All water from the comet with grains is also

vaporized and forms a plume of low density, but the grain material remains on the crater surface

as shown in the lower panel of Fig.5. Thus, despite almost all cometary water vapor being

ejected from the crater, if a comet contains hydrated minerals, they can partly remain in the

crater. This depends on the cometary structure which is not well known. We estimate that comets

may potentially deliver water in the form of hydrated minerals to about 1.5% of all craters.

Fig. 5. Comparison of crater formation after asteroidal and cometary impacts. Plots of density
show the impacts of an asteroid in the upper panel, a comet from water ice in the middle panel,
and a comet as a mixture of water and quartz (as shown in the figure) in the lower panel. A scale
of densities is shown on the upper panel in g/cm3. Target material is hatched. Impactor diameters

14
are 1 km, the impact velocity is 10 km/s, the impact angle is 45°, and time is 0.6 s after the
contact of impactors with the target.

4. Discussion and conclusions

We have obtained that after oblique impacts at velocities below 10 km substantial

fractions of masses of quartz projectiles are heated to temperatures below 1000 K, at which

some, maybe little, portion of phyllosilicates can survive. However, we should extend the results

to the impacts of hydrated chondrites. Shock experiments (Tomioka et al., 2007) with Murchison

meteorite show that Fe-Mg serpentine in the carbonaceous chondrite can survive up to a post-

shock pressure of 36 GPa. At this pressure vaporization of water from hydrous minerals took

place by shock heating. At a pressure of 21 GPa the amount of serpentine in the chondrite matrix

decreased (and accordingly the amount of Si-rich glasses increased), and at 30 GPa serpentine

occurred in minor abundances as relatively coarse grains. Only at 49 GPa the matrix of

Murchison was melted, and the matrix mineralogy and texture have been drastically changed

compared with the samples shocked at lower pressures. As shown in Fig.2, the shock heating of

quartz to a temperature of 1000 K corresponds to shock pressures below 35 GPa, and the heating

of quartz to a temperature of 600 K corresponds to a shock pressure of 22 GPa. As follows from

Fig.1, quartz and Murchison chondrite will be compressed to approximately the same pressures

at a given impact velocity. Therefore, at the most probable impact angle 45°, some amount of

serpentine in hydrated impactors can survive the impacts and remain in a crater up to an impact

velocity of 12 km/s (see Fig. 4).

Some portion of impactor material remains unmelted at sufficiently high speeds (Fig. 3)

when the heating exceeds the threshold of complete dehydration of minerals. If the body is

broken up into relatively large fragments, e.g., ~1 cm, the released water could remain inside

solid fragments, despite the temperature of fragments being above 1200 K. The sizes of some

solid impactor fragments might be large, however, it is difficult to estimate them because the

15
results of laboratory experiments cannot be extended to large impacts – strength of laboratory

projectiles is much higher than strength of asteroids. Most of stony meteoroids are very weak

and break up during the fall in the Earth’s atmosphere at pressures below 10 MPa. For this

reason findings of impactor fragments are rare. The sizes of the largest fragments of an iron

body, which had produced the Barringer crater, were about 30 cm (Kring, 2007), and the

fragments of an L-chondrite impactor from the Morokweng crater (about 70 km in diameter)

reach 25 cm (Maier et al., 2006). However, recovered remnants of impactors are usually much

smaller, and, additionally, carbonaceous chondrites are more fragile than ordinary chondrites and

even more so than iron meteorites. Investigated particles of projectile relics in lunar samples are

small, from 20 to 225 µm (Joy et al., 2012). In the experiments on dehydration of carbonaceous

material the sizes of grains were smaller than 100 µm (Garenne et al., 2014). So it is likely that

the projectile fragments are small and water is released in the form of vapor, but, nevertheless,

we cannot exclude existence of relatively large (e.g., ~10 cm) fragments at low impact velocities.

At an impact velocity of 7 km/s and impact angle 45° about 20% of quartz projectile are

heated to temperatures below 600 K and remain in a crater (Fig.4). These temperatures

correspond to shock pressures below 22 GPa, when serpentine in Murchison chondrite was

decomposed only partly, let it be by 50% (Tomioka et al. (2007) do not give quantitative

estimates). A 10-km-diameter crater produced by the impact of a 1-km-diameter highly hydrated

asteroid will contain 5×1011 kg (37%, Fig.4) of unmelted impactor material which is equivalent

to a layer of impactor material, about 2.5 m thick, spread uniformly over a crater floor. This

impactor material will have more than 2% of chemically bound water, if the impactor initially

had 20%. Vapor of released water runs out from the layer of impactor material and can form a

local cloud because the floor of a fresh impact crater has a sufficiently high temperature. But

some amount of released free water can be locked in a layer of projectile fragments at the crater

floor. Thus, water, which is delivered to the Moon by the impacts of hydrous asteroids, can exist

in high concentrations in residues of meteorite material in the craters. One impact of a 2-km-

16
diameter highly hydrated asteroid at a low velocity near the lunar pole can produce more water

on the crater floor than all cometary impacts in cold traps during 1 Ga.

Importance of asteroidal delivery of water to the Moon depends on the frequency of the

impacts of water-rich objects with relatively low velocities. According to the Catalogue of

Meteorites (Grady, 2000) the falls of CI, CM, and CR meteorites on the Earth make up about 3%

of the total falls of stony bodies. This number may be smaller than the number of carbonaceous

chondrite impacts on the Moon because other stony bodies, such as ordinary chondrites, are

stronger and more easily survive the passage through the Earth’s atmosphere. Asteroids of C-

and P-types, presumably analogues of the dark CI and CM carbonaceous chondrite meteorites,

account for more than half of the main-belt and Trojan asteroids by mass if we exclude the four

most massive asteroids (DeMeo, Carry, 2013). About two-thirds of C-type asteroids have

hydrated silicate surfaces (Jones et al., 1990). Hydrated asteroids have various sizes (DeMeo,

Carry, 2013) and their relative number grows with the distance from the Sun. However, the

population of near-Earth objects (NEO) is replenished primarily from the inner regions of the

main belt. According to the estimates (Bottke et al., 2002), about 24% of NEO population comes

from the central main belt, ~8% comes from the outer main belt, and ~6% comes from the

Jupiter-family comet region. The estimated relative number of C-class asteroids among all Earth-

approaching asteroids is from about 15 % (Lupishko, Lupishko, 2001) to 30% (Shoemaker et al.,

1990). Assuming that two-thirds of them are hydrated (Jones et al., 1990) we estimate that the

relative number of water-bearing asteroidal impacts on the Moon is 10-20%. Low velocity

impacts are not frequent. About 25% of lunar impacts occur at speeds below 12 km/s, about 15%

of the impacts occur at speeds below 10 km/s, and only 1.5% occur at speeds below 6 km/s

(Marchi et al., 2009; Yue et al, 2013). Assuming that some amount of projectile material

containing water survives the impacts at velocities below 10 km/s, we estimate that from 2% to

4.5% of asteroidal craters can contain hydrated meteorite remnants. This chemically combined

water can exist even in the craters illuminated by the Sun.

17
Two experimental impacts have been made into lunar craters with high detected

concentration of hydrogen. The impact of Lunar Prospector (a mass of 160 kg at a velocity of 1.7

km/s) into the shadowed floor of the Shoemaker crater revealed no water (Goldstein et al., 2001).

Such a low velocity is insufficient to release water from a crater if it is in the form of hydrated

minerals. In the LCROSS mission a rocket with a mass of 2400 kg struck a shadowed region

within the lunar south pole crater Cabeus at a velocity of 2.5 km/s, ejecting debris, dust, and

vapor (Colaprete et al., 2010). The total mass of water vapor and water ice within the field of

view of the instrument onboard the second spacecraft was estimated as 155 ± 12 kilograms

(Colaprete et al., 2010). The speed of 2.5 km/s is also insufficient for substantial dehydration of

minerals, and, therefore, it is likely that the LCROSS impact ejected free water (ice) contained in

regolith. As the cometary impacts are rarer and deliver little water to the craters, the ice

contained in the Cabeus crater with greater probability could be derived from hydrated material

of a stony projectile that created this crater.

We conclude that at impact velocities below~10 km/s the bulk of a stony asteroid is

unmelted, and some fraction of hydrous minerals contained in the projectile is heated to

temperatures below 1000 K, decomposes only partly and remains in the crater. Along with the

implantation of hydrogen from solar wind and condensation from water vapor transient

atmospheres produced by high-velocity impacts of comets and asteroids, low-velocity impacts

can deliver water to the whole lunar surface and to the craters. Low-velocity impacts of hydrated

asteroids can produce the highest concentrations of hydrogen in the craters both in the form of

water condensate and chemically combined water. Observable number of these craters should

exist on the Moon.

Note that the impacts with an average velocity about 5 km/s are typical for collisions

among the min belt asteroids. Many large stony asteroids with a dry surface may have spots of

hydrated material on the surface due to collisions with water-rich objects from the outer regions

of the Solar system. Geomorphologic features of water flows on the walls of young impact

18
craters on Vesta (Scully et al., 2015) could be formed by water released from hydrated asteroids

collided with Vesta.

Acknowledgements

This work has been supported by the Russian Foundation for Basic Research, project no.

13-05-00694-a. The authors gratefully thank the editor and reviewers of this paper for their

insightful comments and suggestions.

References

Ahrens, T.J., O’Keefe, J.D., 1977. Equations of state and impact-induced shock-wave

attenuation on the moon. In: Roddy, D.J., Pepin R.O., Merrill, R.B. (Eds,), Impact and

Explosion Cratering. Pergamon Press, NY, pp. 639–656.

Akai, J., 1990. Mineralogical evidence of heating events in Antarctic carbonaceous chondrites,

Yamato-86720 and Yamato-82162. In: Proceedings of the NIPR Symposium on Antarctic

Meteorites 3, pp. 55–68.

Akai, J., 1992. T-T-T diagram of serpentine and saponite, and estimation of metamorphic

heating degree of Antarctic carbonaceous chondrites. In: Proceedings of the NIPR

Symposium on Antarctic Meteorites 5, pp. 120–135.

Anand, M., 2010. Lunar water: A brief review. Earth Moon Planets 107, 65–73.

Anderson, W.W., Ahrens, T.J., 1998. Shock wave equations of state of chondritic meteorites. In:

Schmidt, S.C., Dandelcar, D.P., Forbes, J.W. (Eds.), Shock Compression of Condensed

Matter. AIP Press, Woodbury, NY, pp. 115-118.

Barnes, J.J., Tartèse, R., Anand, M., McCubbin, F.M., Franchi, I.A., Starkey, N.A., Russell, S.S.,

2014. The origin of water in the primitive Moon as revealed by the lunar highlands

samples. Earth Planet. Sci. Lett. 390, 244–252.

19
Bland, P.A., Artemieva, N.A., Collins, G.S., Bottke, W.F., Bussey, D.B.J., Joy, K.H., 2008.

Asteroids on the Moon: Projectile survival during low velocity impact. Lunar Planetary

Science Conference XXXIX, abstract #2045.

Bottke, W.F., Morbidelli, A., Jedicke, R., Petit, J.-M., Levison, H.F., Michel, P., Metcalfe, T.S.,

2002. Debiased orbital and absolute magnitude distribution of the near-Earth objects.

Icarus 156, 399–433.

Clark, R.N., 2009. Detection of adsorbed water and hydroxyl on the Moon. Science 326, 562–

564.

Colaprete, A., Schultz, P., Heldmann, J., Wooden, D., Shirley, M., Ennico, K., Hermalyn, B.,

Marshall, W., Ricco, A., Elphic, R.C., Goldstein, D., Summy, D., Bart, G.D., Asphaug, E.,

Korycansky, D., Landis, D., Sollitt, L., 2010. Detection of water in the LCROSS ejecta

plume. Science 330, 463–468.

Combi, M.R., Harris, W.M., Smyth, W.H., 2004. Gas dynamics and kinetics in the cometary

coma: Theory and observations. In: Festou, M.C., Keller, H.U., Weaver, H.A. (Eds.),

Comets II. University of Arizona Press, Tucson, Arizona, pp. 523–552.

Crawford, I.A., Baldwin, E.C., Taylor, E.A., Bailey, J.A., Tsembelis, K., 2008. On the

survivability and detectability of terrestrial meteorites on the Moon. Astrobiology 8, 242–

252.

DeMeo, F.T., Carry, B., 2013. The taxonomic distribution of asteroids from multi-filter all-sky

photometric surveys. Icarus 226, 723–741.

Dienes, J.K., Walsh, J.M., 1970. Theory of impact: Some general principles and the method of

Eulerian codes. In: Kinslow, R. (Ed.), High-Velocity Impact Phenomena. Academic Press,

New York, pp. 46–104.

Feldman, W.C, Maurice, S., Binder, A.B, Barraclough, B.L., Elphic, R.C., Lawrence, D.J., 1998.

Fluxes of fast and epithermal neutrons from Lunar Prospector: Evidence for water ice at

the lunar poles. Science 281, 1496–1500.

20
Garenne, A., Beck, P., Montes-Hernandez, G., Chiriac, R., Toche, F., Quirico, E., Bonal, L.,

Schmitt, B., 2014. The abundance and stability of “water” in type 1 and 2 carbonaceous

chondrites (CI, CM and CR). Geochim. Cosmochim. Acta 137, 93–112.

Goldstein, D.B., Austin, J.V., Barker, E.S., Nerem, R.S., 2001. Short-time exosphere evolution

following an impulsive vapor release on the Moon. J. Geophys. Res. 106, 32841–32846.

Grady, M.M., 2000. Catalogue of Meteorites. Cambridge Univ. Press.

Greenwood, J.P., Itoh, S., Sakamoto, N., Warren, P., Taylor, L.A., Yurimoto, H., 2011.

Hydrogen isotope ratios in lunar rocks indicate delivery of cometary water to the Moon.

Nature Geoscience 4, 79–82.

Hauri, E.H., Saal, A.E., Rutherford, M.J., VanOrmanc, J.A., 2015. Water in the Moon’s interior:

Truth and consequences. Earth Planet. Sci. Lett. 409, 252–264.

Ivanov, B, 2008. Size-frequency distribution of asteroids and impact craters: Estimates of impact

rate. In Adushkin, V., Nemchinov, I. (Eds.). Catastrophic Events Caused by Cosmic

Objects. Springer, Dordrecht, pp. 91−116.

Ivanova, M.A., Lorenz, C.A., Nazarov, M.A., Brandstaetter, F., Franchi, I.A., Moroz, L.V.,

Clayton, R.N., Bychkov, A.Yu., 2010. Dhofar 225 and Dhofar 735: Relationship to CM2

chondrites and metamorphosed carbonaceous chondrites, Belgica-7904 and Yamato-

86720. Meteorit. Planet. Sci. 45, 1108–1123.

Ivanova, M.A., Lorenz, C.A., Franchi, I.A., Bychkov, A.Y., Post, J.E., 2013. Experimental

simulation of oxygen isotopic exchange in olivine and implication for the formation of

metamorphosed carbonaceous chondrites. Meteorit. Planet. Sci. 48, 2059–2070.

Jeffers, S.V., Manley, S.P., Bailey, M.E., Asher, D.J., 2001. Near-Earth object velocity

distributions and consequences for the Chicxulub impactor. Mon. Not. R. Astron. Soc. 327,

126–132.

21
Jewitt, D.C., 2004. From cradle to grave: The rise and demise of the comets. In: Festou, M.C.,

Keller, H.U., Weaver, H.A. (Eds.), Comets II. University of Arizona Press, Tucson,

Arizona, pp. 659–676.

Jones, T.D., Lebofsky, L.A., Lewis, J.S., Marley, M. S., 1990. The composition and origin of the

C, P, and D asteroids – Water as a tracer of thermal evolution in the outer belt. Icarus 88,

172–192.

Joy, K.H., Zolensky, M,E., Nagashima, K., Huss, G.R., Ross, D.K., McKay, D.S., Kring, D.A.,

2012. Direct detection of projectile relics from the end of the lunar basin-forming epoch.

Science 336, 1426–1429.

Kring, D.A., 2007. Guidebook to the Geology of Barringer Meteorite Crater, Arizona (a.k.a.

Meteor Crater). Lunar and Planetary Institute, Houston, Texas.

Liu, Y., Guan, Y., Zhang, Y., Rossman, G.R., Eiler, J.M., Taylor, L.A., 2012. Direct

measurement of hydroxyl in the lunar regolith and the origin of lunar surface water. Nature

Geoscience 5, 779-782.

Lupishko, D.F., Lupishko, T. A., 2001. On the origins of Earth-approaching asteroids. Sol. Sys.

Res. 35, 227–233.

Mader, C.L., Gibbs, T.R., Hopson, J.W., Marsh, S.P., Hoyt, M.S., Thayer, K.V., 1980. LASL

Shock Hugoniot Data. University of California Press, Berkeley, Los Angeles, London.

Maier, W.D., Andreoli, M.A.G., McDonald, I., Higgins, M.D., Boyce, A.J., Shukolyukov, A.,

Lugmair, G.W., Ashwal, L.D., Gräser, P., Ripley, E.M., Hart, R.J., 2006. Discovery of a

25-cv asteroid clast in the giant Morokweng impact crater, South Africa. Nature 441, 203–

206.

Marchi, S., Mottola, S., Cremonese, G., Massironi, M., Martellato, E., 2009. A new chronology

for the Moon and Mercury. Astrophys. J. 137, 4936–4948.

22
McCubbin, F.M., Steele, A., Hauri, E.H., Nekvasil, H., Yamashita, S., Hemley, R.J., 2010.

Nominally hydrous magmatism in the Moon. Proceedings of the National Academy of

Sciences 107, 11223-11228.

McCubbin F.M., Vander Kaaden, K.E., Tartèse, R., Klima, R.L., Liu, Y., Mortimer, J., Barnes,

J.J., Shearer, C.K., Treiman, A.H., Lawrence, D.J., Elardo, S.M., Hurley, D.M., Boyce,

J.W., Anand, M., 2015. Magmatic volatiles (H, C, N, F, S, Cl) in the lunar mantle, crust,

and regolith: Abundances, distributions, processes, and reservoirs. American Mineralogist

100, 1668–1707.

Melosh, H.J., 1989. Impact Cratering: A Geologic Process. Oxford University Press, New York.

Melosh, H.J., 2007. Hydrocode equation of state for SiO2. Meteorit. Planet. Sci. 42, 2079–2098.

Mitrofanov, I.G., Sanin, A.B., Boynton, W.V., Chin, G.. Garvin, J.B., Golovin, D., Evans, L.G.,

Harshman, K., Kozyrev, A.S., Litvak, M.L., Malakhov, A., Mazarico, E., McClanahan, T.,

Milikh, G., Mokrousov, M., Nandikotkur, G., Neumann, G.A., Nuzhdin, I., Sagdeev, R.,

Shevchenko, V., Shvetsov, V., Smith, D.E., Starr, R., Tretyakov, V.I., Trombka, J.,

Usikov, D., Varenikov, A., Vostrukhin, A., Zuber, M.T., 2010. Hydrogen mapping of the

lunar south pole using the LRO neutron detector experiment LEND. Science 330, 483–

486.

Mitrofanov, I., Litvak, M., Sanin, A., Malakhov, A., Golovin, D., Boynton, W., Droege, G.,

Chin, G., Evans, L., Harshman, K., Fedosov, F., Garvin, J., Kozyrev, A., McClanahan, T.,

Milikh, G., Mokrousov, M., Starr, R., Sagdeev, R., Shevchenko, V., Shvetsov, V.,

Tret'yakov, V., Trombka, J., Varenikov, A., Vostrukhin, A., 2012. Testing polar spots of

water-rich permafrost on the Moon: LEND observations onboard LRO. J. Geophys. Res.

117, E00H27, doi:10.1029/2011JE003956.

Ong, L., Asphaug, E.I., Korycansky, D., Coker, R.F., 2010. Volatile retention from cometary

impacts on the Moon. Icarus 207, 578–589.

23
Pieters, C.M., Goswami, J.N., Clark, R.N., Annadurai, M., Boardman, J., Buratti, B., Combe, J.-

P., Dyar, M.D., Green, R., Head, J.W., Hibbitts, C., Hicks, M., Isaacson, P., Klima, R.,

Kramer, G., Kumar, S., Livo, E., Lundeen, S., Malaret, E., McCord, T., Mustard, J.,

Nettles, J., Petro, N., Runyon, C., Staid, M., Sunshine, J., Taylor, L.A., Tompkins, S.,

Varanasi, P., 2009. Character and spatial distribution of OH/H2O on the surface of the

Moon seen by M3 on Chandrayaan-1. Science 326, 568–572.

Potter, R.W.K., Collins, G.S., 2013. Numerical modeling of asteroid survivability and possible

scenarios for the Morokweng crater-forming impact. Meteorit. Planet. Sci. 48, 744–757.

Prem, P., Artemieva, N.A., Goldstein, D.B., Varghese, P.L., Trafton, L.M., 2015. Transport of

water in a transient impact-generated lunar atmosphere. Icarus 255, 148–158.

Saal, A.E., Hauri, E.H., Cascio, M.L., Van Orman, J.A., Rutherford, M.J., Cooper, R.F., 2008.

Volatile content of lunar volcanic glasses and the presence of water in the Moon’s interior.

Nature 454, 192–195.

Saal, A.E., Hauri, E.H., Van Orman, J.A., Rutherford, M.J., 2013. Hydrogen isotopes in lunar

volcanic glasses and melt inclusions reveal a carbonaceous chondrite heritage. Science,

340, 1317–1320.

Sanin, A.B., Mitrofanov, I.G., Litvak, M.L., Boynton, W.V., Chin, G., Evans, L.G., Golovin,

D.V., Harshman, K., Livengood, T.A., McClanahan, T.P., Malakhov, A.V., Mokrousov,

M.I., Sagdeev, R.Z., Starr, R.D., 2014. Estimation of hydrogen concentration in lunar south

polar regions. Lunar Planetary Science Conference XLV, abstract #1358.

Schmitt, R.T., Deutch, A., Stoffler, D., 1994. Calculation of Hugoniot curves and post-shock

temperatures for H- and L-chondrites. Abstr. Lunar Planetary Science Conference XXV,

1209–1210.

Scully, J.E.C., Russell, C.T., Yin, A., Jaumann, R., Carey, E., Castillo-Rogez, J., McSween,

H.Y., Raymond, C.A., Reddy, V., Le Corre, L., 2015. Geomorphological evidence for

transient water flow on Vesta. Earth Planet. Sci. Lett. 411, 151–163.

24
Shoemaker, E.M., Wolfe, R.M., Shoemaker, C.S., 1990. Asteroid and comet flux in the

neighborhood of Earth. Geological Society of America Special Paper 247, 155–170.

Shuvalov, V.V., 1999. Multi-dimensional hydrodynamic code SOVA for interfacial flows:

Application to the thermal layer effect. Shock Waves 9, 381–390.

Stewart, B D., Pierazzo, E., Goldstein, D.B., Varghese, P.L., Trafton, L.M., 215. Simulations of

a comet impact on the Moon and associated ice deposition in polar cold traps. Icarus 215,

1–16.

Sunshine, J.M., Farnham, T.L., Feaga, L.M., Groussin, O., Merlin, F., Milliken, R.E., A'Hearn,

M,F., 2009. Temporal and spatial variability of lunar hydration as observed by the Deep

Impact Spacecraft. Science 326, 565–568.

Tartèse, R., Anand, M., 2013. Late delivery of chondritic hydrogen into the lunar mantle:

Insights from mare basalts. Earth Planet. Sci. Lett. 361, 480–486.

Tartèse, R., Anand, M., Barnes, J.J., Starkey, N.A., Franchi, I.A., Sano, Y., 2013. The

abundance, distribution, and isotopic composition of Hydrogen in the Moon as revealed by

basaltic lunar samples: Implications for the volatile inventory of the Moon. Geochim.

Cosmochim. Acta, 122, 58-74

Thompson, S.L., Lauson, H.S. 1972. Improvements in the Chart D Radiation-Hydrodynamic

CODE III: Revised Analytic Equations of State, Report SC-RR-71 0714. Sandia National

Laboratory, Albuquerque, NM.

Tillotson, J.H., 1962. Metallic Equations of State for Hypervelocity Impact; General Atomic

Report GA-3216. Advanced Research Project Agency, San Diego.

Tomioka, N., Tomeoka, K., Nakamura-Messenger, K., Sekine, T., 2007. Heating effects of the

matrix of experimentally shocked Murchison CM chondrite: Comparison with

micrometeorites. Meteorit. Planet. Sci. 42, 19–30.

Yue, Z., Johnson, B.C., Minton, D.A., Melosh, H.J., Di, K., Hu, W., Liu, Y., 2013. Projectile

remnants in central peaks of lunar impact craters. Nature Geoscience 6, 435–437.

25

You might also like