ArticlePDF Available

Abstract and Figures

Recent images of the surface of Mercury have revealed an unusual and intriguing landform: sub-kilometre scale, shallow, flat-floored, steep-sided rimless depressions typically surrounded by bright deposits and generally occurring in impact craters. These ‘hollows’ appear to form by the loss of a moderately-volatile substance from the planet’s surface and their fresh morphology and lack of superposed craters suggest that this process has continued until relatively recently (and may be on-going). Hypotheses to explain the volatile-loss have included sublimation and space weathering, and it has been suggested that hollow-forming volatiles are endogenic and are exposed at the surface during impact cratering. However, detailed verification of these hypotheses has hitherto been lacking.
Content may be subject to copyright.
Hollows on Mercury: Materials and mechanisms involved in their
formation
q
Rebecca J. Thomas
a,
, David A. Rothery
a
, Susan J. Conway
a
, Mahesh Anand
a,b
a
Department of Physical Sciences, The Open University, Walton Hall, Milton Keynes MK7 6AA, UK
b
Department of Earth Sciences, Natural History Museum, Cromwell Road, London SW7 5BD, UK
article info
Article history:
Received 30 May 2013
Revised 17 October 2013
Accepted 13 November 2013
Available online 21 November 2013
Keywords:
Mercury, surface
Geological processes
Impact processes
abstract
Recent images of the surface of Mercury have revealed an unusual and intriguing landform: sub-kilome-
tre scale, shallow, flat-floored, steep-sided rimless depressions typically surrounded by bright deposits
and generally occurring in impact craters. These ‘hollows’ appear to form by the loss of a moderately-
volatile substance from the planet’s surface and their fresh morphology and lack of superposed craters
suggest that this process has continued until relatively recently (and may be on-going). Hypotheses to
explain the volatile-loss have included sublimation and space weathering, and it has been suggested that
hollow-forming volatiles are endogenic and are exposed at the surface during impact cratering. However,
detailed verification of these hypotheses has hitherto been lacking.
In this study, we have conducted a comprehensive survey of all MESSENGER images obtained up to the
end of its fourth solar day in orbit in order to identify hollowed areas. We have studied how their location
relates to both exogenic processes (insolation, impact cratering, and solar wind) and endogenic processes
(explosive volcanism and flood lavas) on local and regional scales. We find that there is a weak correlation
between hollow formation and insolation intensity, suggesting formation may occur by an insolation-
related process such as sublimation. The vast majority of hollow formation is in localised or regional
low-reflectance material within impact craters, suggesting that this low-reflectance material is a vola-
tile-bearing unit present below the surface that becomes exposed as a result of impacts. In many cases
hollow occurrence is consistent with formation in volatile-bearing material exhumed and exposed during
crater formation, while in other cases volatiles may have accessed the surface later through re-exposure
and possibly in association with explosive volcanism. Hollows occur at the surface of thick flood lavas
only where a lower-reflectance substrate has been exhumed from beneath them, indicating that this form
of flood volcanism on Mercury lacks significant concentrations of hollow-forming volatiles.
Ó2013 The Authors. Published by Elsevier Inc. All rights reserved.
1. Introduction
The presence of morphologically fresh depressions on the sur-
face of Mercury has been one of the most surprising discoveries
of the MESSENGER (MErcury Surface, Space ENvironment, GEo-
chemistry, and Ranging) spacecraft. Though areas of hollows had
been imaged at low resolution by the Mariner 10 spacecraft in
the 1970s, they appeared only as high-reflectance, spectrally rela-
tively blue patches on the floors of impact craters (BCFDs Bright
Crater Floor Deposits) (Dzurisin, 1977; Robinson et al., 2008; Blew-
ett et al., 2009). When MESSENGER went into orbit in 2011 and ob-
tained higher-resolution images, these were revealed to be clusters
of irregular rimless depressions with flat floors and steep walls
(Fig. 1). These were dubbed ‘hollows’ to distinguish them from dee-
per ‘pits’ with sloping floors, which are proposed to form through
magmatic processes (Gillis-Davis et al., 2009; Kerber et al., 2011).
They range from individual hollows tens of meters across to clus-
ters of hollows tens of kilometres across (Blewett et al., 2011)
and shadow measurements indicate a consistent depth within a
particular host crater in the range of tens of meters (Blewett
et al., 2011; Vaughan et al., 2012). Though their consistent depths
make them flat-floored overall, lumps of material do occur on hol-
low floors that may be degraded remnants of the original surface
(Blewett et al., 2011). The bright deposits that gave BCFDs their
name are revealed from orbit to occur both on hollow floors and
as surrounding haloes.
Hollows appear morphologically fresh and lack superposed im-
pact craters. This implies a young age and suggests hollow forma-
tion may be an on-going process (Blewett et al., 2011). If so, it will
be important to distinguish whether it is a gradual, continual pro-
cess or a more rapid, episodic process.
0019-1035/$ - see front matter Ó2013 The Authors. Published by Elsevier Inc. All rights reserved.
http://dx.doi.org/10.1016/j.icarus.2013.11.018
q
This is an open-access article distributed under the terms of the Creative
Commons Attribution License, which permits unrestricted use, distribution, and
reproduction in any medium, provided the original author and source are credited.
Corresponding author.
E-mail address: rebecca.thomas@open.ac.uk (R.J. Thomas).
Icarus 229 (2014) 221–235
Contents lists available at ScienceDirect
Icarus
journal homepage: www.elsevier.com/locate/icarus
The flat-floored, closed morphology of hollows and the lack of
associated outflow features suggest that they form by the preferen-
tial loss of a volatile component from the surface without melting.
The nature of this material is not known: sulphides or chlorides are
possible candidates (Vaughan et al., 2012; Blewett et al., 2013;
Helbert et al., 2013) but the current resolution of surface elemental
compositional data (Goldsten et al., 2007; Hawkins et al., 2007;
Schlemm et al., 2007; Nittler et al., 2011; Peplowski et al., 2011;
Evans et al., 2012; Weider et al., 2012) is not sufficient to verify this
at the scale of hollows.
Several possible release mechanisms for this volatile substance
have been suggested (Blewett et al., 2011). The feasibility of these
processes is dependent on the nature of the substance lost and the
timescale of hollow formation. In light of the high daytime surface
temperatures at Mercury and the morphological similarity be-
tween hollows and the ‘Swiss cheese terrain’ of Mars (Thomas
et al., 2000), sublimation is a strong candidate. However, various
forms of space weathering are believed to occur at Mercury, and
these may be important release mechanisms if they are relatively
intense in the material where hollows form. Along with thermal
desorption (Madey et al., 1998), photon stimulated desorption
(PSD) releases alkalis from the surface and may be the most effi-
cient form of space weathering supplying these elements to the
exosphere (Cheng et al., 1987; Mura et al., 2009). On a shorter
timescale and at particular localities, physical (Killen et al., 2004)
and perhaps chemical (Potter, 1995) sputtering by the solar wind
may also be important, with the intensity of these processes
depending on the time-variable interaction between the solar wind
and the planet’s magnetic field. Micrometeorite impact vaporisa-
tion also releases material from the surface, and unlike the pro-
cesses mentioned above, penetrates beyond the layer of atoms at
the extreme surface (Killen et al., 2007). The rate of hollow forma-
tion may however be too fast for this to apply (Blewett et al., 2011).
Hollows usually occur in material with a low reflectance rela-
tive to the Hermean average. In some cases this is a regional depos-
it equating to the LRM (low-reflectance material) spectral unit that
has been mapped over large areas of the planetary surface (Denevi
et al., 2009), in others it is a localised deposit, and in a few cases it
is a small ‘dark spot’ with even lower reflectance than LRM (Xiao
et al., 2013). Low-reflectance material may therefore be the vola-
tile-bearing unit that degrades to form hollows. The question arises
of how the volatile component in this material has been able to ac-
cess the surface recently enough to form fresh landforms despite
considerable evidence for global contraction (Strom et al., 1975;
Watters et al., 2009): this stress state would tend to hinder migra-
tion of material through the crust. The correlation of hollow forma-
tion with impact craters strongly suggests that impacts are
involved in bringing the hollow-forming volatiles to the surface.
It has been suggested that this may occur through exposure in cra-
ter walls, floors and ejecta and exhumation in peak structures
(Blewett et al., 2013), or through differentiation of impact melt
(Vaughan et al., 2012).
A deeper understanding of the distribution and mode of occur-
rence of hollows is of great interest because of the probable rela-
tionship between hollows and volatile percentage in Mercury’s
crust, now understood to be higher than previously thought
(Kerber et al., 2009, 2011; Nittler et al., 2011; Peplowski et al.,
2011). We have therefore conducted a full survey of MESSENGER
images of Mercury’s surface. This comprehensive survey has al-
lowed identification of many areas of hollow formation not previ-
ously recognised, building on the global inventory published by
Blewett et al. (2013). We have recorded the extent, location and
associations of the observed hollow clusters, and examined latitu-
dinal and longitudinal variations in their areal extent. We consider
how their occurrence and extent may be controlled by external fac-
tors such as insolation and ion sputtering or endogenic processes
such as the formation of pyroclastic pits or surficial coverage by
thick volcanic plains. On a local scale, we have examined the slope
aspects in locations where hollows occur on slopes, in order to test
whether there is a correlation with insolation intensity, and have
studied the local settings of hollow formation to evaluate possible
exposure mechanisms for hollow-forming volatiles.
2. Methods
2.1. MESSENGER imagery
We examined images taken by MESSENGER’s Mercury Dual
Imaging System (MDIS) (Hawkins et al., 2007, 2009) up to the
end of MESSENGER’s fourth solar day in orbit around Mercury
(product creation times up to March 17, 2013). Monochrome
images were used to identify hollows and study them in detail,
and lower resolution colour composites were used to determine
the spectral character of their associated deposits and substrates.
2.1.1. Monochrome images
We examined all MESSENGER monochrome images with reso-
lutions of less than 180 m/px, excluding images at lower resolu-
tions because they do not reveal the irregular margins, flat-floors
and rimlessness that distinguish hollows from small impact cra-
ters. These images were obtained by the 1.5°field-of-view Narrow
Angle Camera (NAC) and the 748.7 nm filter in the 10.5°field-of-
view Wide Angle Camera (WAC) of MDIS. The highest-resolution
images used were 7.7 m/px and the average resolution of those
available was 106 m/px.
We applied radiometric and photometric correction to all
images using the ISIS3 image processing package of the USGS.
We then overlaid these onto the 250 m/px global monochrome
mosaic version 9 produced by the MESSENGER team (released by
NASA’s Planetary Data System on 8 March 2013) and digitised fea-
tures on this global mosaic.
2.1.2. Colour images
To characterise the spectral type of local and regional sub-
strates, we examined colour composites created by combining data
from three of the twelve spectral filters in the WAC. All major sub-
strates on Mercury have red-sloped reflectance spectra (Denevi
et al., 2009), but the steepness of this slope varies, allowing some
Fig. 1. Irregular, rimless hollows on the floor and terraced wall of an unnamed
impact crater at 46.4°N, 318.7°E. Black arrows indicate individual hollows, white
bracket indicates a cluster (MESSENGER image ID 2760274).
222 R.J. Thomas et al. / Icarus 229 (2014) 221–235
to be classified as red or blue relative to the Hermean average. By
combining reflectance at 996 nm, 749 nm and 433 nm in the red,
green and blue bands, we were able to see these variations and
attribute substrates to the spectral types established by Denevi
et al. (2009), which are believed to indicate real compositional
and geological differences between surface units.
All images at a resolution of less than 1000 m/px were exam-
ined, as was the 1000 m/px global colour mosaic version 3 pro-
duced by the MESSENGER team (released by NASA’s Planetary
Data System on 8 March 2013). The highest-resolution composite
created was 64 m/px and the average resolution was 455 m/px.
2.2. Data collected
2.2.1. Hollows and pits
Data on all non-impact-related depressions visible in the
images were gathered in order to ensure that a distinction was
made between hollows and pits with a probable magmatic
(Gillis-Davis et al., 2009) or pyroclastic (Kerber et al., 2011) origin
and to make spatial comparisons between this activity and hollow
formation. Impact-related craters were distinguished from pits and
hollows on the basis of their circular shape, raised rims and the
characteristic geometry of their ejecta blankets.
The steepness of a depression’s margins and the spectral signa-
ture of its associated deposits were used to distinguish between
hollows and pits: hollows have steep margins leading to flat floors
and bluer deposits while pits have gentler slopes, are deeper, and
any surrounding deposits are redder (Table 1). We identified a
third previously unidentified type of depression that is intermedi-
ate in character between pits and hollows: areas of pitted ground
floored by relatively red deposits. These either lack defined mar-
gins or have steep margins that appear less crisp than those of hol-
lows. Where they have defined margins, they are intermediate in
depth between hollows and pits. The presence of relatively red
deposits in these regions, the lack of relatively blue deposits and
their smoother morphology suggests that these are not hollows.
The similarity of the spectral character of their deposits to those
of pits, which are suggested to be formed by explosive volcanism
(Kerber et al., 2011), may indicate a volcanic origin.
For each depression, we gathered data on its geographical loca-
tion, area, association with tectonic structures such as thrust faults,
the spectral type of the local and regional substrate, and the type of
material hosting it. We identified the spectral type of the regional
substrate by reference to global mapping by Denevi et al. (2009,
2013) and our own observations of colour composite images, dis-
tinguishing between regional low-reflectance material (LRM),
intermediate terrain (IT), high-reflectance plains (HRP) and low-
reflectance blue plains (LBP). On a local scale, we noted the pres-
ence of relatively red material, bright ejecta deposits and localised
low-reflectance material. To record the host material we distin-
guished between the walls, peak structure, ejecta blanket and
smooth or rough floor of craters, and smooth and rough non-crater
surfaces.
We grouped hollows together on the basis of occurrence within
a particular host crater or location within 50 km of each other
where they lie outside craters. We calculated the areal extent of
hollows within each group by mapping them individually and
obtaining the spherical area (area of a polygon without the distor-
tion caused by map projection) using the Graphics and Shapes tool
for ArcGIS (Jenness, 2011).
As several of the proposed formation mechanisms for hollows
are controlled by insolation, we investigated the possibility of pref-
erential hollow formation on sun-facing slopes. We recorded the
aspect of the slope where hollows within a group occurred on
slopes of a particular orientation where that observed orientation
could not be explained by compositional differences or differences
in viewing conditions for nearby slopes at other orientations. This
aspect was taken to be the bearing of a line normal to the horizon-
tal alignment of hollows along the surface of the slope. This was
drawn by eye and rounded to the nearest 5°, as no digital terrain
model of adequate resolution was available.
In order to calculate the depths of hollows, we measured shad-
ows at the margins of hollows in cases where high resolution
(<110 m/px) images were available. A precision of half a pixel
was used to estimate the error. Where multiple images were avail-
able of the same hollow, we used the image with the highest res-
olution and lowest emission angle (angle off nadir) to minimise
error. We also avoided measuring shadows falling on steep slopes.
Because pits are generally larger-scale features than hollows
and since we wanted to investigate whether hollows occur in asso-
ciation with pits, we noted whether the resolution of the available
images of pits would allow identification of hollows, if present.
2.2.2. Impact craters
Where hollows occur in association with an impact crater we
noted the crater diameter as a proxy for depth of excavation. This
was obtained from the Herrick et al. (2011) global crater database
or measured on a sinusoidal projection of the crater if it was not
within that database. We also noted the crater’s degree of degrada-
tion as a proxy for age using the scheme of Barnouin et al. (2012).
Degradation classes range from 1 to 5 (oldest to youngest) and are
defined on the basis of characteristics such as the preservation of
the ejecta blanket, modification of terraces and amount of super-
posed impact craters. Any crater ages mentioned in this work are
based on this scale.
3. Results
We found 445 groups of hollows, covering 57,400 km
2
, which
amounts to 0.08% of the surface area imaged at better than
180 m/px (locations indicated in Supplementary material). These
ranged in areal extent from 0.07 km
2
to 6771 km
2
, with a mean
of 129 km
2
(standard deviation = 475). 140 of these groups were
at locations previously catalogued by Blewett et al. (2013). We also
identified 173 groups of pits and 24 areas of spectrally red pitted
ground (Fig. 2).
Table 1
Characteristics distinguishing pits, hollows and spectrally red pitted ground.
Characteristic Hollows Pits Spectrally red pitted
ground
Wall slope Steep Shallow Lacking or steep
Floor slope Flat, though lumps of material may
occur
Sloping Roughly horizontal but
uneven
Surrounding deposits (when
present)
High-reflectance, relatively blue High-reflectance, relatively red High-reflectance,
relatively red
Depth Tens of meters Can be 1 km or more deep (Rothery et al., 2014; Gillis-Davis
et al., 2009)
Tens of meters
R.J. Thomas et al. / Icarus 229 (2014) 221–235 223
3.1. Global variations in the areal extent of hollows
Hollows occur globally, but are rare in the high reflectance
plains at high northern latitudes and within basins such as Caloris
(160°E, 32°N) and Rembrandt (88°E, 33°N). There is good image
coverage in these regions so this absence is not a product of obser-
vational bias. Though much of the interior of Caloris lacks hollows,
hollows do occur within younger impact craters in its fill. Hollows
also occur at its rim, often in association with pits of probable
pyroclastic origin, and in two sublinear regions outside its north-
west rim.
Though hollows are not observed at high southern latitudes,
this could be largely attributable to observational bias: MESSEN-
GER’s highly elliptical orbit, with an initial periapsis altitude of
200 km at 60°N and an apoapsis altitude of 15,200 km (Hawkins
et al., 2007), means this area is imaged at much lower resolutions
than areas further north.
It has been suggested that hollow formation is controlled by
insolation and that hollows rarely occur on smooth plains sub-
strates (Blewett et al., 2013). In order to assess these hypotheses
with our global dataset, we investigated whether there is a corre-
lation between longitudinal and latitudinal variations in the areal
extent of hollows and in the intensity of insolation and the areal
extent of non-plains substrates.
3.1.1. Longitudinal variation
The elliptical orbit of Mercury leads to a variation in mean inso-
lation along the equator: two ‘hot poles’ (0°E and 180°E) are under
Fig. 2. Global occurrence of hollows, pits and spectrally red pitted ground. Yellow: hollows; black: pits; red: spectrally red pitted ground; grey: area not imaged at <180 m/px.
(Base mosaic: MESSENGER global colour v3.) (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
-160 -140 -120 -100 -80 -60 -40 -20 0 20 40 60 80 100 120 140 160 180
Longitude (°E)
-160 -140 -120 -100 -80 -60 -40 -20 0 20 40 60 80 100 120 140 160 180
Longitude (°E)
A
h
A
np
A
tot
A
c
A
h
A
c
0
0.002
0.004
0.006
0.008
a
b
0
0.002
0.004
0.006
0.008
Fig. 3. Variation in the areal extent of hollows (A
h
) by longitude bin in the region 30°Sto30°N, normalised to (a) area imaged at <180 m/px (A
c
), and (b) A
c
the fraction of the
surface that is not smooth plains (A
np
/A
tot
).
224 R.J. Thomas et al. / Icarus 229 (2014) 221–235
the Sun at perihelion and experience mean temperatures esti-
mated to be 100 K higher than those of two ‘cold poles’ (90°E
and 90°E), under the Sun at aphelion (Melosh and McKinnon,
1988). To investigate the relationship between hollow occurrence
and this longitudinal variation, we calculated the total areal extent
of hollows in 20°bins in a 30°Sto30°N equatorial strip, normalis-
ing the hollowed area to the area that is imaged at <180 m/px
(Fig. 3a). This region has the virtue of being imaged at high resolu-
tion and low to moderate solar incidence angles, which are favour-
able observation conditions for hollows. We found that the areal
extent of hollows is low near the ‘cold poles’, as expected if the
intensity of insolation controls their occurrence. The fraction of
the surface hollowed between the 40°Eto20°E bin and the
60°Eto80°E bin peaks at the ‘hot pole’ at 0°E, but a similar pattern
of increase is not seen around the other ‘hot pole’ at 180°E.
However, plains associated with the Caloris basin occupy a large
part of the equatorial strip from 150°E to 180°E. To test whether
the presence of these plains modifies the pattern of hollow occur-
rence, we normalised the areal extent of hollows to the fraction of
non-plains in each bin, in effect removing the influence of this
parameter from the data (Fig. 3b). We then saw a stronger correla-
tion between the extent of hollows and the intensity of mean inso-
lation, indicating that intensity of insolation controls hollow
occurrence but is not a sufficient condition for their formation on
all substrates.
The large areal extent of hollows in the 60 to 40°E region is a
clear anomaly, neither accounted for by the variation in mean inso-
lation nor by the presence or absence of smooth plains.
3.1.2. Latitudinal variation
The fraction of the surface area imaged at <180 m/px that is hol-
lowed varies widely at different latitudes (Fig. 4a). This is in major
part attributable to observational bias: this fraction is highest at
low and mid-northern latitudes where MESSENGER is closest to
the planet and lowest at high southern latitudes where it is fur-
thest away. At very high latitudes, high solar incidence angles
(Chabot et al., 2013) also preclude identification of hollows where
they occur in craters because large parts of crater interiors are in
shadow.
The lack of hollows at high northern latitudes is likely to be fur-
ther controlled by the presence of a smooth plains substrate here.
Normalising to the fraction of non-plains does not entirely remove
the disparity between hollow occurrence at low and high northern
latitudes (Fig. 4b), but as non-plains areas near the pole are imaged
only with high solar incidence angles, it is possible that the
remaining disparity is due to observational bias.
The areal extent of hollows at low southern latitudes (30°Nto
0°N) is significantly lower than at low northern latitudes (0–30°N),
and this contrast is not removed by normalising to the fraction of
non-plains (Fig. 4b). This suggests a further factor discouraging
hollowing at low southern latitudes or promoting it at low north-
ern latitudes.
3.2. Preferred slope aspect
It has been suggested that hollows preferentially form on sun-
facing slopes and that this is evidence that their formation is linked
to solar heating (Blewett et al., 2013). We found some evidence in
support of this phenomenon. A preferred aspect was found in only
8% of the groups of hollows observed, but in these cases there was
a good correlation to the sun-facing slope (Fig. 5). The small per-
centage of cases in which a preferred aspect was observed is partly
attributable to the fact that only the sun-facing slope is illuminated
in many of the available images of hollows at mid- to high lati-
tudes, and in such cases preferential hollow formation on that
slope was not recorded because lighting conditions were not good
enough to rule out hollow formation on the opposing slope. How-
ever, most hollows are found on flat surfaces or on slopes of oppos-
ing aspects within a group so we do not find that preferential
a
b
0 0.0005 0.001 0.0015 0.002 0.0025 0.003
-60
-30
0
30
60
90
Latitude (°N)
Anp
AhAcAtot
()
/x
()
0 0.0005 0.001 0.0015 0.002 0.0025 0.003
-60
-30
0
30
60
90
Latitude (°N)
AhAc
Fig. 4. Latitudinal variations in areal extent of hollows (A
h
), normalised to (a) the
area imaged at <180 m/px (A
c
) and (b) A
c
the fraction of the surface that is non-
plains (A
np
/A
tot
) within each latitude bin. Hollow extent varies broadly with image
quality, though observational bias does not explain the small extent of hollowing in
the 30°Sto0°N bin.
N
S
EW
(183.9)
10
19
Fig. 5. Aspects of slopes on which hollows preferentially form in the northern
hemisphere (N= 31), showing a correlation with the sun-facing slope. Purple line
indicates the mean, red circumferential line shows one standard deviation. Radial
axis: percentage of the population of hollow groups with a preferred aspect within
the northern hemisphere. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)
R.J. Thomas et al. / Icarus 229 (2014) 221–235 225
formation on sun-facing slopes is a common characteristic of
hollows.
3.3. Hollow depth
We calculated a mean depth for hollows of 47 m (standard devi-
ation 21) on the basis of 108 shadow measurements within 27 hol-
low clusters. Depths ranged from 5 ± 0.75 m to 98 ± 19.5 m. These
results are consistent with but more extensive and representative
than previous studies showing a hollow depth of 30 m in Kertesz
crater (Vaughan et al., 2012) and 44 m in Raditladi basin (Blewett
et al., 2011).
Because of the limited resolution of available images, there is a
large error in depth values and it is not possible to identify signif-
icant depth variations between different substrates (e.g. peak ring
vs. crater floor) or craters of different ages.
3.4. Geological settings
The majority of hollows occur in clusters within impact craters
and upon their proximal ejecta, as discussed by Blewett et al.
(2013). However, our detailed survey also revealed two large areas
of more distributed hollow formation lacking a close relationship
with specific craters, and some association with pyroclastic pits
in non-crater settings.
3.4.1. Association with impact craters
3.4.1.1. Observations. Hollows occur on a variety of crater surfaces.
In simple craters, they commonly occur in a band on the inside rim
of the crater (Fig. 6a). In complex craters, they occur on the walls,
central structures and smooth floor fill (Fig. 6b) and occasionally
on the ejecta blanket. Hollows are commonly clustered, either
loosely with small (<5 km) expanses of non-hollowed surface be-
tween individual hollows (Fig. 1) or more tightly, as in Fig. 6b,
where they form a continuous hollowed area. At the rim of the Cal-
oris basin, many small groups of hollows occur on peaks in the rim
material and associated with probable pyroclastic pits.
Where hollows occur on crater fill, they often cluster around the
central structures or near the walls (Fig. 6c). In old, degraded cra-
ters, they commonly occur on the high inner walls of smaller im-
pacts into the crater or in the hanging walls of crater-crossing
thrust faults (Fig. 6d).
3.4.1.2. Statistical correlation. Hollows occur within impact craters
and their proximal ejecta in 84.5% of cases, and make up 97.5% of
the total global hollowed area. Hollows are therefore strongly asso-
ciated with craters.
If hollow-forming material is brought to the surface at the time
of crater formation, then depending on the duration over which
hollows (once formed) remain visible, a correlation could be ex-
pected between the age of the crater and the areal extent of
Fig. 6. Typical locations of hollows within impact craters: (a) in a curvilinear band at inner rim of a simple crater (58.6°E, 57.4°N, EN0238696485M); (b) hollowing across a
large part of the floor and peak structure of Hopper crater (Mansurian age, 55.8°E, 12.5°N, EN0223616383M); (c) clustered in the area abutting the crater wall of Nampeyo
crater (Mansurian age, 49.9°E, 40.3°N, EN0253678867M); (d) in a young impact into and on a thrust crossing the old, degraded Duccio crater (hollowed areas outlined in
yellow) (Tolstojan age, 52.3°E, 58.2°N, MESSENGER global monochrome mosaic); (e) close-up of hollows in a younger impact crater (EN0223658124M); (f) close-up of
hollows on a thrust (EN0223614937M). North is towards the top of each image. (For interpretation of the references to colour in this figure legend, the reader is referred to the
web version of this article.)
226 R.J. Thomas et al. / Icarus 229 (2014) 221–235
hollowing. We therefore plotted the average extent of hollowing as
a percentage of the crater’s floor area (
p
r
2
where r= crater radius)
in craters of each degradation state (Fig. 7), using degradation state
as a proxy for age (Barnouin et al., 2012). We divided the data into
crater diameter bins to allow for the possibility that a similar deg-
radation state may occur in a shorter period for smaller craters
than larger craters. We found no clear increase in the percentage
of the crater area hollowed in older craters. In fact the average per-
centage hollowed is somewhat lower in older craters than younger
craters (note the logarithmic vertical scale), though it can range up
to 5.5% even in very old craters (degradation class 1, signifying a
Pre-Tolstojan age). Because slope processes and burial by regolith
can potentially obscure the characteristic morphology of hollows
(in particular their steep, sharp margins), hollows seen clearly
now can be assumed either to be still forming or to have ceased
forming in the relatively recent past.
3.4.2. Hollows outside craters
Hollows outside craters make up loose groupings rather than
clusters, and in most cases (excluding those discussed below) the
extent of hollowing within each group is small, averaging
15.2 km
2
(standard deviation = 38.0). This compares to a mean area
of 148.6 km
2
(standard deviation = 514.0) for hollow groups within
and in the proximal ejecta of craters. Hollows outside craters usu-
ally occur on hummocky surfaces or in a linear pattern cross-cut-
ting geological units, suggesting they formed in distal ejecta
(Fig. 8). Some occur around pits with surrounding relatively spec-
trally red deposits, as reported below (Section 3.5).
The detailed examination undertaken by this study has revealed
for the first time two dispersed groupings of hollows covering large
areas (136,000 km
2
and 52,000 km
2
) roughly radial to the Caloris
basin rim and extending to the west and northwest (Fig. 9a). In
the western grouping, hollows with an areal extent totalling
150 km
2
occur on the partially-preserved, heavily-cratered raised
rims of old craters that are floored by smooth plains (Fig. 9b). The
hollows do not appear to be particularly associated with any one
crater but occur wherever this rougher, higher elevation substrate
occurs. This substrate can be classified as low-reflectance material
(LRM), whereas the crater fills are high reflectance plains (HRP).
Hollows in the northwest grouping have an areal extent totalling
498 km
2
. The eastern part of this grouping lies in and around a
wide graben outside the northwest rim of the Caloris basin, which
may have been carved by ejecta during the Caloris impact event
(Fassett et al., 2009). From this area towards the west, hollows oc-
cur on the circum-Caloris low-reflectance blue plains (LBP) at the
margins of a curvilinear unit of high reflectance material that ap-
pears contiguous with a region featuring several broad channels,
possibly lava channels (Byrne et al., 2013) near the margins of
2
12
11
26
33
62
79
56
5
20
0.1
10.0
12345
Degradation class (oldest to youngest)
% of crater floor area hollowed (log km2)
Crater diameter bin
16 − 64 km
64 − 256 km
Fig. 7. Spread in area of hollows as a percentage of the calculated crater floor area against the degradation state of the host crater. Bottom and top of boxes indicate the first
and third quartiles, band inside each box indicates the median, numbers indicate number of observations and are vertically centred to the mean; lines above and below boxes
extend to the most extreme data point where it is no more than 1.5 times the interquartile range; dots indicate outliers.
Fig. 8. Hollows of small areal extent (indicated by white arrows) occur approximately radial to a Mansurian-aged unnamed complex crater with hollows on its peak ring,
floor and terraces (all hollows outlined in yellow). This association suggests they formed in distal ejecta from the crater. (Figure centred at 65.0°E, 44.8°N, image: excerpt
from global monochrome mosaic.) (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
R.J. Thomas et al. / Icarus 229 (2014) 221–235 227
the northern smooth plains (Fig. 9c). In the far northwest, hollows
preferentially form in and around localised very low reflectance
deposits (dubbed ‘dark spots’ by Xiao et al. (2013)), which occur
on a non-plains substrate adjacent to the smooth-floored channels
and to pits that are possible sources of fluid lava that carved the
channels (Byrne et al., 2013). The ‘dark spots’ occur on a rough, cra-
tered substrate that appears superficially smoothed. This smooth-
ing is particularly pronounced in lower elevation areas (Fig. 9d).
3.5. Association with pyroclastic features
Non-impact-related pits have hollows within 50 km of them in
74% of the cases where the resolution of the available images al-
lows hollow identification, showing a strong association. 71% of
these pits have surrounding spectrally-red deposits, proposed to
be pyroclastic in origin (e.g. Kerber et al., 2011). The association be-
tween hollows and pits within craters is stronger than the associ-
ation of hollows and pits outside craters. 77% of pits within craters
have nearby hollows, while only 47% of pits outside craters do. All
of the areas of spectrally red pitted ground that were imaged at a
high-enough resolution to identify hollows did have hollows with-
in 50 km of them.
The reverse relationship is not as strong: only 22% of hollow
groups lie within 50 km of a pit or pitted spectrally red area, 93%
of which have associated bright red deposits indicating pyroclastic
activity. The areal extent of these groups of hollows is higher than
average: they total 52.5% of the total hollowed area and have a
mean extent per group of 307 km
2
(standard deviation = 855)
(compared to a mean of 129 km
2
(standard deviation = 475) for
the total population). Formation of hollows in the vicinity of pits
is more frequent in longitude bins crossing the ‘cold poles’ than
near the ‘hot poles’ (Fig. 10a), but variations in percentage of the
hollowed area within those bins do not follow this pattern
(Fig. 10b).
3.6. Association with regional substrates
Previous studies have indicated an association between hollow
formation and low-reflectance material, both the regional LRM unit
(Blewett et al., 2013) and localised, possibly lower reflectance (Xiao
et al., 2013) deposits. The results of our survey support this: 96% of
the total hollowed area occurs associated with either regional or
localised low-reflectance material. The hollows incise the low-
reflectance material and are floored and/or haloed by bright rela-
tively blue material.
We found that hollows are considerably rarer on plains sub-
strates: only 7% of the hollowed area occurs on high-reflectance
plains (HRP) and 8% on low-reflectance blue plains (LBP). Where
hollows occur on regional HRP, the low-reflectance material is al-
most always present at the surface locally (37 out of 38 cases). Lo-
cal low-reflectance material is also present in 25 out of 33 cases
where hollows occur on regional LBP.
The preference for hollow formation in low-reflectance material
rather than high-reflectance plains is particularly clear where an
impact crater straddles a contact between these two regional sub-
strates: Fig. 11 shows an 80 km diameter crater that intersects the
Fig. 9. (a) Two dispersed groupings of hollows (hollows outlined in yellow) occur to the northwest of the Caloris basin (dashed white line = basin rim). The northwest
grouping extends to a region featuring several named valles, thought to be lava channels (Byrne et al., 2013), several pits (outlined in blue) and a group of areas of spectrally
red pitted ground (outlined in red). Hollows associated with specific impact craters have been omitted from this diagram. Extent of insets indicated by white boxes (excerpt
from the global colour mosaic); (b) hollow formation in the southern grouping occurs on LRM forming degraded crater rims (image ID EW0264188888G); (c) hollows in the
mid-part of the northern grouping form at the margins and on the margin-proximal floor of a smooth, curvilinear unit of HRP (excerpt from the global monochrome mosaic);
(d) hollows (outlined in yellow) occur in ‘dark spots’ on regions of the non-plains surface that are adjacent to smooth channel floors and appear superficially smoothed
(mosaic of image ID EW0231135561G, EW0231135600G and EW0231135586G). (For interpretation of the references to colour in this figure legend, the reader is referred to
the web version of this article.)
228 R.J. Thomas et al. / Icarus 229 (2014) 221–235
LRM rim of Rembrandt basin and its HRP fill. The north and south
halves of the younger crater reflect the spectral properties of these
substrates, and hollow formation is only seen in the low-reflec-
tance substrate. This association supports the theory that the hol-
low-forming volatiles are derived from a constituent of low-
reflectance material.
4. Discussion
The results presented in this study have implications for the
mechanisms that form hollows and bring hollow-forming volatiles
to the surface, and provide clues to the origin of these volatiles.
4.1. Hollow formation mechanisms
4.1.1. Exogenic processes
Previous studies have suggested that hollows may form by sub-
limation or by space weathering processes (Blewett et al., 2011),
such as photon-stimulated desorption or sputtering by the solar
wind.
Sublimation could form hollows if a moderately-volatile sub-
stance that is unstable at the temperatures and pressures at the
surface of Mercury becomes exposed. It would then transition from
solid to gas and be lost to space via the exosphere, leaving a
depression. One reason why this is thought to be a viable mecha-
nism is the morphological similarity between hollows and ‘Swiss
cheese’ terrain in the south polar region of Mars. ‘Swiss cheese’ ter-
rain is believed to form by scarp retreat as CO
2
ice sublimes (Byrne
and Ingersoll, 2003a). The depth of sublimation may be limited by
the thickness of the subliming layer: the CO
2
ice overlies a water
ice layer that is more stable. This may also apply for hollow forma-
0
10
20
30
40
50
60
70
-150 -120 -90 -60 -30 0 30 60 90 120 150 180
% groups of hollows associated
with pits or red pitted ground
Longitude ( E)
°
0
20
40
60
80
100
-150 -120 -90 -60 -30 0 30 60 90 120 150 180
% hollowed area associated
with pits or red pitted ground
a
b
Longitude ( E)
°
Fig. 10. Latitudinal variation in the association of hollows with pits and spectrally red pitted ground. (a) Hollows more commonly occur near these features at ‘cold pole’-
crossing latitudes (90°E and 90°E) than at ‘hot pole’-crossing latitudes (0°E and 180°E) but (b) the percentage of the surface area that is hollowed shows no regular variation.
Fig. 11. Mansurian age crater straddling the southern rim of Rembrandt basin.
Hollows (outlined in yellow) occur in the southern part of the crater, which has a
low-reflectance substrate, and not in the high-reflectance northern part. (a)
Location of the crater at the southern margin of Rembrandt basin (excerpt from
global colour composite, centred at 88.1°E, 37.3°N); (b) the relation of hollows, a
small pit and an pitted red area to the two substrates (colour composite based on
EW0221673142G); (c) sketch map of the area in (b), black outlines: crater terraces
and peak structures; dark grey: hollowed area; light grey: pit; dotted fill: pitted red
area; hatched: LRM surfaces. (For interpretation of the references to colour in this
figure legend, the reader is referred to the web version of this article.)
R.J. Thomas et al. / Icarus 229 (2014) 221–235 229
tion, though unless the volatile component of hollow-forming
material is 100%, accumulation of a residual lag is likely to limit
hollow depth (Blewett et al., 2013).
‘Swiss cheese’ terrain is not a perfect analogue for hollows. The
floors of the depressions in ‘Swiss cheese’ terrain are smooth and
their outlines are more regular and cuspate. The cuspate outlines
are believed to result from the consistency of the solar incidence
angle through the day in polar regions (Thomas et al., 2000; Byrne
and Ingersoll, 2003b). As most of the hollows we have observed are
closer to the equator, it is perhaps not surprising that hollows have
more irregular margins if sublimation is responsible for their for-
mation. The uneven floors of hollows compared to Swiss cheese
terrain suggest that unlike sublimation of CO
2
ice, the formation
of hollows leaves an appreciable lag fraction, and is perhaps
brought to a halt when the surface lag has reached a critical
thickness.
If hollows form by sublimation, it is probable that their occur-
rence would be correlated with local and regional variations in
insolation. Our results are consistent with this. Though few hollow
groups show preferential formation on slopes of a particular orien-
tation, when they do so there is a strong preference for the sun-fac-
ing slope (Fig. 5). The longitudinal variation in hollowing also
supports a correlation with insolation intensity: when the effect
of substrate is removed from a plot of the variation in the extent
of hollowing at equatorial latitudes (Fig. 3b), it varies broadly with
variations in the intensity of insolation. If surface temperature con-
trols hollowing, one would expect to see a greater reduction in the
extent of hollowing with increased latitude than at different longi-
tudes because the difference in maximum surface temperature be-
tween the equator and poles is larger than between points on the
equator (Peplowski et al., 2013). Our data show such a pattern, but
observational biases and differences in substrate mean this is not a
robust result.
Thus we find that hollow formation appears to be correlated
with insolation intensity, but not strongly. In most cases hollows
form on flat surfaces, or else hollows within a group occur on
slopes with a variety of aspects. This may suggest that the thresh-
old above which insolation causes hollow formation is commonly
met on the surface of Mercury.
A correlation of hollow location with insolation intensity does
not uniquely point to sublimation as the formation mechanism.
High insolation also means a higher photon flux, promoting pho-
ton-stimulated desorption (PSD). In this process, UV photons strike
the planetary surface and excite atoms, which are then desorbed
and can be lost to the exosphere. This process is intensified by high
temperatures, possibly due to enhanced diffusion to the topmost
surface (Yakshinskiy and Madey, 2004), and PSD fluxes contribut-
ing to the exosphere are up to three times higher from equatorial
surfaces at perihelion than at aphelion (Lammer et al., 2003). This
is consistent with the disparity of our results at different equatorial
longitudes. However, PSD is a phenomenon of the extreme surface,
affecting the topmost layer of atoms. Unless the volatile elements
that are susceptible to loss by this mechanism can be very effi-
ciently delivered to the surface through any lag components, or un-
less churning of the regolith by impact gardening exposes fresh
materials very efficiently, it is not probable that it plays a major
part in hollow formation.
Another possibility is that hollow formation is enhanced by,
rather than caused by, high daytime insolation, because this leads
to greater diurnal temperature variation. The varying temperature
in the top few tens of cm of the surface (Vasavada et al., 1999)
could set up a circulation system that could concentrate volatiles
sufficiently to allow hollow formation by one or more of the pro-
cesses suggested here.
In light of the extreme conditions at Mercury, at an average of
0.387 AU from the Sun, the solar wind must be considered as a pos-
sible agent to produce hollows. When solar ions strike the planet’s
surface, they may remove material through momentum transfer in
a process known as physical sputtering. The importance of this
process is potentially testable by looking at latitudinal variations
in hollow formation: under normal solar wind conditions, the max-
imum precipitation flux of solar wind ions onto Mercury’s surface
is expected at high latitudes due to their direction along open mag-
netic field lines (Sarantos et al., 2007), while other areas of the pla-
net’s surface are subject to ion bombardment only during relatively
short-lived conditions of higher dynamic pressure (Siscoe and
Christopher, 1975; Kabin et al., 2000; Slavin et al., 2010). This
may mean the effects of physical sputtering are more pronounced
at high latitudes than low latitudes. One would also expect a stron-
ger effect on areas under the Sun at perihelion than at aphelion, as
the flux through open field lines is modelled to vary by a factor of
four and the area open to the solar wind by a factor of two between
these orbital points (Sarantos et al., 2007). However, given the
observational difficulties that hamper identification of hollows at
polar and high southern latitudes and the presence of a plains sub-
strate at high northern latitudes that appears to preclude hollow
formation on compositional grounds, our current data does not al-
low us to confidently compare the extent of hollow formation at
high vs. low latitudes. We do not rule out solar wind sputtering
playing a part in hollow formation, though the stronger evidence
for a correlation with insolation intensity at lower latitudes sug-
gests that surface temperature plays a strong role and that subli-
mation is probably the dominant mechanism.
4.1.2. Endogenic processes
The strong correlation of pyroclastic pits and areas of spectrally
red pitted ground with hollows (Section 3.5) suggests that endo-
genic heat sources may contribute to the heat necessary to release
the volatiles within hollow-forming substrates. If magmatic activ-
ity was contemporaneous with hollow formation at these sites, the
heat of subsurface magma may have mobilised the volatile compo-
nent of the host rock. This component may either ascend to the
surface, condense and later be removed by sublimation, possibly
aided by heat from below, or ascend as a gas and cause hollows
to form by collapse of surface material due to volume loss in the
underlying substrate.
The first hypothesis best fits the evidence, as hollows around
pits have the same morphology as those on crater surfaces. Though
it is possible that hollows and pits are found together only because
they both occur in the same substrates for independent reasons, it
is revealing that the areal extent of hollowing is on average higher
around pyroclastic pits than elsewhere. Additionally, at locations
where there is less insolation and so initiation of hollows would
be relatively more strongly affected by any endogenic component
to volatile mobilisation, a higher proportion of hollow groups occur
near pits. This suggests that proximity to pyroclastic pits leads to
more hollow formation. Also, the percentage of the total hollowed
area within each longitudinal band that is near pits is not strongly
correlated to variations in the intensity of insolation (Fig. 10b),
suggesting that conditions in the vicinity of particular pits exert
a stronger control on the extent of hollowing than do variations
in mean insolation.
The second hypothesis, that hollows form by surface collapse
when hollow-forming volatiles are lost, may be a more suitable
explanation for the shallow areas of pitted ground with red depos-
its. These areas look similar to hollows but are deeper, more un-
even, have a less morphologically-crisp appearance and in all
cases have hollows in their vicinity. They are also all found on
smooth substrates, in most cases crater floors. This juxtaposition
is seen clearly at Rachmaninoff basin (Fig. 12). Here, hollows form
in the low-reflectance material of the crater’s peak ring and walls,
and on the younger volcanic crater fill (Marchi et al., 2011) around
230 R.J. Thomas et al. / Icarus 229 (2014) 221–235
areas of spectrally red pitted ground. The presence of low-reflec-
tance material on the crater’s peak ring and walls suggests this
material also forms the substrate to the volcanic infill. The lava
may have heated the substrate and released its volatile component.
Before the lava fully solidified, disruption of its surface by escaping
volatiles and collapse due to volume loss in the substrate may have
given it a pitted morphology. This can be seen as broadly analogous
to the process by which pitted terrain is suggested to form by the
release of volatiles through impact melt on Mars (Boyce et al.,
2012) and Vesta (Denevi et al., 2012). The spectrally-red deposits
may indicate that magmatic volatiles and entrained juvenile mate-
rial also escaped to the surface along the same pathways as the
hollow-forming volatiles.
Such a process may also explain the morphology of hollowed
areas on crater floors over buried peaks rings, such as Sousa crater
(Fig. 13). The morphology of these areas is similar to that of those
areas of spectrally red pitted ground which lack margins, but they
have no associated red deposits. Here the crater fill (either impact
melt or a later volcanic infill) may have volatilised a component of
the buried low-reflectance peak ring, and released this material to
the surface at the point where the fill is thinnest. This led to col-
lapse of the surface and some formation of crisper hollows where
escaped material condensed on the surface before losing its vola-
tile component via sublimation. This hypothesis is an extension
of that of Blewett et al. (2013) that hollows may form through con-
centration of hollow-forming volatiles by contact heating.
The global distribution of hollows does not simply mirror vari-
ations in insolation, and many hollows occur at a distance from pits
and potential pyroclastic activity and without contact with crater
fills. A further factor plays a stronger controlling role on the forma-
tion of hollows: substrate. This determines the quantity of hollow-
forming volatiles at and near the surface, and the ability of volatiles
to ascend to the surface. We explore this aspect below.
4.2. Means of transfer of hollow-forming volatiles to the surface
4.2.1. Exhumation and exposure by impacts
The strong correlation of hollows with craters suggests a genet-
ic link. The vast majority of hollows lie within impact craters and
their proximal ejecta. Hollows do not form on volcanic plains ex-
cept where these have been breached by later impacts, and most
of the small hollow clusters that occur outside impact craters ap-
Fig. 12. (a) Hollows (outlined in yellow) form in low-reflectance material in the peak ring and walls of Rachmaninoff basin (3.6 Ga old (Marchi et al., 2011), 57.4°E, 27.6°N)
and around bright, relatively-red areas (outlined in red) south of a breach in its peak ring (excerpt from global colour mosaic); (b) the area south of the breach in the peak ring,
where the bright spectrally-red areas are seen to be areas of pitted ground with hollows near their margins (composite of EW0254942264G, EW0254942268F and
EW0254942272I); (c) closeup of an area of pitted ground with steep margins at some points (black arrows) and hollow formation near its margins (white arrows)
(EN0219350311M). (For interpretation of the references to colour in this figure legend, the reader is referred to the web version of this article.)
Fig. 13. Hollows (outlined in yellow) on the LRM peak ring and the crater fill where this overlies the peak ring in Sousa crater (Mansurian age, 0.5°E, 46.8°N). (a) Locations of
hollows within the crater (outlined in yellow) (excerpt from global colour mosaic) and (b) close-up showing the majority of the hollowed region in the crater fill is pitted
ground, with some small crisp hollows (white arrows) (mosaic of EN0251054159M and EN0251054171M). (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
R.J. Thomas et al. / Icarus 229 (2014) 221–235 231
pear to form in impact ejecta. For these reasons, it has previously
been suggested that hollows form in material exposed and ex-
humed by large impacts (Blewett et al., 2011). Larger impacts sam-
ple the crust to great depths compared to smaller impacts,
exposing the strata underlying the impacted surface in their walls
and ejecta, and uplifting material from depth in their central struc-
tures. The evidence supports this as a mechanism for exposure of
hollow-forming material. Hollows occur in low-reflectance mate-
rial within impact craters, particularly on peak rings, which are
the part of a crater that is exhumed from the greatest depth
(Fig. 14). The surface distribution of this low-reflectance material
is consistent with the substrate into which the crater incised
(Fig. 11) as expected if it is exhumed material.
The lack of correlation between increasing crater age and extent
of hollows (Fig. 7) could be seen as weakening this hypothesis: if
hollowing begins at the time of crater formation, older craters
should have a larger extent of hollows. However, firstly, the small
scale of hollows means they may become obscured by later overly-
ing ejecta over time and the areal extent of hollows visible in older
craters may not be indicative of the true cumulative amount of hol-
low formation there. Secondly, if the quantity of material that can
form hollows is limited, hollow formation would eventually cease.
This explains the drop-off in observed hollowed extent in craters of
Calorian age (degradation state 3) and older (Fig. 7).
Estimates of potential burial rates at Mercury’s surface vary:
regolith formation has been estimated at 5–10 m in the last 3–
4Ga(Langevin, 1997), while burial of polar ice deposits has been
modelled to occur at a rate of 0.43 cm/Myr (Crider and Killen,
2005). The former rate would be sufficient to obscure the morphol-
ogy of hollows and the latter to completely bury them in the esti-
mated 3.9 Ga (Neukum et al., 2001)) since the Calorian period.
It has alternatively been suggested (Vaughan et al., 2012) that
hollows form in material differentiated out of impact melt during
crater formation. We find this to be an unconvincing explanation:
hollows are found in small ejecta deposits distal from their host
craters (Fig. 8) and on a variety of steep surfaces. While differenti-
ation is viable in pooled melt, it is difficult to envision in settings
such as these.
4.2.2. Post-impact exposure
The presence of hollows in even very old craters (Fig. 7) indi-
cates that in these cases a process is operating that replenishes hol-
low-forming material at the surface of craters long after crater
formation. Our observation that hollows in older craters are found
in the walls of younger craters and on thrust faults (Fig. 6d) sug-
gests that these are the agents of this late exposure. If hollow-
forming material was exposed to surface conditions during the for-
mation of the original crater, hollow formation may have ceased
prior to its depletion in the near-surface due to deposition of a
lag or burial by ejecta. Small new craters, crater-crossing thrust
faults and fractures at fault-bend folds may have later exposed it
to surface conditions, at which time fresh hollows formed.
Such processes may operate on a regional scale to produce the
western broad area of distributed hollowing outside the Caloris ba-
sin (Fig. 9a). The low-reflectance deposits here are close enough to
Caloris to have been deposited as ejecta from that impact, and may
have originally been as volatile-rich as the extensively-hollowed,
very dark LRM deposits that are exhumed by younger impacts into
the Caloris fill. The heavily-cratered appearance of the hollowed
surfaces suggests they are old and that any initial hollow formation
in them would have long ceased, but smaller impacts and possibly
mass wasting may continue to expose new volatile-bearing mate-
rial to surface conditions and initiate new hollows.
The association of hollows with pyroclastic pits within craters
could indicate that the structures associated with this volcanism
aided the release of hollow-forming volatiles from depth. Pyroclas-
tic pits occur primarily in parts of impact craters that are underlain
by planes of weakness such as wall terraces and central structures.
This suggests that crater-related faults act as conduits for the re-
lease of volatiles or volatile-bearing magma towards the surface,
possibly aided by on-going fault movement in response to global
contraction (Klimczak et al., 2013). The same may be true for hol-
low-forming volatiles. The evidence does not, however, allow a
definite identification of this phenomenon. For example, Fig. 15
shows hollows around a pyroclastic pit in the north-west rim of
a younger crater that intersects the wall of an older crater, and also
hollows in small areas to its south. This could be explained by the
migration of hollow-forming volatiles up the same crater-wall
faults as were exploited by the probable pyroclastic volcanism.
However, exhumation is a viable alternative explanation: hollow-
forming volatiles may have been present in the deeply-excavated
wall material of the older crater and then been exposed by the
younger crater. The small cluster of distal hollows may be located
in the ejecta from this impact. In this scenario, the association of
hollows with pyroclastic volcanism may be partly due to the spa-
tial coincidence of deep fractures that are conducive to magma as-
cent with volatile-bearing wall rocks, and possibly, if hollow
formation and volcanism were contemporaneous, partly due to in-
creased heat flow in this area enhancing upward diffusion of vola-
tiles or hollow formation by sublimation (Section 4.1).
Fig. 14. Localisation of hollows (outlined in yellow) and low-reflectance deposits in the peak ring of Renoir basin (Tolstojan age, 51.8°E, 18.3°N). Dashed white line shows
the outer rim of the basin, where no low-reflectance material or hollows are observed. Left: excerpt from global monochrome mosaic; right: mosaic of colour composites
based on EW0253851174G, EW0253851412G and EW0241374406G. (For interpretation of the references to colour in this figure legend, the reader is referred to the web
version of this article.)
232 R.J. Thomas et al. / Icarus 229 (2014) 221–235
4.2.3. Exposure outside craters
A non-impact-related process is necessary to explain the expo-
sure of hollow-forming material to produce the broad area of dis-
tributed hollowing along a curvilinear unit of HRP to the northwest
of Caloris. These hollows form in LBP and on LRM outcrops stand-
ing above the plains. We propose that the HRP unit is a lava flow,
and that contact heating by this material caused concentration of
volatiles from within the LBP and LRM substrates at the surface
(in a process similar to that suggested for other incidences of hol-
lowing by Blewett et al. (2013)), after which hollows formed by
sublimation.
The formation of hollows in small ‘dark spots’ at the far north-
west of this grouping (Fig. 9d) is more enigmatic. These ‘dark
spots’ occur on a non-plains substrate in which lower elevation
areas appear anomalously smooth, adjacent to broad, smooth-
floored channels and to the pits that have been suggested as
the source of voluminous lavas that carved those channels (Byrne
et al., 2012; Hurwitz et al., 2013). It is possible that the surface
here appears somewhat smoothed because it has been draped
by a thin layer of lava. This may have covered a volatile-bearing
substrate to a shallow depth. Subsequently either a high regional
heatflow connected to magmatic activity volatilised the underly-
ing substrate or burial was sufficiently shallow (<1 m (Vasavada
et al., 1999)) for penetration of solar heating do so. Pressure built
up and finally volatiles were released through fractures in the
overlying material. This is consistent with the hypothesis that
‘dark spots’ form during intense outgassing during hollow forma-
tion (Xiao et al., 2013). This process may have directly produced
hollows in the manner of fumerolic vents, or through deposition
of the volatile-rich dark material on the surface followed by its
sublimation.
4.3. Nature of the hollow-forming material
Hollows are commonly found in low-reflectance material. Con-
versely, they are found in high-reflectance smooth plains only
where low-reflectance material is locally present such as where
younger impact craters exhume it from beneath the volcanic fill
of the Caloris basin.
This strongly suggests that the volatile material responsible for
hollow formation is not present in high-reflectance flood lavas, but
is a component of low-reflectance material. The nature of low-
reflectance deposits on Mercury is not yet established, although
it has been suggested that the regional LRM unit could be primary
crust (Rothery et al., 2010) or a cumulate darkened by Fe- or Ti-
bearing or other opaque oxides (Denevi et al., 2009; Riner et al.,
2009). Space weathering complicates the determination of the
composition of LRM on the basis of reflectance: the creation of
nanophase iron during space weathering leads to more darkening
of rock types initially richer in iron, so the albedo of any rock is
the product of its composition and mineralogy, its exposure time
and its susceptibility to space weathering (Riner and Lucey, 2012).
The presence of low-reflectance material in some craters of a
particular diameter (and thus excavation depth) and absence in
others shows that low-reflectance material is not present globally
at a specific depth, so variations in the igneous and/or tectonic his-
tory of different parts of Mercury’s crust appear to play a role in its
occurrence. For example, the formation of the Caloris impact basin
may have exhumed large amounts of this material, providing the
substrate for the broad regions of dispersed hollow formation to
its northwest.
The relatively lower reflectance of some localised deposits
around hollows (Xiao et al., 2013) may suggest that before LRM
has been hollowed it contains an additional darkening agent, and
that this is the hollow-forming volatile. This may explain the pres-
ence of bright material in haloes around hollows and on their
floors, which could be a brighter residue formed by the removal
of a spectrally dark component (Blewett et al., 2013). Alternatively,
the high reflectivity of these deposits may be due to an unusual
texture or small grain size (Blewett et al., 2013). The diffused mar-
gins of the haloes suggest two possible emplacement mechanisms
for the bright material: (a) ballistic ejection of a bright component
as a result of high-energy escape of the hollow-forming volatiles,
or (b) diffusive alteration of hollow wall rock as a result of chemical
reactions during hollow formation. In both cases, either the com-
positional or physical characteristics of these bright deposits could
potentially be the cause of their high reflectance.
The composition of the darker, volatile component is as yet un-
known, but will perhaps be resolved when BepiColombo with its
higher-resolution visible-NIR, thermal infrared and X-ray spec-
trometers (Rothery et al., 2010) arrives at Mercury in the coming
decade.
Fig. 15. (a) Hollows (outlined in yellow) occur at the north-west edge of a Calorian age crater at 3.6°E, 25.6°N, around a pyroclastic pit (outlined in orange) and scattered
towards the south. This may be due to explosive escape of hollow-forming volatiles up the same conduits as used by the pyroclastic volcanism, or, if the two phenomena were
contemporaneous, hollow formation in crater deposits and ejecta intensified in the region of volcanism by endogenic heat flow (MESSENGER global monochrome mosaic); (b)
colour composite of the superposed crater in the north (based on EW0225312562G); (c) sketch map of (b), hatched area: low-reflectance material; dotted area: bright ejecta;
pink area: bright ‘red’ deposits; orange area: pit; yellow area: hollowed area; black outline: crater walls. (For interpretation of the references to colour in this figure legend,
the reader is referred to the web version of this article.)
R.J. Thomas et al. / Icarus 229 (2014) 221–235 233
5. Conclusions
1. In a global survey of the surface of Mercury, we found that the
shallow rimless depressions known as hollows cover
57,400 km
2
, which is 0.08% of the total surface imaged at bet-
ter than 180 m/px.
2. A weak overall correlation was found between hollow occur-
rence and insolation, as well as a possible correlation with sub-
surface heat sources. Both suggest a thermal control on hollow
formation, thus supporting sublimation as the primary hollow-
forming mechanism.
3. In most cases it appears probable that material containing hol-
low-forming volatiles was exposed and exhumed from depth by
large impacts.
4. Some small impact craters and thrust faults within older craters
also have hollows, hence these structures may expose subsur-
face hollow-forming material or facilitate the migration of vol-
atiles to the surface. This suggests that some volatiles remain in
the near-subsurface even after hollow formation has ceased at
the surface.
5. Hollows do not occur in volcanic plains but are found mostly in
low-reflectance material. This suggests that this low-reflectance
material has a volatile component, and that hollows are formed
by loss of that component. The widespread occurrence of hol-
lows suggests that this material is similarly widespread within
the crust of Mercury.
Acknowledgments
Rebecca Thomas acknowledges support via a PhD grant from
the Science and Technology Facilities Council (UK) and David Roth-
ery acknowledges support from the UK Space Agency in prepara-
tion for the BepiColombo mission. Excerpts from global mosaics
used in figures in this paper are credited to NASA/Johns Hopkins
University Applied Physics Laboratory/Carnegie Institution of
Washington, and the NAC and WAC images are credited to NASA/
JPL-Caltech. We thank David Blewett and Laura Kerber for their
constructive and insightful reviews that helped improve the qual-
ity of the manuscript. We also thank Oded Aharonson for his edito-
rial handling.
Appendix A. Supplementary material
Supplementary data associated with this article can be found, in
the online version, at http://dx.doi.org/10.1016/j.icarus.2013.11.
018.
References
Barnouin, O.S., Zuber, M.T., Smith, D.E., Neumann, G.a., Herrick, R.R., Chappelow, J.E.,
Murchie, S.L., Prockter, L.M., 2012. The morphology of craters on Mercury:
Results from MESSENGER flybys. Icarus 219 (1), 414–427. http://dx.doi.org/
10.1016/j.icarus.2012.02.029.
Blewett, D.T. et al., 2009. Multispectral images of Mercury from the first
MESSENGER flyby: Analysis of global and regional color trends. Earth Planet.
Sci. Lett. 285 (3–4), 272–282. http://dx.doi.org/10.1016/j.epsl.2009.02.021.
Blewett, D.T. et al., 2011. Hollows on Mercury: MESSENGER evidence for
geologically recent volatile-related activity. Science 333 (6051), 1856–1859.
http://dx.doi.org/10.1126/science.1211681.
Blewett, D.T. et al., 2013. Mercury’s hollows: Constraints on formation and
composition from analysis of geological setting and spectral reflectance. J.
Geophys. Res. Planets 118 (5), 1013–1032.
Boyce, J.M., Wilson, L., Mouginis-Mark, P.J., Hamilton, C.W., Tornabene, L.L., 2012.
Origin of small pits in martian impact craters. Icarus 221 (1), 262–275. http://
dx.doi.org/10.1016/j.icarus.2012.07.027.
Byrne, S., Ingersoll, AP., 2003a. A sublimation model for martian south polar ice
features. Science 299 (5609), 1051–1053. http://dx.doi.org/10.1126/
science.1080148.
Byrne, S., Ingersoll, A.P., 2003b. Martian climatic events on timescales of centuries:
Evidence from feature morphology in the residual south polar ice cap. Geophys.
Res. Lett. 30 (13), 2–5. http://dx.doi.org/10.1029/2003GL017597.
Byrne, P.K. et al., 2012. An assemblage of lava flow features on Mercury. J. Geophys.
Res. Planets. http://dx.doi.org/10.1002/JGRE.20052.
Byrne, P.K. et al., 2013. An assemblage of lava flow features on Mercury. J. Geophys.
Res. Planets 118 (6), 1303–1322. http://dx.doi.org/10.1002/jgre.20052.
Chabot, N.L. et al., 2013. Craters hosting radar-bright deposits in Mercury’s north
polar region: Areas of persistent shadow determined from MESSENGER images.
J. Geophys. Res. Planets 118 (1), 26–36. http://dx.doi.org/10.1029/
2012JE004172.
Cheng, A.F., Johnson, R.E., Krimigis, S.M., Lanzerotti, L.J., 1987. Magnetosphere,
exosphere, and surface of Mercury. Icarus 71, 430–440. http://dx.doi.org/
10.1016/0019-1035(87)90038-8.
Crider, D., Killen, R.M., 2005. Burial rate of Mercury’s polar volatile deposits.
Geophys. Res. Lett. 32 (12), L12201. http://dx.doi.org/10.1029/2005GL022689.
Denevi, BW. et al., 2009. The evolution of Mercury’s crust: A global perspective from
MESSENGER. Science 324 (5927), 613–618. http://dx.doi.org/10.1126/
science.1172226.
Denevi, B.W. et al., 2012. Pitted terrain on Vesta and implications for the presence of
volatiles. Science 338 (6104), 246–249. http://dx.doi.org/10.1126/
science.1225374.
Denevi, B.W. et al., 2013. The distribution and origin of smooth plains on Mercury. J.
Geophys. Res. Planets 118 (5), 891–907. http://dx.doi.org/10.1002/jgre.20075.
Dzurisin, D., 1977. Mercurian bright patches: Evidence for physio-chemical
alteration of surface material? Geophys. Res. Lett. 4 (10), 383–386.
Evans, L.G. et al., 2012. Major-element abundances on the surface of Mercury:
Results from the MESSENGER Gamma-Ray Spectrometer. J. Geophys. Res. 117,
E00L07. http://dx.doi.org/10.1029/2012JE004178.
Fassett, C.I. et al., 2009. Caloris impact basin: Exterior geomorphology, stratigraphy,
morphometry, radial sculpture, and smooth plains deposits. Earth Planet. Sci.
Lett. 285 (3–4), 297–308. http://dx.doi.org/10.1016/j.epsl.2009.05.022.
Gillis-Davis, J.J. et al., 2009. Pit-floor craters on Mercury: Evidence of near-surface
igneous activity. Earth Planet. Sci. Lett. 285 (3–4), 243–250. http://dx.doi.org/
10.1016/j.epsl.2009.05.023.
Goldsten, J.O. et al., 2007. The MESSENGER Gamma-Ray and Neutron Spectrometer.
Space Sci. Rev. 131 (1–4), 339–391. http://dx.doi.org/10.1007/s11214-007-
9262-7.
Hawkins, S.E. et al., 2007. The Mercury Dual Imaging System on the MESSENGER
Spacecraft. Space Sci. Rev. 131 (1–4), 247–338. http://dx.doi.org/10.1007/
s11214-007-9266-3.
Hawkins, S.E. et al., 2009. In-flight performance of MESSENGER’s Mercury Dual
Imaging System. Proc. SPIE Int. Soc. Opt. Eng. SPIE7441. http://dx.doi.org/
10.1117/12.826370.
Helbert, J., Maturilli, A., D’Amore, M., 2013. Visible and near-infrared reflectance
spectra of thermally processed synthetic sulfides as a potential analog for the
hollow forming materials on Mercury. Earth Planet. Sci. Lett. 369–370, 233–
238. http://dx.doi.org/10.1016/j.epsl.2013.03.045.
Herrick, R.R., Curran, L.L., Baer, A.T., 2011. A Mariner/MESSENGER global catalog of
mercurian craters. Icarus 215 (1), 452–454. http://dx.doi.org/10.1016/
j.icarus.2011.06.021.
Hurwitz, D.M. et al., 2013. Investigating the origin of candidate lava channels on
Mercury with MESSENGER data: Theory and observations. J. Geophys. Res.
Planets 118 (3), 471–486. http://dx.doi.org/10.1029/2012JE004103.
Jenness, J., 2011. Tools for graphics and shapes: extension for ArcGIS. Jenness
Enterp.
Kabin, K., Gombosi, T.I., Dezeeuw, D.L., Powell, K.G., 2000. Interaction of Mercury
with the solar wind. Icarus 143 (2), 397–406. http://dx.doi.org/10.1006/
icar.1999.6252.
Kerber, L., Head, J.W., Solomon, S.C., Murchie, S.L., Blewett, D.T., Wilson, L., 2009.
Explosive volcanic eruptions on Mercury: Eruption conditions, magma volatile
content, and implications for interior volatile abundances. Earth Planet. Sci. Lett.
285 (3–4), 263–271. http://dx.doi.org/10.1016/j.epsl.2009.04.037.
Kerber, L. et al., 2011. The global distribution of pyroclastic deposits on Mercury:
The view from MESSENGER flybys 1–3. Planet. Space Sci. 59 (15), 1895–1909.
http://dx.doi.org/10.1016/j.pss.2011.03.020.
Killen, R.M., Sarantos, M., Reiff, P.H., 2004. Space weather at Mercury. Adv. Space
Res. 33 (11), 1899–1904. http://dx.doi.org/10.1016/j.asr.2003.02.020.
Killen, R. et al., 2007. Processes that promote and deplete the exosphere of Mercury.
Space Sci. Rev. 132 (2–4), 433–509. http://dx.doi.org/10.1007/s11214-007-
9232-0.
Klimczak, C. et al., 2013. The role of thrust faults as conduits for volatiles on
Mercury. Lunar Planet. Sci. 44. Abstract 1390.
Lammer, H., Wurz, P., Patel, M.R., Killen, R., Kolb, C., Massetti, S., Orsini, S., Milillo, A.,
2003. The variability of Mercury’s exosphere by particle and radiation induced
surface release processes. Icarus 166 (2), 238–247. http://dx.doi.org/10.1016/
j.icarus.2003.08.012.
Langevin, Y., 1997. The regolith of Mercury: Present knowledge and implications for
the Mercury Orbiter mission. Planet. Space Sci. 45 (1), 31–37.
Madey, T.E., Yakshinskiy, BV., Ageev, V.N., Johnson, RE., 1998. Desorption of alkali
atoms and ions from oxide surfaces: Relevance to origins of Na and K in
atmospheres of Mercury and the Moon. J. Geophys. Res. 103 (E3), 5873–5887.
Marchi, S., Massironi, M., Cremonese, G., Martellato, E., Giacomini, L., Prockter, L.,
2011. The effects of the target material properties and layering on the crater
chronology: The case of Raditladi and Rachmaninoff basins on Mercury. Planet.
Space Sci. 59 (15), 1968–1980. http://dx.doi.org/10.1016/j.pss.2011.06.007.
234 R.J. Thomas et al. / Icarus 229 (2014) 221–235
Melosh, H.J., McKinnon, W.B., 1988. The tectonics of Mercury. Mercury, 374–400.
Mura, A., Wurz, P., Lichtenegger, H.I.M., Schleicher, H., Lammer, H., Delcourt, D.,
Milillo, A., Orsini, S., Massetti, S., Khodachenko, M.L., 2009. The sodium
exosphere of Mercury: Comparison between observations during Mercury’s
transit and model results. Icarus 200 (1), 1–11. http://dx.doi.org/10.1016/
j.icarus.2008.11.014.
Neukum, G., Oberst, J., Ho, H., Wagner, R., Ivanov, B.A., 2001. Geologic evolution and
cratering history of Mercury. Planet. Space Sci. 49, 1507–1521.
Nittler, LR. et al., 2011. The major-element composition of Mercury’s surface from
MESSENGER X-ray spectrometry. Science 333 (6051), 1847–1850. http://
dx.doi.org/10.1126/science.1211567.
Peplowski, P.N. et al., 2011. Radioactive elements on Mercury’s surface from
MESSENGER: Implications for the planet’s formation and evolution. Science
333, 1850–1852. http://dx.doi.org/10.1126/science.1211576.
Peplowski, P.N. et al., 2013. Enhanced sodium abundance in Mercury’s north polar
region revealed by the MESSENGER Gamma-Ray Spectrometer. Icarus. http://
dx.doi.org/10.1016/j.icarus.2013.09.007.
Potter, A.E., 1995. Chemical sputtering could produce sodium vapor and ice on
Mercury. Geophys. Res. Lett. 22 (23), 3289–3292.
Riner, M.a., Lucey, P.G., 2012. Spectral effects of space weathering on Mercury: The
role of composition and environment. Geophys. Res. Lett. 39, L12201. http://
dx.doi.org/10.1029/2012GL052065.
Riner, M.a., Lucey, P.G., Desch, S.J., McCubbin, F.M., 2009. Nature of opaque
components on Mercury: Insights into a mercurian magma ocean. Geophys.
Res. Lett. 36 (L02201). http://dx.doi.org/10.1029/2008GL036128.
Robinson, M.S. et al., 2008. Reflectance and color variations on Mercury: Regolith
processes and compositional heterogeneity. Science 321 (5885), 66–69. http://
dx.doi.org/10.1126/science.1160080.
Rothery, D. et al., 2010. Mercury’s surface and composition to be studied by
BepiColombo. Planet. Space Sci. 58 (1–2), 21–39. http://dx.doi.org/10.1016/
j.pss.2008.09.001.
Rothery, D.A., Thomas, R.J., Kerber, L., 2014. Prolonged eruptive history of a
compound volcano on Mercury: Volcanic and tectonic implications. Earth
Planet. Sci. Lett. 385, 59–67.
Sarantos, M., Killen, R.M., Kim, D., 2007. Predicting the long-term solar wind ion-
sputtering source at Mercury. Planet. Space Sci. 55 (11), 1584–1595. http://
dx.doi.org/10.1016/j.pss.2006.10.011.
Schlemm, C.E. et al., 2007. The X-ray spectrometer on the MESSENGER Spacecraft.
Space Sci. Rev. 131 (1–4), 393–415. http://dx.doi.org/10.1007/s11214-007-
9248-5.
Siscoe, G., Christopher, L., 1975. Variations in the solar wind stand-off distance at
Mercury. Geophys. Res. Lett. 2 (4), 158–160.
Slavin, J.a. et al., 2010. MESSENGER observations of extreme loading and unloading
of Mercury’s magnetic tail. Science 329 (5992), 665–668. http://dx.doi.org/
10.1126/science.1188067.
Strom, R.G., Trask, N.J., Guest, J.E., 1975. Tectonism and volcanism on Mercury. J.
Geophys. Res. 80 (17), 2478–2507. http://dx.doi.org/10.1029/JB080i017p02478.
Thomas, PC. et al., 2000. North–south geological differences between the residual
polar caps on Mars. Nature 404 (6774), 161–164. http://dx.doi.org/10.1038/
35004528.
Vasavada, A.R., Paige, D.A., Wood, S.E., 1999. Near-surface temperatures on Mercury
and the Moon and the stability of polar ice deposits 1. Icarus 141, 179–193.
Vaughan, W.M. et al., 2012. Hollow-forming layers in impact craters on Mercury:
Massive sulfide or chloride deposits formed by impact melt differentiation?
Lunar Planet. Sci. 43, 1187.
Watters, T.R. et al., 2009. The tectonics of Mercury: The view after MESSENGER’s
first flyby. Earth Planet. Sci. Lett. 285 (3–4), 283–296. http://dx.doi.org/10.1016/
j.epsl.2009.01.025.
Weider, S.Z. et al., 2012. Chemical heterogeneity on Mercury’s surface revealed by
the MESSENGER X-Ray Spectrometer. J. Geophys. Res. 117, 1–15. http://
dx.doi.org/10.1029/2012JE004153.
Xiao, Z. et al., 2013. Dark spots on Mercury: A distinctive low-reflectance material
and its relation to hollows. J. Geophys. Res. Planets 118, 1–14. http://
dx.doi.org/10.1002/jgre.20115.
Yakshinskiy, B.V., Madey, T.E., 2004. Photon-stimulated desorption of Na from a
lunar sample: Temperature-dependent effects. Icarus 168 (1), 53–59. http://
dx.doi.org/10.1016/j.icarus.2003.12.007.
R.J. Thomas et al. / Icarus 229 (2014) 221–235 235
... However, there is some evidence for weak absorption features associated with Mercury's hollows and low-reflectance material. Hollows are rimless, geologically recent depressions tens of kilometers across and tens of meters deep most commonly found on crater floors hypothesized to be formed by surface volatile loss Lucchetti et al., 2018;Thomas et al., 2014). Low-reflectance material is hypothesized to be graphite-rich remnants of Mercury's primary crust (Klima et al., 2018;Vander Kaaden & McCubbin, 2015). ...
... Mercury's surface is enriched in sulfur compared with Earth (Nittler et al., 2011) and so it is conceivable that similar sublimate deposition might have been an important process on Mercury, although the exact speciation of any sublimates would most likely differ. Pitted ground on Mercury is postulated to form when lavas flowed over volatile-rich substrates, causing pits to form in the lava surfaces (Thomas et al., 2014;Wright, Byrne, & Rothery, 2021). It is possible that there are more lava-substrate interactions on Mercury that have a spectral signature with no presently observable morphological expression. ...
... More powerfully, by combining morphostratigraphic units and SUs we been able to subdivide smooth materials in contact with each other on a quantitative basis, allowing us to distinguish between Rachmaninoff's interior impact melt (sp 10 ) and subsequent volcanic plains (cfs 6 ). The spectrally-defined contact between sp 10 and cfs 6 coincides with the northern boundary of Suge Facula, which corroborates the hypothesis of Wright, Byrne, and Rothery (2021) that Suge Facula preserves a pitted ground texture (Jozwiak et al., 2018;Thomas et al., 2014) because lavas emanating from the center of Rachmaninoff terminated here and did not sufficiently bury the volatilebearing substrate. This also supports the hypothesis that spectral variation within the center of Rachmaninoff might be the result of variable lava-substrate interactions, but while a spectral signature is preserved the lavas in the center of the crater are too thick for the pitted ground texture to be preserved (Wright, Byrne, & Rothery, 2021). ...
Article
Full-text available
Plain Language Summary Geological maps of rocks on Earth include information about what landforms the rocks exhibit, what they are made from, their orientation, and more. Many such details require ground observations, but these are generally not available for planetary geological maps, which rely on spacecraft data. Spacecraft images can be used to map planetary surface textures (geomorphology), but they can also be used to measure surfaces' responses to light (reflectance or emission spectra), which contain information about what the surface rocks are made from, their physical properties (e.g., grain size, roughness, porosity), and how long they have been exposed at the surface. We have combined earlier, independent geomorphic and spectral maps of the Rachmaninoff impact basin on Mercury to create a new “geostratigraphic” map that is more like a geological map that could be made of Earth. The new map highlights places that in the original geomorphic map would have been mapped all as a single unit, but are divisible based on spectral variations, attributable to differences in what the rocks are made from. This allows us to reconstruct a more detailed geological history of the region. Our method can be applied to other regions on Mercury and to other planetary surfaces.
... Hollows are irregularly shaped, flat-floored, shallow depressions that occur predominantly within impact crater structures (crater floors, walls and rims, central peaks, and ejecta) and are considered very young in age (Blewett et al. 2011(Blewett et al. , 2013(Blewett et al. , 2018Thomas et al. 2014Thomas et al. , 2016. They are often, but not always, found in association with low-reflectance materials (Thomas et al. 2014(Thomas et al. , 2016, a unit considered rich in graphite . ...
... Hollows are irregularly shaped, flat-floored, shallow depressions that occur predominantly within impact crater structures (crater floors, walls and rims, central peaks, and ejecta) and are considered very young in age (Blewett et al. 2011(Blewett et al. , 2013(Blewett et al. , 2018Thomas et al. 2014Thomas et al. , 2016. They are often, but not always, found in association with low-reflectance materials (Thomas et al. 2014(Thomas et al. , 2016, a unit considered rich in graphite . Their morphology suggests that they formed via sublimation of volatiles through a number of mechanisms, including solar heating and heating via contact with magmatic materials or impact melt (Blewett et al. 2011(Blewett et al. , 2013(Blewett et al. , 2018Thomas et al. 2014Thomas et al. , 2016Phillipps et al. 2021). ...
... They are often, but not always, found in association with low-reflectance materials (Thomas et al. 2014(Thomas et al. , 2016, a unit considered rich in graphite . Their morphology suggests that they formed via sublimation of volatiles through a number of mechanisms, including solar heating and heating via contact with magmatic materials or impact melt (Blewett et al. 2011(Blewett et al. , 2013(Blewett et al. , 2018Thomas et al. 2014Thomas et al. , 2016Phillipps et al. 2021). Examination and modeling of the MESSENGER MDIS color imaging and MASCS Vis-NIR observations, coupled with comparisons with laboratory spectral measurements, indicate that the volatile species involved in hollows formation are predominately sulfides, such as CaS, MgS, and NaS (e.g., Barraud et al. 2023), though chlorides have also been considered (e.g., Lucchetti et al. 2021). ...
Article
Full-text available
The fate of Mercury’s exospheric volatiles and, in a lesser way, of the refractory particles absorbed in the first few centimeters of the surface both depend highly on the temperature profile with depth and its diurnal variation. In this paper, we review several mechanisms by which the surface temperature might control the surface/exosphere interface. The day/night cycle of the surface temperature and its orbital variation, the temperature in the permanent shadow regions, and the subsurface temperature profiles are key thermal properties that control the fate of the exospheric volatiles through the volatile ejection mechanisms, the thermal accommodation, and the subsurface diffusion. Such properties depend on the solar illumination from large to small scales but also on the regolith structure. The regolith is also space-weathered by the thermal forcing and by the thermal-mechanical processing. Its composition is changed by the thermal conditions. We conclude by discussing key characteristics that need to be investigated theoretically and/or in the laboratory: the dependency of the surface spectra with respect to temperature, the typical diffusion timescale of the volatile species, and the thermal dependency of their ejection mechanisms.
... They are rimless and shallow, and many have high-reflectance and diffuse bright halos (Blewett et al., 2011). The data from MESSENGER's Mercury flybys and especially the orbital phase of the mission revealed that hollows occur in a variety of settings and around craters as well as in locations not directly linked with impact structures (e.g., Blewett et al., 2013;Thomas et al. 2014). Here, we Dominici crater (20 km diameter, 1°N 323°E, see the inset of Fig. 1) (Vilas et al., 2016;Blewett et al., 2013;Thomas et al., 2016). ...
... Global occurrence of hollows, pits and spectrally red pitted ground over MESSENGER MDIS global mosaic (9 th version), modified byThomas et al. (2014): inset displays Dominici crater (1°N 323°E). Also, in Dominici crater is possible to observe the high reflectance features. ...
Article
The temperature excursion variation of Mercury's surface may significantly change crystal structure of surfacecomprising minerals. The thermal stability of oldhamite (CaS) was investigated to validate its presence on the Mercury's surface. In particular, X-Ray Powder Diffraction (XRPD) and Thermogravimetric Analyses (TGA) on synthetic powder calcium sulfide (Alfa Aesar) were performed with the aim of confirming its stability up to 723.15 K, the highest temperature that is recorded for the surface of Mercury. Our results by XRPD and TGA results confirmed that CaS phase is stable within the daily temperature excursion on Mercury surface. Thermal expansion analyses determined the thermal expansion volume coefficient of αV = 4.03 × 10−5 K−1. The results of this work support the presence of Ca-sulfide phases on Mercury's surface and provide valid tools for interpreting the data that will be collected by the BepiColombo space mission (European Space Agency and Japanese Aerospace Exploration Agency) to Mercury.
... The sublimation-driven processes described above on 67P have both planetary and cometary analogs. Mars's scalloped terrains and swiss cheese terrains (Lefort et al., 2010;Morgenstern et al., 2007), Pluto's pits (Howard et al., 2017), Triton's depressions in its southern hemisphere terrains (Hansen et al., 2021) and hollows on Mercury (R. J. Thomas et al., 2014) all exhibit striking similarities to scarp fronts observed on 67P, wherein volatile sublimation liberates refractory grains, driving the back-wasting of scarps. Scarp retreat is also proposed as a mechanism for the growth of Titan's small lake depressions (Hayes, 2016) and occurs across Earth where failure at a cliff base, combined with efficient transport of the produced sediment, forms a characteristic shape (Howard, 1995). ...
Article
Full-text available
Comets are active geological worlds with primitive surfaces that have been shaped to varying degrees by sublimation‐driven sediment transport and mass wasting processes. Rosetta's rendezvous with comet 67P/Churyumov‐Gerasimenko (67P) in 2014 provided data with the necessary spatial and temporal resolutions to observe many evolutionary processes on micro‐gravity worlds. Rosetta's observations have thus far revealed that many changes to the surface occurred within 67P's smooth terrains, vast sedimentary deposits that blanket a significant fraction of the nucleus. Understanding the global context of these changes, and therefore the sediment transport pathways that govern the evolution of 67P's surface, requires a thorough description of their changing morphologies and an evaluation of existing global‐scale spatial and temporal trends. Accordingly, we present a time‐resolved synthesis of erosion and deposition activity on comet 67P as it passed through its 13 August 2015 perihelion from September 2014 to August 2016. Our mapping results indicate that, around perihelion, sediment is globally redistributed inter‐regionally from 67P's more active south to the north. Equally important, however, are local, topographically influenced sediment transport processes, with large volumes of sediment moving intra‐regionally over sub‐kilometer distances. We also show evidence for regions of near‐zero net erosion/deposition between approximately 30°N–60°N latitude, which may act as terminal sedimentary sinks, with the remobilization of these materials hindered by multiple factors. Our work therefore provides the most complete mapping of sediment transport processes and pathways across 67P, a critical step toward understanding the global landscape evolution of both 67P and other comets.
Article
Using MESSENGER Mercury Dual Imaging System data, we produced 3 new maps of Sibelius Crater, Mercury. Geomorphological and spectral maps were combined into a single hybrid map containing units associated with ejecta deposits, crater floor landforms and impact melt ponding. Spatial measurement of these units show that ∼50% of the mapped melt pond area lies within a large, degraded impact crater (crater B), beyond the significantly lower northern Sibelius rim, with a potential melt flow to a smaller, degraded impact structure further north (crater C). Freshly processed spectral data from the 8-color Map Projected Multispectral Reduced Data Record (MDR) data highlight the emplacement of multiple uplifted ejecta units with distinct spectral properties. A new, high resolution digital elevation model was created to help define and analyse crater floor uplift features, disrupted crater rims and to create detailed cross sections. These illustrate a proposed location of B's central uplift structure exposed in the northern wall slopes of Sibelius. Small features at the limit of visibility, such as a groove possibly associated with a rolling or sliding mega-boulder, and lobate melt flow on the crater floor with accompanying channel opening, are highlighted for future investigations by BepiColombo's instruments once it reaches orbit.
Article
MESSENGER Gamma-Ray Spectrometer measurements demonstrate that the abundance of Na varies across the surface of Mercury. The maximum Na/Si abundance ratio of 0.20 ± 0.03 by weight (∼5 wt% Na) is observed at high northern latitudes and is significantly larger than the equatorial Na/Si ratio of 0.11 ± 0.01 (∼2.6 wt% Na). Comparisons of forward-modeled surface distributions with the gamma-ray measurements suggest that the observed distribution of Na can be explained by differences in elemental composition between the volcanic smooth plains units and heavily cratered terrain. The comparison improves when thermally driven depletion of Na from areas near Mercury’s hot poles is included. When combined with other MESSENGER data sets, these results indicate that the smooth plains units include substantial abundances of alkali feldspars. Thermal depletion of Na from the hot poles without an assumed underlying compositional variability can also reproduce the measured Na/Si distribution, but that mechanism fails to account for other MESSENGER observations that support the presence of higher abundances of feldspars in the smooth plains units.
Article
A 27×13 km27×13 km ‘rimless depression’ 100 km inside the southwest rim of the Caloris basin is revealed by high resolution orbital imaging under a variety of illuminations to consist of at least nine overlapping volcanic vents, each individually up to 8 km in diameter. It is thus a ‘compound’ volcano, indicative of localised migration of the site of the active vent. The vent floors are at a least 1 km below their brinks, but lack the flat shape characteristically produced by piston-like subsidence of a caldera floor or by flooding of a crater bottom by a lava lake. They bear a closer resemblance to volcanic craters sculpted by explosive eruptions and/or modified by collapse into void spaces created by magma withdrawal back down into a conduit. This complex of overlapping vents is at the summit of a subtle edifice at least 100 km across, with flank slopes of about only 0.2 degrees, after correction for the regional slope. This is consistent with previous interpretation as a locus of pyroclastic eruptions. Construction of the edifice could have been contributed to by effusion of very low viscosity lava, but high resolution images show that the vent-facing rim of a nearby impact crater is not heavily embayed as previously supposed on the basis of lower resolution flyby imaging. Contrasts in morphology (sharpness versus blurredness of the texture) and different densities of superposed sub-km impact craters inside each vent are consistent with (but do not prove) substantial differences in the age of the most recent activity at each vent. This suggests a long duration of episodic magmagenesis at a restricted locus. The age range cannot be quantified, but could be of the order of a billion years. If each vent was fed from the same point source, geometric considerations suggest a source depth of at least 50 km. However, the migration of the active vent may be partly controlled by a deep-seated fault that is radial to the Caloris basin. Other rimless depressions in this part of the Caloris basin fall on or close to radial lines, suggesting that elements of the Pantheon Fossae radial fracture system that dominates the surface of the central portion of the Caloris basin may continue at depth almost as far as the basin rim.
Article
Orbital images acquired by the MErcury, Surface, Space ENvironment, GEochemistry, and Ranging (MESSENGER) spacecraft reveal a distinctive low-reflectance material on the surface of Mercury. Such material occurs in small, isolated, and thin surficial units. We term these features dark spots. Dark spots have the lowest average reflectance yet documented on the planet. In every case observed at sufficiently high resolution, dark spots feature hollows at their centers. Not all hollows, however, are surrounded by a dark spot. Dark spots have been found on low-reflectance smooth plains, intercrater plains, heavily cratered terrain, and impact craters at almost all longitudes on Mercury, but they have not been documented on high-reflectance smooth plains material. Dark spots are one of the youngest endogenic features on Mercury, and some postdate craters with distinctive rays. Sulfides may be the phase responsible for the low albedo of dark spot material. We propose that dark spots form during the initial stages of hollow formation, perhaps in a manner similar to intense outgassing events that feature exit velocities in excess of 100 m/s. Such outgassing could contemporaneously produce a depression that constitutes an embryonic hollow. Under this scenario, dark spot material is subsequently removed or modified by regolith gardening or other surface processes on time scales shorter than the lifetime of the central hollow.
Article
The frequent spatial association of volcanic pits with thrust faults on Mercury is motivation to study if faults functioned as conduits for volatile-rich magmas.
Article
Radar-bright features near Mercury's poles were discovered in Earth-based radar images and proposed to be water ice present in permanently shadowed areas. Images from MESSENGER's one-year primary orbital mission provide the first nearly complete view of Mercury's north polar region, as well as multiple images of the surface under a range of illumination conditions. We find that radar-bright features near Mercury's north pole are associated with locations persistently shadowed in MESSENGER images. Within 10° of the pole, almost all craters larger than 10 km in diameter host radar-bright deposits. There are several craters located near Mercury's north pole with sufficiently large diameters to enable long-lived water ice to be thermally stable at the surface within regions of permanent shadow. Craters located farther south also host radar-bright deposits and show a preference for cold-pole longitudes; thermal models suggest that a thin insulating layer is required to cover these deposits if the radar-bright material consists predominantly of long-lived water ice. Many small (<10 km diameter) and low-latitude (extending southward to 66°N) craters host radar-bright material, and water ice may not be thermally stable in these craters for ~1 Gy, even beneath an insulating layer. The correlation of radar-bright features with persistently shadowed areas is consistent with the deposits being composed of water ice, and future thermal modeling of small and low-latitude craters has the potential to further constrain the nature, source, and timing of emplacement of the radar-bright material.
Article
Orbital images from the MESSENGER spacecraft show that ~27% of Mercury's surface is covered by smooth plains, the majority (>65%) of which are interpreted to be volcanic in origin. Most smooth plains share the spectral characteristics of Mercury's northern smooth plains, suggesting they also share their magnesian alkali-basalt-like composition. A smaller fraction of smooth plains interpreted to be volcanic in nature have a lower reflectance and shallower spectral slope, suggesting more ultramafic compositions, an inference that implies high temperatures and high degrees of partial melting in magma source regions persisted through most of the duration of smooth plains formation. The knobby and hummocky plains surrounding the Caloris basin, known as Odin-type plains, occupy an additional 2% of Mercury's surface. The morphology of these plains and their color and stratigraphic relationships suggest that they formed as Caloris ejecta, although such an origin is in conflict with a straightforward interpretation of crater size-frequency distributions. If some fraction is volcanic, this added area would substantially increase the abundance of relatively young effusive deposits inferred to have more mafic compositions. Smooth plains are widespread on Mercury, but they are more heavily concentrated in the north and in the hemisphere surrounding Caloris. No simple relationship between plains distribution and crustal thickness or radioactive element distribution is observed. A likely volcanic origin for some older terrain on Mercury suggests that the uneven distribution of smooth plains may indicate differences in the emplacement age of large-scale volcanic deposits rather than differences in crustal formational process.
Article
Volcanic plains identified on Mercury are morphologically similar to lunar mare plains but lack constructional and erosional features that are prevalent on other terrestrial planetary bodies. We analyzed images acquired by the MESSENGER spacecraft to identify features on Mercury that may have formed by lava erosion. We used analytical models to estimate eruption flux, erosion rate, and eruption duration to characterize the formation of candidate erosional features, and we compared results with analyses of similar features observed on Earth, the Moon, and Mars. Results suggest that lava erupting at high effusion rates similar to those required to form the Teepee Butte Member of the Columbia River flood basalts (0.1-1.2 × 106 m3 s-1) would have been necessary to form wide valleys (>15 km wide) observed in Mercury's northern hemisphere, first by mechanical erosion to remove an upper regolith layer, then by thermal erosion once a lower rigid layer was encountered. Alternatively, results suggest that lava erupting at lower effusion rates similar to those predicted to have formed Rima Prinz on the Moon (4400 m3 s-1) would have been required to form, via thermal erosion, narrower channels (<7 km wide) observed on Mercury. Although these results indicate how erosion might have occurred on Mercury, the observed features may have formed by other processes, including lava flooding terrain sculpted during the formation of the Caloris basin in the case of the wide valleys, or impact melt carving channels into impact ejecta in the case of the narrower channels.
Article
contrast to other terrestrial planets, Mercury does not possess a great variety of volcanic features, its history of volcanism instead largely manifest by expansive smooth plains. However, a set of landforms at high northern latitudes on Mercury resembles surface flow features documented on Earth, the Moon, Mars, and Venus. The most striking of such landforms are broad channels that host streamlined islands and that cut through the surrounding intercrater plains. Together with narrower, more sinuous channels, coalesced depressions, evidence for local flooding of intercrater plains by lavas, and a first-order analysis of lava flow rates, the broad channels define an assemblage of flow features formed by the overland flow of, and erosion by, voluminous, high-temperature, low-viscosity lavas. This interpretation is consistent with compositional data suggesting that substantial portions of Mercury's crust are composed of magnesian, iron-poor lithologies. Moreover, the proximity of this partially flooded assemblage to extensive volcanic plains raises the possibility that the formation of these flow features may preface total inundation of an area by lavas emplaced in a flood mode and that they escaped complete burial only due to a waning magmatic supply. Finally, that these broad channels on Mercury are volcanic in nature yet resemble outflow channels on Mars, which are commonly attributed to catastrophic water floods, implies that aqueous activity is not a prerequisite for the formation of such distinctive landforms on any planetary body.