03.11.2014 Views

Untitled

Untitled

Untitled

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Bulletin of the Czech Geological Survey, Vol. 77, No. 4, 243–252, 2002<br />

© Czech Geological Survey, ISSN 1210-3527<br />

Asteroids: Their composition and impact threat<br />

THOMAS H. BURBINE<br />

Department of Mineral Sciences, National Museum of Natural History, Smithsonian Institution, Washington, DC 20560-0119, USA; e-mail:<br />

burbine.tom@nmnh.si.edu<br />

Abstract. Impacts by near-Earth asteroids are serious threats to life as we know it. The energy of the impact will be a function of the mass of<br />

the asteroid and its impact velocity. The mass of an asteroid is very difficult to determine from Earth. One way to derive a near-Earth object’s mass is<br />

by estimating the object’s density from its surface composition. Reflectance spectra are the best way to determine an object’s composition since many<br />

minerals (e.g. olivine, pyroxene, hydrated silicates) have characteristic absorption features. However, metallic iron does not have characteristic absorption<br />

bands and is very hard to identify from Earth. For a particular size, asteroids with compositions similar to iron meteorites pose the biggest<br />

impact threat since they have the highest densities, but they are expected to be only a few percent of the impacting population. Knowing an asteroid’s<br />

composition is also vital for understanding how best to divert an incoming asteroid.<br />

Key words: Earth, asteroids, impact features, meteorites, mineral composition, geologic hazards<br />

Introduction<br />

As we enter a new millennium, we are constantly being<br />

bombarded with news of close encounters with near-Earth<br />

objects (NEOs). Observers have discovered over 2,000 near-<br />

Earth asteroids (NEAs); luckily none are known to be on a<br />

collision course with the Earth. Comets are also serious impact<br />

threats as shown by the collision of Shoemaker-Levy-9<br />

with Jupiter in 1994. Since the discovery of the iridium<br />

anomaly at the Cretaceous-Tertiary (K-T) boundary layer by<br />

Alvarez et al. (1980), there has been a considerable discussion<br />

of the possibility and consequences of such an impact<br />

(e.g. Chapman and Morrison 1994, Adushkin and Nemchinov<br />

1994, Toon et al. 1997, Garshnek et al. 2000).<br />

As we discover more Earth-approaching asteroids, we<br />

are also learning more about their compositions and structure.<br />

Charge-coupled devices (CCDs) now allow us to obtain<br />

visible and near-infrared telescopic spectra of<br />

near-Earth asteroids that are a few hundred meters or<br />

smaller in diameter (e.g. Binzel et al. 2001a). Technological<br />

improvements to the radio telescope at the Arecibo Observatory<br />

have allowed similarly sized objects to be<br />

characterized by radar (e.g. Ostro et al. 2002). Spacecraft<br />

missions now show asteroids to be geologic bodies with a<br />

variety of morphologic features. More meteorites are discovered<br />

every year and are being extensively studied in the<br />

laboratory with more precise analytical techniques.<br />

However as we continue our research on asteroids, a<br />

number of questions should be asked. “How do asteroid<br />

compositions affect the impact threat?” “How well can we<br />

determine asteroid compositions from Earth?” This paper<br />

will review what we currently know about asteroid compositions<br />

and how it affects the impact threat. Since we<br />

have not yet returned any samples from any asteroids, our<br />

knowledge of asteroid compositions is derived from analyses<br />

of meteorites, remote sensing observations from Earth,<br />

and spacecraft missions.<br />

Impact threat<br />

I will first briefly review what we know about the number<br />

of near-Earth objects and the energy and effects of impacts.<br />

The near-Earth object population is defined as small<br />

bodies with perihelion distance less than 1.3 AU (astronomical<br />

units) and aphelion distance greater than 0.983<br />

AU (Morbidelli et al. 2002). These objects are primarily<br />

thought to be asteroids ejected from the main belt; however,<br />

a few extinct comets probably exist in the population.<br />

Recent estimates of the numbers of NEAs larger than one<br />

kilometer in diameter vary from 855 (±110) (Morbidelli et<br />

al. 2002) to 1227 (uncertainties of +170 and –90) (Stuart<br />

2001).<br />

The kinetic energy (E) of an incoming asteroid is<br />

(1/2)mv 2 where m is the mass and v is the velocity. Mass is<br />

a function of the density (ρ) and volume (V) of the object.<br />

Since the energy of an impact is usually given as megatons<br />

of TNT, the kinetic energy equation can be written (Morbidelli<br />

et al. 2002) as<br />

E = 62.5 ρd 3 v 2<br />

where E is in megatons, ρ is in g/cm 3 , d is the diameter of<br />

the impactor in kilometers, and v is in km/s. The average<br />

impact velocity for asteroids with Earth are ~ 20 km/s (e.g.<br />

Hughes, 1998). Comets have lower densities (estimated to<br />

be around ~ 1 g/cm 3 ), but some long-period comets have<br />

much higher average impact velocities (~ 55 km/s) (Marsden<br />

and Steel 1994). Since water covers approximately<br />

three-fourths of the Earth’s surface, “large” impacts are<br />

likely to cause tsunamis (e.g. Paine 1999), giant tidal<br />

waves created by sudden disturbances.<br />

The best-characterized localized catastrophe is the impact<br />

at Tunguska where an object exploded 5–10 km in the<br />

air over an uninhabited region of Siberia (e.g. Chyba et al.<br />

1993, Vasilyev 1998). The energy was estimated to be<br />

~ 10–20 megatons and devastated an area of ~ 2000 km 2<br />

of forest area. In comparison, the energy is ~ 1,000 times<br />

243


Thomas H. Burbine<br />

larger than the bomb dropped on Hiroshima and the area is<br />

~ 4 times larger than New York City.<br />

The best-characterized “large” impact is the Chicxulub<br />

crater found in the Yucatan peninsula (e.g. Hildebrand et<br />

al. 1998). Hildebrand et al. (1998) argues that the crater is<br />

180 km in diameter, but other researchers (e.g. Morgan et<br />

al. 1998) argue that the actual crater size could be as large<br />

as 270 km. The energy of the impactor for producing the<br />

Chicxulub crater is estimated to be ~ 10 8 megatons (e.g.<br />

Toon et al. 1997). This impact occurred ~ 65 million years<br />

ago, which is the same age as the K-T extinction. The K-<br />

T boundary marks the demise of the dinosaurs as the dominant<br />

animal species on Earth (e.g. Milner 1998).<br />

We currently have evidence for over 160 impact craters<br />

on Earth over the last ~ 2 billion years that range from 15<br />

meters to 300 kilometers in diameter (Earth Impact<br />

Database 2002). Many more impacts have occurred over<br />

this time with evidence for these impacts wiped out due to<br />

erosion by water and wind. Cratering rates for the Earth<br />

(e.g. Neukum and Ivanov, 1994) can be estimated from determinations<br />

of the NEO population, the lunar cratering<br />

rate (where the effects of the Earth’s atmosphere needed to<br />

be added), and crater counts on terrestrial cratons.<br />

Neukum and Ivanov (1994) estimate that craters one kilometer<br />

in size or greater form on the Earth every ~ 1600<br />

years and craters 100 kilometers in size or greater form<br />

every 27 million years.<br />

Morbidelli et al. (2002) have done a recent study of<br />

collision probabilities of NEOs with the Earth. They estimate<br />

that a 1000 megaton impact, which would produce<br />

large-scale regional damage and a crater ~ 5 km in size<br />

(e.g. Hughes 1998), should occur every 63,000 ± 8,000<br />

years. They estimate that known NEOs carry only ~ 18%<br />

of the overall collision probability.<br />

The consequences of any impact on Earth will be a<br />

function of the mass (density times volume) and impact<br />

velocity plus the location, date, and time of the impact on<br />

Earth. Astrometric observations, which determine the<br />

location of an object in the sky, can help determine an asteroid’s<br />

orbit with a high degree of certainty. These observations<br />

can be used to predict the probability that an object<br />

will strike the Earth. If an object is on a collision course<br />

these astrometric observations can be used to predict the<br />

impact velocity and the location of the impact. Radar observations<br />

(e.g. Ostro et al. 2002) can image an asteroid (if<br />

it is large enough and close enough to Earth) and then determine<br />

its diameter, which can be used to calculate its<br />

volume.<br />

However, the determination of an asteroid’s mass is<br />

much more difficult from Earth. For the largest asteroids,<br />

masses can be determined by asteroid-asteroid interactions<br />

or their perturbations on Mars (Britt et al. 2002). To determine<br />

a near-Earth asteroid’s mass, the object needs to have<br />

a natural body revolving around it (e.g. Margot et al. 2002)<br />

or a spacecraft encounter (e.g. Veverka et al. 1999, Yeomans<br />

et al. 2000).<br />

However in the absence of such a mass determination,<br />

the best way to estimate a near-Earth asteroid’s mass<br />

would be to determine its surface composition, which<br />

would give insight into the object’s density. (Since near-<br />

Earth asteroids are fragments of much larger asteroids, we<br />

assume that the surface composition is representative of<br />

the object as a whole.) The mass could then be estimated<br />

by multiplying the inferred density by the object’s volume.<br />

Such an estimate would only give an upper limit on an object’s<br />

mass, since many asteroids are thought to be rubble<br />

piles with significant amounts of macroporosity (e.g. Britt<br />

and Consolmagno 2001, Britt et al. 2002). The estimated<br />

macroporosities for ~ 20 asteroids range from 0 to ~ 80%<br />

(Britt et al. 2002).<br />

Meteorites<br />

Except for ~ 50 samples from the Moon and Mars, meteorites<br />

appear to be fragments of sub-planetary sized bodies<br />

(asteroids) that formed ~ 4.56 billion years ago.<br />

Meteorites can basically be broken into two types: those that<br />

experienced heating but not melting (chondrites) (e.g.<br />

Brearley and Jones 1998) and those that experienced melting<br />

and differentiation (achondrites, stony-irons, irons) (e.g.<br />

Mittlefehldt et al. 1998). Silicate-rich meteorites are often<br />

referred to as stony and would include all chondrites and<br />

achondrites. Meteorites that are similar in terms of petrologic,<br />

mineralogical, bulk chemical, and isotopic properties<br />

are separated into groups (Table 1). In general, groups contain<br />

five or more members to allow for the compositional<br />

characteristics of the group to be adequately characterized.<br />

Currently, 13 groups (Table 1) of chondritic meteorites<br />

have been defined. Chondritic groups are subdivided according<br />

to petrologic type (1–6) with 1 being the most<br />

aqueously altered, 3.0 being the least altered, and 6 being<br />

heated to the highest temperatures. Mineralogically, meteorites<br />

of petrologic type 1 and 2 (CI, CM, CR) have compositions<br />

dominated by phyllosilicates. Visually, these<br />

meteorites are extremely dark. The other types of carbonaceous<br />

chondrites (CH, CV, CO, CK) tend to have<br />

mineralogies dominated by mafic silicates. The R chondrites<br />

are dominated by olivine. Ordinary chondrites (H,<br />

L, LL) are mixtures of mafic silicates and metallic iron,<br />

while enstatite chondrites are composed of enstatite (virtually<br />

FeO-free pyroxene) and metallic iron. In addition to<br />

these 13 well-defined groups, ~ 14 chondritic grouplets or<br />

unique meteorites (Meibom and Clark 1999, Weisberg et<br />

al. 2001, Mittlefehldt, 2002) have been recognized.<br />

The differentiated meteorites range from those that experienced<br />

only limited differentiation (primitive achondrites)<br />

to those (differentiated achondrites, stony-irons,<br />

irons) that were produced by extensive melting, melt migration,<br />

and fractional crystallization. These processes<br />

produce a wide variety of lithologies.<br />

Many of the differentiated meteorites are thought to<br />

sample the crusts (howardites, eucrites, and diogenites or<br />

HEDs; angrites), core-mantle boundaries (pallasites), and<br />

cores (irons) of differentiated bodies. Eucrites contain primarily<br />

plagioclase and both low-Ca and high-Ca pyrox-<br />

244


Asteroids: Their composition and impact threat<br />

enes, while diogenites are predominantly magnesian orthopyroxene.<br />

Howardites are breccias of eucritic and diogenitic<br />

material. Pallasites are mixtures of metallic iron<br />

and olivine. We do not appear to be sampling the mantles<br />

of these differentiated bodies (e.g. Burbine et al. 1996);<br />

however, we do appear to be sampling the mantles of veryreduced<br />

differentiated bodies in the form of aubrites.<br />

Iron meteorites are composed of metallic iron with<br />

5–20% Ni plus accessory phases such as sulfides,<br />

schreibersite, and silicate inclusions. Iron meteorites are<br />

classified according to siderophile (“iron-loving”) element<br />

(Ga, Ge, Ir, Ni) concentrations. Of the thirteen groups of<br />

iron meteorites, ten (IC, IIAB, IIC, IID, IIF, IIIAB, IIIE,<br />

IIIF, IVA, IVB) have fractional crystallization trends suggestive<br />

of prolonged cooling as expected from the cores of<br />

differentiated bodies. Three of the groups (IAB, IIE, and<br />

IIICD) do not display well-developed fractional crystallization<br />

trends and contain abundant silicate inclusions,<br />

which argues that they are not core fragments. Approximately<br />

90 irons (Grady 2000) are not classified as members<br />

of the 13 groups and are labeled anomalous. These<br />

ungrouped irons are believed to require 50-70 distinct parent<br />

bodies (Wasson 1995, Burbine et al. 2002b).<br />

Other types of differentiated meteorites have undergone<br />

varying amounts of melting. These include a number<br />

of primitive achondritic meteorites (acapulcoite-lodranites,<br />

winonaites), which are samples of partially differentiated<br />

asteroids. These meteorites are mixtures of olivine,<br />

pyroxene, and metallic iron. Mesosiderites are breccias<br />

composed of HED-like clasts of basaltic material mixed<br />

with metallic clasts. One possible scenario for the formation<br />

of the mesosiderites is the disruption of an asteroid<br />

with a molten core (Scott et al. 2001). Rounding out the<br />

differentiated meteorites are the olivine-dominated brachinites<br />

and the carbon-rich ureilites, whose origins are<br />

still being debated (e.g. Mittlefehldt et al. 1998).<br />

Meteorite densities and strengths<br />

Consolmagno and Britt (1998), Flynn et al. (1999), and<br />

Wilkinson and Robinson (2000) have recently done studies<br />

of meteorite bulk densities. The only CI chondrite<br />

measured had a bulk density of 1.58 g/cm 3 while the bulk<br />

densities of CM chondrites were 2.08–2.37 g/cm 3 . CO<br />

chondrites (2.36–2.98 g/cm 3 ) and CV chondrites<br />

(2.60–3.18 g/cm 3 ) tended to have slightly higher bulk densities.<br />

The only enstatite chondrite measured had a bulk<br />

density of 3.36 g/cm 3 . On average, ordinary chondrites<br />

(3.05–3.75 g/cm 3 for the most reliable measurements)<br />

have the highest bulk densities of the chondrites.<br />

HEDs have bulk densities of 2.99–3.29 g/cm 3 . Stonyirons<br />

such as mesosiderites (4.16–4.22 g/cm 3 ) and pallasites<br />

(4.82–4.97 g/cm 3 ) have higher bulk densities due to<br />

the presence of significant amounts of metallic iron. As expected,<br />

iron meteorites tend to have the highest bulk densities<br />

of all meteorite types (6.99–7.59 g/cm 3 for relatively<br />

unweathered specimens).<br />

Table 1. Meteorite groups<br />

Groups Composition* Fall percentages # (%)<br />

Carbonaceous chondrites<br />

CI phy, mag 0.5<br />

CM phy, toch, ol, 1.7<br />

CR phy, px, ol, met 0.3<br />

CO ol, px, CAIs, met 0.5<br />

CH px, met, ol, Only finds<br />

CV ol, px, CAIs 0.6<br />

CK ol, CAIs 0.3<br />

Enstatite chondrites<br />

EH enst, met, sul, plag, ± ol 0.8<br />

EL enst, met, sul, plag 0.7<br />

Ordinary chondrites<br />

H ol, px, met, plag, sul 34.1<br />

L ol, px, plag, met, sul 38.0<br />

LL ol, px, plag, met, sul 7.9<br />

R chondrites ol, px, plag, sul 0.1<br />

Primitive achondrites<br />

Acapulcoites a px, ol, plag, met, sul 0.1<br />

Lodranites a px, ol, met, ± plag, ± sul 0.1<br />

Winonaites ol, px, plag, met 0.1<br />

Differentiated achondrites<br />

Angrites TiO 2-rich aug, ol, plag 0.1<br />

Aubrites enst, sul 0.1<br />

Brachinites ol, cpx, ± plag Only finds<br />

Diogenites b opx 1.2<br />

Eucrites b pig, plag 2.7<br />

Howardites b eucritic-diogenitic breccia 2.1<br />

Ureilites ol, px, graph 0.5<br />

Stony-irons<br />

Mesosiderites basalt-met breccia 0.7<br />

Pallasites ol, met ol, met 0.5<br />

Irons c met, sul, schreib 4.2<br />

Table is revised from tables found in Burbine et al. (2002b).<br />

* Minerals or components are listed in decreasing order of average abundance.<br />

Abbreviations: ol – olivine, px – pyroxene, opx – orthopyroxene,<br />

pig – pigeonite, enst –enstatite, aug – augite, cpx – clinopyroxene, plag<br />

– plagioclase, mag – magnetite, met – metallic iron, sul – sulfides, phy –<br />

phyllosilicates, toch – tochilinite, graph – graphite, CAIs – Ca-Al-rich<br />

refractory inclusions, schreib – schreibersite, ± – may be present<br />

# Fall percentages are calculated from the 942 classified falls that are listed<br />

in Grady (2000), Grossman (2000), and Grossman and Zipfel (2001).<br />

a Acapulcoites and lodranites appear to come from the same parent body<br />

(e.g. Mittlefehldt et al. 1998).<br />

b Howardites, eucrites, and diogenites (HEDs) appear to come from the<br />

same parent body (e.g. Mittlefehldt et al. 1998).<br />

c There are 13 iron meteorite groups plus ~100 ungrouped irons.<br />

In terms of physical strength, most chondritic meteorites<br />

can be easily crushed into powders with a mortar<br />

and pestle. Meteorites containing phyllosilicates tend to be<br />

the most fragile with the CI chondrites being the weakest<br />

of these objects with laboratory crushing strengths of 1–10<br />

bars (e.g. Lewis 2000). What cannot be easily pulverized<br />

is metallic iron. Metallic iron is ductile and tends to flatten<br />

and elongate while being crushed at room temperatures. It<br />

is unclear how ductile metallic iron is at the colder temperatures<br />

found in the asteroid belt.<br />

Iron meteorites are extremely strong with strengths of<br />

approximately 3.5 kbars (e.g. Lewis, 2000). The much<br />

stronger physical strength of irons allows them to survive<br />

in space much longer than stony bodies as seen by their<br />

longer cosmic ray exposure ages. Cosmic ray exposure<br />

ages record the time an object has spent as a meter-sized<br />

245


Thomas H. Burbine<br />

Normalized Reflectance<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0.2 0.6 1 1.4 1.8 2.2 2.6<br />

Wavelength (µm)<br />

or less body in space or within a few meters of the surface.<br />

Irons have cosmic ray exposure ages ranging from hundreds<br />

of millions to a few billion years (e.g. Voshage and<br />

Feldman 1979) while stony meteorites tend to have much<br />

shorter exposure ages (


Asteroids: Their composition and impact threat<br />

Asteroid observations<br />

This section will detail what we currently think we understand<br />

about asteroid compositions. Asteroid spectral<br />

surveys (e.g. Zellner et al. 1985, Bus and Binzel 2002a)<br />

have primarily been done in the visible due to the peaking<br />

of the illuminating solar flux and the relative transparency<br />

of the atmosphere at these wavelengths. Near-infrared observations<br />

(1-3.5 µm) have now become easier to obtain<br />

due to the advent of SpeX (an infrared spectrograph at the<br />

Infrared Telescope Facility on Mauna Kea) (e.g. Binzel et<br />

al. 2001b).<br />

Asteroids are generally grouped into classes (Table 2)<br />

based on their visible spectra (~ 0.4 to ~ 0.9–1.1 µm) and<br />

visual albedo (when available). The most widely used taxonomy<br />

(Tholen 1984) classifies objects observed in the<br />

eight-color asteroid survey (ECAS) (Zellner et al. 1985).<br />

Bus and Binzel (2002a) did a CCD spectral survey of over<br />

1300 objects and developed an expanded taxonomy (Bus<br />

and Binzel 2002b) with many more classes and subclasses<br />

to represent the diversity of spectral properties seen in<br />

higher-resolution spectra. Representative spectra of a<br />

number of Bus and Binzel (2002a, 2002b) asteroid classes<br />

are shown in Figure 2.<br />

As stated earlier, to accurately determine an asteroid’s<br />

mineralogy, observations are needed in the near-infrared<br />

since many minerals have diagnostic features in this wavelength<br />

region. Near-infrared asteroid spectra (Gaffey et al.<br />

1993, Rivkin et al. 2000, Burbine and Binzel 2002) at a variety<br />

of wavelengths have shown that each class tends to<br />

contain a wide variety of mineralogies. Only qualitative<br />

mineralogical descriptions will be given for asteroid classes<br />

discussed in this paper; quantitative analyses of asteroid<br />

compositions (determining the proportion and composition<br />

of different mineral species) are very difficult (e.g.<br />

Clark, 1995) since an asteroid’s reflectance spectrum is a<br />

function of a number of factors such as mineralogy, mineral<br />

chemistry, particle size, and temperature.<br />

Asteroid taxonomy is based on astronomically observed<br />

parameters without regard to composition. However, many<br />

asteroid classes have been given letter designations that<br />

“imply” specific compositions. This includes the letter “M”<br />

for metallic, “S” for siliceous (or stony or stony-iron), and<br />

“C” for carbonaceous. (Bus and Binzel (2002b) do not differentiate<br />

between E, M, and P asteroids and call them all X<br />

objects since there spectra are similar and albedo is not used<br />

in their taxonomy.) Since each asteroid class only groups asteroids<br />

with specific spectral characteristics, not all objects<br />

in a class have to have similar surface mineralogies. For example,<br />

it is unknown how many M-class asteroids actually<br />

have surfaces similar to iron meteorites. For example, enstatite<br />

chondrites (primarily mixtures of metallic iron and<br />

enstatite) also have relatively featureless spectra (Gaffey<br />

1976) and similar albedos to metallic iron. Also, a subgroup<br />

of M-class asteroids (called the W class) have distinctive 3<br />

µm features (e.g. Rivkin et al. 2000), which apparently indicate<br />

hydrated minerals and surface compositions inconsistent<br />

with metallic iron.<br />

Radar observations have been used to estimate the metal<br />

contents of asteroids (Ostro et al., 2002). Radar experiments<br />

of asteroids usually measure the distribution of echo<br />

power reflected from an object in time delay and Doppler<br />

Table 2. Asteroid Classes<br />

Class characteristics<br />

A Distinctive olivine absorption features<br />

B Weak UV feature, blue past 0.4 µm, subclass of C types; low albedo (generally less than 0.1)<br />

C a Weak UV feature, flat to reddish past 0.4 µm; low albedo (generally less than 0.10)<br />

D Very red spectrum; low albedo (usually around 0.05 or less)<br />

E Flat to slightly red, featureless spectrum; high albedo (> 0.30); usually associated with aubrites<br />

F Very weak UV feature, flat to bluish past 0.4 µm, subclass of C types; low albedo (< 0.10)<br />

G Strong UV feature, flat past 0.4 µm, subclass of C class; usually have strong 3 µm features; low albedo (< 0.10)<br />

J Stronger 1 µm feature than V types; appear to have compositions similar to the HEDs<br />

K Spectrum intermediate between S and C asteroids; usually associated with CO3/CV3 chondrites<br />

L b Very strong UV feature and then becoming approximately flat<br />

M Flat to slightly red, featureless spectrum; moderate albedo (0.10-0.30)<br />

O Weak UV feature out to 0.44 µm, very strong 1 µm feature; type spectrum is 3628 Božněmcová<br />

P Flat to slightly red, featureless spectrum; low albedo (usually around 0.05 or less)<br />

Q Strong UV and 1 µm feature; spectrum similar to ordinary chondrites<br />

R Strong UV and 1 µm feature; type spectrum is 349 Dembowska<br />

S c Strong UV feature; usually has 1 µm feature, indicating olivine and/or pyroxene; moderate albedo (0.10–0.30)<br />

T Weak UV feature, reddish past 0.4 µm; low albedo (< 0.10)<br />

V Distinctive 1 and 2 µm features due to pyroxene; appear to have compositions similar to the HEDs<br />

W M-class visible spectrum with a 3 µm absorption feature<br />

Spectrum similar to E, M, or P types, but no albedo information<br />

X d<br />

This table is revised from tables in Wetherill and Chapman (1988), Pieters and McFadden (1994), and Bus and Binzel (2002b). The term “red” refers<br />

to reflectances increasing with increasing wavelength and the term “blue” refers to reflectances decreasing with increasing wavelength.<br />

a Bus and Binzel (2001b) subdivided the C-class objects into the C, Cb, Cg, Ch, and Cgh subclasses based on CCD spectra. Bus and Binzel (2001b)<br />

do not define the F and G classes.<br />

b Bus and Binzel (2001b) subdivided the L-class objects into the L and Ld classes based on CCD spectra.<br />

c Gaffey et al. (1993) subdivided the S-class objects into the S(I) to S(VII) subclasses on the basis of both visible and near-infrared spectra. Bus and<br />

Binzel (2002b) have subdivided the S class into the S, Sa, Sk, Sl, Sq, and Sr subclasses based on CCD spectra.<br />

d Bus and Binzel (2002b) subdivided the X-class class objects into the X, Xc, Xe, and Xk subclasses based on CCD spectra.<br />

247


Thomas H. Burbine<br />

Normalized Reflectance<br />

2.8<br />

2.3<br />

1.8<br />

1.3<br />

1929 Kollaa (V)<br />

1542 Schalen (D)<br />

6 Hebe (S)<br />

221 Eos (K)<br />

64 Angelina (Xe)<br />

19 Fortuna (Ch)<br />

16 Psyche (X)<br />

0.8<br />

0.4 0.5 0.6 0.7 0.8 0.9 1<br />

Wavelength (µm)<br />

Fig. 2. Reflectance spectra of a number of Bus and Binzel (2002a,<br />

2002b) asteroid classes. All spectra are normalized to unity at 0.55 µm<br />

and offset in reflectance from each other. Error bars are one sigma. All<br />

spectra are available at the website http://smass.mit.edu.<br />

frequency in the opposite sense (OC) of circular polarization<br />

and the same sense. OC radar albedos are equal to the<br />

OC radar cross section divided by the target’s projected<br />

area. For homogeneous and particulate surfaces, radar<br />

albedos are functions of the near-surface bulk density,<br />

which is related to both the solid-rock density and the surface<br />

porosity of the object. Increasing the solid-rock density<br />

or decreasing the surface porosity would increase the<br />

radar albedo of an object. Without knowing the porosity of<br />

the surface, it is impossible to conclusively determine the<br />

surface assemblage of an object. M-asteroids with the<br />

highest radar albedos (0.6–0.7) could have surfaces of<br />

metallic iron with “lunar-like” porosities (35-55%) or solid<br />

enstatite-chondritic material with little to no porosity. It<br />

is unclear if such solid surfaces of enstatite chondrite material<br />

could exist on the surface of an asteroid. M asteroids<br />

tend to have higher radar albedos than C or S asteroids<br />

(Magri et al. 1999), apparently implying that they are richer<br />

in metallic iron.<br />

S asteroids tend to have 1 µm absorption features,<br />

which indicates assemblages containing Fe 2+ -bearing silicates<br />

such as olivine and/or pyroxene. On the basis of<br />

high-resolution CCD spectra, Bus and Binzel (2002b) subdivide<br />

the S asteroids into a number of subtypes (S, Sa, Sk,<br />

Sl, Sq, and Sr). The subscript represents that the objects is<br />

intermediate in spectral properties between that class (A,<br />

K, L, Q, and R) and the S class; however, it is currently<br />

unclear if each of these subclasses is grouping objects with<br />

similar compositions. From near-infrared spectra, Gaffey<br />

et al. (1993) finds that S asteroids tended to have compositions<br />

that ranged from olivine-rich (which he defined as<br />

S(I)) to pyroxene-rich (S(VII)). To classify these objects,<br />

they use the band area ratio (ratio of the area of Band I to<br />

the area of Band II) and the Band I minimum, which are<br />

both function of the olivine/pyroxene abundance (Cloutis<br />

et al., 1986). The band parameters of S(IV)-objects appear<br />

consistent with ordinary chondrites, but also consistent<br />

with other types of meteorites such as ureilites or acapulcoites/lodranites.<br />

S asteroids are the most abundant type of classified asteroid<br />

since they tend to be found in the inner main belt<br />

(closer to the Sun) and have higher albedos than C-types,<br />

making them brighter and easier to discover and observe.<br />

The biggest question concerning S asteroids is what fraction<br />

is compositionally similar to ordinary chondrites. S<br />

asteroids are spectrally redder than ordinary chondrites<br />

and tend to have weaker absorption bands (Figure 3). It<br />

has long been argued (e.g. Wetherill and Chapman 1988)<br />

whether these spectral differences are due to inherent compositional<br />

differences or simply due to an alteration<br />

process that can redden ordinary chondrite material. Analyses<br />

of lunar regolith (e.g. Pieters et al. 2000) and alteration<br />

experiments (e.g. Sasaki et al. 2001) appear to show<br />

that this reddening on asteroidal surfaces could be due to<br />

surface alteration processes (e.g. micrometeorite impacts,<br />

solar wind sputtering) that produce vapor-deposited coatings<br />

of nanophase iron.<br />

C-type asteroids (including the B, C, F, G, and P classes)<br />

tend to have relatively featureless spectra in low-resolution<br />

(photometric) surveys, which was consistent with<br />

carbonaceous meteorites. Higher-resolution spectral surveys<br />

(e.g. Bus and Binzel 2002b) have shown that almost<br />

half of observed C-type asteroids have a 0.7 µm feature.<br />

Observations (Jones et al. 1990) indicate that approximately<br />

two-thirds of all observed C-type asteroids have<br />

3 µm features.<br />

A-class asteroids tend to have strong UV and 1 µm features<br />

that appear similar to those of olivine. Near-infrared<br />

spectra (Figure 3) of these objects (Bell et al. 1988) confirm<br />

that these surfaces contain significant amounts of<br />

olivine as seen by the three distinctive olivine bands that<br />

make up the 1 µm feature. Two types of meteorites (brachinites<br />

and pallasites) have silicate mineralogies dominated<br />

by olivine and have been postulated to have<br />

compositions similar to the A asteroids.<br />

V- and J-type asteroids are the asteroids whose mineralogies<br />

appear to be the best-determined by remote sensing.<br />

These objects (including the 500-km diameter 4 Vesta<br />

and much smaller objects with diameters of 10 km or less)<br />

have visible and near-infrared spectra (Figure 3) (e.g.<br />

McCord et al. 1970, Binzel and Xu 1993, Burbine et al.<br />

2001b) similar to the HEDs, which have very distinctive<br />

spectral features due to pyroxene. The smaller V- and J-objects<br />

(called Vestoids) have been found in the Vesta family<br />

and between Vesta and the 3:1 and the ν 6 resonances. The<br />

presence of only one “large” body (4 Vesta) in the main<br />

belt with a spectrum similar to HEDs argues that Vesta is<br />

248


Asteroids: Their composition and impact threat<br />

the parent body of the HEDs (e.g. Consolmagno and<br />

Drake, 1977).<br />

E asteroids, due to their relatively featureless spectra and<br />

high albedos (greater than 0.3), have been interpreted as<br />

having surfaces composed predominately of an essentially<br />

iron-free silicate (e.g. Zellner et al. 1977). The most obvious<br />

meteoritic analog is the aubrites (enstatite achondrites),<br />

which are igneous meteorites composed primarily of essentially<br />

iron-free enstatite (Watters and Prinz 1979). However,<br />

the identification of an absorption feature in the 3 µm wavelength<br />

region of a number of main-belt E-asteroid spectra<br />

(Rivkin et al., 1995) has been interpreted as indicating hydrated<br />

minerals on the surfaces of some of these objects,<br />

which is inconsistent with an igneous origin. A feature centered<br />

at ~ 0.5 µm has also been identified in a number of E-<br />

class asteroids (Bus 1999, Fornasier and Lazzarin 2001) and<br />

these objects have been classified as Xe by Bus and Binzel<br />

(2002b) (Figure 2). This feature is believed to be due to a<br />

sulfide (Burbine et al. 2002a), which is commonly found in<br />

aubrites (Watters and Prinz 1979).<br />

The surface mineralogies of many other asteroid classes<br />

are thought to be somewhat understood. The K-class asteroids<br />

tend to have visible and near-infrared spectral<br />

properties similar to CO3/CV3 chondrites (Bell 1988,<br />

Burbine et al. 2001a, Burbine and Binzel 2002). Q asteroids<br />

(most notably 1862 Apollo) tend to have spectral<br />

properties similar to ordinary chondrites. D and P objects,<br />

which tend to be found in the outer main belt, are thought<br />

to have very primitive, organic-rich surfaces due to their<br />

relatively featureless and red spectra (e.g. Vilas and Gaffey,<br />

1989). R asteroids (most notably 349 Dembowska)<br />

appear to be mixtures of olivine and pyroxene (Gaffey et<br />

al., 1989). T asteroids tend to be very dark and featureless<br />

and Hiroi and Hasegawa (2002) have noted the spectral<br />

similarity of unusual carbonaceous chondrite Tagish Lake<br />

to T asteroids.<br />

The surface compositions of a few asteroid classes are<br />

unclear, but do appear to contain silicates. O-asteroid 3628<br />

Božněmcová has a 1 µm feature unlike any known meteorite<br />

(Burbine and Binzel 2002). L class objects have been<br />

newly defined by Bus and Binzel (2002b) and appear intermediate<br />

in spectral properties in the visible between the<br />

K and some S asteroids.<br />

Spacecraft missions<br />

The first dedicated asteroidal mission was the NEAR-<br />

Shoemaker spacecraft rendezvous with S-asteroid 433<br />

Eros (e.g. McCoy et al. 2002) and its flyby of C-asteroid<br />

253 Mathilde (e.g. Veverka et al. 1999). NEAR-Shoemaker<br />

Spacecraft obtained high-resolution images, reflectance<br />

spectra of different lithologic units, bulk density, magnetic<br />

field measurements, and bulk elemental compositions of<br />

433 Eros. Eros had a density of ~ 2.7 g/cm 3 , consistent<br />

with a silicate-rich assemblage while Mathilde had an extremely<br />

low density of 1.3 g/cm 3 .<br />

Eros was classified as an S(IV) object (Murchie and<br />

Normalized Reflectance<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

Bouvante (eucrite)<br />

1929 Kollaa (V)<br />

Ehole (H5)<br />

6 Hebe (S)<br />

289 Neneta (A)<br />

0.5<br />

0.4 0.8 1.2 1.6 2 2.4<br />

Wavelength (µm)<br />

Fig. 3. Reflectance spectra of S-asteroid 6 Hebe versus H5 chondrite<br />

Ehole, A-asteroid 289 Nenetta, and 1929 Kollaa versus eucrite Bouvante.<br />

The 6 Hebe and 289 Nenetaa spectra are a combination of data from<br />

Binzel and Bus (2002a) and Bell et al. (1988) while the 1929 Kollaa<br />

spectrum is a combination of data from Binzel and Bus (2002a) and Burbine<br />

and Binzel (2002). All spectra are normalized to unity at 0.55 µm<br />

and offset in reflectance from each other. Error bars are one sigma. All<br />

the Binzel and Bus (2002a) and the Burbine and Binzel (2002) spectra<br />

are available at the website http://smass.mit.edu. The Bouvante spectrum<br />

is from Burbine et al. (2001b) and was taken at Brown University’s<br />

RELAB facility.<br />

Pieters 1996) and it was hoped that the NEAR-Shoemaker<br />

data would help “solve” the S-asteroid/ordinary chondrite<br />

question. The average olivine to pyroxene<br />

composition derived from band area ratios (McFadden et<br />

al. 2001) and elemental ratios (Mg/Si, Fe/Si, Al/Si, and<br />

Ca/Si) derived from X-ray data (Nittler et al. 2001) of Eros<br />

are consistent (McCoy et al. 2001) with ordinary chondrite<br />

compositions. However, the S/Si ratio derived from X-ray<br />

data (Nittler et al. 2001) and the Fe/O and Fe/Si ratios derived<br />

from gamma-ray data (Evans et al. 2001) are significantly<br />

depleted relative to ordinary chondrites. McCoy et<br />

al. (2001) believe that the best meteoritic analogs to Eros<br />

are an ordinary chondrite, whose surface mineralogy has<br />

been altered by the depletion of metallic iron and sulfides,<br />

or a primitive achondrite, derived from a precursor assemblage<br />

of the same mineralogy as one of the ordinary chondrite<br />

groups.<br />

What is the percentage of near-Earth asteroids<br />

with iron meteorite compositions?<br />

What is extremely difficult from Earth to determine is<br />

if an object has a composition similar to iron meteorites.<br />

This is a problem since for a particular size, the biggest<br />

249


Thomas H. Burbine<br />

devastation among asteroid impacts will occur for these<br />

objects. Metallic iron has no distinctive absorption features<br />

and radar observations are often inconclusive since<br />

we do not know the surface porosity. There is some evidence<br />

that these objects do exist in the near-Earth asteroid<br />

population. M-type near-Earth asteroid (6178 1986 DA)<br />

(Tedesco and Gradie 1987), has one of the highest radar<br />

albedos (~ 0.6) of any observed asteroid, strongly implying<br />

a metallic iron surface (Ostro et al. 1991).<br />

Fall percentages give a probability of 4% of an iron<br />

meteorite being seen to land on the Earth and a 94% probability<br />

of seeing a stony or stony-iron meteorite fall. It is<br />

unclear if these fall statistics can be extrapolated to hundreds<br />

of meters to tens of kilometer-sized bodies in the<br />

near-earth population; however, compositions interpreted<br />

from the classifications derived from NEA spectral surveys<br />

roughly correspond to these percentages. Binzel et al.<br />

(2001a) have published the spectra and classifications of<br />

48 near-Earth asteroids. Approximately 90% of the objects<br />

(S, B, C, L, Q, V, O, K) had compositions consistent with<br />

silicate-dominated mineralogies if our compositional interpretations<br />

of these asteroid classes are correct. Approximately<br />

10% of the objects were classified as X types. As<br />

stated before, only some percentage of the X types are<br />

probably similar in composition to iron meteorites.<br />

This estimated percentage of iron projectiles is roughly<br />

consistent with numbers derived from impact craters.<br />

Koerbel (1998) lists 41 impact craters with diameters<br />

greater than 100 meters where the compositional characteristics<br />

of the impactor have been tentatively identified.<br />

He listed 14 (34%) as being due to iron meteorite projectiles<br />

with 3 impacts (7%) being due to either chondritic or<br />

iron-meteorite assemblages. The rest of the impacts appeared<br />

to be due either to chondritic, achondritic, stone, or<br />

stony-iron objects. However, the impacts due to iron meteorite<br />

assemblages are predominately associated with the<br />

smallest craters. Ten of the eleven craters smaller than 1.2<br />

kilometers in diameter are all thought to be due to iron meteorite<br />

projectiles. This preponderance of iron meteorite<br />

impacts among small craters is due to the atmospheric selection<br />

effect (e.g. Melosh 1989) that allows smaller iron<br />

meteorite projectiles (since they are denser compared to<br />

chondritic ones) to reach the ground and produce hypervelocity<br />

impacts. For the 30 craters larger than 2 km in diameter,<br />

~ 25% could be due to iron (or stony-iron)<br />

projectiles. However, only one of the eleven craters greater<br />

than 24 kilometers in diameter listed by Koeberl (1998) is<br />

believed to be due to an asteroid with a composition similar<br />

to iron meteorites.<br />

The meteorite fall statistics, classification of NEAs,<br />

and the characterization of the impactors that produced<br />

large craters all tend to argue that iron meteorite projectiles<br />

are a small percentage of the impacting population.<br />

All these lines of evidence argue that iron meteorite projectiles<br />

are a few percent and certainly less than 10% of<br />

near-Earth asteroids. Silicate-dominated assemblages are<br />

much more likely to strike the Earth.<br />

Is it important to know the composition<br />

of an impacting asteroid?<br />

To estimate the effects of an impact, the mass needs to<br />

be known. Since most meteorites that land on the Earth<br />

have bulk densities between 2 and 4 g/cm 3 , we can estimate<br />

pretty well an upper limit on an incoming object’s<br />

mass by using a density of 4 g/cm 3 and multiplying it by<br />

the object’s volume computed from its diameter. Also<br />

making this mass estimate an upper limit is the fact that asteroids<br />

may have significant macroporosity.<br />

But in a real-world situation where a sizable asteroid<br />

(e.g. a kilometer in diameter) that will be extremely destructive<br />

(~ 5,000 MT) is known to be a collision course<br />

with Earth, we will care more about diverting the object<br />

than knowing exactly how destructive it will be. Compositional<br />

information will be vital for determining how best to<br />

divert an object (e.g. Ahrens and Harris 1994). Deflection<br />

techniques such as impacting an asteroid with a spacecraft<br />

or a nuclear explosion will work best with knowledge of<br />

the surface composition. We need to know how the surface<br />

will be affected by an impact or blast and this information<br />

can only be derived from compositional studies. For example,<br />

an asteroid with a composition similar to a carbonaceous<br />

chondrite would be expected to fracture much<br />

more easily than an object with a composition similar to<br />

iron meteorites.<br />

Conclusions<br />

It is certain that the Earth will be hit in the future by an<br />

asteroid. The only question is “When?” For this impacting<br />

object, compositional studies will be vital for trying to determine<br />

how destructive the impact will be and for diverting<br />

the object.<br />

Acknowledgments. The author would like to thank Clark Chapman<br />

and Guy Consolmagno for very thoughtful reviews and the editorial work<br />

of Roman Skála. Almost all of the meteorite spectra in this paper were<br />

measured at Brown University’s Keck/NASA Reflectance Experiment<br />

Laboratory (RELAB), which is a multi-user facility supported by NASA<br />

grant NAG5-3871.<br />

References<br />

Adams J. B. (1975): Interpretation of visible and near-infrared diffuse reflectance<br />

spectra of pyroxenes and other rock-forming minerals. In:<br />

Karr C. III (ed.) Infrared and Raman Spectroscopy of Lunar and Terrestrial<br />

Minerals. Academic Press, Inc, New York, pp. 91–116.<br />

Adushkin V. V., Nemchinov I. V. (1994): Consequences of impacts of<br />

cosmic bodies on the surface of the Earth. In: Gehrels T. (ed.)<br />

Hazards due to Comets and Asteroids. University of Arizona Press,<br />

Tucson, pp. 721–778<br />

Ahrens T. J., Harris A. W. (1994): Deflection and fragmentation of near-<br />

Earth asteroids. In: Gehrels T. (ed.) Hazards due to Comets and Asteroids.<br />

University of Arizona Press, Tucson, pp. 897–927.<br />

Alvarez L., Alvarez W., Asaro F., Michel H. V. (1980): Extraterrestrial cause<br />

for the Cretaceous-Tertiary extinction. Science 208, 1095–1108.<br />

Bell J. F. (1988): A probable asteroidal parent body for the CV or CO<br />

chondrites (abstract). Meteoritics 23, 256–257.<br />

250


Asteroids: Their composition and impact threat<br />

Bell J. F., Owensby P. D., Hawke B R., Gaffey M. J. (1988): The 52-color<br />

asteroid survey: Final results and interpretation (abstract). Lunar<br />

Planet. Sci. XIX, 57–58.<br />

Binzel R. P., Xu S. (1993): Chips off of asteroid 4 Vesta: Evidence for the<br />

parent body of basaltic achondrite meteorites. Science 260, 186–191.<br />

Binzel, R. P., Harris A. W., Bus S. J., Burbine T. H. (2001a): Spectral<br />

properties of near-Earth objects: Palomar and IRTF results for 48 objects<br />

including spacecraft targets (9969) Braille and (10302) 1989<br />

ML. Icarus 151, 139–149.<br />

Binzel R. P., Rivkin A. S., Bus S. J., Sunshine J. M., Burbine T. H.<br />

(2001b): MUSES-C target asteroid 1998 SF36: A reddened ordinary<br />

chondrite. Meteoritics & Planet. Sci. 36, 1167–1172.<br />

Brearley A. J., Jones R. H. (1998): Chondritic meteorites. In: Papike J. J.<br />

(ed.) Reviews in Mineralogy, Mineralogical Society of America,<br />

Washington, pp. 3-1 to 3-398.<br />

Britt D. T., Consolmagno, G. J. (2001): Modeling the structure of high<br />

porosity asteroids. Icarus 152, 134–139.<br />

Britt D. T., Yeoman D., Housen K., Consolmagno G. (2002): Asteroid<br />

density, porosity, and structure. In: Bottke W. F. Jr. et al. (eds) Asteroids<br />

III. University of Arizona Press, Tucson, pp. 485–500.<br />

Burbine T. H. (1998): Could G-class asteroids be the parent bodies of the<br />

CM chondrites? Meteoritics & Planet. Sci. 33, 253–258.<br />

Burbine T. H., Binzel R. P. (2002): Small main-belt asteroid spectroscopic<br />

survey in the near-infrared. Icarus 158, 468–499.<br />

Burbine T. H., Binzel R. P., Bus S. J., Clark B. E. (2001a): K asteroids<br />

and CO3/CV3 chondrites. Meteoritics & Planet. Sci. 36, 245–253.<br />

Burbine T. H., Buchanan P. C., Binzel R. P., Bus S. J., Hiroi T., Hinrichs<br />

J. L., Meibom A., McCoy T. J. (2001b): Vesta, Vestoids, and the<br />

HEDs: Relationships and the origin of spectral differences. Meteoritics<br />

& Planet. Sci. 36, 761–781.<br />

Burbine T. H., McCoy T. J., Binzel R. P. (2001c): Spectra of angrites and<br />

possible parent bodies (abstract). Lunar and Planetary Science<br />

XXXIII, Abstract #1857, Lunar and Planet. Inst., Houston, Texas,<br />

USA (CD-ROM).<br />

Burbine T. H., Meibom A., Binzel R. P. (1996): Mantle material in the<br />

main belt: Battered to bits? Meteoritics & Planet. Sci. 31, 607–620.<br />

Burbine T. H., McCoy T. J., Nittler L. R., Benedix G. K., Cloutis E. A.,<br />

Dickinson T. L. (2002a): Spectra of extremely reduced assemblages:<br />

Implications for Mercury. Meteoritics & Planet. Sci. 37, 1233–1244.<br />

Burbine T. H., McCoy T. J., Meibom A., Gladman B., Keil K. (2002b):<br />

Meteoritic parent bodies: Their number and identification. In: Bottke<br />

W. F. Jr. et al. (eds) Asteroids III. University of Arizona Press, Tucson,<br />

pp. 653–667.<br />

Bus S. J. (1999): Compositional structure in the asteroid belt: Results of<br />

a spectroscopic survey. Ph.D. Thesis, Massachusetts Institute of<br />

Technology, Cambridge.<br />

Bus S. J., Binzel R. P. (2002a): Phase I of the Small Main-belt Asteroid<br />

Spectroscopic Survey: The observations. Icarus 158, 106–145.<br />

Bus S. J., Binzel R. P. (2002b): Phase II of the Small Main-Belt Asteroid<br />

Spectroscopic Survey: A feature-based taxonomy. Icarus 158,<br />

146–177.<br />

Chapman C. R., Morrison D. (1994): Impacts on the Earth by asteroids<br />

and comets: Assessing the hazard. Nature 367, 33–40.<br />

Chyba C. F., Thomas P. J., Zahnle K. J. (1993): The 1908 Tunguska explosion:<br />

Atmospheric disruption of a stony asteroid. Nature 361,<br />

40–44.<br />

Clark B. E. (1995): Spectral mixing models of S-type asteroids. J. Geophys.<br />

Res. 100, 14443–14456.<br />

Cloutis E. A., Gaffey M. J. (1991): Pyroxene spectroscopy revisited:<br />

Spectral-compositional correlations and relationship to geothermometry.<br />

J. Geophys. Res. 96, 22809–22826.<br />

Cloutis, E. A., Gaffey M. J., Jackowski, T. L., Reed K. L. (1986): Calibrations<br />

of phase abundance, composition, and particle size distribution<br />

for olivine-orthopyroxene mixtures from reflectance spectra. J.<br />

Geophys. Res. 91, 11641–11653.<br />

Cloutis E. A., Gaffey M. J., Smith D. G. W., Lambert R. St. J. (1990): Reflectance<br />

spectra of “featureless” materials and the surface mineralogies<br />

of M- and E-class asteroids. J. Geophys. Res. 95, 281–293.<br />

Consolmagno G. J., Britt D. T. (1998): The density and porosity of meteorites<br />

from the Vatican collection. Meteoritics & Planet. Sci.33,<br />

1231–1241.<br />

Consolmagno G. J., Drake M. J. (1977): Composition and evolution of<br />

the eucrite parent body: Evidence from rare earth elements.<br />

Geochim. Cosmochim. Acta 41, 1271–1282.<br />

Earth Impact Database, 2002. <br />

(Accessed: 16 Sept. 2002).<br />

Evans L. G., Starr R. D., Brückner J., Reedy R. C., Boynton W. V., Trombka<br />

J. I., Goldsten J. O., Masarik J., Nittler L. R., McCoy T. J.<br />

(2001): Elemental composition from gamma-ray spectroscopy of the<br />

NEAR-Shoemaker landing site on 433 Eros. Meteoritics & Planet.<br />

Sci. 36, 1639–1660.<br />

Flynn G. J., Moore L. B., Klöck W.(1999): Density and porosity of stone<br />

meteorites: Implications for the density, porosity, cratering, and collisional<br />

disruption of asteroids. Icarus 142, 97–105.<br />

Fornasier S., Lazzarin M. (2001): E-type asteroids: Spectroscopic investigation<br />

on the 0.5 µm absorption band. Icarus 152, 127–133.<br />

Gaffey M. J. (1976): Spectral reflectance characteristics of the meteorite<br />

classes. J. Geophys. Res. 81, 905–920.<br />

Gaffey M. J., Bell J. F., Cruikshank D. P. (1989): Reflectance spectroscopy<br />

and asteroid surface mineralogy. In: Binzel R. P. et al. (eds)<br />

Asteroids II. University of Arizona Press, Tucson, pp. 98–127.<br />

Gaffey M. J., Bell J. F., Brown R. H., Burbine T. H., Piatek J. L., Reed<br />

K. L., Chaky D. A. (1993): Mineralogical variations within the S-<br />

type asteroid class. Icarus 106, 573–602.<br />

Garshnek V., Morrison D., Burkle F. M. Jr. (2000): The mitigation, management,<br />

and survivability of asteroid/comet impact with Earth.<br />

Space Policy 16, 213–222.<br />

Grady M. M. (2000): Catalogue of Meteorites. Cambridge University<br />

Press, Cambridge.<br />

Grossman J. N. (2000): The Meteoritical Bulletin, No. 84, 2000 August.<br />

Meteoritics & Planet. Sci. 35, A199–A225.<br />

Grossman J. N., Zipfel J. (2001): The Meteoritical Bulletin, No. 85, 2001<br />

September. Meteoritics & Planet. Sci. 36, A293–A322.<br />

Hildebrand A. R., Pilkington M., Ortiz-Aleman C., Chavez R. E., Urrutia-Fucugauchi<br />

J., Connors M., Graniel-Castro E., Camara-Zi A.,<br />

Halpenny J. F., Niehaus D. (1998): Mapping Chicxulub crater structure<br />

with gravity and seismic reflection data. In: Grady M. M. et al.<br />

(eds): Meteorites: Flux with Time and Impact Effects. Geological<br />

Society, London, pp. 155–176.<br />

Hinrichs J. L., Lucey P. G., Robinson M. S., Meibom A., Krot A. N.<br />

(1999): Implications of temperature-dependent near-IR spectral<br />

properties of common minerals and meteorites for remote sensing of<br />

asteroids. Geophys. Res. Lett. 26, 1661–1664.<br />

Hiroi T., Hasegawa S. (2002): Revisiting the search for the parent body<br />

of the Tagish Lake meteorite – Case of a T/D asteroid 308 Polyxo<br />

(abstract). Antarctic Meteorites XXVII, 32–33.<br />

Hughes D. W. (1998): The mass distribution of crater-producing bodies.<br />

In: Grady M. M. et al. (eds) Meteorites: Flux with Time and Impact<br />

Effects. Geological Society, London, pp. 31–42.<br />

Jones T. D., Lebofsky L. A., Lewis J. S., Marley M. S. (1990): The composition<br />

and origin of the C, P, and D asteroids: Water as a tracer of<br />

thermal evolution in the outer belt. Icarus 88,172–192.<br />

Koeberl C. (1998): Identification of meteoritic components in impactites.<br />

In: Grady M. M. et al. (eds) Meteorites: Flux with Time and Impact<br />

Effects. Geological Society, London, pp. 133–153.<br />

Lewis J. S. (2000): Comet and asteroid impact hazards on a populated<br />

Earth: Computer modeling. Academic Press, San Diego.<br />

Magri C., Ostro S. J., Rosema K. D., Thomas M. L., Mitchell D. L., Campbell<br />

D. B., Chandler J. F., Shapiro I. I., Giorgini J. D., Yeomans D. K.<br />

(1999): Mainbelt asteroids: Results of Arecibo and Goldstone radar observations<br />

of 37 objects during 1980–1985. Icarus 140, 379–407.<br />

Margot J. L., Nolan M. C., Benner L. A. M., Ostro S. J., Jurgens R. F.,<br />

Giorgini J. D., Slade M. A., Campbell D. B. (2002): Binary asteroids<br />

in the near-Earth population. Science 296, 1445–1448.<br />

Marsden B. G., Steel D. I. (1994): Warning times and impact probabilities<br />

for long-period comets. In: Gehrels T. (ed.) Hazards due to<br />

Comets and Asteroids. University of Arizona Press, Tucson, pp.<br />

221–239.<br />

Marti K., Graf T. (1992): Cosmic-ray exposure history of ordinary chondrites.<br />

Ann. Rev. Earth Planet. Sci. 20, 221–243.<br />

McCord T. B., Adams J. B., Johnson T. V. (1970): Asteroid Vesta: Spectral<br />

reflectivity and compositional implications. Science 168,<br />

1445–1447.<br />

McCoy T. J., Burbine T. H., McFadden L. A., Starr R. D., Gaffey M. J.,<br />

251


Thomas H. Burbine<br />

Nittler L. R., Evans L. G., Izenberg N., Lucey P. G., Trombka J. I., Bell<br />

J. F. III, Clark B. E., Clark P. E., Squyres S. W., Chapman C. R., Boynton<br />

W. V., Veverka J. (2001): The composition of 433 Eros: A mineralogical-chemical<br />

synthesis. Meteoritics & Planet. Sci. 36, 1661–1672.<br />

McCoy T. J., Robinson M. S., Nittler L. R., Burbine T. H. (2002): The<br />

Near Earth Asteroid Rendezvous mission to asteroid 433 Eros: A<br />

milestone in the study of asteroids and their relationship to meteorites.<br />

Chemie der Erde 62, 89–121.<br />

McFadden L. A., Wellnitz D. D., Schnaubelt M., Gaffey M. J., Bell J. F. III,<br />

Izenberg N., Chapman C. R., Murchie S. (2001): Mineralogical interpretation<br />

of reflectance spectra of Eros from NEAR near-infrared spectrometer<br />

low phase flyby. Meteoritics & Planet. Sci. 36, 1711–1726.<br />

Meibom A., Clark B. E. (1999): Evidence for the insignificance of ordinary<br />

chondritic material in the asteroid belt. Meteoritics & Planet.<br />

Sci. 34, 7–24.<br />

Melosh H. J. (1989): Impact Cratering: A Geologic Process. Oxford<br />

University Press, New York.<br />

Melosh H. J., Tonks W. B. (1993): Swapping rocks: Ejection and exchange<br />

of surface material among the terrestrial planets (abstract).<br />

Meteoritics & Planet. Sci. 28, 398.<br />

Milner A. C. (1998): Timing and causes of vertebrate extinction across<br />

the Cretaceous-Tertiary boundary. In: Grady M. M. et al. (eds) Meteorites:<br />

Flux with Time and Impact Effects. Geological Society,<br />

London, pp. 247–257.<br />

Mittlefehldt D. W. (2002): Geochemistry of the ungrouped carbonaceous<br />

chondrite Tagish Lake, the anomalous CM chondrite Bells, and comparison<br />

with CI and CM chondrites. Meteoritics & Planet. Sci. 37,<br />

703–712.<br />

Mittlefehldt D. W., McCoy T. J., Goodrich C. A., Kracher A. (1998):<br />

Non-chondritic meteorites from asteroidal bodies. In: Papike J. J.<br />

(ed.) Reviews in Mineralogy, Mineralogical Society of America,<br />

Washington, pp. 4-1 to 4-195.<br />

Morbidelli A., Jedicke R., Bottke W. F., Michel P., Tedesco E. F. (2002):<br />

From magnitudes to diameters: The albedo distribution of near Earth<br />

objects and the Earth collision hazard. Icarus 158, 329–342.<br />

Morgan J., Warner M., Grieve R. (2002): Geophysical constraints on the<br />

size and structure of the Chicxulub impact crater. In: Grady M. M. et<br />

al. (eds.) Meteorites: Flux with Time and Impact Effects. Geological<br />

Society, London, pp. 39–46.<br />

Murchie S. L., Pieters C. M. (1996): Spectral properties and rotational spectral<br />

heterogeneity of 433 Eros. J. Geophys. Res. 101, 2201–2214.<br />

Neukum G., Ivanov B. A. (1994): Crater size distributions and impact<br />

probabilities on Earth from lunar, terrestrial-planet, and asteroid cratering<br />

data. In: Gehrels T. (ed.) Hazards due to Comets and Asteroids.<br />

University of Arizona Press, Tucson, pp. 359–416.<br />

Nittler L. R., Starr R. D., Lim L., McCoy T. J., Burbine T. H., Reedy R.<br />

C., Trombka J. I., Gorenstein P., Squyres S. W., Boynton W. V.,<br />

McClanahan T. P., Bhangoo J. S., Clark P. E., Murphy M. E., Killen<br />

R. (2001): X-ray fluorescence measurements of the surface elemental<br />

composition of asteroid 433 Eros. Meteoritics & Planet. Sci. 36,<br />

1673–1695.<br />

Ostro S. J., Campbell D. B., Chandler J. F., Hine A. A., Hudson R . S.,<br />

Rosema K. D., Shapiro I. I. (1991): Asteroid 1986 DA: Radar evidence<br />

for a metallic composition. Science 252, 1399–1404.<br />

Ostro S. J., Hudson R. S., Benner L. A. M., Giorgini J. D., Magri C., Margot<br />

J. L., Nolan M. (2002): Asteroid radar astronomy. In: Bottke W.<br />

F. Jr. et al. (eds) Asteroids III. University of Arizona Press, Tucson,<br />

pp. 151–168.<br />

Paine M. P. (1999): Asteroid impacts: The extra hazard due to tsunami.<br />

Science of Tsunami Hazards 17, 155–172.<br />

Pieters C. M., McFadden L. A. (1994): Meteorite and asteroid reflectance<br />

spectroscopy: Clues to the early solar system processes. Ann. Rev.<br />

Earth Planet. Sci. 22, 457–497.<br />

Pieters C. M., Taylor L. A., Noble S. K., Keller L. P., Hapke B., Morris<br />

R. V., Allen C. C., McKay D. S., Wentworth S. (2000): Space weathering<br />

on airless bodies: Resolving a mystery with lunar samples. Meteoritics<br />

& Planet. Sci. 35, 1101–1107.<br />

Rivkin A. S., Lebofsky L. A., Clark B. E., Howell E. S., Britt D. T.<br />

(2000): The nature of M-class asteroids from 3-µm observations.<br />

Icarus 145, 351–368.<br />

Sasaki S., Nakamura K. Hamabe Y., Kurahashi E., Hiroi T. (2001): Production<br />

of iron nanoparticles by laser irradiation in a simulation of<br />

lunar-like space weathering. Nature 410, 555–557.<br />

Sato K., Miyamoto M., Zolensky M. E. (1997): Absorption bands near three<br />

micrometers in diffuse reflectance spectra of carbonaceous chondrites:<br />

Comparison with asteroids. Meteoritics & Planet. Sci. 32, 503–507.<br />

Scherer P., Schultz L. (2000): Noble gas record, collisional history, and<br />

pairing of CV, CO, CK, and other carbonaceous chondrites. Meteoritics<br />

& Planet. Sci. 35, 145–153.<br />

Scott E. R. D. (2001): Formation of mesosiderites by fragmentation and<br />

reaccretion of a large differentiated asteroid. Meteoritics & Planet.<br />

Sci. 36, 869–881.<br />

Sears D. W. G. (1998): The case for rarity of chondrules and CAI in the<br />

early solar system and some implications for astrophysical models.<br />

Astrophys. J. 498, 773–778.<br />

Singer R. B., Roush T. L. (1985): Effects of temperature on remotely sensed<br />

mineral absorption features. J. Geophys. Res. 90, 12434–12444.<br />

Stuart J. S. (2001): Near-Earth asteroid population from the LINEAR<br />

survey. Science 294, 1691–1693.<br />

Tedesco E. F., Gradie J. (1987): Discovery of M class objects among the<br />

near-Earth asteroid population. Astron. J. 93, 738–746.<br />

Tholen D. J. (1984): Asteroid taxonomy from cluster analysis of photometry.<br />

Ph.D. Thesis, Univ. Arizona, Tucson.<br />

Toon O. B., Zahnle K., Morrison D., Turco R. P., Covey C. (1997): Environmental<br />

perturbations caused by the impacts of asteroids and<br />

comets. Rev. Geophysics 35, 41–78.<br />

Vasilyev N. V. (1998): The Tunguska Meteorite problem today. Planetary<br />

& Space Sciences 46, 129–150.<br />

Veverka J, Thomas P., Harch A., Clark B., Bell J. F. III., Carcich B.,<br />

Joseph, J., Murchie S., Izenberg N., Chapman C., Merline W., Malin<br />

M., McFadden L., Robinson M. (1999): NEAR encounter with asteroid<br />

253 Mathilde: Overview. Icarus 140, 3–16.<br />

Vilas F., Gaffey M. J. (1989): Phyllosilicate absorption features in mainbelt<br />

and outer-belt asteroid reflectance spectra. Science 246,<br />

790–792.<br />

Voshage H., Feldman H. (1979): Investigations on cosmic-ray produced<br />

nuclides in iron meteorites, 3. Exposure ages, meteoroid sizes and<br />

sample depths determined by mass spectrometric analyses of potassium<br />

and rare gases. Earth Planet. Sci. Lett. 45, 293–308.<br />

Wasson J. T. (1995): Sampling the asteroid belt: How biases make it difficult<br />

to establish meteorite-asteroid connections (abstract). Meteoritics<br />

30, 595.<br />

Watters T. R., Prinz M. (1979): Aubrites: Their origin and relationship to<br />

enstatite chondrites. Proc. Lunar Planet. Sci. Conf. 10th, 1073–1093.<br />

Weisberg M. K., Prinz M., Clayton R. N., Mayeda T. K., Sugiura N.,<br />

Zashu S., Ebihara M. (2001): A new metal-rich chondrite grouplet.<br />

Meteoritics & Planet. Sci. 36, 401–418.<br />

Wetherill G. W., Chapman C. R. (1988): Asteroids and meteorites. In:<br />

Kerridge J. F., Matthews M. S. (eds) Meteorites and the Early Solar<br />

System. University of Arizona Press, Tucson, pp. 35–67.<br />

Wilkison S. L., Robinson M. S. (2000): Bulk density of ordinary chondrite<br />

meteorites and implications for asteroidal internal structure.<br />

Meteoritics & Planet. Sci. 35, 1203–1213.<br />

Yeomans D. K., Antreasian P. G., Barriot J. -P., Chesley S. R., Dunham<br />

D. W., Farquhar R. W., Giorgini J. D., Helfrich C. E., Konopliv A. S.,<br />

McAdams J. V., Miller J. K., Owen W. M Jr., Scheeres D. J., Thomas<br />

P. C., Veverka J., Williams B. G. (2000): Radio science results during<br />

the NEAR-Shoemaker rendezvous with Eros. Science 289,<br />

2085–2088.<br />

Zellner B., Leake M., Morrison D., Williams J. G. (1977): The E asteroids<br />

and the origin of the enstatite achondrites. Geochim. Cosmochim.<br />

Acta 41, 1759–1767.<br />

Zellner B., Tholen D. J., Tedesco E. F. (1985): The eight-color asteroid<br />

survey: Results for 589 minor planets. Icarus 61, 355–416.<br />

Handling editor: Roman Skála<br />

252


Bulletin of the Czech Geological Survey, Vol. 77, No. 4, 253–263, 2002<br />

© Czech Geological Survey, ISSN 1210-3527<br />

The recognition of terrestrial impact structures<br />

ANN M. THERRIAULT – RICHARD A. F. GRIEVE – MARK PILKINGTON<br />

Natural Resources Canada, Booth Street, Ottawa, Ontario, KIA 0ES Canada; e-mail: ATherria@NRCan.gc.ca<br />

Abstract. The Earth is the most endogenically active of the terrestrial planets and, thus, has retained the poorest sample of impacts that have<br />

occurred throughout geological time. The current known sample consists of approximately 160 impact structures or crater fields. Approximately 30%<br />

of known impact structures are buried and were initially detected as geophysical anomalies and subsequently drilled to provide geologic samples. The<br />

recognition of terrestrial impact structures may, or may not, come from the discovery of an anomalous quasi-circular topographic, geologic or geophysical<br />

feature. In the geologically active terrestrial environment, anomalous quasi-circular features, however, do not automatically equate with an<br />

impact origin. Specific samples must be acquired and the occurrence of shock metamorphism, or, in the case of small craters, meteoritic fragments,<br />

must be demonstrated before an impact origin can be confirmed. Shock metamorphism is defined by a progressive destruction of the original rock and<br />

mineral structure with increasing shock pressure. Peak shock pressures and temperatures produced by an impact event may reach several hundreds of<br />

gigaPascals and several thousand degrees Kelvin, which are far outside the range of endogenic metamorphism. In addition, the application of shockwave<br />

pressures is both sudden and brief. Shock metamorphic effects result from high strain rates, well above the rates of normal tectonic processes.<br />

The well-characterized and documented shock effects in quartz are unequivocal indicators and are the most frequently used indicator for terrestrial impact<br />

structures and lithologies.<br />

Key words: Earth, impact structures, shock metamorphism, melting, glasses, shatter cones, geophysical anomaly<br />

Introduction<br />

On Earth, compared to the other terrestrial planets, the<br />

very active geologic environment tends to modify and destroy<br />

the impact crater record. Approximately 160 impact<br />

structures or crater fields are currently known on Earth.<br />

Impact involves the transfer of massive amounts of energy<br />

to a spatially limited area of the Earth’s surface, in an extremely<br />

short time interval. As a consequence, local geology<br />

of the target area is of secondary importance. The<br />

effects of impact are, however, scale-dependent and show<br />

progressive changes with increasing energy of the impact<br />

event. The net result is that, impacts of similar scale produce<br />

similar first-order geological and geophysical effects.<br />

Thus, general observations can be derived with respect to<br />

the appearance and geological and geophysical signatures<br />

of terrestrial impact structures, in specific size ranges.<br />

Terrestrial impact structures were first recognized by<br />

their bowl-like shape and meteorite fragments found in their<br />

vicinity or within them (the classic example being Meteor<br />

or Barringer Crater, Arizona). In the 1960’s, petrographic<br />

studies of rocks from impact structures defined a series of<br />

unique characteristics produced by a style of deformation<br />

called shock metamorphism (e.g. French and Short 1968).<br />

Shock metamorphic effects include shatter cones (e.g. Dietz<br />

1947, Milton 1977), the only macroscopic diagnostic shock<br />

effect observed at terrestrial impact structures, a number of<br />

microscopic effects in minerals, some of which are diagnostic<br />

of shock, and impact melting.<br />

The aim of this paper is to summarize the morphology<br />

and geoscientific aspects of terrestrial impact structures<br />

and provide a general description of shock-metamorphic<br />

effects.<br />

Morphology<br />

On most planetary bodies, well-preserved impact<br />

structures are recognized by their characteristic morphology<br />

and morphometry. The basic shape of an impact structure<br />

is a depression with an upraised rim. Detailed<br />

appearance, however, varies with crater diameter. With increasing<br />

diameter, impact structures become proportionately<br />

shallower and develop more complicated rims and<br />

floors, including the appearance of central peaks and interior<br />

rings. Impact craters are divided into three basic morphologic<br />

subdivisions: simple craters, complex craters,<br />

and basins (Dence 1972, Wood and Head 1976).<br />

Small impact structures have the form of a bowl-shaped<br />

depression with an upraised rim and are known as simple<br />

craters (Fig. 1). The exposed rim, walls, and floor define the<br />

so-called apparent crater. At the rim, there is an overturned<br />

flap of ejected target materials, which displays inverted<br />

stratigraphy, with respect to the original target materials.<br />

Beneath the floor is a lens of brecciated target material that<br />

is roughly parabolic in cross-section (Fig. 2). This breccia<br />

lens is allochthonous and polymict, with fractured blocks of<br />

various target materials. In places, near the top and the base,<br />

the breccia lens may contain highly shocked, and possibly<br />

melted, target materials. Beneath the breccia lens, parautochthonous,<br />

fractured rocks define the walls and floor<br />

of what is known as the true crater. In the case of terrestrial<br />

simple craters, the depth to the base of the breccia lens (i.e.,<br />

the base of the true crater) is roughly twice that of the depth<br />

to the top of the breccia lens (i.e., the base of the apparent<br />

crater, Fig. 2). Shocked rocks in the parautochthonous materials<br />

of the true crater floor are confined to a small central<br />

volume at the base of the true crater.<br />

253


Ann M. Therriault – Richard A. F. Grieve – Mark Pilkington<br />

a<br />

Figure 1. (a) Oblique aerial view of 1.2 km diameter, 50,000 years old simple crater, Meteor or Barringer Crater, Arizona, U.S.A. (b) Vertical aerial<br />

view of 3.8 km diameter, 450 ± 30 million years old, Brent Crater, Ontario, Canada. Note how this ancient crater has no rim, has been filled by sediments<br />

and lakes and is a generally subtle topographic feature.<br />

b<br />

With increasing diameter, simple craters show increasing<br />

evidence of wall and rim collapse and evolve into complex<br />

craters (Fig. 3). The transition diameter varies<br />

between planetary bodies and is, to a first approximation,<br />

an inverse function of planetary gravity (Pike 1980).<br />

Other variables, such as target material and possibly projectile<br />

type and velocity, play a lesser role, so that the transition<br />

diameter varies over a small range. The most<br />

obvious effect of secondary variables appears on Earth,<br />

where there are major areas of both sedimentary and crystalline<br />

rocks at the surface. Complex craters on Earth first<br />

occur at diameters greater than 2 km in layered sedimentary<br />

target rocks but not until diameters of 4 km or greater<br />

in stronger, more coherent, igneous or metamorphic, crystalline<br />

target rocks (Dence 1972).<br />

With a central topographic peak or peaks, a broad, flat<br />

floor, and terraced, inwardly slumped rim areas (Fig. 4),<br />

complex craters are a highly modified craterform compared<br />

to simple craters. The rim of a typical complex<br />

D<br />

<br />

yyyyyyyyy<br />

a D t<br />

<br />

yyyyyyyyy<br />

<br />

yyyyyyyyy Slump<br />

Autochthonous<br />

target rocks<br />

Simple crater – final form<br />

D<br />

Basal melt pool<br />

(minor fall back)<br />

breccia fill<br />

Figure 2. Schematic cross-section of a simple crater. D is the diameter<br />

and d a and d t are the depths of the apparent and true crater, respectively.<br />

See text for details.<br />

crater is a structural feature corresponding to a series of<br />

fault terraces. Interior to the rim lays an annular trough,<br />

which is partially filled by a sheet of impact-melt rock<br />

and/or polymict allochthonous breccia (Fig. 4). Only in<br />

the central area of the crater is there evidence of substantial<br />

excavation of target materials. This region is structurally<br />

complex and, in large part, occupied by a central<br />

peak, which is the topographic manifestation of a much<br />

broader and extensive area of uplifted rocks that occurs<br />

beneath the center of complex craters. Readers interested<br />

in the details of cratering mechanics at simple and complex<br />

structures are referred to Melosh (1989) and references<br />

therein.<br />

With increasing diameter, a fragmentary ring of interior<br />

peaks appears, marking the transition from craters to<br />

basins. While a single interior ring is required to define a<br />

basin, basins have been subdivided, with increasing diameter,<br />

on other planetary bodies, into central-peak basins,<br />

with both a peak and ring; peak ring basins, with only a<br />

ring; and multi-ring basins, with two or more interior rings<br />

(Wood and Head 1976). There have been claims that the<br />

largest known terrestrial impact structures have multi-ring<br />

forms, e.g. Chicxulub, Mexico (Sharpton et al. 1993),<br />

Sudbury, Canada (Stöffler et al. 1994, Spray and Thompson<br />

1995) and Vredefort, South Africa (Therriault et al.<br />

1997). Although certain of their geological and geophysical<br />

attributes form annuli, it is not clear that these correspond,<br />

or are related in origin, to the obvious<br />

topographical rings observed, for example, in lunar multiring<br />

basins (Spudis 1993, Grieve and Therriault 2000).<br />

Most terrestrial impact structures are affected by erosion.<br />

In extreme cases, the craterform has been completely<br />

removed. In such cases, recognition of structural and<br />

254


The recognition of terrestrial impact structures<br />

a<br />

Figure 3. (a) Oblique aerial photograph of the Gosses Bluff impact structure, Australia.<br />

Note that all that is visible of this originally 22 km, 142.5 ± 0.8 million<br />

years old structure is a 5 km annulus of hills, representing the eroded remains of<br />

a central uplift. See text for details. (b) Shuttle photograph of the Manicouagan<br />

impact structure, Canada, 100 km in diameter and 214 ± 1 million years old. Note<br />

that the annular trough (with a diameter of ~ 65 km) is filled by water.<br />

b<br />

geological effects of impact in the target rocks is essential<br />

to the identification of an impact structure rather than the<br />

presence of a characteristic craterform. For example,<br />

Gosses Bluff, Australia has a positive topographical form<br />

consisting of an annular ring of hills, approximately 5 km<br />

in diameter (Fig. 3). The ring consists of erosionally resistant<br />

beds from within the original central uplifted area<br />

of a complex impact structure. The original craterform,<br />

which has an estimated diameter of approximately 22 km<br />

(Milton et al. 1996), has been removed by erosion. There<br />

are several other impact structures, which have some form<br />

of rings, e.g. Manicouagan, Canada (Floran and Dence<br />

1976), Haughton, Canada (Robertson and Sweeney 1983),<br />

but it is not clear whether these are primary forms or secondary<br />

features, with some relation to primary structural<br />

features (Grieve and Head 1983).<br />

There are also other subtleties to the character of<br />

craterforms in the terrestrial record that do not appear on<br />

the other terrestrial planets. A number of relatively young,<br />

and, therefore, only slightly eroded, complex impact structures<br />

(e.g. Haughton, Canada; Ries Germany; Zhamanshin,<br />

Kazakhstan) do not have an emergent central peak or<br />

other interior topographical expression of a central uplift<br />

(Garvin and Schnetzler 1994). These structures are in<br />

mixed targets of platform sediments overlying crystalline<br />

basement. Although there are no known comparably<br />

young complex structures entirely in crystalline targets,<br />

the buried and well-preserved Boltysh structure, Ukraine,<br />

which is of comparable size, has a central peak (emergent<br />

from the crater-fill), similar to the appearance of lunar central<br />

peak craters. This difference in form is probably a target<br />

rock effect but it has not been studied in detail.<br />

The morphology of impact craters formed in marine<br />

environment is also quite distinct. These impact structures<br />

are characterized by a broad and shallow brim at the periphery<br />

of the crater, extensive infilling, and prominent<br />

fault blocks floored by apparent low-angle décollement<br />

surfaces at the periphery of the crater (e.g. Tsikalas et al.<br />

1999, Ormö and Lindström 2000). The extensive infilling<br />

is most likely due to large amounts of ejecta and crater<br />

wall material transported into the excavated crater by the<br />

collapse of the impact-induced water cavity and the subsequent<br />

rapid surge of sea water (Tsikalas et al. 1999, Ormö<br />

and Lindström 2000). The 40-km-diameter Mjølnir submarine<br />

impact structure in the Barents Sea, for example,<br />

consists of a central region of deep excavation surrounded<br />

by a shallow excavated shelf, without a raised crater rim<br />

(Tsikalas et al. 1998, 1999). This morphology is also observed<br />

at the 13.5-km-diameter Lockne impact structure,<br />

Sweden (Lindström et al. 1996).<br />

Attempts to define morphometric relations, particularly<br />

depth-diameter relations, for terrestrial impact structures<br />

have had limited success, because of the effects of<br />

erosion and, to a lesser degree, post-impact sedimentation.<br />

Unlike depth, the variation of stratigraphic uplift (SU, Fig.<br />

4) with diameter at complex impact structures is fairly<br />

d t<br />

d a<br />

Complex structure – final form<br />

Central Uplift Area<br />

Dcp<br />

D<br />

Melt/allochthonous<br />

breccia sheet<br />

"Autochthonous"<br />

crater floor<br />

SU<br />

Figure 4. Schematic cross-section of complex impact structure. Notation<br />

as in Figure 2, with SU corresponding to structural uplift and D cp to the<br />

diameter of the central uplift. Note preservation of beds in outer annular<br />

trough of the structure, with excavation limited to the central area. See<br />

text for details.<br />

255


Ann M. Therriault – Richard A. F. Grieve – Mark Pilkington<br />

well constrained with SU = 0.86D 1.03 (n = 24), where n is<br />

the number of data points (Grieve and Pilkington 1996).<br />

Similarly, the diameter of the central uplift area (D cp , Fig.<br />

4), at its maximum radial expression, is constrained by D cp<br />

= 0.31 D 1.02 (n = 44) (Therriault et al. 1997).<br />

Geophysics of impact structures<br />

Geophysical anomalies over terrestrial impact structures<br />

vary in their character and, in isolation, do not provide<br />

definitive evidence for an impact origin. About 30 per<br />

cent of known terrestrial impact structures are buried by<br />

post-impact sediments. Geophysical methods resulted in<br />

their initial discovery and subsequent drilling provided<br />

geologic samples, which confirmed their impact origin.<br />

Interpretation of a single geophysical data set over a suspected<br />

impact structure can be ambiguous (for example,<br />

Hildebrand et al. 1998, Sharpton et al. 1993). When combined,<br />

however, with complementary geophysical methods<br />

and the existing database over other known impact structures,<br />

a more definite assessment can be made (e.g. Ormö<br />

et al. 1999).<br />

Since potential-field data are available over large areas,<br />

with almost continuous coverage, gravity and magnetic<br />

observations have been the primary geophysical indicators<br />

used for evaluating the occurrence of possible terrestrial<br />

impact structures. Seismic data, although providing much<br />

better spatial resolution of subsurface structure, is used<br />

less, because it is less generally available. Electrical methods<br />

have been used even less (e.g. Henkel 1992). Given<br />

space limitations and some lack of specificity of the geophysical<br />

attributes of terrestrial impact craters, they are<br />

generally discussed here and the reader is referred to the<br />

most recent synthesis in Grieve and Pilkington (1996).<br />

Gravity signature<br />

The most notable geophysical signature associated<br />

with terrestrial impact structures is a negative gravity<br />

anomaly. These gravity lows are generally circular, extending<br />

to, or slightly beyond, the crater rim, and are due<br />

to lithological and physical changes associated with the<br />

impact process. In well-preserved impact structures, lowdensity<br />

sedimentary infill of the topographic depression of<br />

the crater contributes to the gravity low. In complex impact<br />

structures, relatively lower density impact-melt sheets<br />

also contribute to the negative gravity effect. However,<br />

such lithological effects are minor compared to density<br />

contrasts induced by fracturing and brecciation of the target<br />

rocks.<br />

In general, the amplitude of the maximum negative<br />

gravity anomaly associated with impact structures increases<br />

with the final crater diameter (Dabizha and Fedynsky<br />

1975, Dabizha and Feldman 1982). Over simple craters, a<br />

circular bowl-shaped negative anomaly is observed;<br />

whereas most of larger complex impact structures, greater<br />

than 30 km in diameter, tend to exhibit a central gravity<br />

high. Based on data from 58 terrestrial impact structures,<br />

Pilkington and Grieve (1992) showed that erosional level<br />

has only a secondary effect on gravity anomaly size.<br />

It is important to note that due to differences in target<br />

lithologies, large variations in gravity signature are observed<br />

between structures of similar sizes. In general,<br />

structures formed in sedimentary lithologies produce<br />

smaller anomalies than similar sized ones formed in crystalline<br />

rocks. Structures formed in unconsolidated sediments<br />

in continental shelf areas may not produce<br />

detectable negative gravity anomalies but are marked only<br />

by a central gravity high.<br />

Magnetic signature<br />

In general, magnetic anomalies associated with terrestrial<br />

impact structures are more complex than gravity<br />

anomalies. This observation reflects the greater variation<br />

possible in the magnetic properties of rocks. The dominant<br />

effect over impact structures is a magnetic low or subdued<br />

zone ranging in amplitude from tens to a few hundred nanotesla<br />

that is commonly manifested as a truncation of the<br />

regional magnetic fabric (Dabizha and Fedynsky 1975,<br />

Clark 1983). Magnetic lows are best defined over simple<br />

and some small complex craters, where the anomaly is<br />

smooth and simple; whereas at larger impact structures,<br />

the magnetic low can be modified by the presence of<br />

shorter-wavelength, large-amplitude, localized anomalies<br />

that usually occur at or near the centre of the structure.<br />

No correspondence exists between the magnetic anomaly<br />

character and crater morphology of impact structures.<br />

Moreover, the presence of a central gravity high does not<br />

imply the existence of a central magnetic anomaly. There<br />

are several structures with no obvious magnetic signature.<br />

Shock effects, thermal effects or chemical effects may<br />

cause magnetic anomalies related to impact. Shock effects<br />

in impact structures can serve to increase or decrease magnetization<br />

levels. Thermal effects may result in the production<br />

of non-magnetic impact glasses (Pohl 1971) or in<br />

resetting magnetic minerals through thermoremanent<br />

magnetization in the direction of the Earth’s magnetic field<br />

at the time of impact. Chemical effects may result in the<br />

production of new magnetic phases, through elevated<br />

residual temperatures and hydrothermal alteration, leading<br />

to the acquisition of a chemical remanent magnetization in<br />

the direction of the ambient field.<br />

Seismic signature<br />

Reflection seismic surveys allow for detailed imaging<br />

of impact structure morphology and delineating zones of<br />

incoherent reflections that are characteristic of brecciation<br />

and fracturing. The disturbance of coherent subsurface reflectors<br />

is most prominent in the central uplift of complex<br />

structures and decreases outward and downward from this<br />

zone (Brenan et al. 1975). Reflection data can provide es-<br />

256


The recognition of terrestrial impact structures<br />

timates of such morphological parameters as the dimensions<br />

of the central uplift, annular trough and faulted<br />

blocks at the structural rim of complex structures (e.g.<br />

Morgan et al. 2002). The depth to horizontal reflectors that<br />

exist below the crater floor can be used to determine the<br />

amount of structural uplift.<br />

Electrical signature<br />

The presence of fluids in impact-induced fractures and<br />

pore spaces leads to decreased resistivity levels that can be<br />

mapped effectively by various electrical methods. The<br />

conductivity of rocks is heavily dependent on their water<br />

content: < 1% change in water content can produce more<br />

than an order of magnitude change in conductivity. The<br />

degree of fragmentation determines the amount and distribution<br />

of fluids within the rock and hence, its electrical<br />

properties.<br />

Where a distinct contrast exists between the allochthonous<br />

breccia deposits and the underlying autochthonous<br />

target rocks, electrical profiling using resistivity sounding<br />

can map the structure of the true crater floor (e.g. Vishnevsky<br />

and Lagutenko 1986). In order to determine the<br />

deeper electrical structure associated with impact, magnetotelluric<br />

surveys have been carried out (e.g. Zhang et al.<br />

1988, Campos-Enriquez et al. 1997).<br />

Geology of impact structures<br />

Although an anomalous circular topographic, structural,<br />

or geological feature may indicate the presence of an<br />

impact structure, there are other endogenic geological<br />

processes that can produce similar features in the terrestrial<br />

environment. An obvious craterform is an excellent indicator<br />

of a possible impact origin; particularly, if it has<br />

the appropriate morphometry, but as noted, such features<br />

are rare and short-lived in the terrestrial environment. The<br />

burden of proof for an impact origin generally lies with the<br />

documentation of the occurrence of shock-metamorphic<br />

effects.<br />

Few structures preserve physical evidence of the impacting<br />

body. Such structures are limited to small, young,<br />

simple structures, where the impacting body (or, more<br />

commonly, fragments of it) has been slowed by atmospheric<br />

deceleration and impacts at less than cosmic velocity.<br />

These are restricted generally to the impact of iron or<br />

stony-iron meteorites. Stony meteorites are weaker than<br />

their iron-bearing counterparts and small stones are generally<br />

crushed as a result of atmospheric interaction (Melosh<br />

1981). Larger impacting bodies (>100–150 m in diameter)<br />

survive atmospheric passage with undiminished impact<br />

velocity. Consequently, the peak shock pressures upon impact<br />

are sufficient, in most cases, to result in the melting<br />

and vaporisation of the impacting body, destroying it as a<br />

physical entity.<br />

On impact, the bulk of the impacting body’s kinetic<br />

Temperature (°C)<br />

10 000<br />

1 000<br />

100<br />

P-T field of<br />

endogenic<br />

metamorphism<br />

Shatter<br />

cones<br />

Rock<br />

melting<br />

Fused<br />

glasses<br />

Diaplectic<br />

glasses<br />

Planar<br />

features<br />

Vaporization<br />

Pressure post-shock<br />

temperature curve for<br />

shock metamorphism<br />

of granitic rocks<br />

1 10 100 1 000<br />

Pressure, GPa<br />

Figure 5. Temperature and pressure range of shock metamorphic effects<br />

compared to that of endogenic metamorphism. Planar features include<br />

planar deformation features (PDFs) and planar fractures (PFs). Scale is<br />

log-log. See text for details.<br />

energy is transferred to the target by means of a shock<br />

wave. This shock wave imparts kinetic energy to the target<br />

materials, which leads to the formation of a crater. It also<br />

increases the internal energy of the target materials, which<br />

leads to the formation of so-called shock-metamorphic effects.<br />

The details of the physics of impact and shockwave<br />

behavior can also be found in Melosh (1989), and references<br />

therein.<br />

Shock metamorphism is the progressive breakdown in<br />

the structural order of minerals and rocks due to the passage<br />

of a high-pressure shock wave and requires pressures<br />

and temperatures well above the pressure-temperature<br />

field of endogenic terrestrial metamorphism (Fig. 5). The<br />

dependence on high pressures for the formation of shockmetamorphic<br />

effects has been shown by their duplication<br />

in nuclear and chemical explosion craters, and in laboratory<br />

shock recovery experiments (e.g. Hörz 1968, Müller<br />

and Hornemann 1969, Borg 1972). Minimum shock pressures<br />

required for the production of diagnostic shockmetamorphic<br />

effects are 5–10 GPa for most silicate<br />

minerals. Strain rates produced by impact cratering<br />

process are of the order of 10 6 s -1 to 10 9 s -1 (Stöffler and<br />

Langenhorst 1994), many orders of magnitude higher than<br />

typical tectonic strain rates (10 -12 s -1 to 10 -15 s -1 ; e.g. Twiss<br />

and Moores 1992), and shock-pressure duration is measured<br />

in seconds, or less, in even the largest impact events<br />

(Melosh 1989). These physical conditions are not reproduced<br />

by endogenic geologic processes. They are unique<br />

to impact and, unlike endogenic terrestrial metamorphism,<br />

disequilibrium and metastability are common phenomena<br />

in shock metamorphism.<br />

The extreme pressures and high strain rates of shock<br />

deformation are fundamental differences from normal endogenic<br />

causes of compression (Ashworth and Schneider<br />

1985, Goltrant et al. 1991, 1992, Langenhorst 1994). A<br />

shock wave passing through a heterogeneous rock mass<br />

undergoes numerous modifications, as it interacts with<br />

grain boundaries, fractures, foliations, and different mineral<br />

species with different shock impedances within the<br />

257


Ann M. Therriault – Richard A. F. Grieve – Mark Pilkington<br />

Figure 6. Outcrop (~ 80 m high) of coherent impact melt rock at the Mistastin<br />

complex impact structure, Canada.<br />

rock. There is, thus, local variations in shock pressure. Petrographic<br />

study indicates that shock pressures may vary by<br />

a factor of two or more over distances ranging from millimeters<br />

to meters in outcrops (Grady 1977). Hence, each<br />

individual mineral grain experiences its own particular<br />

shock history based upon its physical properties and its relationship<br />

to both the adjacent grains and the overall structural<br />

character of the rock. A maximum shock effect in<br />

grains of a particular mineral species in a hand specimen<br />

may, thus, be a means of measuring relative deformation<br />

intensities throughout an impact structure. For example,<br />

shock pressures of at least ~ 5 GPa are required to produce<br />

PFs in quartz and greater than 10 GPa to produce PDFs in<br />

quartz or feldspars. This variation of shock deformation of<br />

important rock-forming minerals of the target rocks with<br />

increasing shock pressures have been used to delineate<br />

zones of shock metamorphism in the floor of a number of<br />

impact structures, e.g. Charlevoix, Canada (Robertson<br />

1968), Brent, Canada (Dence 1968), Ries, Germany (von<br />

Engelhardt and Stöffler 1968), and Manicouagan, Canada<br />

(Dressler 1990), with the intensity of deformation decreasing<br />

from the center outwards.<br />

The exact physical conditions on impact are a function<br />

of the specific impact parameters. The density of the impacting<br />

body and the target, and the impact velocity determine<br />

the peak pressure on impact. The shock wave<br />

attenuates with distance from the impact point with the kinetic<br />

energy of the impact event determining the absolute<br />

radial distance in the target at which a specific shock pressure<br />

is achieved and, thus, which specific shock-metamorphic<br />

effects occur. Shock-metamorphic effects are well<br />

described in papers by Chao (1967), Bunch (1968), Stöffler<br />

(1971, 1972, 1974), Stöffler and Langenhorst (1994),<br />

Grieve et al. (1996), French (1998), Langenhorst and<br />

Deutsch (1998), and Langenhorst (this volume). They are<br />

discussed here only in general terms.<br />

Impact melting<br />

During compression, considerable pressure-volume<br />

work is done and the pressure release occurs adiabatically.<br />

Heating of the target rocks, thus, occurs as not all this<br />

pressure-volume work is recovered upon pressure release<br />

and results in irreversible waste heat. Above 60 GPa, the<br />

waste heat is sufficient to cause whole-rock melting and,<br />

and at higher pressures, vaporisation of a certain volume<br />

of target rocks (Melosh 1989). This volume is a function<br />

of the impact velocity, physical properties of the impacting<br />

body and target, and, most importantly, the size of the impacting<br />

body (Grieve and Cintala 1992).<br />

Impact melt lithologies may occur as glass bombs in<br />

crater ejecta (von Engelhardt 1990), as dykes within the<br />

crater floor and walls, as glassy to crystalline pools and<br />

lenses within the breccia lenses of simple craters, or as coherent<br />

annular sheets (Fig. 6) lining the floor of complex<br />

craters and stratigraphically located immediately above<br />

breccias and/or brecciated basement rocks and overlain by<br />

breccias.<br />

When crystallized, impact-melt sheets have igneous<br />

textures, but tend to be heavily charged with clastic debris<br />

a<br />

b<br />

Figure 7. Photomicrographs of far-from-equilibrium textures examples in impact melts: (a) plagioclase crystals with swallow-tail texture, Boltysh impact<br />

melt sheet, Ukraine, plane light, field-of-view = 2.28 mm; (b) pyroxene-plagioclase spherulitic texture, Vredefort Granophyre impact melt dyke,<br />

South Africa, plane light, field-of-view = 5 mm.<br />

258


The recognition of terrestrial impact structures<br />

Figure 8. Photomicrograph of fused glass (lechatelierite), Ries, Germany,<br />

plane light, field-of-view = 2.5 mm.<br />

towards their lower and upper contacts. They may, therefore,<br />

have a textural resemblance to endogenic igneous<br />

rocks. Impact melts are superheated, reaching thousands<br />

of degrees Kelvin. Temperature differences with host<br />

rocks may result in rapid cooling of the melt leading to farfrom-equilibrium<br />

textures (Fig. 7). Grain-size in thick impact-melt<br />

sheets increases inwards from the contacts, but,<br />

in general, impact-melt rocks are usually fine-grained to<br />

glassy. An important textural property of impact-melt<br />

rocks is the presence of mineral and rock fragments, which<br />

have undergone shock metamorphism of different degrees,<br />

and have been variously reworked by the melt. The size of<br />

such fragments ranges from millimeters to several hundreds<br />

of meters, and gradational changes in inclusion content<br />

are observed in thick melt sheets, varying from one to<br />

several tens of percent (e.g. von Engelhardt 1984), with<br />

highest concentrations towards their lower and upper contacts.<br />

Impact-melt rocks can have an unusual chemistry compared<br />

with endogenic volcanic rocks, as their composition<br />

depends on the wholesale melting of a mix of target rocks,<br />

as opposed to partial melting and/or fractional crystallization<br />

relationships for endogenous igneous rocks. The composition<br />

of impact-melt rocks is characteristic of the target<br />

rocks and may be reproduced by a mixture of the various<br />

country rock types in their appropriate geological proportions.<br />

Such parameters as 87 Sr/ 86 Sr and 143 Nd/ 144 Nd ratios<br />

may also reflect the pre-existing target rocks within the<br />

impact-melt rocks composition (Jahn et al. 1978, Faggart<br />

et al. 1985). In general, unlike endogenous magmatic rock<br />

masses of comparable size (up to a few hundred meters<br />

thick), even relatively thick impact-melt sheets are chemically<br />

homogeneous over distances of millimeters to kilometers.<br />

In cases where the target rocks are not<br />

homogeneously distributed, this observation may not hold<br />

true, such as for Manicouagan, Canada (Grieve and Floran<br />

1978), Chicxulub (Kettrup et al. 2000) and Popigai (Kettrup<br />

et al. 2002). Differentiation is not a characteristic of<br />

relatively thick coherent impact-melt sheets (with the exception<br />

of the extremely thick, ~ 2.5 km, Sudbury Igneous<br />

Complex, Sudbury Structure, Canada; Ostermann 1996,<br />

Ariskin et al. 1999, Therriault et al. 2002).<br />

Enrichments above target rock levels in siderophile elements<br />

and Cr have been identified in some impact-melt<br />

rocks. These are due to an admixture of up to a few percent<br />

of meteoritic material from the impacting body. In<br />

some melt rocks, the relative abundances of the various<br />

siderophiles have constrained the composition of the impacting<br />

body to the level of meteorite class, (e.g. East<br />

Clearwater, Canada, was formed by a C1 chondrite, Palme<br />

et al. 1979). In other melt rocks, no siderophile anomaly<br />

has been identified. This may be due to the inhomogeneous<br />

distribution of meteoritic material within the impact-melt<br />

rocks and sampling variations (Palme et al.<br />

1981) or to differentiated, and, therefore, relatively nonsiderophile-enriched<br />

impacting bodies, such as basaltic<br />

achondrites. More recently, high precision osmium-isotopic<br />

analyses have been used to detect a meteoritic signature<br />

at terrestrial impact structures (e.g. Koeberl et al.<br />

1994). Unfortunately, Re-Os systematics are, in themselves,<br />

not an effective discriminator between meteorite<br />

classes.<br />

Fused glasses and diaplectic glasses<br />

In general, shock fused minerals are characterized<br />

morphologically by flow structures and vesiculation (Fig.<br />

8). Peak pressures required for shock melting of single<br />

crystals are in the order of 40 to 60 GPa (Stöffler 1972,<br />

1974), for which postshock temperatures (> 1000 °C) exceed<br />

the melting points of typical rock-forming minerals<br />

(Fig. 5). At these conditions, the minerals in the rock will<br />

melt immediately and independently after the passage of<br />

the shock wave. This melt has approximately the same<br />

composition as the original mineral before any flow or<br />

mixing takes place, and the melt regions are initially distributed<br />

through the rock in the same manner as the original<br />

mineral grains (French 1998). Melting is mineral<br />

selective, producing unusual textures in which one or more<br />

minerals show typical melting features; whereas, others,<br />

even juxtaposed ones, do not. One of the most common<br />

fused glasses observed at terrestrial impact structures is<br />

that of quartz, i.e. lechatelierite (e.g. Fig. 8).<br />

Conversion of minerals to an isotropic, dense, glassy<br />

phase at peak pressures of 30 to 50 GPa (Fig. 5) and temperatures<br />

well below their normal melting point is a shock<br />

metamorphic effect unique to framework silicates. These<br />

phases are called diaplectic (from the Greek “destroyed by<br />

striking”) glasses, which are produced by breakdown of<br />

long-range order of the crystal lattice without fusion.<br />

Although diaplectic forms may occur as the direct result of<br />

compression by the shock wave, they are probably more<br />

commonly produced by inversion from a high-pressure<br />

crystalline phase, which is unstable in the postshock P-T<br />

environment (Robertson 1973). Based on shock recovery<br />

experiments, the formation of diaplectic glass occurs between<br />

30 and 45 GPa for feldspar and 35 to 50 GPa for<br />

quartz (e.g. Stöffler and Hornemann 1972). The morphology<br />

of the diaplectic glass is the same as the original mineral<br />

crystal and shows no evidence of fluid textures (e.g.<br />

Grieve et al. 1996). Diaplectic glasses have densities low-<br />

259


Ann M. Therriault – Richard A. F. Grieve – Mark Pilkington<br />

Figure 9. Photomicrograph of partial conversion to maskelynite of plagioclase<br />

feldspar crystals, Manicouagan, Canada, cross-polarized light,<br />

field-of-view = 5 mm.<br />

er than the crystalline form from which they are derived,<br />

but higher than thermally melted glasses of equivalent<br />

composition (e.g. Stöffler and Hornemann 1972, Langenhorst<br />

and Deutsch 1994). With increasing pressure, the<br />

bulk density of diaplectic glass decreases. This decrease is<br />

due in part to progressively greater portions of the mineral<br />

having been converted to low density, disordered phases,<br />

but also to the fact that diaplectic phases exist in a<br />

sequence of intermediate structural states, whose refractive<br />

index and density decrease with increasing pressure<br />

and temperature (Stöffler and Hornemann 1972). The refractive<br />

index of diaplectic glasses is also generally higher<br />

than for synthetic, or thermally melted, glasses of<br />

equivalent composition (e.g. Robertson 1973, Grieve et al.<br />

1996). However, in the case of K-feldspar, its diaplectic<br />

glass has a slightly lower refractive index than the fused<br />

feldspar glass (Stöffler and Hornemann 1972). Maskelynite,<br />

the diaplectic form of plagioclase (Fig. 9), is the<br />

most common example from terrestrial rocks; diaplectic<br />

glasses of quartz (Chao 1967) and of alkali feldspar<br />

(Bunch 1968) are also reported but in lesser abundance.<br />

Diaplectic glasses of different minerals can exist adjacent<br />

to one another without mixing (e.g. Robertson 1973).<br />

High-pressure polymorphs<br />

Shock can result in the formation of metastable polymorphs,<br />

such as stishovite and coesite from quartz (Chao<br />

et al. 1962, Langenhorst this volume) and diamond and<br />

lonsdaleite from graphite (Grieve and Masaitis 1996, Masaitis<br />

1998, Langenhorst this volume). Coesite and diamond<br />

are also products of endogenic terrestrial geological<br />

processes, including high-grade metamorphism, but the<br />

paragenesis and, more importantly, the geological setting<br />

are completely different from that in impact events.<br />

Under high pressure, the mineral lattice is unstable and<br />

is converted to a more stable configuration. Such transformation<br />

begins at ~ 11.5 GPa for K-feldspars (Robertson<br />

1973) and at ~ 12 GPa for quartz (De Carli and Milton<br />

1965). With increasing pressure, a greater proportion of<br />

the mineral is converted to a high-pressure polymorph until<br />

complete transformation is achieved at ~ 30 GPa for<br />

feldspars (Ahrens et al. 1969) and ~ 35 GPa for quartz<br />

(Stöffler and Langenhorst 1994). Neither the high-pressure<br />

phase of K-feldspar, thought to be the dense hollandite-type<br />

structure with Al and Si in octahedral<br />

co-ordination, nor an equivalent plagioclase polymorph<br />

have been recovered from shock experiments or identified<br />

in non-impact terrestrial rocks (Robertson 1973). It would<br />

appear that these phases are very unstable in postshock environments<br />

and, more likely, invert to more disordered,<br />

metastable phases. The high-pressure polymorphs of<br />

quartz (i.e. stishovite and coesite) have only rarely been<br />

produced by laboratory shock recovery experiments (cf.<br />

Stöffler and Langenhorst 1994). Contrary to what is expected<br />

from equilibrium phase diagram, stishovite is<br />

formed at lower pressures (12–30 GPa) than coesite<br />

(30–50 GPa; Stöffler and Langenhorst 1994) in impact<br />

events. This is mainly due to the fact that stishovite is<br />

formed during shock compression, whereas, coesite crystallizes<br />

during pressure release. In terrestrial impact structures,<br />

these polymorphs occur in small or trace amounts as<br />

very fine-grained aggregates and are formed by partial<br />

transformation of the host quartz. In crystalline or dense<br />

rocks, coesite is found in quartz with planar deformation<br />

features (PDFs) and strongly lowered refractive index and,<br />

more commonly, in diaplectic glass; whereas, in porous<br />

sandstone, coesite co-exists with > 80% of quartz displaying<br />

planar fractures (PFs) and diaplectic quartz glass<br />

(Grieve et al. 1996). Stishovite occurs most commonly in<br />

quartz with PDFs and less frequently in diaplectic glass<br />

(Stöffler 1971). For details on the characteristics of coesite<br />

and stishovite, the reader is referred to Stöffler and Langenhorst<br />

(1994) and references therein.<br />

Planar microstructures<br />

The most common documented shock-metamorphic<br />

effect is the occurrence of planar microstructures in tectosilicates,<br />

particularly quartz (Hörz 1968). The utility of<br />

planar microstructures in quartz reflects the ubiquitous nature<br />

of the mineral and its stability, including the stability<br />

of the microstructures themselves, in the terrestrial environment,<br />

and the relative ease with which they can be documented.<br />

For details, the reader is referred to the<br />

accompanying paper by Langenhorst. Recent reviews of<br />

the nature of the shock metamorphism of quartz, with an<br />

emphasis on the nature and origin or planar microstructures<br />

in experimental and natural impacts, can be found in<br />

Stöffler and Langenhorst (1994) and Grieve et al. (1996).<br />

Planar deformation features (PDFs) in minerals are<br />

produced under pressures of ~ 10 to ~ 35 GPa (Fig. 5).<br />

Planar fractures (PFs) form under shock pressures ranging<br />

from ~ 5 GPa up to ~ 35 GPa (Stöffler 1972, Stöffler and<br />

Langenhorst 1994).<br />

260


The recognition of terrestrial impact structures<br />

Shatter cones<br />

The only known diagnostic shock effect that is megascopic<br />

in scale is the occurrence of shatter cones (Dietz<br />

1968). Shatter cones are unusual, striated, and horse-tailed<br />

conical fractures ranging from millimeters to meters in<br />

length produced in rocks by the passage of a shock wave<br />

(e.g. Sagy et al. 2002). The striated surfaces of shatter<br />

cones are positive/negative features and the striations are<br />

directional, i.e., they appear to branch and radiate along<br />

the surface of the cone. The acute angle of this distinctive<br />

pattern points toward the apex of the cone and the shatter<br />

cones themselves generally point upward with their axes<br />

lying at any angle to the original bedding. Once the host<br />

rocks are graphically restored to their original impact position,<br />

shatter cones indicate the point of impact.<br />

Shatter cones are initiated most frequently in rocks that<br />

experienced moderately low shock pressures, 2–6 GPa<br />

(Fig. 5), but have been observed in rocks that experienced<br />

~25 GPa (Milton 1977). These conical striated fracture<br />

surfaces are best developed in fine-grained, structurally<br />

isotropic lithologies, such as carbonates and quartzites.<br />

They do occur in coarse-grained crystalline rocks but are<br />

less common and poorly developed. They are generally<br />

found as individual or composite groups of partial to complete<br />

cones (Fig. 10) in place in the rocks below the crater<br />

floor, especially in the central uplifts of complex impact<br />

structures, and rarely in isolated rock fragments in breccia<br />

units. Shatter cones are used as a diagnostic field criterion<br />

to identify impact structures (e.g. Dietz 1947, Milton<br />

1977).<br />

Conclusion<br />

The detailed study of impact events on Earth is a relatively<br />

recent addition to the spectrum of studies engaged in<br />

by the geological sciences. More than anything, it was<br />

preparations for and, ultimately, the results of the lunar<br />

and the planetary exploration program that provided the<br />

initial impetus and rationale for their study. Some recent<br />

discoveries have resulted from the occurrence or re-examination<br />

of unusual lithologies, rather than an obvious circular<br />

geological or topographic feature. For example,<br />

unusual breccias at Gardnos, Norway and Lockne, Sweden<br />

had been known for some time, but their shock-metamorphic<br />

effects were documented only recently, and they<br />

are now associated with the remnants of impact structures<br />

(French et al. 1997, Lindström and Sturkell 1992).<br />

The level of knowledge concerning individual terrestrial<br />

impact structures is highly variable. In some cases, it<br />

is limited to the original discovery publication. In terms of<br />

understanding the terrestrial record, this is compensated,<br />

to some degree, by the fact that impact structures with similar<br />

dimensions and target rocks have the same major characteristics.<br />

Nevertheless, there is still much to be learned<br />

about impact processes from terrestrial impact structures,<br />

Figure 10. Complete shatter cone in limestone, Cap de la Corneille,<br />

Charlevoix, Canada.<br />

particularly with respect to details of the third dimension.<br />

This is the property that is unobtainable from impact structures<br />

on other bodies in the solar system, where it must be<br />

studied by remote-sensing methodologies.<br />

Apart from increasing our understanding of impact<br />

processes, the study of terrestrial impact structures has influenced<br />

the siting of significant economic deposits<br />

(Grieve and Masaitis 1994, Donofrio 1997, Grieve 1997).<br />

In addition, the documentation of the terrestrial impact<br />

record provides a direct measure of the cratering rate on<br />

Earth and, thus, a constraint on the hazard that impact<br />

presents to human civilization (Gehrels 1994). The K/T<br />

impact may have resulted in the demise of the dinosaurs as<br />

the dominant land-life form and, thus, permitted the ascendancy<br />

of mammals and, ultimately, humans. It is, however,<br />

inevitable that human civilization, if it persists long<br />

enough, will be subjected to an impact-induced environmental<br />

crisis of potentially immense proportions.<br />

Acknowledgements. We would like to thank J. Ormö and<br />

A. Deutsch for their critical reviews. This paper is GSC contribution<br />

No. 2002141.<br />

References<br />

Ahrens T. J., Petersen C. F., Rosenberg J. T. (1969): Shock compression<br />

of feldspars. J. Geophys. Res. 74, 2727–2746.<br />

Ariskin A. A, Deutsch A., Ostermann M. (1999): The Sudbury “Igneous”<br />

Complex: simulating phase equilibria and in situ differentiation for<br />

two proposed parental magmas. In: Dressler B. O. and Sharpton V.<br />

L. (eds) Large Meteorite Impacts and Planetary Evolution II. Geol.<br />

Soc. Amer. Spec. Pap. 339, pp. 373–387.<br />

Ashworth J. R., Schneider H. (1985): Deformation and transformation in<br />

experimentally shock-loaded quartz. Phys. Chem. Miner. 11,<br />

241–249.<br />

Borg I. Y. (1972): Some shock effects in granodiorite to 270 kbar at the<br />

Piledriver site. In: Head H. C. et al. (eds) Flow and fracture of rock.<br />

Amer. Geophys. Union Monograph, Washington, D.C., pp. 293–311.<br />

Brenan R. L., Peterson B. L., Smith H. J. (1975): The origin of Red Wing<br />

Creek structure: McKenzie County, North Dakota. Wyoming Geol.<br />

Ass. Earth Sci. Bull. 8, 1–41.<br />

Bunch T. E. (1968): Some characteristics of selected minerals from<br />

craters. In: French B. M. and Short N. M. (eds) Shock Metamorphism<br />

of Natural Materials. Mono Book Corp., Baltimore, pp.<br />

413–432.<br />

261


Ann M. Therriault – Richard A. F. Grieve – Mark Pilkington<br />

Campos-Enriquez J. O., Arzate J. A., Urrutia-Fucugauchi J., Delgado-<br />

Rodriguez O. (1997): The subsurface structure of the Chicxulub<br />

Crater (Yucatan, Mexico): preliminary results of a magnetotelluric<br />

study. The Leading Edge 16, 1774–1777.<br />

Chao E. C. T. (1967): Shock effects in certain rock-forming minerals.<br />

Science 156, 192–202.<br />

Chao E. C. T., Fahey J. J., Littler J., Milton D. J. (1962): Stishovite, SiO 2,<br />

a very high pressure new mineral from Meteor Crater, Arizona. J.<br />

Geophys. Res. 67, 419–421.<br />

Clark J. F. (1983): Magnetic survey data at meteoritic impact sites in<br />

North America. Geomagnetic Service of Canada, Earth Physics<br />

Branch, Open File, 83-5, 1–32.<br />

Dabizha A. I., Fedynsky V. V. (1975): The Earth’s “star wounds” and<br />

their diagnosis by geophysical methods. Zemlya i Vselennaya 3,<br />

56–64. (in Russian)<br />

Dabizha A. I., Feldman V. I. (1982): The geophysical properties of some<br />

astroblemes in the USSR. Meteoritika 40, 91–101. (in Russian)<br />

De Carli P. S., Milton D. J. (1965): Stishovite, synthesis by shock wave.<br />

Science 147, 144–145.<br />

Dence M. R. (1968): Shock Zoning at Canadian Craters: Petrography and<br />

Structural Implications. In: French B. M. and Short N. M. (eds)<br />

Shock Metamorphism of Natural Materials. Mono Book Corp., Baltimore,<br />

pp. 169–184.<br />

Dence M. R. (1972): The nature and significance of terrestrial impact<br />

structures. 24 th Inter. Geol. Congr. Section 15, 77–89.<br />

Dietz R. (1947): Meteorite impact suggested by orientation of shatter<br />

cones at the Kentland, Indiana, disturbance. Science 105, 42–43.<br />

Dietz R. S. (1968): Shatter cones in cryptoexplosion structures. In:<br />

French B. M. and Short N. M. (eds) Shock Metamorphism of Natural<br />

Materials. Mono Book Corp., Baltimore, pp. 267–285.<br />

Donofrio R. R. (1977): Survey of hydrocarbon-producing impact structures<br />

in North America: exploration results to date and potential for<br />

discovery in Precambrian basement rock. In: Johnson K. S. and<br />

Campbell J. A. (eds) Ames structure in northwest Oklahoma and<br />

similar features: Origin and petroleum production. Oklahoma Geol.<br />

Surv. Circular 100, pp. 17–29.<br />

Dressler B. (1990): Shock metamorphic features and their zoning and<br />

orientation in the Precambrian rocks of the Manicouagan Structure,<br />

Quebec, Canada. Tectonophysics. 171, 229–245.<br />

Faggart B. E., Basu A. R., Tatsumoto M. (1985): Origin of the Sudbury<br />

complex by meteoritic impact: Neodymium isotopic evidence.<br />

Science 230, 436–439.<br />

Floran R. J., Dence M. R. (1976): Morphology of the Manicouagan ringstructure,<br />

Quebec, and some comparisons with lunar basins and<br />

craters. Proc. 7 th Lunar Sci. Conf., 2845–2865.<br />

French B. M. (1998): Traces of Catastrophe: A Handbook of Shock-Metamorphic<br />

Effects in Terrestrial Meteorite Impact Structures. LPI<br />

Contribution No. 954, Lunar and Planetary Institute, Houston.<br />

French B. M., Short N. M. (1968): Shock Metamorphism of Natural Materials.<br />

Mono Book Corp., Baltimore.<br />

French B., Koeberl C., Gilmour I., Shirey S. B., Dons J. A., Naterstad J.<br />

(1997): The Gardnos impact structure, Norway: Petrology and geochemistry<br />

of target rock and impactites. Geochim. Cosmochim. Acta<br />

61, 873–904.<br />

Garvin J. B., Schnetzler C. C. (1994): The Zhamanshin impact structure:<br />

A new class of complex crater? In: Dressler B. O. et al. (eds) Large<br />

Meteorite Impacts and Planetary Evolution. Geol. Soc. Amer. Spec.<br />

Pap. 293, pp. 249–257.<br />

Gehrels T. (1994): Hazards Due to Comets and Asteroids. University of<br />

Arizona Press, Tuscon.<br />

Goltrant O., Cordier P., Doukhan J.-C. (1991): Planar deformation features<br />

in shocked quartz: a transmission electron microscopy investigation.<br />

Earth Planet. Sci. Lett. 106, 103–115.<br />

Goltrant O., Leroux H., Doukhan J.-C., Cordier P. (1992): Formation<br />

mechanism of planar deformation features in naturally shocked<br />

quartz. Phys. Earth Planet. Int. 74, 219–240.<br />

Grady D. E. (1977): Processes occurring in shock wave compression of<br />

rocks and minerals. In: Managhnani M. H. and Akimoto S.-I. (eds)<br />

High pressure research: Applications in geophysics. Academic Press,<br />

New York, pp. 389–438.<br />

Grieve R. A. F. (1997): Terrestrial impact structures: basic characteristics<br />

and economic significance with emphasis on hydrocarbon production.<br />

In Johnson K. S. and Campbell J. A. (eds) Ames structure<br />

in northwest Oklahoma and similar features: Origin and petroleum<br />

production. Oklahoma Geol. Surv. Circular 100, pp. 3–16.<br />

Grieve R. A. F., Cintala M. J. (1992): An analysis of differential impact<br />

melt – crater scaling and implications for the terrestrial impact<br />

record. Meteoritics 27, 526–539.<br />

Grieve R. A. F., Floran R. J. (1978): Manicouagan impact melt, Quebec<br />

2. Chemical interrelations with basement and formational processes.<br />

J. Geophys. Res. 83, 2761–2771.<br />

Grieve R. A. F., Head J. W. (1983): The Manicouagan impact structure:<br />

An analysis of its original dimensions and form. J. Geophys. Res.<br />

Suppl. 88, A807–A818.<br />

Grieve R. A. F., Langenhorst F., Stöffler D. (1996): Shock metamorphism<br />

of quartz in nature and experiment: II. Significance in geoscience.<br />

Meteoritics Planet. Sci. 31, 6–35.<br />

Grieve R. A. F., Masaitis V. L. (1994): The economic potential of terrestrial<br />

impact craters. Inter. Geol. Rev. 36, 105–151.<br />

Grieve R. A. F., Masaitis V. L. (1996): Impact diamonds. In: LeCheminant<br />

A. N. et al. (eds) Searching for diamonds in Canada. Geol. Surv.<br />

Can. Open-File 3228, pp. 183–186.<br />

Grieve R. A. F., Pilkington M. (1996): The signature of terrestrial impacts.<br />

AGSO J. Austr. Geol. Geophys. 16, 399–420.<br />

Grieve R. A. F., Therriault A. M. (2000): Vredefort, Sudbury, Chicxulub:<br />

Three of a kind? Ann. Rev. Earth Planet. Sci. 28, 305–338.<br />

Henkel H. H. (1992): Geophysical aspects of impact craters in eroded<br />

shield environments, with special emphasis on electric resistivity.<br />

Tectonophysics 216, 63–90.<br />

Hildebrand A. R., Pilkington M., Otriz-Aleman C., Chavez R., Urrutia-<br />

Fucugauchi J., Connors M, Graniel-Castro E., Camaro-Zi A.,<br />

Halpenny J., Niehaus D. (1998): Mapping Chicxulub crater structure<br />

with gravity and seismic reflection data. In: Grady M. M. et al. (eds)<br />

Meteorites: Flux with time and impact effects. Geol. Society (London)<br />

Spec. Publ. 140, pp. 155–176.<br />

Hörz F. (1968): Statistical measurements of deformation structures and<br />

refractive indices in experimentally shock loaded quartz. In: French<br />

B. M. and Short N. M. (eds) Shock Metamorphism of Natural Materials.<br />

Mono Book Corp., Baltimore, pp. 243–253.<br />

Jahn B., Floran R. J., Simonds C. H. (1978): Rb-Sr isochron age of the<br />

Manicouagan melt sheet, Quebec, Canada. J. Geophys. Res. 83,<br />

2799–2803.<br />

Kettrup B., Deutsch A., Ostermann M., Agrinier P. (2000): Chicxulub impactities:<br />

geochemical clues to the precursor rocks. Meteoritics<br />

Planet. Sci. 35, 1129–1138.<br />

Kettrup B., Deutsch A., Masaitis V. L. (2002): Homogeneous impact<br />

melts produced by a heterogeneous target? Sr-Nd isotopic evidence<br />

from the Popigai crater, Russia. Geochim. Cosmochim. Acta, (in<br />

press)<br />

Koeberl C., Shirey S. B., Reimold W. U. (1994): Re-Os isotope systematics<br />

as a diagnostic tool for the study of impact craters. Lunar Planet.<br />

Inst. Contrib. 825, 61–63.<br />

Langenhorst F. (1994): Shock experiments on α- and β-quartz: II. Modelling<br />

of lattice expansion and amorphization. Earth Planet. Sci. Lett.<br />

128, 683–698.<br />

Langenhorst F. (2002): Shock metamorphism of some minerals: Basic introduction<br />

and microstructural observations. Bull. Czech Geol. Surv.<br />

77, this volume.<br />

Langenhorst F., Deutsch A. (1994): Shock experiments on α- and β-<br />

quartz: I. Optical and density data. Earth Planet. Sci. Lett. 125,<br />

407–420.<br />

Langenhorst F., Deutsch A. (1998): Mineralogy of Astroblemes – Terrestrial<br />

Impact Craters. In: Marfunin A. S. (ed.) Advanced Mineralogy,<br />

Vol. 3, Mineral Matter in Space, Mantle, Ocean Floor, Biosphere,<br />

Environmental Management, Jewelry, Chapter 1.10, pp. 95–119.<br />

Lindström M., Sturkell E. F. F. (1992): Geology of the early Paleozoic<br />

Lockne impact structure, central Sweden. Tectonophysics 216,<br />

169–185.<br />

Lindström M., Sturkell E. F. F., Törnberg R., Ormö J. (1996): The marine<br />

impact crater at Lockne, central Sweden. GFF 118, 193–206.<br />

Masaitis V. L. (1998): Popigai crater: Origin and distribution of diamondbearing<br />

impactites. Meteoritics Planet. Sci. 33, 349–359.<br />

262


The recognition of terrestrial impact structures<br />

Melosh H. J. (1981): Atmosphere breakup of terrestrial impactors. In:<br />

Schultz P. H. and Merrill P. B. (eds) Multi-Ring Basins. New York,<br />

pp. 29–35.<br />

Melosh H. J. (1989): Impact Cratering: A Geologic Process. Oxford<br />

University Press, New York.<br />

Milton D. J. (1977): Shattercones - an outstanding problem in shock mechanics.<br />

In: Roddy D. J. et al. (eds) Impact and Explosion Cratering.<br />

New York, pp. 703–714.<br />

Milton D. J., Glikson A. Y., Brett R. (1996): Gosses Bluff - a latest Jurassic<br />

impact structure central Australia. Part 1: geological structure,<br />

stratigraphy and origin. AGSO J. Austr. Geol. Geophys. 16,<br />

453–486.<br />

Morgan J., Warner M., Grieve R. (2002): Geophysical constraints on the<br />

size and structure of the Chicxulub impact crater. In: Koeberl C. and<br />

MacLeod K. G. (eds) Catastrophic Events and Mass Extinctions: Impacts<br />

and Beyond. Boulder, Colorado. Geol. Soc. Amer. Spec. Pap.<br />

356, pp. 39–46.<br />

Müller W. F., Hornemann U. (1969): Shock-induced planar deformation<br />

structures in experimentally shock-loaded olivines and in olivines<br />

from chondritic meteorites. Earth Planet. Sci. Lett. 7, 251–264.<br />

Ormö J., Lindström M. (2000): When a cosmic impact strikes the seabed.<br />

Geol. Mag. 137, 67–80.<br />

Ormö J., Sturkell E. F. F., Blomqvist G., Törnberg R. (1999): Mutually<br />

constrained geophysical data for the evaluation of a proposed impact<br />

structure: Lake Hummeln, Sweden. Tectonophysics 311, 155–177.<br />

Ostermann M. (1996): Die Geochemie der Impaktschmelzdecke (Sudbury<br />

Igneous Complex) im Multiring-Becken Sudbury. PhD thesis,<br />

Univ. Münster.<br />

Palme H., Goebel E., Grieve R. A. F. (1979): The distribution of volatile<br />

and siderophile elements in the impact melt of East Clearwater (Quebec).<br />

Proc. 10 th Lunar Planet. Sci. Conf., 2465–2492.<br />

Palme H., Grieve R. A. F., Wolf R. (1981): Identification of the projectile<br />

at Brent crater, and further considerations of projectile types at terrestrial<br />

craters. Geochim. Cosmochim. Acta 45, 2417–2424.<br />

Pike R. J. (1980): Formation of complex impact craters: Evidence from<br />

Mars and other planets. Icarus 43, 1–19.<br />

Pilkington M, Grieve R. A. F. (1992): The geophysical signature of terrestrial<br />

impact craters. Rev. Geophys. 30, 161–181.<br />

Pohl J. (1971): On the origin of the magnetization of impact breccias on<br />

Earth. Z. Geophys. 37, 549–555.<br />

Robertson P. B. (1968): La Malbaie Structure, Quebec – A Palaeozoic<br />

Meteorite Impact Site. Meteoritics 4, 1–24.<br />

Robertson P. B. (1973): Shock metamorphism of potassic feldspars. PhD<br />

thesis, Univ. of Durham.<br />

Robertson P. B., Sweeney J. F. (1983): Haughton impact structure: Structural<br />

and morphological aspects. Can. J. Earth Sci. 20, 1134–1151.<br />

Sagy A., Reches Z., Fineberg J. (2002): Dynamic fracture by large extraterrestrial<br />

impacts as the origin of shatter cones. Nature 418,<br />

310–313.<br />

Sharpton V. L., Burke K., Camargo-Zanoguera A., Hall S. A., Lee D. S.,<br />

Marin L. E., Suarez-Reynoso G., Quezaela-Muneton J. M., Spudis P.<br />

D., Urrita-Fucugauchi J. (1993): Chicxulub multiring impact basin:<br />

Size and other characteristics derived from gravity analysis. Science<br />

261, 1564–1567.<br />

Spray J. G., Thompson L. M. (1995): Friction melt distribution in terrestrial<br />

multi-ring impact basins. Nature 373, 130–132.<br />

Spudis P. D. (1993): The Geology of Multi-ring Impact Basins. Cambridge<br />

University Press, Cambridge.<br />

Stöffler D. (1971): Progressive metamorphism and classification of<br />

shocked and brecciated crystalline rocks in impact craters. J. Geophys.<br />

Res. 76, 5541–5551.<br />

Stöffler D. (1972): Deformation and transformation of rock-forming<br />

minerals by natural and experimental shock processes. I. Behavior of<br />

minerals under shock compression. Fortschr. Mineral. 49, 50–113.<br />

Stöffler D. (1974): Deformation and transformation of rock-forming<br />

minerals by natural and experimental shock processes. II. Physical<br />

properties of shocked minerals. Fortschr. Mineral. 51, 256–289.<br />

Stöffler D., Hornemann U. (1972): Quartz and feldspar glasses produced<br />

by natural and experimental shock. Meteoritics 7, 371–394.<br />

Stöffler D., Langenhorst F. (1994): Shock metamorphism of quartz in nature<br />

and experiment: I. Basic observation and theory. Meteoritics 29,<br />

155–181.<br />

Stöffler D., Deutsch A., Avermann M., Bischoff L., Brockmeyer P., Buhl<br />

D., Lakomy R., Müller-Mohr V. (1994): The formation of the Sudbury<br />

Structure, Canada: Towards a unified impact model. Geol. Soc.<br />

Amer. Spec. Pap. 293, 303–318.<br />

Therriault A. M., Fowler A. D., Grieve R. A. F. (2002): The Sudbury Igneous<br />

Complex: A differentiated impact melt sheet. Econ. Geol. 97, in press.<br />

Therriault A. M., Grieve R. A. F., Reimold W. U. (1997): Original size of<br />

the Vredefort Structure: Implications for the geological evolution of<br />

the Witwatersrand Basin. Meteoritics Planet. Sci. 32, 71–77.<br />

Tsikalas F., Gudlaugsson S. T., Faleide J. I. (1998): The anatomy of a<br />

buried complex impact structure: The Mjølnir Structure, Barents<br />

Sea. J. Geophys. Res. 103, 30,469–30,483.<br />

Tsikalas F., Gudlaugsson S. T., Faleide J. I., Eldholm O. (1999): Mjølnir<br />

Structure, Barents Sea: A marine impact crater laboratory. In:<br />

Dressler B. O. and Sharpton V. L. (eds) Large Meteorite Impacts and<br />

Planetary Evolution II. Boulder, Colorado, Geol. Soc. Amer. Spec.<br />

Pap. 339, pp.193–204.<br />

Twiss R. J., Moores E. M. (1992): Structural Geology. W.H. Freeman and<br />

Company, New York.<br />

Vishnevsky S. A., Lagutenko V. N. (1986): The Ragozinka astrobleme:<br />

An Eocene crater in the central Ural. Akad. Nauk. SSSR 14, 1–42.<br />

(in Russian)<br />

von Engelhardt W. (1984): Melt products from terrestrial impact structures.<br />

Proc. 27 th Intern. Geol. Congr. 19, 149–163.<br />

von Engelhardt W. (1990): Distribution, petrography and shock metamorphism<br />

of the ejecta of the Ries crater in Germany – A review.<br />

Tectonophysics 171, 259–273.<br />

von Engelhardt W., Stöffler D. (1968): Stages of Shock Metamorphism<br />

in the Crystalline Rocks of the Ries Basin, Germany. In: French B.<br />

M. and Short N. M. (eds) Shock Metamorphism of Natural Materials.<br />

Mono Book Corp., Baltimore, pp. 159–168.<br />

Wood C. A., Head J. W. (1976): Comparison of impact basins on Mercury,<br />

Mars and the Moon. Proc. 7 th Lunar Sci. Conf., 3629–3651.<br />

Zhang P., Rasmussen T. M., Pedersen L. B. (1988): Electric resistivity<br />

structure of the Siljan impact region. J. Geophys. Res. 93,<br />

6486–6501.<br />

Handling editor: Roman Skála<br />

263


Bulletin of the Czech Geological Survey, Vol. 77, No. 4, 265–282, 2002<br />

© Czech Geological Survey, ISSN 1210-3527<br />

Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

FALKO LANGENHORST<br />

Bayerisches Geoinstitut, University of Bayreuth, Universitätsstr. 30, D-95447 Bayreuth, Germany; e-mail: Falko.Langenhorst@uni-bayreuth.de<br />

Abstract. Minerals show a unique behaviour when subjected to shock waves. The ultradynamic loading to high pressures and temperatures<br />

causes deformation, transformation and decomposition phenomena in minerals that are unequivocal indicators of impact events. This paper introduces<br />

into the basics of shock compression, required to understand the formation and experimental calibration of these shock effects in minerals, and particularly<br />

focuses on the recent advances in the field of shock metamorphism achieved by the application of transmission electron microscopy (TEM).<br />

TEM studies underline that the way minerals respond to shock compression largely depends on their crystal structures and chemical compositions, as<br />

is illustrated here on the basis of four minerals: quartz, olivine, graphite and calcite.<br />

The crystal structure of a mineral exerts an important control on the shock-induced deformation phenomena, comprising one- to two-dimensional<br />

lattice defects, such as dislocations, mechanical twins, planar fractures, and amorphous planar deformation lamellae. For example, dislocations cannot<br />

be activated in quartz due to the strong covalent bonding, whereas the island silicate olivine easily deforms by dislocation glide.<br />

Transformation phenomena include phase transitions to (diaplectic) glass and/or high-pressure polymorphs. TEM studies reveal that high-pressure<br />

polymorphs such as coesite, stishovite and ringwoodite are liquidus phases, which form upon decompression by crystallization from high-pressure<br />

melts. The graphite-to-diamond transition is however a rare example for a solid-state transformation, taking place by a martensitic shear mechanism.<br />

Shock-induced decomposition reactions are typical of volatile-bearing minerals and liberate toxic gases that, in case of large impacts, may affect<br />

Earth’s climate. Shock experiments show that degassing of calcite does not take place under high pressure but can massively occur after decompression<br />

if the post-shock temperature is sufficiently high. A recombination reaction happens however if CaO and CO 2 are not physically separated.<br />

Key words: impact features, shock metamorphism, shock waves, minerals, crystal structure, defects, experimental study<br />

Introduction<br />

Number of craters (>1 km) per km 2<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0.00<br />

4 3 2 1 0<br />

Age (Ga)<br />

Apollo Luna Earth<br />

Fig. 1. Cratering statistics of the Moon-Earth system for the last 4 Ga<br />

years [data from Stöffler and Ryder (2001) and regression curve after<br />

Neukum et al. (2001)].<br />

Collisions of solid bodies played a crucial role in the<br />

formation of Earth and its subsequent evolution. The Proto-Earth<br />

accreted by collisions of planetesimals (Wetherill<br />

1984) and underwent, in Archean times, a very heavy<br />

bombardment (Melosh 1989). The high collision rate at<br />

the beginning of Earth was simply the consequence of the<br />

chaotic states in the orbital movements of early solid bodies.<br />

Subsequently, this early collision rate distinctly declined<br />

until about 2.5 Ga, and since then we have an<br />

essentially constant flux of bodies colliding with Earth<br />

(Fig. 1).<br />

This knowledge of the collision history of Earth is<br />

however not exclusively based on observations of terrestrial<br />

impact craters, as one would possibly expect, but it<br />

largely relies on the cratering history of the Moon (Chapman<br />

and Morrison 1994, Neukum et al. 2001, Stöffler and<br />

Ryder 2001). Due to the absence of an atmosphere (and<br />

hence weathering) and the early cessation of volcanic activity<br />

on Moon, impact craters accumulated, in nearly<br />

undisturbed fashion, particularly in the lunar highlands,<br />

where crater densities are highest. On the geologically<br />

very active Earth, the morphologic landforms of impact<br />

craters easily disappear beyond recognition due to erosion,<br />

plate tectonics and other exogenic and endogenic processes.<br />

Although four Archean impact layers have been reported<br />

(Byerly et al. 2002), we have no knowledge of an<br />

Archean crater (Fig. 2). According to the cratering statistics<br />

(Fig. 1), one would however expect that most craters<br />

formed during this early episode of Earth (Grieve and<br />

Shoemaker 1994).<br />

265


Falko Langenhorst<br />

Identified craters per 1 Ma<br />

1.0<br />

0.5<br />

0.0<br />

Cenozoic (0–65 Ma)<br />

Mesozoic (65–250 Ma)<br />

Impact events are however not only expressed at the<br />

large scale in the form of circular craters but they manifest<br />

also at microscopic scales in minerals. This article focuses<br />

on these microscopic traces in minerals termed shock or<br />

shock-metamorphic effects. Shock effects in minerals are<br />

an unequivocal fingerprint and, often, the last remnants of<br />

impact events. Even if the crater is completely erased,<br />

shocked minerals can be preserved in distal or global ejecta<br />

horizons, still providing evidence of the impact and its<br />

age. In case of the Cretaceous/Tertiary extinction (Alvarez<br />

et al. 1980), the discovery of shocked minerals in the KT<br />

boundary layer increased, for example, the efforts to find<br />

the corresponding impact structure (Bohor et al. 1984 and<br />

1987), subsequently identified as the Chicxulub crater<br />

(Hildebrand et al. 1991).<br />

In the last 10–15 years we have obtained new insights<br />

and a better understanding of the shock behaviour of minerals,<br />

mainly due to the increased use of transmission electron<br />

microscopy (TEM). In terms of spatial resolution, this<br />

technique is well superior to optical microscopy. Therefore,<br />

it became possible to decipher the nature of shock<br />

phenomena already known from optical studies (e.g. planar<br />

deformation features (PDFs)), to discover so-far unknown<br />

effects, and to understand their mechanisms of<br />

Paleozoic (250–570 Ma)<br />

Geological Era<br />

Fig. 2. The number of identified impact craters per Ma plotted for different<br />

eras in the Earth history. Note that despite the high cratering activity<br />

in the Archean we have no knowledge of an impact crater from this<br />

time period (cf. Fig. 1, data from http://gdcinfo.agg.nrcan.gc.ca/crater/<br />

world_craters_e.html).<br />

Proterozoic (570–2500 Ma)<br />

Archean (2500–4550 Ma)<br />

formation and subsequent alteration. This article does not<br />

aim to give a comprehensive review on shocked minerals;<br />

for a more detailed compilation the reader is referred to:<br />

e.g., French and Short (1968), Stöffler (1972, 1974), Stöffler<br />

and Langenhorst (1994), French (1998), Deutsch and<br />

Langenhorst (1998), Langenhorst and Deutsch (1998). It<br />

is merely an attempt to utilize instructive examples to explain<br />

the specific response of minerals to shock compression,<br />

which is distinctly different from the transformation<br />

and deformation behaviour of minerals in other geologic<br />

processes, principally of lower strain rate.<br />

Principles of shock waves<br />

From a basic understanding of shock waves it will be<br />

immediately clear why minerals respond differently to impact<br />

than to other geologic processes. It will also help to<br />

understand the mechanics of the large-scale crater-forming<br />

process (Roddy et al. 1977, Melosh 1989). A shock wave<br />

is an extreme compression wave that propagates with supersonic<br />

velocity, abruptly compresses, heats, and plastically<br />

deforms solid matter. In this respect, shock waves are<br />

fundamentally different from seismic (elastic) waves (e.g.<br />

Duvall and Fowles 1963).<br />

A shock wave is produced by the impact of a high-velocity<br />

projectile or the detonation of an explosive (Roddy<br />

et al. 1977, Asay and Shahinpoor 1993). In these processes,<br />

the time of load is extremely short and, therefore, the<br />

initial stress wave steepens immediately to an almost<br />

atomically sharp discontinuity, which separates highly<br />

compressed from uncompressed material. This so-called<br />

“shock front” represents consequently an infinite discontinuity<br />

in all state parameters (Duvall and Fowles 1963,<br />

Melosh 1989): pressure P, temperature T, specific volume<br />

V, and energy E. Minerals and rocks undergo some kind of<br />

a physical “shock”, because they have to instantly adapt to<br />

extreme P,T conditions with strain rates on the order of 10 6<br />

to 10 9 s -1 . Additionally, a shock wave is very short-lived<br />

with a typical pulse duration of about 1 second for a natural<br />

impact of a 10 km diameter projectile. It is due to these<br />

two time parameters, high strain rates and short shock duration,<br />

that minerals cannot respond by equilibrium reactions<br />

but rather by the activation of fast deformation and<br />

transformation mechanisms.<br />

The unusual properties of shock waves can be illustrated<br />

by a simple train model (Fig. 3). The cartoon shows<br />

the collision of a moving train (projectile) with a standing<br />

train (target). The collision leads to deformation and compression<br />

of the hitting locomotive and the wagons in two<br />

opposite directions. The two boundaries between compressed<br />

and uncompressed wagons are the shock fronts,<br />

which propagate with the shock velocity U. An additional<br />

effect of the compression is material flow in the direction<br />

of the target (see the displacement of the boundary between<br />

the hitting locomotive and the last target wagon).<br />

The material moves behind the shock wave into the target,<br />

266


time<br />

Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

at a particle velocity u p , which is distinctly smaller than<br />

the shock velocity U. As the process proceeds, more and<br />

more wagons are engulfed by the compression until the<br />

shock wave in the projectile is reflected as rarefaction<br />

wave at the free rear side of the projectile (Fig. 3). This<br />

leads to a backward acceleration of wagons and represents<br />

the ejection of material in natural impact cratering. As the<br />

rarefaction wave is propagating in already compressed<br />

material, it is faster than the shock wave in the target. Consequently,<br />

the shock wave will be overtaken at some depth<br />

by the rarefaction wave, i.e. it will decay.<br />

Both shock U and particle u p velocities characterise the<br />

state of material under shock compression and are related<br />

to pressure P, density ρ, and energy E via the Hugoniot<br />

equations (Duvall and Fowles 1963, Melosh 1989):<br />

ρ 0 U = ρ (U – u p ) (1)<br />

P – P 0 = ρ 0 U u p (2)<br />

P u p = ρ 0 U (u p2 /2) + ρ 0 U (E – E 0 ) (3)<br />

These equations express the conservation of mass, momentum,<br />

and energy across the shock front. By combining<br />

equations (1) to (3) the velocities can be eliminated to give<br />

the so-called Rankine-Hugoniot equation:<br />

E – E 0 = (P – P 0 ) (V 0 – V)/2 with V = 1/ρ (4)<br />

This is an equation of state, describing the physical<br />

states that can be achieved by shock waves of variable intensity<br />

in any solid. Commonly, the shock equation of state<br />

of a particular material is experimentally determined by<br />

measuring particle u p and shock U velocities. Since the zero-pressure<br />

densities ρ 0 of minerals are usually well known,<br />

pressures P and densities ρ (or specific volumes V) can be<br />

calculated with the above mentioned Hugoniot equations.<br />

As is usual for all equations of state, the shock data of<br />

a particular material are displayed in a pressure-specific<br />

volume plot, defining the so-called Hugoniot curve. The<br />

Hugoniot curves of rock-forming minerals are commonly<br />

characterised by kinks, i.e., discontinuous changes in the<br />

curve slope. One discontinuity is typically at 5–10 GPa,<br />

the so-called Hugoniot elastic limit, which is the yield<br />

strength of the material. Discontinuities at higher pressures<br />

were generally assumed to result from phase transformations<br />

to high-pressure polymorphs (e.g. Ahrens and<br />

Rosenberg 1968, Jackson and Ahrens 1979, Ahrens et al.<br />

1976, Melosh 1989). This interpretation is certainly correct<br />

for materials such as iron and graphite, which undergo<br />

martensitic (displacive) transformations, fast enough to<br />

take place under shock compression. It is however an inappropriate<br />

assumption for silicate materials such as<br />

quartz, feldspars and olivine. These minerals can only be<br />

converted into high-pressure polymorphs via reconstructive<br />

mechanisms. These mechanisms are however too<br />

time-consuming to occur under shock compression.<br />

Therefore, shock experiments on these minerals usually<br />

result in the formation of dense (diaplectic) glass with possibly<br />

five- and/or six-fold coordinated silicon or finely recrystallised<br />

low-pressure phase.<br />

U<br />

V 0 , E 0<br />

P 0 , T 0<br />

U<br />

P 1 , T 1<br />

V 1 , E 1<br />

shock front<br />

u p<br />

material<br />

front<br />

u p<br />

Experimental simulation<br />

shock front<br />

rarefaction front<br />

ejection<br />

Fig. 3. A train model illustrating the formation and propagation of shock<br />

and rarefaction waves and the associated material movement (from Langenhorst<br />

2000).<br />

Shock experiments are indispensable for understanding<br />

the formation of shock effects in minerals, because<br />

they are performed under controlled laboratory conditions<br />

and because they provide shocked minerals in their initial,<br />

unmodified state. This has two basic advantages. First,<br />

shock effects can be calibrated as function of pressure (or<br />

other factors) and the resultant barometers can then, with<br />

some care, be applied to nature (Stöffler and Langenhorst<br />

1994). Secondly, the study of pristine shock effects helps<br />

to understand possible post-shock modifications (annealing,<br />

chemical alteration etc.) in the natural impact environment<br />

(Grieve et al. 1996).<br />

The major disadvantage of shock experiments is their<br />

short pulse duration (< 1µs), which is at least 6 orders of<br />

magnitude shorter than the pressure pulse in a natural<br />

bolide impact (~ 1 s, Langenhorst et al. 2002a). The short<br />

pulse duration in an experiment is simply the result of the<br />

small size of the projectile. The pressure duration corresponds<br />

approximately to the time that a shock wave needs<br />

to travel through the projectile and back (i.e. 2 projectile<br />

diameters, Fig. 3).<br />

It may be some surprise, but pioneering, experimental<br />

work on impact processes started with Alfred Wegener<br />

(1921), the founder of plate tectonics. He simulated impact<br />

events by throwing cement powder in a tablespoon<br />

with his bare hand onto a target, which was also composed<br />

of cement powder. He was able to produce circular craters<br />

with central uplifts, resembling in shape and proportions<br />

those observed on the Moon. Since then, a large variety of<br />

sophisticated shock techniques has been developed to<br />

simulate cratering mechanics and shock metamorphism of<br />

267


Falko Langenhorst<br />

a<br />

b<br />

booster<br />

laser<br />

d<br />

D<br />

high explosive<br />

flyer plate<br />

spacing ring<br />

cover layer<br />

specimen<br />

Fe cylinder<br />

vacuum<br />

chamber<br />

lens<br />

mirrors<br />

sample<br />

steel blocks<br />

Fig. 4. Two different experimental designs to produce shock waves: (a) High-explosive device used at the Ernst-Mach-Institut, Efringen-Kirchen, Germany<br />

(modified after Langenhorst and Deutsch 1994) and (b) a laser irradiation setup with the sample being clamped into an Al block. The laser is focused<br />

on a thin Al foil, acting as flyer plate (modified after Langenhorst et al. 1999a and 2002a).<br />

minerals (e.g. French and Short 1968, Roddy et al. 1977,<br />

Asay and Shahinpoor 1993, Davison et al. 2002).<br />

Most experiments related to cratering mechanics employ<br />

spherical projectiles that produce and excavate some<br />

crater in an infinite half space medium via a spherically<br />

expanding shock wave. In contrast, most experiments related<br />

to shock metamorphism employ flat-plate projectiles<br />

that drive a planar shock wave through a similarly planar<br />

target; the thickness of the latter is typically less than projectile<br />

thickness to avoid measurable pressure decay<br />

across the sample (Barker et al. 1993). The target is usually<br />

a metallic container, encapsulating the sample in a fashion<br />

to allow its partial or complete recovery. The<br />

experiments may differ widely in the type of accelerating<br />

system (Fig. 4): powder and light-gas guns, high-explosive<br />

charges, electric discharge guns, and laser irradiation techniques<br />

have been used and tested to successfully reproduce<br />

shock effects in minerals (e.g. Milton and DeCarli<br />

1963, Müller and Hornemann 1969, Gratz et al. 1992,<br />

Stöffler and Langenhorst 1994, Langenhorst et al. 1999a,<br />

2002a). The principle of the electric gun is to vaporise a<br />

thin metal foil by rapid electric discharge of a capacitor.<br />

The high electrical current leads to the instantaneous vaporisation<br />

of the foil and the production of a shock wave<br />

(Langenhorst et al. 2002a). Laser irradiation experiments<br />

can be performed either with or without a projectile. In the<br />

latter case, the beam is directly focused on the sample surface<br />

(Langenhorst et al. 2002a), whereby the absorption of<br />

the laser energy generates rapidly exploding plasma that<br />

subsequently induces a shock wave. Such plasma techniques<br />

are capable to produce the highest shock pressures<br />

with an unbelievable world record of 750 Mbar (Cauble et<br />

al. 1993). However, higher shock pressures are commonly<br />

at the expense of shorter shock durations and smaller sample<br />

volumes. This is because higher impact velocities and<br />

hence higher shock pressures can only be achieved by reducing<br />

the size and weight of the projectile. Typical pressure<br />

pulses in laser irradiation, electric discharge and highexplosive<br />

experiments are approximately 1 ns, 10 ns, and<br />

1 µs with shocked sample volumes on the order of 0.1, 1,<br />

and 100 mm 3 , respectively (Langenhorst et al. 2002a).<br />

To determine pressures in shock experiments it is necessary<br />

to measure the velocities associated with the shock<br />

waves (projectile v, particle u and shock U velocities), using<br />

electrical pin contact, optical interferometry (VISAR)<br />

or similar techniques (Hornemann and Müller 1971, Barker<br />

et al. 1993). It is then possible to calculate the pressures<br />

temperature (°C)<br />

3000<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

quartz<br />

eclogite<br />

liquid<br />

coesite<br />

release<br />

paths<br />

stishovite<br />

porous sandstone<br />

0<br />

1 10<br />

100<br />

shock pressure (GPa)<br />

Fig. 5. Phase diagram depicting the pressure-temperature conditions<br />

reached in quartz, olivine (solid line), and porous sandstone (long dashed<br />

line) by shock compression (data after Wackerle 1962, Kieffer et al.<br />

1976, and Holland and Ahrens 1997). The release paths hold for porous<br />

quartz, which first melts on loading and then solidifies on cooling as coesite<br />

or stishovite. The equilibrium phase boundaries between quartz, coesite,<br />

stishovite and liquid are drawn as dashed lines. The grey<br />

pressure-temperature field represents the conditions reached in regional<br />

metamorphism; the solid line within this field depicts the pressure-temperature<br />

path of a diamond-bearing gneiss (Stöckhert et al. 2001).<br />

quartz<br />

olivine<br />

268


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

by aid of the known Hugoniot relations of the materials involved<br />

(Marsh 1980). Thus, the determination of shock<br />

pressure is fundamentally different from the approach<br />

used in static compression techniques (e.g. piston cylinder,<br />

multi-anvil or diamond anvil cell experiments), in which<br />

the pressures are calibrated via properties and/or phase<br />

transformations of reference materials (e.g. Rubie 1999).<br />

The determination of temperatures is a less precise and<br />

more difficult undertaking. So far, pyrospectrometric techniques<br />

have been used to measure both shock and postshock<br />

temperatures (e.g. Raikes and Ahrens 1979, Holland<br />

and Ahrens 1997; Fig. 5). The knowledge of temperatures<br />

is however of fundamental importance for the determination<br />

of the melting temperatures of materials at very high<br />

pressures. For example, the precise measurement of the<br />

melting curve of iron enables the temperature at the<br />

boundary between Earth’s inner (solid) and outer (liquid)<br />

core to be determined (Brown and McQueen 1986), an<br />

important fix point of the geotherm.<br />

calcite graphite olivine quartz<br />

shock effects<br />

mechanical twins<br />

PF<br />

PDF<br />

mosaicism<br />

diaplectic glass \ lechatelierite<br />

stishovite \ coesite<br />

dislocations<br />

PF and PDF<br />

mosaicism<br />

recrystallisation and staining<br />

ringwoodite<br />

kink bands<br />

diamond<br />

mechanical twins<br />

dislocations<br />

decomposition and recombination products<br />

shock pressure (GPa)<br />

0 10 20 30 40 50 60<br />

0 10 20 30 40 50 60<br />

Shock effects in minerals<br />

Fig. 6. Compilation of the pressure intervals over which certain shock effects<br />

are formed in quartz, olivine, graphite and calcite (modified after<br />

Stöffler and Langenhorst 1994 and Langenhorst and Deutsch 1998). The<br />

diagram is based on shock experiments with non-porous samples.<br />

Shock-induced physical and chemical changes in minerals<br />

are collectively called shock effects or shock-metamorphic<br />

effects. This term is relatively broad and covers<br />

any type of shock-induced change, such as formation of<br />

lattice defects, phase transformations, decomposition reactions<br />

and resultant changes in physical and chemical properties.<br />

A great diversity of natural shock effects has been<br />

described on the basis of spectroscopic techniques, X-ray<br />

diffraction, and optical microscopy (Hörz and Quaide<br />

1973, Stöffler 1972 and 1974, Schneider 1978, Boslough<br />

et al. 1989). The physical nature of some of the effects was<br />

not completely clear and others were not even known until<br />

TEM was used to characterize the mineralogical effects<br />

at the nanometer scale. Based on TEM observations, the<br />

following simple subdivision of shock effects and processes<br />

occurring during shock compression and decompression<br />

has been proposed (Langenhorst and Deutsch 1998):<br />

(1) Deformation: formation of dislocations, planar microstructures<br />

(planar fractures and planar deformation<br />

features), mechanical twins, kink bands, and mosaicism<br />

(2) Phase transformations into high-pressure phases and<br />

diaplectic glass<br />

(3) Decomposition into a solid residue and a gaseous<br />

phase<br />

(4) Melting and vaporisation of entire mineral (subsequently<br />

quenched as shock-fused glass or polycrystalline<br />

aggregates)<br />

This simple list does not contain the post-shock thermal<br />

effects forming distinctly after the impact. In a wide<br />

sense, these effects may also be regarded as shock indicators,<br />

although they are not primary effects of shock compression<br />

and decompression and simply result from the<br />

high temperatures prevailing after an impact event. Examples<br />

of diagnostic post-shock effects are the formation of<br />

Ballen quartz and checkerboard feldspars in impact melts<br />

(Carstens 1975, Bischoff and Stöffler 1984) or the crystallization<br />

of highly oxidised Ni-rich spinels (magnesioferrites)<br />

in microtektites (Robin et al. 1992).<br />

We will focus here exclusively on the shock effects<br />

listed above. The deformation and transformation effects<br />

largely form during the compression phase of shock<br />

waves, whereas decomposition, melting and vaporisation<br />

are temperature dominated processes, which take place<br />

during the decompression phase when pressure declines to<br />

a larger extent than temperature. There is no single mineral<br />

that shows all of these effects (Fig. 6). The response of<br />

a mineral to shock compression largely depends on its<br />

crystal structure and composition. A mineral with strong<br />

three-dimensional covalent bonding between polyhedra in<br />

the crystal structure can, for example, not respond by dislocation<br />

glide. Also, minerals without volatile components<br />

cannot respond by decomposition reactions such as dehydration<br />

and decarbonation. To highlight these aspects and<br />

to illustrate the various types of shock effects in minerals<br />

we will concentrate on four minerals, differing largely in<br />

terms of crystal structure and composition.<br />

Quartz<br />

Among the rock-forming minerals, quartz (SiO 2 ) displays<br />

the widest variety of shock effects. It possesses a<br />

crystal structure, consisting of three-dimensionally linked,<br />

corner-shared SiO 4 -tetrahedra with strong covalent bonding.<br />

In the low-pressure regime (< 30 GPa), quartz can<br />

269


Falko Langenhorst<br />

(101 — 1)<br />

a<br />

50 µm<br />

0.2 µm<br />

b<br />

(0001)<br />

c<br />

0.5 µm<br />

100 nm<br />

d<br />

Fig. 7. Shock effects in quartz: (a) Optical micrograph of planar deformation features in shocked quartz in a garnet gneiss from the Popigai crater,<br />

Siberia, (b) bright-field TEM image of fresh, amorphous PDFs in shocked quartz from the Massignano outcrop, Italy (see Langenhorst 1996), (c) Darkfield<br />

TEM image of a mechanical Brazil twin in shocked quartz from the Mjølnir crater, Barents Sea and (d) dark-field scanning TEM image of numerous<br />

stishovite needles embedded in a glassy shock vein of the Zagami meteorite (see Langenhorst and Poirier 2000).<br />

therefore not react by dislocation glide, particularly because<br />

the activation and glide of dislocations is controlled<br />

by the diffusion of water-related defects (see McLaren<br />

1991 for “hydrolytic weakening”), which is a rather slow<br />

process compared to the short time frame of shock compression.<br />

Instead, quartz develops so-called planar microstructures<br />

(Grieve et al. 1990), which comprise planar fractures<br />

(PF), planar deformation features (PDFs), and mechanical<br />

twins (Stöffler and Langenhorst 1994). The mechanical<br />

twins were previously also regarded as PDFs, because a distinction<br />

on the basis of optical microscopy is impossible.<br />

PFs can basically be regarded as high-pressure cleavage<br />

planes that are only activated by dynamic shock loading;<br />

they are relatively widely spaced (> 5–20 µm). On the<br />

contrary, quartz has no cleavage at ambient conditions and<br />

fails by a glassy-like parting.<br />

PDFs were previously also called planar elements (Engelhardt<br />

and Bertsch 1969) and are thin (< 200–300 nm)<br />

amorphous lamellae with the same composition as the host<br />

crystal (Fig. 7b); at the TEM scale, their spacing (< 1 µm)<br />

is much smaller than that of PFs (Müller 1969, Ashworth<br />

and Schneider 1985, Gratz et al. 1992, Langenhorst 1994).<br />

PDF orientations are crystallographically controlled and<br />

show a pressure-dependent variation. Most PDFs are oriented<br />

parallel to rhombohedral planes such as {101 — 3} and<br />

{101 — 2}; other less abundant PDF orientations are given in<br />

Table 1 (see also Stöffler and Langenhorst 1994). In the<br />

low-pressure regime (< 30 GPa), PDF orientations are regarded<br />

as the most robust barometer, because even postshock<br />

annealing and alteration merely converts the PDFs<br />

into the “decorated” type (French 1969; Fig. 7a), but PDF<br />

orientations remain unchanged. The decoration of PDFs in<br />

naturally shocked quartz is due to recrystallization of the<br />

270


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

Table 1. Compilation of the most abundant PDF orientations in shocked quartz.<br />

Miller indices {h k i l} Azimuth angle Pole distance ρ Symbol Form<br />

(0001) – 0° c basal pinacoid<br />

{101 — 3}, {011 — 3} p, n 30° 22.95° ω, ω’ rhombohedron<br />

{101 — 2}, {011 — 2} p, n 30° 32.42° π, π’ rhombohedron<br />

{101 — 1}, {011 — 1} p, n 30° 51.79° r, z rhombohedron<br />

{101 — 0} 30° 90° m hexagonal prism<br />

{404 — 1v, {044 — 1} p, n 30° 78.87° t rhombohedron<br />

{516 — 0}, {61 — 5 — 0} r, l 40° 90° k ditrigonal prism<br />

{516 — 1}, {61 — 5 — 1} p, n, r, l 40° 82.07° x trigonal<br />

{61 — 5 — 1}, {156 — 1} trapezoedron<br />

{314 — 1}, {43 — 1 — 1} p, n, r, l 45° 77.91° – trigonal<br />

{41 — 3 — 1}, {134 — 1} trapezoedron<br />

{213 — 1}, {32 — 1 — 1} p, n, r, l 50° 73.71° – trigonal<br />

{31 — 2 — 1}, {123 — 1} trapezoedron<br />

{112 — 2}, {21 — 1 — 2} r, l 60° 47.73° ξ trigonal dipyramid<br />

{112 — 1}, {21 — 1 — 1} r, l 60° 65.56° s trigonal dipyramid<br />

{112 — 0}, {21 — 1 — 0} r, l 60° 90.0° a trigonal prism<br />

{224 — 1}, {42 — 2 — 1} r, l 60° 77.20° trigonal dipyramid<br />

p = positive, n = negative, r = right, l = left<br />

amorphous lamellae and resultant exsolution<br />

of water. The exsolution of water can<br />

be attributed to the different solubility of<br />

water in silica glass and crystalline quartz.<br />

Therefore, the fluids mobilized in impact<br />

events can initially be dissolved in the<br />

amorphous PDFs but subsequent recrystallization<br />

of the glass expels the water in<br />

form of tiny voids (Goltrant et al. 1991,<br />

1992, Leroux and Doukhan 1996, Grieve et<br />

al. 1996, Langenhorst and Deutsch 1998).<br />

Sub-planar features of tectonic origin<br />

such as the so-called Böhm lamellae have<br />

been erroneously assigned as PDFs (Ernstson<br />

et al. 1985, Vrána 1987; Fig. 8a). TEM<br />

studies have deciphered the nature of these<br />

tectonic features, as being subgrain boundaries<br />

(Cordier et al. 1994, Langenhorst and<br />

Deutsch 1996, Joreau et al. 1997a). Subgrain<br />

boundaries are dislocation arrays,<br />

separating a crystal into sub-grains that are<br />

slightly tilted (< 5°) with respect to each other. They are<br />

the result of slow plastic deformation and recovery of deformed<br />

quartz in a tectonic environment. The deformation<br />

proceeds by the activation and emission of dislocations<br />

coupled with simultaneous or subsequent climb and recovery<br />

of dislocations into sub-grain boundaries (Fig. 8b;<br />

Poirier 1985). Water can easily penetrate along these internal<br />

boundaries leading to a decoration with water bubbles.<br />

A careful optical inspection of the suspected planar<br />

features in quartz will immediately reveal whether they are<br />

of endogenic or exogenic origin. Sub-grain boundaries<br />

show a sub-parallel arrangement but are not perfectly planar,<br />

unlike shock-produced PDFs. The spacing of tectonic<br />

features (usually > 5–10 µm) is larger than that of shockproduced<br />

PDFs (< 1 µm). Under crossed Nicols, the tectonically<br />

deformed quartz grains show undulatory extinction<br />

and can exhibit an anomalous optic axial angle of up<br />

to 10°. In contrast, shocked quartz usually shows a patchy<br />

extinction pattern, with extinct areas in different parts of<br />

the crystal. This behaviour is the so-called mosaicism. In a<br />

tectonic source rock, not all of the deformed quartz grains<br />

contain sub-grain boundaries in a sub-planar arrangement;<br />

many quartz grains may even be devoid of sub-grain<br />

boundaries. On the other hand, it would be rather unusual<br />

for an impact rock that only one or few quartz grains contain<br />

PDFs. If the quartz grains have experienced the same<br />

pressure-temperature conditions, then at least most of<br />

them should contain PDFs (Hörz 1968, Engelhardt and<br />

Bertsch 1969). The size and orientation of quartz grains<br />

with respect to the shock front has however an influence<br />

on the development of PDFs (Walzebuck and Engelhardt<br />

a<br />

50 µm 3 µm<br />

b<br />

Fig. 8. (a) Optical micrograph and (b) corresponding bright-field TEM image of subgrain boundaries in a tectonically deformed quartz from Azuara,<br />

Spain (see Langenhorst and Deutsch 1996).<br />

271


Falko Langenhorst<br />

1979). These effects may therefore be responsible for a<br />

heterogeneous PDF distribution throughout quartzose<br />

rocks, particularly at pressures required for incipient PDF<br />

formation (~ 10 GPa).<br />

Another important outcome of TEM studies was the<br />

discovery of mechanical Brazil twins (Leroux et al. 1994),<br />

which are exclusively oriented parallel to the basal plane<br />

(0001). Numerous partial dislocations in the twin boundaries<br />

indicate the mechanical nature of the twins (Fig. 7c).<br />

In contrast, the commonly known Brazil twins are hydrothermally<br />

grown twins, which lack partial dislocations<br />

and are always oriented parallel to (101 — 1) (McLaren and<br />

Pithkethly 1982, Langenhorst and Poirier 2002). Static deformation<br />

experiments have shown that a high shear stress<br />

of 3 to 4 GPa has to be applied to generate the mechanical<br />

Brazil twins parallel to (0001) (McLaren et al. 1967). In<br />

crustal rocks, such conditions are only met by an impact<br />

event. At the optical scale, the mechanical twins become<br />

only visible if they are decorated with water bubbles. The<br />

mechanical Brazil twins are more resistant to post-shock<br />

annealing and alteration. Long-term regional metamorphism<br />

with complete recrystallization of the shocked rock<br />

or melting is probably the only way to erase them. At the<br />

Vredefort structure, most of the PDFs in shocked quartz<br />

are lost by post-shock overprint but the Brazil twins are<br />

still present. This discovery could explain why the PDF<br />

orientations indicate an apparent decrease in shock pressure<br />

toward the center of the crater (Schreyer 1983, Nicolaysen<br />

and Reimold 1990).<br />

In the high-pressure regime (> 25–30 GPa), quartz is<br />

dominated by phase transformations either to diaplectic<br />

glass or to high-pressure polymorphs. Diaplectic glass is a<br />

densified glass, preserving the shape and sometimes even<br />

internal features of precursor grains (Engelhardt et al.<br />

1967, Stöffler 1984). X-ray studies demonstrate that the<br />

transition from crystalline quartz to diaplectic glass is<br />

marked by increasing mosaicism coupled with decreasing<br />

domain size (Dachille et al. 1968, Hörz and Quaide 1973,<br />

Hanns et al. 1978). The transition is also characterized by<br />

a decrease of refractive indices and densities but diaplectic<br />

glass has a density about 5% higher than that of synthetic<br />

silica glass (Langenhorst and Deutsch 1994). It is<br />

still not fully understood on whether the transformation<br />

represents a solid-state collapse (Stöffler 1984) or quenching<br />

of a liquid under high pressure (Langenhorst 1994).<br />

The latter seems to be more likely because diaplectic glass<br />

is often associated with coesite that crystallized from highpressure<br />

melt (see below).<br />

Solid-state phase transformations of quartz into the<br />

high-pressure polymorphs coesite and stishovite are reconstructive.<br />

This means that time is needed for the cooperative<br />

movement and diffusion of atoms. Consequently,<br />

the short duration of shock compression impedes a direct<br />

solid-state transformation of quartz into the high-pressure<br />

polymorphs. Instead, quartz has to be melted under high<br />

pressure and the high-pressure phases can then readily<br />

crystallize during the decompression phase (see release<br />

paths in Fig. 5). The latter means that coesite and<br />

stishovite form in their stability fields at pressures, which<br />

are smaller than the maximum shock pressure achieved<br />

during the compression phase. The very high temperatures<br />

needed for melting are reached locally within shocked<br />

rocks, e.g. at pores or along shock veins, which result from<br />

shear-induced frictional melting (Kieffer et al. 1976, Langenhorst<br />

and Poirier 2000). The crystallization of highpressure<br />

polymorphs from melt is much faster than a<br />

reconstructive solid-state transformation because the<br />

liquidus temperatures at high pressures are well above<br />

2000°C, where kinetics is very fast. Furthermore, it is<br />

known from NMR studies that compressed silicate melts<br />

contain five- and six-fold coordinated silicon (Xue et al.<br />

1989, Stebbins and Poe 1999), which should facilitate per<br />

se the crystallization of high-pressure polymorphs from<br />

such dense melts. Indeed, coesite and stishovite have always<br />

been found in conjunction with silica or rock glass<br />

(Kieffer et al. 1976, White 1993, Leroux et al. 1994, Langenhorst<br />

and Poirier 2000). Polycrystalline coesite aggregates<br />

occur in diaplectic silica glass (Hörz 1965, Stöffler<br />

1971), e.g. at the Ries and Popigai craters. The rapid crystallization<br />

led to the formation of numerous (100) rotation<br />

twins with the composition plane (010) (Leroux et al.<br />

1994, Grieve et al. 1996). These twins are known to represent<br />

growth defects (Bourret et al. 1986). Stishovite has<br />

been identified in the porous Coconino sandstone from the<br />

Barringer crater (Kieffer et al. 1976), in thin pseudotachylite<br />

veins in shocked basement rocks of the Vredefort<br />

structure, South Africa (Martini 1991, White 1993), as<br />

well as in shock veins in the Martian meteorite Zagami<br />

(Fig. 7d, Langenhorst and Poirier 2000). Recently, the discovery<br />

of post-stishovite polymorphs has been reported, as<br />

well (ElGoresy et al. 2000), but the high beam sensitivity<br />

prevented a thorough characterization of the phases (Sharp<br />

et al. 1999).<br />

Olivine<br />

Olivine is an island silicate with isolated SiO 4 tetrahedra<br />

that are joined by divalent cations. As a consequence,<br />

it is easier than in quartz to break bonds and to activate and<br />

move dislocations in this crystal structure. In the low-pressure<br />

regime, olivine deforms thus by dislocation glide. The<br />

Burgers vector b is always [001] (= c) but depending on<br />

the orientation of the olivine to the shock front, various<br />

slip planes can be activated, all sharing the c direction as<br />

zone axis (Fig. 9a): (010), (100), (110), (hk0) and symmetrically<br />

equivalent planes (Ashworth and Barber 1975,<br />

Langenhorst et al. 1995, Leroux et al. 1996, Joreau et al.<br />

1997b, Langenhorst and Greshake 1999). The dislocation<br />

densities can be as high as 2 × 10 14 m –2 (Madon and Poirier<br />

1983, Langenhorst et al. 1995, Leroux 2001), even in experimentally<br />

shocked olivine (Langenhorst et al. 1999a<br />

and 2002b). Experiments have also demonstrated that dislocations<br />

can propagate at approximately half the sound<br />

velocity (~ 3 km/s, Langenhorst et al. 1999a) and are<br />

272


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

0.5 µm 0.2 µm<br />

a b<br />

Fig. 9. (a) Dark-field TEM image of numerous c dislocations and a planar fracture in a shocked olivine grain from the Tenham chondrite, (b) Brightfield<br />

TEM image of ringwoodite in a shock vein of the ordinary chondrite Acfer 90072. Note the numerous stacking faults parallel to {110} planes.<br />

probably hampered to reach this maximum speed due to<br />

interactions between themselves. The sources of dislocations<br />

seem to be the tips of forming planar fractures. When<br />

the planar fractures are formed, a high stress field is created<br />

at their tips, from which the dislocations are first emitted<br />

as loops. Since the edge component is distinctly faster<br />

than the screw component, the dislocations become very<br />

straight during further propagation, with the long dislocation<br />

line representing the screw segment.<br />

The PFs in olivine are indicative of shock as they are<br />

oriented parallel to rational crystallographic planes that<br />

are not known as normal cleavage planes of olivine. The<br />

PFs are typically parallel to low index planes (Müller and<br />

Hornemann 1969, Snee and Ahrens 1975, Reimold and<br />

Stöffler 1978, Bauer 1979, Langenhorst et al. 1995, Stöffler<br />

et al. 1991): (100), (010), (001), (130), and (110), belonging<br />

to pinacoidal and prismatic forms. Both, the<br />

internal fragmentation of olivine by fracturing and the<br />

high density of dislocations contribute to the patchy extinction<br />

behaviour under crossed Nicols, known as mosaicism.<br />

Observations on naturally and experimentally shocked<br />

olivine reveal that mosaicism increases distinctly in the<br />

high-pressure regime (> 25–30 GPa; Carter et al. 1968,<br />

Müller and Hornemann 1969, Snee and Ahrens 1975,<br />

Reimold and Stöffler 1978, Bauer 1979, Schmitt 2000).<br />

Additional shock phenomena in this pressure regime are<br />

staining, recrystallization, and transformation into the<br />

high-pressure polymorphs wadsleyite and ringwoodite<br />

(Fig. 9b). Although the production of glass has been reported<br />

in one experimental study (Jeanloz et al. 1977), diaplectic<br />

or shock-fused glasses are generally unknown for<br />

olivine.<br />

Brownish staining of olivine has been described for experimentally<br />

and naturally shocked olivine (Stöffler et al.<br />

1991, Schmitt 2000) but the reason for the colorization is<br />

unknown; a careful microstructural characterization is certainly<br />

required to unravel the nature of the staining effect.<br />

The term recrystallisation decribes the formation of<br />

fine-grained olivine (1–2 µm) aggregates, mostly in the<br />

vicinity of shock veins. Shock-induced recrystallisation of<br />

olivine is generally assumed to be a solid-state process<br />

(Carter et al. 1968, Ashworth and Barber 1975, Bauer<br />

1979, Stöffler et al. 1991). This interpretation actually<br />

makes sense, because recovery and polygonization of dislocations<br />

are expected to occur at elevated post-shock temperatures,<br />

resulting in fine-grained strain-free olivine<br />

aggregates with lattice-preferred orientation of subgrains<br />

(Lally et al. 1976).<br />

In analogy to quartz, the high-pressure transformations<br />

require first the production of an olivine melt, which has<br />

then to rapidly crystallize upon decompression as highpressure<br />

polymorphs. Such rapid crystallization is so far<br />

only manifested in shock veins of ordinary chondrites, i.e.<br />

in thin localized melt zones that are first formed by shearinduced<br />

frictional heating (Stöffler et al. 1991) and are<br />

then rapidly quenched by the surrounding cold host rock.<br />

The shock veins contain aggregates of tiny (~1 µm)<br />

wadsleyite and ringwoodite grains, together with other<br />

high-pressure phases (Binns et al. 1969, Putnis and Price<br />

1979, Madon and Poirier 1983, Langenhorst et al. 1995,<br />

Chen et al. 1996). Olivine melt crystallizing at pressures<br />

beyond the ringwoodite stability field yields for example<br />

an assemblage of silicate perovskite and magnesiowüstite<br />

(Sharp et al. 1997, Tomioka and Fujino 1997), the expected<br />

mineral assemblage in the Earth’s lower mantle. The<br />

coexistence of high-pressure phases that should not coexist<br />

under equilibrium indicates that the crystallization of<br />

shock veins probably happened during decompression, i.e.<br />

the phases crystallized progressively from the melt while<br />

the pressure declined.<br />

In ordinary chondrites, wadsleyite and ringwoodite are<br />

both characterized by numerous stacking faults. Wadsleyite<br />

develops faults parallel to the (010) plane (Madon<br />

and Poirier 1983) and ringwoodite parallel to {110} planes<br />

(Fig. 9b; Langenhorst et al. 1995, Chen et al. 1996). The<br />

273


Falko Langenhorst<br />

a<br />

Graphite<br />

b<br />

Diamond<br />

A<br />

C<br />

A<br />

c<br />

B<br />

B<br />

<br />

C<br />

C<br />

A<br />

Fig. 10. Crystal structures of (a) hexagonal (ABAB…) graphite and (b) cubic (ABCABC…) diamond showing the different stacking sequences in the<br />

[0001] and [111] directions, respectively.<br />

stacking faults are regarded as growth defects, i.e. they are<br />

an inevitable result of the short crystallization times. The<br />

time for crystallisation of shock veins depends primarily<br />

on the thickness of the vein and the initial temperature difference<br />

between vein and adjacent cold host rock. Recent<br />

calculations suggest that the times for crystallisation of<br />

high-pressure polymorphs are much shorter than the shock<br />

duration (Langenhorst and Poirier 2000).<br />

In terrestrial impact rocks, the high-pressure polymorphs<br />

of olivine have not been found yet, probably because<br />

of the lack of well-preserved, olivine-rich target<br />

rocks with thin shock veins.<br />

Graphite<br />

Graphite is composed of pure carbon and is an example<br />

of a mineral with a pronounced sheet structure (Fig.<br />

10a). Within the sheets, carbon atoms form hexagonal<br />

rings with very strong covalent bonds. Weak van-der-<br />

Waals bonding prevails between the sheets. For such a layered<br />

structure, it is characteristic to react to shock<br />

compression by kink or twin operations. In the low-pressure<br />

regime (< 25–30 GPa), graphite is usually assumed to<br />

kink but there are no exact measurements of rotation angles<br />

between different parts of shock-folded graphite crystals<br />

to exclude the possibility of twin operations (Stöffler<br />

1972).<br />

Graphite is of most interest for its phase transformation<br />

to diamond. In the context of impact events, the formation<br />

of diamond is a rare example for a solid-state transformation.<br />

Another mineralogical example for such a transformation<br />

is the conversion of zircon into the scheelite<br />

structured high-pressure polymorph reidite (Glass et al.<br />

2002). Two atomic operations are necessary to convert<br />

graphite into diamond (Fig. 10). The hexagonal carbon<br />

layers have to be brought together by compression along<br />

the c axis and the stacking sequence has to be changed<br />

from a hexagonal to a cubic array by shearing of the<br />

hexagonal carbon layers in their a-a plane. Neglecting the<br />

van-der-Waals bonds, it is not necessary to break any<br />

strong bonds within the layers. Such shear-induced transitions<br />

are called martensitic transformations. Even in short<br />

shock experiments it is possible to produce this martensitic<br />

transformation; this was first demonstrated by DeCarli<br />

and Jamieson (1961).<br />

In light of this first shock synthesis, the long-known diamonds<br />

in iron meteorites and ureilites (Erofeev and<br />

Lachinov 1888, Foote 1891) were reinterpreted as shock<br />

products (Lipschutz 1964). On Earth, a large number of<br />

impact craters and the K/T boundary are yet known as find<br />

locations of impact diamonds (Masaitis et al. 1990 and<br />

2000, Valter et al. 1992, Koeberl et al. 1997, Hough et al.<br />

1997). They were first discovered in the 100 km sized<br />

Popigai structure, Siberia (Masaitis et al. 1972), which is<br />

regarded as largest diamond deposit on Earth (Deutsch et<br />

al. 2000). In the Ries, impact diamonds were already<br />

found in 1978 by Rost et al., but it was not before 1995<br />

that this finding was noticed by a broader scientific community<br />

(Masaitis et al. 1995, Hough et al. 1995, El Goresy<br />

et al. 2001).<br />

The source rocks of diamonds from terrestrial impact<br />

craters are usually graphite-bearing gneisses or other crystalline<br />

rocks (Masaitis et al. 1990, ElGoresy et al. 2001).<br />

Graphite in these rocks was transformed within the very<br />

short time of shock compression. As a consequence of the<br />

short transformation time, impact diamonds are very defect-rich<br />

and inherited some features of the precursor<br />

graphite. They are birefringent (Fig. 11a), retain the tabular<br />

hexagonal and sometimes even preserve spectacular<br />

growth twins rotated about the c axis of graphite (Fig.<br />

11b). Therefore impact diamonds are regarded as para- or<br />

pseudomorphs after graphite and are called apographitic<br />

diamonds (Masaitis et al. 1990). At the TEM scale, these<br />

impact diamonds contain numerous kink or twin bands<br />

parallel to (h h 2 — h — l) planes of precursor graphite (Fig. 11c;<br />

Langenhorst et al. 1999b). The bands were generated<br />

274


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

b<br />

0.1 mm a<br />

c<br />

Fig. 11. (a) Optical micrograph of a tabular impact diamond from the<br />

Ries crater, Germany. The anomalous interference colours are due to internal<br />

strain; crossed Nicols; (b) secondary electron scanning image of<br />

impact diamond from the Popigai impact crater, showing inherited twins<br />

of the precursor graphite, (c) Dark-field TEM image of an impact diamond<br />

from Popigai crater, showing mechanical twin bands that are inherited<br />

from precursor graphite.<br />

when the shock wave was transmitted into the graphite, i.e.<br />

directly before the phase transformation (Langenhorst<br />

2000). TEM studies failed, so far, to detect lonsdaleite although<br />

X-ray diffraction techniques indicate the presence<br />

of this high-pressure polymorph with a hexagonal stacking<br />

sequence (Frondel and Marvin 1967, Hannemann et al.<br />

1967, Masaitis et al. 1990). Instead, one observes that the<br />

diamonds are very disordered and contain numerous<br />

stacking faults, changing locally the cubic into a hexagonal<br />

stacking sequence (Langenhorst 2000). Diffuse X-ray<br />

scattering on these stacking faults is a way to explain the<br />

extra-peaks in X-ray diffraction patterns.<br />

Strongly corroded impact diamonds that are intergrown<br />

with moissanite (SiC) have been found at the Ries<br />

impact crater, Germany. Based on this finding, Hough et<br />

al. (1995) concluded that the diamonds formed by vapour<br />

condensation (so-called chemical vapour deposition<br />

(CVD) mechanism). Meanwhile, this idea has been discarded,<br />

because the moissanite may have formed by reaction<br />

of diamond with silica-rich impact melt (Langenhorst<br />

et al. 1999b). This reaction also forms the basis for the industrial<br />

production of SiC (Mehrwald 1992). Incorporation<br />

of impact diamonds into hot impact melt additionally<br />

provides an elegant explanation for the corroded surfaces.<br />

Calcite<br />

Calcite possesses a crystal structure with planar CO 3<br />

groups that are bridged via Ca atoms. The CO 3 groups are<br />

arranged in layers normal to the c-axis. The bonding between<br />

Ca cations and CO 3 groups is rather weak, allowing<br />

this structure to cleave as rhombohedra and to deform by<br />

mechanical twinning and dislocation glide (Nicolas and<br />

Poirier 1976). Static deformation experiments have shown<br />

that calcite can basically develop three types of mechanical<br />

twins (Barber and Wenk 1979a): e = {011 — 8}, r = {101 — 4}<br />

and f = {011 — 2} using the hexagonal unit cell setting with<br />

a = 4.99 Å and c = 17.06 Å. Little is known about the<br />

shock deformation of calcite but these twin laws apparently<br />

operate also under low to moderate shock pressures<br />

(Fig. 12a; Barber and Wenk 1979b, Langenhorst et al.<br />

2002a). At slightly higher pressures, the microstructure of<br />

shocked calcite becomes more dominated by dislocations<br />

(Fig. 12b; Barber and Wenk 1976, Langenhorst et al.<br />

2002a), occurring in a high density of up to 10 14 m -2 .<br />

X-ray line broadening observed for naturally shocked calcite<br />

might be due to such high dislocation densities (Skála<br />

and Jakeš 1999).<br />

Effects in the high-pressure regime are of fundamental<br />

importance, as calcite is known to decompose into solid<br />

CaO and gaseous CO 2 . The devolatilization of carbonates<br />

by impacts is considered to be important for the evolution<br />

of Earth’s atmosphere, climate, and life (Silver and<br />

Schultz 1980, Crutzen 1987). For example, the Chicxulub<br />

impact possibly perturbed the atmosphere with large<br />

amounts of CO 2 and SO x , changing its radiative balance<br />

and causing acid rain (Emiliani et al. 1981, Prinn and Fegley<br />

1987). This may have played an important role in the<br />

275


Falko Langenhorst<br />

3 µm 1 µm<br />

a<br />

b<br />

Fig. 12. Shock effects in calcite: (a) bright-field TEM image of crossing deformation twins in calcite shocked to 85 GPa using a high-explosive setup;<br />

(b) dark-field TEM image of numerous dislocations in a laser-shocked calcite (see Langenhorst et al. 2002a).<br />

mass extinction scenario at the Cretaceous-Tertiary<br />

boundary (Alvarez et al. 1980).<br />

Numerous shock experiments have recently been conducted<br />

to address the question of shock-induced devolatilization<br />

(Martinez et al. 1995, Skála et al. 2001,<br />

Ivanov et al. 2002, Langenhorst et al. 2002a). The experiments<br />

indicate that strongly shocked calcite can melt at<br />

high pressure but degassing probably takes only place<br />

after decompression if the post-shock temperatures are<br />

sufficiently high. This result is strengthened by theoretical<br />

calculations of the phase diagram of calcite (Ivanov and<br />

Deutsch 2002) and the discovery of quenched carbonate<br />

melts in suevites of the Ries and Chicxulub craters (Graup<br />

1999, Jones et al. 2000). Furthermore, experiments indicate<br />

that back reactions between the decomposition products,<br />

CO 2 and CaO, are fast and efficient and therefore<br />

have to be taken into account in any quantification of impact-released<br />

CO 2 (Agrinier et al. 2001). To avoid the back<br />

reaction it is necessary to spatially separate the decomposition<br />

products, which is another complexity in a natural<br />

environment.<br />

Shock-melted calcite develops upon quenching a<br />

feathery texture (Jones et al. 2000). Under the optical microscope,<br />

the feathery texture appears as aggregates of radiating,<br />

elongated calcite crystals. At the TEM scale,<br />

quench crystals of calcite are usually devoid of lattice defects,<br />

indicating that they went through the liquid state<br />

(Langenhorst et al. 2002a). On the other hand, incipient<br />

decomposition in shocked calcite is manifested by the<br />

presence of numerous tiny dislocation loops, indicative for<br />

the mobilization of CO 2 (Langenhorst et al. 2002a).<br />

Estimation of shock conditions<br />

A prerequisite for deciphering the shock-metamorphic<br />

history of minerals and their host rocks is an assessment of<br />

their pressure-temperature-time paths. It is important to remember<br />

here that polymict impact breccias consist of rock<br />

and mineral fragments that have suffered different degrees<br />

of shock metamorphism. Therefore it is necessary to consider<br />

each fragment separately.<br />

The estimation of the shock duration is rather difficult,<br />

since no mineralogical speedometer is available. However,<br />

one can obtain a rough idea of the shock pulse if the size<br />

of the projectile is known from e.g. crater size. The shock<br />

pulse corresponds substantially to the time that the shock<br />

wave needs to propagate to the rear surface of the projectile<br />

and back to the point of impact.<br />

The estimation of temperatures is a difficult task as<br />

well, because it is influenced by a number of factors such<br />

as porosity. In the context of shock metamorphism, one<br />

can, in principle, distinguish three different temperatures:<br />

(1) pre-shock, (2) shock and (3) post-shock temperatures.<br />

The pre-shock temperature is not directly related to shock<br />

metamorphism but elevated pre-shock temperatures have<br />

the effect to lower the threshold pressure for certain shockmetamorphic<br />

effects (Huffman et al. 1989, Langenhorst et<br />

al. 1992). It can be elevated due to regional metamorphism<br />

at the time of impact and the temperature could be determined<br />

via classical petrologic concepts (mineral assemblages,<br />

element partitioning between coexisting minerals).<br />

It is also known that PDF orientations in quartz significantly<br />

change at elevated pre-shock temperature, when<br />

quartz is shocked in the β high-temperature structure<br />

(> 573°C; Langenhorst and Deutsch 1994, Grieve et al.<br />

1996). Quartz experimentally shocked in the β stability<br />

field contains PDFs parallel to planes of hexagonal pyramids,<br />

whereas shocked α-quartz exhibits PDFs that only<br />

belong to one rhombohedron (for further details see Langenhorst<br />

and Deutsch 1994).<br />

Shock and post-shock temperatures are directly related<br />

to the magnitude of shock compression. The shock temperature<br />

is the maximum temperature achieved during<br />

shock compression, whereas the post-shock temperature is<br />

the temperature prevailing directly after decompression.<br />

276


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

For compact materials, the pressure-dependence of shock<br />

and post-shock temperatures is fairly well known through<br />

calculations and pyrospectrometric measurements (e.g.,<br />

Wackerle 1962, Raikes and Ahrens 1969, Martinez et al.<br />

1995, Holland and Ahrens 1997; Fig. 5). Therefore, temperatures<br />

can be estimated from these reference data, if the<br />

shock pressure has been determined by applying a shock<br />

barometer.<br />

It is more difficult to assess the shock and post-shock<br />

temperatures in porous rocks/minerals, because porosity<br />

has the general effect to significantly increase both temperatures.<br />

Shock compression leads however to compaction<br />

of rocks and loss of pore space, making it difficult<br />

to assess the initial porosity of rocks. Therefore, one can<br />

only apply simple arguments to estimate the temperatures<br />

from observations. For example, in case of mineral melting,<br />

the post-shock temperature has certainly exceeded the<br />

melting point of the mineral at ambient pressure.<br />

The estimation of shock pressure relies on calibration<br />

data obtained in well-controlled shock experiments. Shock<br />

experiments have provided quantitative information on (1)<br />

co(existence) of shock effects in certain pressure ranges,<br />

(2) changes in physical (bulk) properties (e.g. refractive<br />

index and density), (3) variations in PDF orientations (Fig.<br />

13), and (4) degree of mosaicism (Stöffler 1972, Stöffler et<br />

al. 1988, Hanns et al. 1978, Stöffler and Langenhorst<br />

1994, Grieve et al. 1996, Langenhorst and Deutsch 1998).<br />

To obtain a rough estimate of the shock pressure, it is<br />

often enough to use the petrographic microscope and to<br />

observe the shock effects in coexisting minerals. Data on<br />

the (co)existence of pressure-specific shock effects is<br />

available for most rock-forming minerals; compilations<br />

can be found in (Fig. 5): Stöffler (1972), Stöffler et al.<br />

(1988) and (1991), Bischoff and Stöffler (1992), Langenhorst<br />

and Deutsch (1998).<br />

Better constraints on shock pressures can be obtained<br />

if the refractive indices or densities of quartz or feldspars<br />

were measured, using a spindle stage and a density gradient<br />

column (Fig. 13; Medenbach 1985, Langenhorst<br />

1993). The change in these physical properties reflects the<br />

gradual conversion of quartz or feldspars into diaplectic<br />

glasses. This technique applies only over a small pressure<br />

range and requires that the glass is unaffected by postshock<br />

annealing or alteration.<br />

A more robust and widely used technique is to determine<br />

PDF orientations in quartz with an universal-stage<br />

(see appendix and Fig. 13, Stöffler and Langenhorst 1994,<br />

Grieve et al. 1996). Statistical data on PDF orientations in<br />

quartz have been provided, as a function of pressure, by<br />

various U-stage studies (Hörz 1968, Müller and Defourneaux<br />

1969, Robertson and Grieve 1977, Langenhorst<br />

and Deutsch 1994). These measurements reveal, for example,<br />

that, at pressures of about 20 GPa, PDFs are predominantly<br />

oriented parallel to {101 — 3} planes but change at<br />

higher pressure to {101 —<br />

2} orientations. A practical description<br />

on how to identify PDF orientations with an U-<br />

stage is given in the appendix of this paper.<br />

Mean refractive index<br />

1.55<br />

1.50<br />

1.45<br />

PDFs<br />

(1013)<br />

(1011)<br />

(1122)<br />

PDFs<br />

(1012)<br />

(1013)<br />

(1011)<br />

(1122)<br />

X-ray diffractometer studies on shocked rock-forming<br />

minerals reveal an increasing degree of mosaicism with increasing<br />

shock pressure as expressed by decreasing diffraction<br />

peak amplitude and pronounced line-broadening<br />

(Hörz and Quaide 1973, Hanns et al. 1978). Although this<br />

technique is capable of yielding relatively accurate pressure<br />

determinations, it has never been widely applied because<br />

it requires fresh, unaltered samples.<br />

Concluding remarks<br />

transformation regime<br />

PDFs<br />

(1012)<br />

(1013)<br />

20 25 30 35<br />

Shock pressure (GPa)<br />

diaplectic<br />

glass<br />

Fig. 13. Refractive index, density and combinations of PDF orientations<br />

of experimentally shocked quartz as function of pressure (from Langenhorst<br />

and Deutsch 1994).<br />

In the past decade, we have made considerable<br />

progress in understanding the nature and formation mechanisms<br />

of shock effects in minerals. These advances are, to<br />

a large extent, due to TEM studies of shocked minerals,<br />

providing a thorough characterization of the defect microstructures.<br />

Minerals show a unique response when subjected to<br />

strong shock waves. The effects range from deformation<br />

and transformation phenomena to decomposition, melting<br />

and vaporisation. Which shock effects are activated at<br />

what pressures and temperatures depends on the crystal<br />

structure and chemical composition of the mineral studied.<br />

Most shock effects have been reproduced in short laboratory<br />

experiments with various designs, although the<br />

shock pulses in experiments can be more than 6 orders of<br />

magnitude shorter than those prevailing in nature. Experimental<br />

limits are however reached if more time-consuming<br />

processes are simulated, such as reconstructive phase<br />

transformations to high-pressure silicates. On the other<br />

hand, the experimental simulation is very successful in reproducing<br />

deformation defects in shocked minerals. Furthermore,<br />

experiments led also to a better understanding of<br />

post-shock modifications of mineralogical shock effects in<br />

the natural environment. Finally, the shock signature of<br />

minerals is not only an unequivocal indicator for hypervelocity<br />

impact but can also be used to obtain quantitative<br />

constraints on the pressures and temperatures in natural<br />

impacts.<br />

2.6<br />

2.4<br />

2.2<br />

Density (g/cm 2 )<br />

277


Falko Langenhorst<br />

Acknowledgements. I am grateful for the financial support<br />

provided by the Deutsche Forschungsgemeinschaft, which funded part of<br />

this work (grants DE 401/15, HO 1446/3, LA 830/4). I am also indebted<br />

to many colleagues who provided samples or collaborated with me on the<br />

subject of shock metamorphism: A. Bischoff, M. Boustie, A. Deutsch,<br />

J.-C. Doukhan, H. Dypvik, U. Hornemann, B. Ivanov, V. L. Masaitis,<br />

J.-P. Poirier, G. Shafranovsky, and D. Stöffler. I also wish to thank<br />

F. Hörz, H. Leroux, and R. Skala for their constructive reviews and<br />

J. Hopf for technical assistance.<br />

References<br />

Agrinier P., Deutsch A., Schärer U., Martinez I. (2001): Fast back-reactions<br />

of shock-released CO 2 from carbonates: an experimental approach.<br />

Geochim. Cosmochim. Acta, 65, 2615–2632.<br />

Ahrens T. J., Rosenberg J. T. (1968): Shock metamorphism: experiments<br />

on quartz and plagioclase. In: French B. M., Short N. M. (eds) Shock<br />

metamorphism of natural materials, Mono Book Corp., Baltimore,<br />

pp. 59–81.<br />

Ahrens T. J., Tsay F. D., Live D. H. (1976): Shock-induced fine-grained<br />

recrystallisation of olivine: Evidence against subsolidus reduction of<br />

Fe 2+ . Proc. Lunar Sci. Conf. 7 th , 1143–1156.<br />

Alvarez L. W., Alvarez W., Asaro F., Michel H. V. (1980): Extraterrestrial<br />

cause for the Cretaceous-Tertiary extinction. Science, 208,<br />

1095–1108.<br />

Asay J. R., Shahinpoor M. (1993): High-Pressure Shock Compression of<br />

Solids. Spinger, New York, Berlin, Heidelberg.<br />

Ashworth J. R., Barber D. J. (1975): Electron petrography of shock-deformed<br />

olivine in stony meteorites. Earth Planet. Sci. Lett., 27,<br />

43–50.<br />

Ashworth J. R., Schneider H. (1985): Deformation and transformation in<br />

experimentally shock-loaded quartz. Phys. Chem. Min., 11,<br />

241–249.<br />

Bauer J. F. (1979): Experimental shock metamorphism of mono- and<br />

polycrystalline olivine: A comparative study. Proc. Lunar Planet.<br />

Science Lett. 10th, 2573–2576.<br />

Barber D. J., Wenk H. R. (1976): Defects in deformed calcite and carbonate<br />

rocks. In: Wenk H.-R. (ed.) Electron Microscopy in Mineralogy.<br />

Springer, Berlin Heidelberg, New York, pp. 428–442.<br />

Barber D. J., Wenk H. R. (1979a): Deformation twinning in calcite,<br />

dolomite, and other rhombohedral carbonates. Phys. Chem. Min., 5,<br />

141–165.<br />

Barber D. J., Wenk H. R. (1979b): On geological aspects of calcite microstructure.<br />

Tectonophysics, 54, 45–60.<br />

Barker L. M., Shahinpoor M., Chhabildas L. C. (1993): Experimental<br />

and diagnostic techniques. In: Asay J. R. and Shahinpoor M. (eds)<br />

High-Pressure Shock Compression of Solids, Springer, New York,<br />

Berlin, Heidelberg, pp. 43–74.<br />

Binns R. A., Davis R. J., Reed S. J. B. (1969): Ringwoodite, natural<br />

(Mg,Fe) 2SiO 4 spinel in the Tenham meteorite. Nature, 221, 943–944.<br />

Bischoff A., Stöffler D. (1984): Chemical and structural changes induced<br />

by thermal annealing of shocked feldspar inclusions in impact melt<br />

rocks from Lappajärvi crater, Finland. J. Geophys. Res., 89,<br />

B645–B645.<br />

Bischoff A., Stöffler D. (1992): Shock metamorphism as a fundamental<br />

process in the evolution of planetary bodies: Information from meteorites.<br />

Eur. J. Mineral., 4, 707–755.<br />

Bloss F. D. (1961): An Introduction to the methods of optical crystallography.<br />

Holt, Rinehart and Winston, New York.<br />

Bohor B. F., Foord E. E., Modreski P. J., Triplehorn D. M. (1984): Mineralogical<br />

evidence for an impact event at the Cretaceous-Teriary<br />

boundary. Science, 224, 867–869.<br />

Bohor B. F., Modreski P. J., Foord E. E. (1987): Shocked quartz in the<br />

Cretaceous/Tertiary boundary clays: Evidence for global distribution.<br />

Science, 236, 705–708.<br />

Boslough M. B., Asay J. R. (1993): Basic principles of shock compression.<br />

In: Asay J. R., Shahinpoor M. (eds) High-Pressure Shock Compression<br />

of Solids. Springer, New York, Berlin, Heidelberg, pp. 7–42.<br />

Boslough M. B., Cygan R. T., Kirkpatrick R. J., Montez B. (1989): NMR<br />

spectroscopic analysis of experimentally shocked quartz and the formation<br />

of diaplectic glass. Lunar Planet. Science Conf. 20, 97.<br />

Bourret A., Hinze E., Hochheimer H. D. (1986): Twin structure in coesite<br />

studied by high resolution electron microscopy. Phys. Chem. Minerals,<br />

13, 206–212.<br />

Brown J. M., McQueen R. G. (1986): Phase transitions, Grüneisen parameter<br />

and elasticity for shocked iron between 77 GPa and 400<br />

GPa. J. Geophys. Res., 91, 7485–7494.<br />

Byerly G. R., Lowe D. R., Wooden J. L., Xie X. (2002): An Archean impact<br />

layer from the Pilbara and Kapvaal cratons. Science, 297,<br />

1325–1327.<br />

Carstens H. (1975): Thermal history of impact melt rocks in the<br />

Fennoscandian shield. Contrib. Mineral. Petrol., 50, 145–155.<br />

Carter N. L., Raleigh C. B., DeCarli P. S. (1968): Deformation of olivine<br />

in stony meteorites. J. Geophys. Res., 73, 5439–5461.<br />

Cauble R., Phillion D. W., Hoover T. J., Holmes N. C., Kilkenny J. D.,<br />

Lee R. W. (1993): Demonstration of 0.75 Gbar planar shocks in X-<br />

ray driven colliding foils. Phys. Rev.Lett., 70, 4, 2102–2105.<br />

Chapman C. R., Morrison D. (1994): Impacts on the earth by asteroids<br />

and comets: Assessing the hazard. Nature, 367, 33–40.<br />

Chen M., Sharp T. G., ElGoresy A. E., Wopenka B., Xie X. (1996): The<br />

majorite-pyrope + magnesiowüstite assemblage: constraints on the<br />

history of shock veins in chondrites. Science, 271, 1570–1573.<br />

Cordier P., Vrána S., Doukhan J. C. (1994): Shock metamorphism in<br />

quartz at Sevetin and Susice (Bohemia)? A TEM investigation. Meteoritics,<br />

29, 98–99.<br />

Crutzen P. J. (1987): Acid rain at the K/T boundary. Nature, 330,<br />

108–109.<br />

Dachille F., Gigl P., Simons P. Y. (1968): Experimental and analytical<br />

studies of crystalline damage useful for the recognition of impact<br />

structures. In: French B. M., Short N. M. (eds) Shock Metamorphism<br />

of Natural Materials, Mono Book Corp., Baltimore, pp.<br />

555–570.<br />

Davison L., Horie Y., Sekine T. (2002): High-pressure shock compression<br />

of solids V-shock chemistry with applications to meteorite impacts.<br />

Springer, New York, pp. 245.<br />

DeCarli P. S., Jamieson J. C. (1961): Formation of diamond by explosive<br />

shock. Science, 133, 1821–1822.<br />

Deutsch A., Langenhorst F. (1998): Mineralogy of astroblemes – terrestrial<br />

impact craters. In: Marfunin. A. S. (ed.) Advanced Mineralogy<br />

3, Springer, Berlin, pp. 76–95.<br />

Deutsch A., Masaitis V. L., Langenhorst F., Grieve R. A. F. (2000): Popigai,<br />

Siberia - well preserved giant impact structure, national treasury,<br />

and world’s geological heritage. Episodes, 23, 3–11.<br />

Duvall G. E., Fowles G. R. (1963): Shock waves. In: Bradley R. S. (ed.)<br />

High Pressure Physics and Chemistry 2, Academic Press Inc., London,<br />

New York, pp. 209–291.<br />

El Goresy A., Dubrovinsky L., Sharp T. G., Saxena S. K., Chen M.<br />

(2000): A monoclinic post-stishovite polymorph of silica in the Shergotty<br />

meteorite. Science, 288, 1632–1634.<br />

El Goresy, A., Gillet P., Chen M., Künstler F., Graup G., Stähle V. (2001):<br />

In situ discovery of shock-induced graphite-diamond phase transition<br />

in gneisses from the Ries crater, Germany. Amer. Miner., 86,<br />

611–621.<br />

Emiliani C., Kraus E. B., Shoemaker E. M. (1981): Sudden death at the<br />

end of the Mesozoic. Earth Planet. Sci. Lett., 55, 317-334.<br />

Emmons R. C. (1943): The universal stage. Geolog. Soc. Am., Mem. 8.<br />

Engelhardt von W., Bertsch W. (1969): Shock induced planar deformation<br />

structures in quartz from the Ries crater, Germany. Contrib.<br />

Mineral. Petrol., 20, 203-234.<br />

Engelhardt von W., Arndt J., Stöffler D., Müller W. F., Jeziorkowski H.,<br />

Gubser R. A. (1967): Diaplektische Gläser in den Breccien des Ries<br />

von Nördlingen als Anzeichen für Stoßwellenmetamorphose. Contr.<br />

Miner. Petrol., 15, 93-102.<br />

Erofeev M. V., Lachinov P. A. (1888): The description of the Novo Urey<br />

meteorite. Proceedings of the Emperor’s Saint-Petersburg Mineralogical<br />

Society, pp. 13 (in Russian).<br />

Ernstson K., Hammann W., Fiebag J., Graup G. (1985): Evidence of an<br />

impact origin for the Azuara structure (Spain). Earth Planet. Science<br />

Letters, 74, 361–370.<br />

Federov von E. (1896): Universalmethode und Feldspathstudien. Zeits.<br />

Krist. Miner., 26, pp. 225.<br />

278


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

Foote A. E. (1891): New locality for meteoritic iron with a preliminary<br />

notice of the discovery of diamonds in the iron. Amer. J. Sci, 42,<br />

413–417.<br />

French B. M. (1969): Distribution of shock-metamorphic features in the<br />

Sudbury Basin, Ontario, Canada. Meteoritics, 4, 173–174.<br />

French B. M. (1998): Traces of catastrophe – A handbook of shock-metamorphic<br />

effects in terrestrial meteorite impact structures. LPI contribution<br />

No. 954, Lunar and Planetary Institute, Houston, 120 pp.<br />

French B. M., Short N. M. (1968): Shock metamorphism of natural materials.<br />

Mono Book Corp, Baltimore, Maryland.<br />

Frondel C., Marvin U. B. (1967): Lonsdaleite, a hexagonal polymorph of<br />

diamond. Nature, 214, 587–589.<br />

Glass B. P., Liu S., Leavens P. B. (2002): Reidite: An impact-produced<br />

high-pressure polymorph of zircon found in marine sediments.<br />

Amer. Mineral., 87, 562–565.<br />

Goltrant O., Cordier P., Doukhan J. C. (1991): Planar deformation features<br />

in shocked quartz: a transmission electron microscopy investigation.<br />

Earth Planet. Sci. Letters, 106, 103–115.<br />

Goltrant O., Leroux H., Doukhan J.C., Cordier P. (1992): Formation<br />

mechanism of planar deformation features in naturally shocked<br />

quartz. Phys. Earth Planet. Int., 74, 219–240.<br />

Gratz A. J., Nellis W. J., Christie J. M., Brocious W., Swegle J., Cordier<br />

P. (1992): Shock metamorphism of quartz with initial temperatures<br />

–170 to +1000°C. Phys. Chem. Miner., 19, 267–288.<br />

Graup G. (1999): Carbonate-silicate liquid immiscibility upon impact<br />

melting: Ries crater, Germany. Meteoritics & Planetary Science, 34,<br />

425–438.<br />

Grieve R. A. F., Sharpton V. L., Stöffler D. (1990): Shocked minerals and<br />

the K/T controversy. EOS, 71, 1792–1793.<br />

Grieve R. A. F., Shoemaker E. M. (1994): The record of past impacts on<br />

earth. In: Gehrels T. (ed.) Hazards due to comets and asteroids. Univ.<br />

of Arizona Press, Tucson, AZ, London, pp. 417–462.<br />

Grieve R. A. F., Langenhorst F., Stöffler D. (1996): Shock metamorphism<br />

of quartz in nature and experiment: II. Significance in geoscience.<br />

Meteoritics & Planet. Sci., 31, 6–35.<br />

Hannemann R. E., Strong H. M., Bundy F. P. (1967): Hexagonal diamonds<br />

in meteorites: implications. Science, 155, 995–997.<br />

Hanns R. E., Montague B. R., Davis M. K., Galindo C., Hörz F. (1978):<br />

X-ray diffractometer studies of shocked materials. Proc. Lunar Planet.<br />

Sci., Conf. 9th, 2773–2787.<br />

Hildebrand A. R., Penfield G. T., Kring D. A., Pilkington M., Camargo<br />

A. Z., Jacobsen S. B., Boyton W. V. (1991): Chicxulub crater: A possible<br />

Cretaceous/Tertiary boundary impact crater on the Yucatan<br />

Peninsula, Mexico. Geology, 19, 867–871.<br />

Holland K. G., Ahrens T. J. (1997): Melting of (Mg,Fe) 2SiO 4 at the<br />

core-mantle boundary of the earth. Science, 275, 1623–1625.<br />

Hornemann U., Müller W. F. (1971): Shock-induced deformation twins<br />

in clinopyroxene. Neues Jb. Mineral. Mh., 6, 247–256.<br />

Hörz F. (1965): Untersuchungen an Riesgläsern. Beiträge Mineral. Petrographie,<br />

11, 621–661.<br />

Hörz F. (1968): Statistical measurements of deformation structures and<br />

refractive indices in experimentally shock-loaded quartz. In: French<br />

B. M., Short N. M. (eds) Shock metamorphism of natural materials.<br />

Mono Book Corp, Baltimore, Maryland, pp. 243–254.<br />

Hörz F., Quaide W. L. (1973): Debye-Scherrer investigations of experimentally<br />

shocked silicates. The Moon, 6, 45–82.<br />

Hough R. M., Gilmour I., Pillinger C. T., Arden J. W., Gilkes K. W. R.,<br />

Yuan J., Milledge H. J. (1995): Diamond and silicon carbide in impact<br />

melt rock from the Ries impact crater. Nature, 378, 41–44.<br />

Hough R. M., Pillinger C. T., Gilmour I., Langenhorst F., Montanari A.<br />

(1997): Diamonds from the iridium rich K-T boundary layer at Arroyo<br />

el Mimbral, Tamaulipas, Mexico. Geology, 11, 1019–1022.<br />

Huffman A. R., Brown J. M., Carter N. L. (1989): Temperature dependence<br />

of shock-induced microstructures in tectosilicates. In: Schmidt<br />

S. C., Johnson J. N., Davison L. W. (eds) Shock Compression of<br />

Condensed Matter. Elsevier Science Publishers B.V., Amsterdam,<br />

pp. 649–652.<br />

Ivanov B. A., Deutsch A. (2002): The phase diagram of CaCO 3 in relation<br />

to shock compression and decomposition. Phys. Earth Planet.<br />

Interiors, 129, 131–143.<br />

Ivanov B. A., Langenhorst F., Deutsch A., Hornemann U. (2002): How<br />

strong was the impact-induced CO 2 degassing in the K/T event?<br />

Numerical modeling of laboratory experiments. In: Koeberl Ch.,<br />

Mac Leod K. G. (eds) Catastrophic events and Mass Extinctions: Impact<br />

and beyond. Geol. Soc. Amer. Spec. Pap., 356, pp. 587–594.<br />

Jackson I., Ahrens T. J. (1979): Shock wave compression of single-crystal<br />

forsterite. J. Geophys. Res., 84, 3039–3048.<br />

Jeanloz R. (1980): Shock effects in olivine and implications for Hugoniot<br />

data. J. Geophys. Res., 85, 3163–3176.<br />

Jeanloz R., Ahrens T. J., Lally J. S., Nord G. L., Christie J. M., Heuer A.<br />

H. (1977): Shock-produced olivine glass: First observation. Science,<br />

197, 457–459.<br />

Jones A. P., Claeys P., Heuschkel S. (2000): Impact melting of carbonates<br />

from the Chicxulub Crater. In: Gilmour I., Koeberl C. (eds) Impacts<br />

and the early Earth, Lecture Notes in Earth Sciences 91, Springer,<br />

Berlin-Heidelberg-New York, pp. 343–361.<br />

Joreau P., Reimold W. U., Robb L. J., Doukhan J. C. (1997a): A TEM<br />

study of deformed quartz grains from volcaniclastic sediments associated<br />

with the Bushveld Complex, South Africa. Eur. J. Mineral., 9,<br />

393–401.<br />

Joreau P., Leroux H., Doukhan J. C. (1997b): A transmission electron microscope<br />

investigation of shock metamorphism in olivine of the<br />

Ilafegh 013 chondrite. Meteoritics & Planetary Sci., 32, 309–316.<br />

Kieffer S. W., Phakey P. P., Christie J. M. (1976): Shock processes in<br />

porous quartzite: transmission electron microscope observations and<br />

theory. Contrib. Mineral. Petrol., 59, 41–93.<br />

Koeberl C., Masaitis V. L., Shafranovsky G. I., Gilmour I., Langenhorst<br />

F., Schrauder M. (1997): Diamonds from the Popigai impact structure,<br />

Russia. Geology, 25, 967–970.<br />

Lally J. S., Christie J. M., Nord G. L., Heuer A. H. (1976): Deformation,<br />

recovery and recrystallization of lunar dunite 72417. Proc. Lunar<br />

Sci. Conf. 7th, 1845–1863.<br />

Langenhorst F., Deutsch A., Hornemann U., Stöffler D. (1992): Effect of<br />

temperature on shock metamorphism of single crystal quartz. Nature,<br />

356, 507–509.<br />

Langenhorst F. (1993): Eine modifizierte Dichtegradientenkolonne zur<br />

präzisen Dichtebestimmung an Einzelkörnern. Ber. Dtsch. Mineral.<br />

Ges. 1, Beih Eur. J. Mineral., 5, 244.<br />

Langenhorst F. (1994): Shock experiments on α- and β- quartz: II. X-ray<br />

investigations. Earth Planet. Sci. Letters, 128, 638–698.<br />

Langenhorst F. (1996): Characteristics of shocked quartz in late Eocene<br />

impact ejecta from Massignano (Ancona, Italy): Clues to shock conditions<br />

and source crater. Geology, 24, 487–490.<br />

Langenhorst F. (2000): Stoßwellenmetamorphose von Mineralen: Neue<br />

Erkenntnisse durch Transmissionselektronenmikroskopie. Habilitation<br />

thesis, University of Bayreuth, Germany.<br />

Langenhorst F., Deutsch A. (1994): Shock experiments on pre-heated α-<br />

and β- quartz: I. Optical and density data, Earth Planet. Science Letters,<br />

125, 407–420.<br />

Langenhorst F., Deutsch A. (1996): The Azuara and Rubielos structures,<br />

Spain; twin impact craters or alpine thrust systems? TEM investigations<br />

on deformed quartz disprove shock origin. Lunar Planet.<br />

Science Conf. 27, 725–726.<br />

Langenhorst F., Deutsch A. (1998): Minerals in terrestrial impact structures<br />

and their characteristic features. In: Marfunin A. S. (ed.) Advanced<br />

Mineralogy 3, Springer, Berlin, pp. 95–119.<br />

Langenhorst F., Greshake A. (1999): A TEM study of Chassigny: evidence<br />

for strong shock metamorphism. Meteoritics Planet. Sci., 34,<br />

43–48.<br />

Langenhorst F., Poirier J. P. (2000): Anatomy of black veins in Zagami:<br />

Clues to the formation of high-pressure phases. Earth Planet. Sci.<br />

Lett., 184, 37–55.<br />

Langenhorst F., Poirier J. P. (2002): Transmission electron microscopy of<br />

coesite inclusions in the Dora Maira high-pressure metamorphic pyrope-quartzite.<br />

Earth Planet. Sci. Lett., 203, 793–803.<br />

Langenhorst F., Joreau P. Doukhan J. C. (1995): Thermal and shock<br />

metamorphism of the Tenham meteorite: a TEM examination.<br />

Geochim. Cosmochim. Acta, 59, 1835–1845.<br />

Langenhorst F., Boustie M., Migault A., Romain J. P. (1999a): Laser<br />

shock experiments with nanoseconds pulses: A new tool for the reproduction<br />

of shock defects in olivine. Earth Planet. Sci. Lett., 173,<br />

333–342.<br />

279


Falko Langenhorst<br />

Langenhorst F., Shafranovsky G. I., Masaitis V. L., Koivisto M. (1999b):<br />

Discovery of impact diamonds in a Fennoscandian crater and evidence<br />

for their genesis by solid-state transformation. Geology, 27,<br />

747–750.<br />

Langenhorst F., Boustie M., Deutsch A., Hornemann U., Matignon Ch.,<br />

Migault A., Romain J. P. (2002a): Experimental techniques for the<br />

simulation of shock metamorphism: A case study on calcite. In: Davison<br />

L., Horie Y., Sekine T. (eds) High-pressure shock compression<br />

of solids V-Shock chemistry with applications to meteorite impacts.<br />

Springer, New York, pp. 1–27.<br />

Langenhorst F., Poirier J.-P., Deutsch A., Hornemann U. (2002b): Experimental<br />

approach to generate shock veins in single-crystal olivine by<br />

shear melting. Meteoritics & Planet. Sci., 37, 1541–1554.<br />

Leroux H. (2001): Microstructural shock signatures of major minerals in<br />

meteorites. Eur. J. Mineral., 13, 253–272.<br />

Leroux H., Doukhan J.-C. (1996): A transmission electron microscope<br />

study of shocked quartz from the Manson impact structure. Geol.<br />

Soc. Am. Spec. Paper, 302, 267–274.<br />

Leroux H., Reimold W. U., Doukhan J. C. (1994): A TEM investigation<br />

of shock metamorphism in quartz from the Vredefort dome, South<br />

Africa. Tectonophysics, 230, 223–239.<br />

Leroux H., Doukhan J. C., Guyot F. (1996): An analytical electron microscopy<br />

(AEM) investigation of opaque inclusions in some type 6<br />

ordinary chondrites. Meteoritics & Planetary Sci., 31, 767–776.<br />

Lipschutz M. E. (1964): Origin of diamonds in the Ureilites. Science,<br />

143, 1431–1434.<br />

Madon M., Poirier J. P. (1983): Transmission electron microscope observation<br />

of α, β and γ (Mg,Fe) 2SiO 4 in shocked meteorites: planar defects<br />

and polymorphic transitions. Phys. Earth Planet. Int., 33,<br />

31–44.<br />

Marsh S. P. (1980): LASL Shock Hugoniot Data. University of California<br />

Press, Berkeley, California.<br />

Martinez I., Deutsch A., Schärer U., Ildefonse Ph., Guyot F., Agrinier P.<br />

(1995): Shock recovery experiments on dolomite and thermodynamical<br />

modeling of impact-induced decarbonation. J. Geophys Res.,<br />

100, B8, 15,465–15,476.<br />

Martini J. E. J. (1991): The nature, distribution and genesis of the coesite<br />

and stishovite associated with the pseudotachylite of the Vredefort<br />

Dome, South Africa. Earth Planet. Sci. Letters, 103, 285–300.<br />

Masaitis V. L., Futergendler D. I., Gnevushev M. A. (1972): Diamonds in<br />

impactites of the Popigay meteorite crater. Zap. Vsesoyuznogo Mineralogicheskogo<br />

Obshchestva 101, 108–112.<br />

Masaitis V. L., Shafranovsky G. I., Ezersky V. A., Reshetnyak N. B.<br />

(1990): Impact diamonds in ureilites and impactites. Meteoritika 49,<br />

180–195.<br />

Masaitis V. L., Shafranovsky G. I., Federova I. G. (1995): The<br />

apographitic impact diamonds from astroblemes Ries and Popigai.<br />

Proc. Russ. Min. Soc., 4, 12–18.<br />

Masaitis V. L., Shafranovsky G. I., Grieve R. A. F., Langenhorst F.,<br />

Peredery W. V., Therriault A. M., Balmasov E. L., Federova I. G.<br />

(2000): Discovery of impact diamonds in black Onaping suevites,<br />

Sudbury structure, Ontario, Canada. In: Dressler B. O., Sharpton V.<br />

L. (eds) Large meteorite impacts and planetary evolution II, Geol.<br />

Soc. Am. Spec. Paper, 339, pp. 317–321.<br />

McLaren A. C. (1991): Transmission electron microscopy of minerals<br />

and rocks. Cambridge Univ. Press, Cambridge, U.K.<br />

McLaren A. C., Pithkethly D. R. (1982): The twinning microstructure<br />

and growth of amethyst quartz. Phys. Chem. Minerals, 8, 128–135.<br />

McLaren A. C., Retchford J. A., Griggs D. T., Christie J. M. (1967):<br />

Transmission electron microscope study of Brazil twins and dislocations<br />

experimentally produced in natural quartz. Phys. Status Solidi,<br />

19, 631–644.<br />

Medenbach O. (1985): A new microrefractometer spindle stage and its<br />

application. Fortschritte Miner., 63, 111–133.<br />

Mehrwald K.-H. (1992): History and economic aspects of industrial SiC<br />

manufacture. Berichte der Deutschen Keramischen Gesellschaft, 69,<br />

72–81.<br />

Melosh H. J. (1989): Impact cratering – A geological process. Oxford<br />

Univ. Press, New York.<br />

Milton D. J., DeCarli P. S. (1963): Maskelynite: formation by explosive<br />

shock. Science, 140, 670–671.<br />

Müller W. F. (1969): Elektronenmikroskopischer Nachweis amorpher<br />

Bereiche in stoßwellenbeanspruchtem Quarz. Naturwiss., 56, 279.<br />

Müller W. F., Defourneaux W. (1969): Deformationsstrukturen in Quarz<br />

als Indikator für Stoßwellen: eine experimentelle Untersuchung an<br />

Quarzeinkristallen. Z. Geophys., 34, 483–504.<br />

Müller W. F., Hornemann U. (1969): Shock-induced planar deformation<br />

structures in experimentally shock-loaded olivines and in olivines<br />

from chondritic meteorites. Earth Planet. Sci. Letters, 7, 251–264.<br />

Neukum G., Ivanov B. A., Hartmann W. K. (2001): Cratering records in<br />

the inner solar system in relation to the lunar reference system.<br />

Space Sci. Rev., 96, 55–86.<br />

Nicolas A., Poirier J. P. (1976): Crystalline plasticity and solid state flow<br />

in metamorphic rocks. Wiley, New York, 444 pp.<br />

Nicolaysen L. O., Reimold W. U. (1990): Cryptoexplosions and Catastrophes<br />

in the Geological Record, with a special focus on the Vredefort<br />

structure. Tectonophysics, 171, pp. 422.<br />

Nikitin W. W. (1936): Die Federow-Methode. Borntraeger, Berlin, pp.<br />

109.<br />

Phillips W. R. (1971) Mineral optics – principles and techniques. Freeman,<br />

San Francisco, pp. 249.<br />

Poirier J. P. (1985): Creep of crystals. Cambridge University Press, Cambridge.<br />

Prinn R. G., Fegley B., Jr. (1987): Bolide impacts, acid rain and biospheric<br />

traumas at the Cretaceous-Tertiary boundary. Earth Planet.<br />

Sci. Lett., 83, 409–418.<br />

Putnis A., Price G. D. (1979): High pressure (Mg,Fe) 2SiO 4 phases in the<br />

Tenham chondritic meteorite. Nature, 280, 217–218.<br />

Raikes S. A., Ahrens T. J. (1979): Post-shock temperatures in minerals.<br />

Geophys. J. R. Astron. Soc., 58, 717–747.<br />

Reimold W. U., Stöffler D. (1978): Experimental shock metamorphism<br />

of dunite. Proc. Lunar Planet. Sci. Lett. 9th, 2805–2824.<br />

Reinhard M. (1931): Universal Drehtischmethoden. Verlag Von B. Wepf<br />

und Cie, Basel, pp. 119.<br />

Robertson P. B., Grieve R. A. F. (1977): Shock attenuation at terrestrial<br />

impact structures. In: Roddy D. J., Pepin P. O., Merill R. B. (eds) Impact<br />

and explosion cratering. Pergamon Press, New York, pp.<br />

687–702.<br />

Robin E., Bonté Ph., Froget L., Jéhanno C., Rocchia R. (1992): Formation<br />

of spinels in cosmic objects during atmospheric entry: a clue to<br />

the Cretaceous-Tertiary boundary event. Earth Planet. Sci. Lett., 108,<br />

181–190.<br />

Roddy D. J., Pepin R. O., Merrill R. B. (1977): Impact and explosion cratering.<br />

Pergamon Press, New York, pp.1301.<br />

Rost R., Dolgov Y., A., Vishnevskiy S., A. (1978): Gases in inclusions of<br />

impact glass in the Ries crater, West Germany, and finds of highpressure<br />

carbon polymorphs. Doklady Akademii Nauk SSSR, 241,<br />

165–168.<br />

Rubie D. (1999): Characterising the sample enviroment in multianvil<br />

high-pressure experiments. Phase Transitions, 68, 431–451.<br />

Schmitt R. T. (2000): Shock experiments with the H6 chondrite Kernouvé:<br />

pressure calibration of microscopic shock effects. Meteoritics &<br />

Planet. Sci., 35, 545–560.<br />

Schneider H. (1978): Infrared Spectroscopic Studies of Experimentally<br />

Shock-Loaded Quartz. Meteoritics, 13, 227.<br />

Schreyer W. (1983): Metamorphism and fluid inclusions in the basement<br />

of the Vredefort Dome, South Africa: Guidelines to the origin of the<br />

structure. J. Petrol., 24, 26–47.<br />

Sharp T. G., Lingemann C. M., Dupas C., Stöffler D. (1997): Natural occurrence<br />

of MgSiO 3-ilmenite and evidence for MgSiO 3-perovskite<br />

in a shocked L chondrite. Science, 277, 352–355.<br />

Sharp T. G., ElGoresy A., Wopenka B., Chen M. (1999): A poststishovite<br />

SiO 2 polymorph in the meteorite Shergotty: implications<br />

for impact events. Science, 284, 1511–1513.<br />

Silver L. T., Schultz P. H. (1982): Geological implications of impacts of<br />

large asteroids and comets on the Earth. Geol. Soc. Amer. Spec. Pap.,<br />

190, Boulder.<br />

Skála R., Jakeš P. (1999): Shock-induced effects in natural calcite-rich<br />

targets as revealed by X-ray powder diffraction. Geol. Soc. Amer.<br />

Spec. Pap., 339, 205–214.<br />

Skála R., Ederová J., Matějka P., Hörz F. (2001): Mineralogical studies<br />

of experimentally shocked dolomite: Implications for the outgassing<br />

280


Shock metamorphism of some minerals: Basic introduction and microstructural observations<br />

of carbonates. In: Koeberl Ch., Mac Leod K. G. (eds) Catastrophic<br />

events and Mass Extinctions: Impact and beyond. Geol. Soc. Amer.<br />

Spec. Pap., 356, pp. 571–586.<br />

Snee L. W., Ahrens T. J. (1975): Shock induced deformation features in<br />

terrestrial peridot and lunar dunite. Proc. Lunar Planet. Sci. Conf.<br />

6th, 833–842.<br />

Stebbins J. F., Poe B. T. (1999): Pentacoordinate silicon in high-pressure<br />

crystalline and glassy phases of calcium disilicate (CaSi 2O 5). Geophys.<br />

Res. Lett., 26, 2521–2523.<br />

Stöckhert B., Duyster J., Trepmann C., Massonne H.-J. (2001): Microdiamond<br />

daughter crystals precipitated from supercritical COH + silicate<br />

fluids included in garnet, Erzgebirge, Germany. Geology, 29,<br />

391–394.<br />

Stöffler D. (1971): Coesite and stishovite: Identification and formation<br />

conditions in shock-metamorphosed rocks. J. Geophys. Res., 76,<br />

5474–5488.<br />

Stöffler D. (1972): Deformation and transformation of rock-forming<br />

minerals by natural and experimental shock processes: I. Behaviour<br />

of minerals under shock compression. Fortschritte der Mineralogie,<br />

49, 50–113.<br />

Stöffler D. (1974): Deformation and transformation of rock-forming<br />

minerals by natural and experimental shock processes: II. Physical<br />

properties of shocked minerals. Fortschritte der Mineralogie, 51,<br />

256–289.<br />

Stöffler D. (1984): Glasses formed by hypervelocity impact. J. Non-<br />

Cryst. Solids, 67, 465–502.<br />

Stöffler D., Langenhorst F. (1994): Shock metamorphism of quartz in nature<br />

and experiment: I. Basic observation and theory. Meteoritics,<br />

29, 155–181.<br />

Stöffler D., Ryder G. (2001): Stratigraphy and isotope ages of lunar geologic<br />

units: chronological standard for the inner solar system. Space<br />

Science Reviews, 96, 9–54.<br />

Stöffler D., Bischoff A., Buchwald V., Rubin A.,E. (1988): Shock effects<br />

in meteorites. In: Kerridge J. F., Matthews M. S. (eds) Meteorites<br />

and the Early Solar System. Univ. of Arizona press, Tuscon, pp.<br />

165–202.<br />

Stöffler D., Keil K., Scott E. R. D. (1991): Shock metamorphism of ordinary<br />

chondrites. Geochim. Cosmochim. Acta, 55, 3845–3867.<br />

Tomioka N., Fujino K. (1997): Natural (Mg,Fe)SiO 3-ilmenite and -perovskite<br />

in the Tenham meteorite. Science, 277, 1084–1086.<br />

Valter A. A., Yeremenko G. K., Kwasnitsa V. N., Polkanov Y. A. (1992):<br />

Shock-metamorphosed carbon minerals. Nauka Press, Kiev (in<br />

Russian).<br />

Vrána S. (1987): The Sevetin astrobleme, southern Bohemia, Czechoslovakia.<br />

Geol. Rundschau, 76, 505–528.<br />

Wackerle J. (1962): Shock wave compression of quartz. J. Appl. Phys.,<br />

33, 922–937.<br />

Walzebuck J. P., von Engelhardt W. (1979): Shock deformation of quartz<br />

influenced by grain size and shock direction: observations on quartzplagioclase<br />

rocks from the basement of the Ries crater, Germany.<br />

Contr. Mineral. Petrol., 70, 267–271.<br />

Wegener A. (1921): Die Entstehung der Mondkrater. Friedrich Vieweg &<br />

Sohn, Braunschweig, pp. 48.<br />

Wetherill G. W. (1984): Accumulation of terrestrial planets and implications<br />

concerning lunar origin. In: Hartmann W. K., Phillips R. J.,<br />

Taylor G. J. (eds) Origin of the Moon. Lunar and Planet. Sci. Inst.,<br />

Houston, pp. 519–550.<br />

White J. C. (1993): Shock-induced melting and silica polymorph formation,<br />

Vredefort structure, South Africa. In: Boland J. A., FitzGerald<br />

J. D. (eds) Defects and processes in solid state: geoscience applications.<br />

Elsevier, New York, pp. 69–84.<br />

Xue X., Stebbins J. F., Kanzaki M., Trønnes R. G. (1989): Silicon coordination<br />

and speciation changes in a silicate liquid at high pressures.<br />

Science, 245, 962–964.<br />

Handling editor: Roman Skála<br />

The following is a simple description on how to determine the crystallographic<br />

orientation of PDFs in shocked quartz grains, using an universal<br />

stage (von Federov 1896, Reinhard 1931, Nikitin 1936, Emmons<br />

1943, Phillips 1971). Although the use of universal stages has considerably<br />

declined in structural geology, it remains the only method, which<br />

provides statistical data on the orientation of PDFs. Other techniques<br />

such as TEM are also capable of measuring PDF orientations but TEM<br />

measurements are too time-consuming to obtain statistically meaningful<br />

data.<br />

There are four fundamental practical steps in the determination of<br />

PDF orientations in shocked quartz. One has to locate the spatial orientation<br />

of (1) the optic axis (parallel to c axis) and (2) the normals to the<br />

PDF planes in the grain studied. Once these directions are known from<br />

measurements, they are plotted and transformed in a stereographic Wulff<br />

net (3) and can then be indexed by comparison with the standard stereographic<br />

projection of quartz (4).<br />

(1) Since quartz is an uniaxial (trigonal) mineral it is only possible to locate<br />

the c axis optically. The a axes are in the plane perpendicular to<br />

c but their exact positions within this plane are not measurable. This<br />

causes a problem for indexing a PDF plane because the angle to the<br />

c axis can be measured but not the angle to the a axes. To circumvent<br />

this problem in part one needs at least two or more crossing PDF sets<br />

in a single quartz grain. To explain on how to locate the c axis of<br />

quartz, we will follow here the Reinhard notation of rotation axes for<br />

a 4-axis or 5-axis universal stage (Reinhard 1931, Fig. 14). Depending<br />

on the orientation of the quartz grain studied, the optic axis can<br />

be brought either parallel to the M (microscope) axis or parallel to<br />

the K (control) axis (E-W). Hence the first step is to find out whether<br />

the optic axis of the quartz grain is highly inclined to (polar position)<br />

or lies almost within the plane (equatorial position) of the thin section.<br />

Polar position: Rotate about the M axis to the diagonal position (45°<br />

to E-W) and then rotate about the H axis until the grain extincts under<br />

crossed Nicols. If it remains light, the grain is in the equatorial position<br />

APPENDIX: How to determine PDF orientations ?<br />

(see below). Rotate about the M axis to restore the 0° position, but keep<br />

H at its inclined position. The grain will become light again (unless the<br />

optic axis is already parallel to M). Now rotate about the N axis to<br />

achieve again the extinction position. Finally, repeat all the steps until the<br />

grain remains extinct when rotated about the M axis.<br />

Equatorial position: Rotate first about the N axis until the quartz<br />

grain extincts and its optic axis is in the E-W plane. This is the case if<br />

the grain becomes light by a further rotation about the K axis. Rotate<br />

then about the H axis until the extinction position of the grain is restored.<br />

The grain will again become light by a further rotation about the<br />

K axis in opposite direction (unless the optic axis is already parallel to<br />

E-W). Finally, repeat all the steps until the grain remains extinct when<br />

tilted about the K axis. The equatorial position is the more common situation<br />

in PDF measurements, because PDF poles mostly form a small<br />

angle to the c axis.<br />

(2) To determine the orientation of a PDF plane it is necessary to bring<br />

its normal parallel to E-W. This is achieved by rotating about the N<br />

axis until the trace of the PDF plane is parallel to N-S. Rotate then<br />

about the H axis until the trace becomes as sharp as possible. One<br />

can test whether the plane is exactly vertical by defocusing the PDF.<br />

It is vertical if the trace of the PDF does not move. Another way to<br />

test for the vertical position is to tilt about the K axis. If the PDF<br />

plane is exactly vertical, its trace will remain in N-S orientation.<br />

(3) The two important rotation angles to be plotted in the stereonet are the<br />

azimuth angle (i.e. rotation about N) and the ρ angle (i.e. rotation<br />

about H). When plotting the values for the optic axis, it has to be remembered<br />

whether it was tilted vertically or horizontally. Additionally,<br />

one needs to keep in mind the sense of tilting about H. Detailed<br />

descriptions on the plotting of U-stage data can be found in Bloss<br />

(1961) and Phillips (1971). Once the orientation of the optic axis and<br />

the normals to PDF planes are plotted, they need to be transformed<br />

into the standard stereographic projection with the c axis in the center<br />

(Fig. 15). The pole of the optic axis is transformed along the<br />

equatorial line of the Wulff net toward the center, and the normals to<br />

281


Falko Langenhorst<br />

M<br />

A<br />

West<br />

K<br />

H<br />

K<br />

N<br />

K<br />

East<br />

1 (0001)<br />

2 {1013}<br />

3 {1012}<br />

4 {1011}<br />

5 {1010}<br />

6 {1122}<br />

7 {1121}<br />

8 {2131}<br />

9 {5161}<br />

10 {1120}<br />

a 3<br />

a 2<br />

left<br />

negative<br />

Fig. 14. Sketch of the rotation axes on a 5-axis universal stage. The notation<br />

of axes is according to Reinhard (1931): M = microscope axis,<br />

A = auxiliary axis, N = normal axis, H = horizontal axis, and K = control<br />

axis.<br />

a 1<br />

right<br />

positive<br />

Fig. 16. Standard stereographic projection of quartz with the c axis in the<br />

center. The circles have a 5° radius and mark the positions of the most<br />

abundant PDF orientations (modified after Stöffler and Langenhorst<br />

1994). Since quartz is trigonal, one can distinguish positive and negative<br />

(e.g., rhombohedra) as well as right and left (e.g. pyramids) forms.<br />

FN'<br />

ρ<br />

∆<br />

FN'<br />

' – transformed positions<br />

OA – optic axis<br />

FN – face normal<br />

∆ – azimuth difference<br />

ρ – pole distance<br />

ϕ – azimuth angle<br />

ρ OA<br />

OA'<br />

FN<br />

ρ FN'<br />

ρ OA<br />

ϕ FN<br />

– ϕ OA<br />

OA<br />

a 3<br />

a 1<br />

(1012)<br />

(1013)<br />

(1011)<br />

(2112)<br />

a 2<br />

{1013}<br />

{1011}<br />

{0112}<br />

{1122}<br />

120°<br />

Fig. 17. Example for PDF orientations in quartz experimentally shocked<br />

to 25 GPa (from Langenhorst and Deutsch 1994). The right-hand<br />

diagram depicts schematically the full symmetry of coexisting forms.<br />

Fig. 15. An example for the transformation of a PDF pole (FN) and the<br />

optic axis (OA) into the standard stereographic projection of quartz with<br />

the optic axis (=c axis) in the center.<br />

the PDF planes are transformed along small circles by the same angle<br />

(see also von Engelhardt and Bertsch 1969). You can read now<br />

the angle between the PDF normals and the c axis. Alternatively, one<br />

can calculate this angle by the following equation:<br />

cos ρ FN’ = cos ρ FN cos ρ OA + sin ρ FN sin ρ OA cos (ϕ FN – ϕ OA) (5)<br />

In the literature, one will often find histograms in which this angle<br />

is plotted as function of the frequency of PDFs. However, as the a-axes<br />

cannot be located by polarizing microscopy, the angle to the c axis alone<br />

is insufficient for unequivocal indexing of PDF planes. This problem<br />

cannot be solved if the quartz grain of interest contains only one set of<br />

PDFs. However, if the grain contains at least two sets of PDFs, the next<br />

analytical step probably yields a reliable indexing result.<br />

(4) In this final step, the transformed stereoplot is compared to the standard<br />

stereographic projection of quartz, displaying the orientation of<br />

known PDF planes (Fig. 16). The latter are drawn as 5° circles, representing<br />

the estimated errors in the measurements. Table 1 shows<br />

that not only the pole distance varies for the crystallographic PDF<br />

planes but also the azimuth angles, e.g. rhombedral forms lie 30° off<br />

the pyramidal forms because they do not belong to the same crystallographic<br />

zone. This is the basis for indexing the PDF planes. In<br />

practice, the stereographic projection of measured PDF poles is rotated<br />

until all poles fall into the circles of the standard stereographic<br />

projection (Fig. 16). Usually, there should be only one indexing<br />

solution; if the measurements are not precise enough, some PDFs<br />

can remain unindexed. The stereographic projection yields not only<br />

the Miller indices of PDF planes but also provides information on<br />

the coexistence of positive and negative or right and left forms (Langenhorst<br />

and Deutsch 1994). Figure 17 shows an example for typical<br />

PDF combinations in quartz shocked at a pressure of 25 GPa. It<br />

demonstrates that the positive rhombohedra {101 — 1} and {101 — 3} are<br />

combined with the negative rhombohedron {01 — 12} and the right<br />

pyramid {112 — 2}. It is of course not possible to distinguish between<br />

a negative and a positive form but when indexing a stereogram one<br />

has to make an arbitrary choice for the first PDF pole (positive or<br />

negative) and can then consistently index the following poles.<br />

282

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!