Tiwari - 2022 - JGR-Planets - Vesicular Olivine and Oyroxene

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

RESEARCH ARTICLE Vesicular Olivines and Pyroxenes in Shocked Kamargaon L6

10.1029/2022JE007420
Chondrite: Implications for Primary Volatiles and Its Multiple
Key Points:
• W e report the first occurrence of
Impacts History
vesicular olivine and pyroxene in an Kishan Tiwari1  , Sujoy Ghosh1  , Masaaki Miyahara2  , and Dwijesh Ray3 
ordinary chondrite
• Vesicles possibly formed due to 1
Department of Geology and Geophysics, Indian Institute of Technology, Kharagpur, India, 2Graduate School of Advanced
release of sulfur vapor, evaporation
Science and Engineering, Hiroshima University, Higashihiroshima, Japan, 3Planetary Sciences Division, Physical Research
of moderately volatile elements, and
vaporization of silicate grains Laboratory, Ahmedabad, India
• Kamargaon meteorite record
multiple impacts in the form of
textures of contrasting magnitude Abstract  Abundant vesicles are found in terrestrial rocks, basaltic meteorites, and carbonaceous chondrites
on inter-granular and intra-granular
which testify to the presence of significant quantities of volatiles and favorable conditions for vesiculation.
scales
Furthermore, vesicular olivine has been reported in terrestrial rocks and carbonaceous chondrites. However,
vesicles in the ordinary chondrites are rare due to their low volatile content and obliteration by the late impact
Supporting Information:
events. Here, we report the first evidence of vesicular olivine (Fo76) and pyroxene (En73–81Fs17–26Wo01–02)
Supporting Information may be found in
the online version of this article.
in an ordinary chondrite. The vesicular texture in the shocked Kamargaon L6 chondrite possibly formed
due to localized melting during a shock event and subsequent degassing of volatiles after decompression.
We propose three possible mechanisms for vesicle formation in the Kamargaon meteorite: (a) liberation of
Correspondence to:
K. Tiwari,
S2 vapor, (b) evaporation of moderately volatile elements (MVEs) like Na and Mn, and (c) vaporization of
kishantiwari@iitkgp.ac.in olivine and pyroxene, by constraining the primary abundance of volatiles based on the observed volume
of  vesicles. We suggest that all three or any combination of these mechanisms could be responsible for vesicle
Citation: formation. Segmented texture in olivine is also observed in the shock vein (SV) of the Kamargaon chondrite.
Tiwari, K., Ghosh, S., Miyahara, M., The segmentation has developed due to the formation of sub-grain boundaries during the recovery process
& Ray, D. (2022). Vesicular olivines when grains were subjected to localized shear stress. Average shock pressure and temperature conditions in
and pyroxenes in shocked Kamargaon
L6 chondrite: Implications for primary
the SV are ∼24–25 GPa and ∼2310–2633K, respectively. The thermal model of the SV cooling gives the
volatiles and its multiple impacts history. crystallization time of ∼50 ms and shock pulse duration of ∼2s.
Journal of Geophysical Research:
Planets, 127, e2022JE007420. https://doi. Plain Language Summary  Shocked meteorites are extraterrestrial rocks that experience transient
org/10.1029/2022JE007420
high-pressure and high-temperature conditions caused by planetary collisions in outer space and contain
Received 9 JUN 2022 abundant fractures subsequently filled with frictional melt known as shock veins. Formation and preservation
Accepted 8 NOV 2022 of vesicles are common in terrestrial rocks. However, vesicles are rare in the chondritic meteorites due to their
low volatile content and destruction of these vesicles by the late impact events on the parent bodies. This study
Author Contributions: reports the first occurrence of vesicular olivine and pyroxene in an ordinary chondrite. We used the volume
Conceptualization: Kishan Tiwari, Sujoy of vesicles present in the Kamargaon meteorite to estimate the abundance of elements such as sulfur, sodium,
Ghosh manganese, and iron, whose degassing resulted in the formation of vesicles. These vesicles coexist with other
Formal analysis: Kishan Tiwari, Sujoy
Ghosh, Masaaki Miyahara, Dwijesh Ray textures like dissociated olivine and high-pressure phases in our sample. These different textures and phases
Funding acquisition: Sujoy Ghosh correspond to contrasting magnitudes of pressure and temperature of formation. Their coexistence in the same
Methodology: Kishan Tiwari sample and some cases within the same grain indicates that the Kamargaon meteorite has experienced multiple
Resources: Dwijesh Ray
Supervision: Sujoy Ghosh impacts of different magnitudes. We propose that the Kamargaon meteorite came from an asteroid of ∼6.4 km
Visualization: Kishan Tiwari in size which collided with another asteroid body with a velocity of ∼2.3 km/s.
Writing – original draft: Kishan Tiwari
Writing – review & editing: Kishan
Tiwari, Sujoy Ghosh, Masaaki Miyahara,
Dwijesh Ray
1. Introduction
Impact events on the surface of planets and asteroids are one of the most fundamental processes responsible for
the formation and evolution of the celestial bodies in the solar system. Shock metamorphism in meteorites takes
place as a result of such impacts among asteroid parent bodies and on larger rocky planetary bodies (e.g., Earth,
Mars, and Moon). The generation and propagation of shock waves create high pressure and friction along numer-
ous fractures developed during the shock event that give rise to high-temperature and frictional melting (Stöffler
et  al.,  2018), and consequently, the shock veins (SVs) in the parent body are formed. Various high-pressure
© 2022. American Geophysical Union. minerals have been reported in and around the SVs in different shocked meteorites, particularly in the ordinary
All Rights Reserved. chondrites as they are most abundant of all the meteorite classes on the Earth (Nakamura et al., 2011) and thus

TIWARI ET AL. 1 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

have been extensively studied (e.g., Ghosh et al., 2021; Hu & Sharp, 2017; Miyahara et al., 2016, 2021; Ohtani
et al., 2004; Sharp & DeCarli, 2006; Tomioka & Miyahara, 2017).

The stability range of high-pressure minerals found in and around SVs in chondrites either formed by the
solid-state transformation or crystallization of chondritic melts in the SVs, has been used to constrain the pres-
sure and temperature conditions of their formation (Chen et al., 1996; Miyahara et al., 2017; Ohtani et al., 2004;
Xie et al., 2006). The mineralogy, composition, grain size, and microtextures developed in melting experiments
of mantle peridotite and chondrite compositions (Agee et al., 1995; Zhang & Herzberg, 1994) are similar to the
high-pressure minerals crystallize in the SVs within shocked chondrites at high pressure and temperature condi-
tions (Xie et al., 2006). These observations allow us to use melting phase relations derived from static high-pressure
melting experiments (Agee et al., 1995; Zhang & Herzberg, 1994) to constrain the pressure-temperature (P-T)
conditions of SV crystallization (Chen et  al.,  1996). Estimation of pressure, temperature, and time history of
shock events recorded in the SVs in the meteorites have been used to interpret the impact history of their parent
bodies (Ohtani et  al.,  2004; Sharp & DeCarli,  2006; Stöffler et  al.,  1991). Additionally, the shocked meteor-
ites play a key role to understand the phase transformation mechanisms. For example, deciphering if the new
phase (high-pressure polymorph) has nucleated homogeneously (i.e., throughout the parent grain) or heteroge-
neously (i.e., intracrystalline nucleation mechanism along defects or grain boundary nucleation) with coherent
or incoherent crystallographic relation with the parent phase and if further growth was interface controlled or
diffusion controlled (e.g., Chen et al., 2006, 2007; Gillet et al., 2007; Kato et al., 2017; Sharp & DeCarli, 2006).
Therefore, comprehensive study of shocked meteorites allows us to probe the Earth and planetary interiors
where high-pressure and high-temperature conditions prevail. One should, however, proceed with caution while
comparing the phase transformation processes in the shocked meteorites with that occurring in the planetary
interiors as the strain rate in the impact events are several orders of magnitude higher than those existing at the
mantle depths (Gillet et  al.,  2007; Langenhorst,  2002). This difference in strain rate may cause variations in
phase transformation mechanisms and the P-T conditions at which such transformations may occur (e.g., Chen
et al., 2004; Cordier et al., 2012). Nevertheless, discoveries of new high-pressure phases in the shocked meteorite
samples continue to serve as the anchor points for static high-pressure experiments to better constrain the P-T
conditions of natural phase transitions (e.g., Chen et al., 2003).

Contrary to the high-pressure minerals in the shocked meteorites, the formation of vesicles (bubbles) requires
high-temperature and low-pressure conditions for degassing of volatiles in the chondritic parent bodies.
Further impact or heating events in the parent bodies may obliterate these vesicular textures due to remelting
and high-pressure conditions. A low volatile content further makes it difficult for the formation of vesicles in
the ordinary chondrite parent bodies. Due to these reasons, vesicular olivines and pyroxenes have never been
reported in the ordinary chondrites. These observations indicate that the amount and species of volatiles pres-
ent in the ordinary chondrites and the prevailing physical conditions like pressure, temperature, and gravity
put constraints on the vesicle formation. On the other hand, vesicles are readily found in terrestrial rocks (e.g.,
Oppenheimer et al., 2003), basaltic achondrites such as the eucrites and the shergottites (e.g., He et al., 2015;
McCoy et al., 2006) and carbonaceous chondrites (e.g., Lunning et al., 2016). Similarly, vesicular olivines have
also been reported in terrestrial rocks, for example, komatiite (Dann, 2001) and carbonaceous chondrites (Ohnishi
et al., 2007; Tomeoka et al., 2001), but are relatively rare than vesicles present in melt breccia and matrix of SVs.

The Kamargaon meteorite is a heavily shocked L6 chondrite, which fell on November 13, 2015 near the town of
Kamargaon in the Indian state of Assam (Goswami et al., 2016; Ray et al., 2017). Previous studies in the Kamar-
gaon chondrite reported the textures and occurrence of different low-pressure silicate phases, metal, troilite, and
its radiogenic and cosmic-ray exposure age (Goswami et al., 2016; Ray et al., 2017). Recently, shock-induced
incongruent melting of olivine dissociated into magnesiowüstite and orthoenstatite was reported in an SV of the
Kamargaon L6 chondrite (Tiwari et al., 2021). In this study, we carefully examined the SV of the Kamargaon
chondrite in detail to understand the dissociation and melting textures displayed by the silicate phases (olivine
and pyroxene) present and their formation mechanisms which provide further clues to estimate the shock condi-
tions in the chondrite parent body. Here, we found the first occurrence of vesicular olivine and pyroxene in
an ordinary chondrite. We also investigated the occurrence of high-pressure phases, and back transformation
textures in an SV of the Kamargaon L6 chondrite and calculated the shock-pulse duration required to produce the
mineral assemblage. Based on these calculations, we further calculated the impact velocity and parent body size
of the Kamargaon meteorite.

TIWARI ET AL. 2 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

2.  Experimental Methods


A small chip of the Kamargaon chondrite was mounted in a low-viscosity epoxy resin and its surface was
polished with diamond paste. The experimental methods have been described in detail by Tiwari et al. (2021) and
only essential details are repeated here. Textural and mineralogy of the Kamargaon chondrite were made using
Scanning Electron Microscope (SEM) equipped with an X-ray energy-dispersive spectrometer (EDS), a JEOL
JSM 6490. Fine textural observations were conducted with Field-Emission Gun Scanning Electron Microscope
(FEG-SEM), JEOL JSM-7000F with an accelerating voltage of 15 kV. Chemical compositions of the constituent
phases were performed with wavelength dispersive spectrometers (WDS) employing a Cameca-SX 100 electron
microprobe with an acceleration voltage of 15 kV. Feldspar grains were analyzed with a beam current of 10 nA
and a defocused beam (5–10 μm) while all other phases were analyzed with a beam current of 15 nA and a focused
beam. The average detection limits of the analyses are as follows: 120 ppm for SiO2, CaO, MgO; 75 ppm for
Al2O3; 150 ppm for P2O5, V2O3; 165 ppm for Cr2O3; 180 ppm for MnO, TiO2; 220 ppm for FeO; and 260 ppm for
NiO. Raman spectra of constituent minerals in the Kamargaon meteorite were measured by a Horiba Jobin-Yvon
LabRam HR800 laser micro-Raman spectrometer with a liquid nitrogen-cooled charge-coupled device (CCD)
detector. A microscope was used to focus the excitation laser beam (a He-Ne laser, 633 nm line). Two slices
were excavated with a focused ion beam (FIB) system, JEOL-9320FIB. A gallium ion beam was accelerated at
30 kV during sputtering of the slices and the slices were approximately 100 nm in thickness. A JEOL JEM-2100F
field-emission transmitted electron microscopy (FE-TEM) operating at 200 kV with an EDS detector system was
used for conventional TEM observation to characterize the microstructures of the minerals and selected area
electron diffraction analyses. Chemical composition of individual minerals was determined by STEM-EDS.

3. Results
3.1. Petrography

The host rock of Kamargaon chondrite consists of olivine (Fo73–74), low-calcium (Ca) pyroxene
(En77–80Fs19–22Wo1–2), high-Ca pyroxene (En45–46Fs9–10Wo44–46), plagioclase (Ab62–70An18–23Or12–15) (Table  1),
iron-nickel alloy (kamacite and taenite), troilite plus a minor amount of phosphate and chromite. The studied
sample of Kamargaon meteorite shows a typical equilibrated chondrite texture and contains some chondrule
fragments typically 0.5–1 mm in diameter. Relict outlines of pre-existing chondrules are visible only at a few
locations; however, the matrix portion has been completely recrystallized. Chondrules are usually composed of
olivine and pyroxene with a smaller amount of feldspar and Fe-Ni metal. Backscattered electron (BSE) images
of the phases present in the chondritic portion apparently do not show variation in greyscale contrast within
individual grains and between grains of the same phase which is indicative of their homogeneous major element
compositions. Olivine and pyroxene are subhedral to euhedral with a grain size up to 100 μm in diameter and
display numerous shock-induced irregular fractures. Plagioclase grains are irregular in shape with a size range
from ∼20 to 100 μm where smaller grains (<35 μm) are generally euhedral whereas larger grains are subhedral to
completely anhedral. The olivine contains 73–74 mol% forsterite which is in the compositional range of L-group
ordinary chondrites (Weisberg et al., 2006). Many Shock-induced fractures in the host rock portion are filled by
quenched metal-sulfide melt (Figures 1, 2b–2e).

The host rock shows a pervasive SV of up to 1.6 mm thickness passing through the middle of the sample (Figure 1),
enclosing coarse-grained subrounded, rounded, and elongated fragments of host minerals in a fine-grained matrix
with a few irregular shaped metal-sulfide eutectic intergrowths (Figure 2a). Plagioclase grains of the SV have
been amorphized. The fine-grained matrix of the SV consists of euhedral and rounded grains (1–5 μm) of olivine,
pyroxene with interstitial Fe-Ni metal, and troilite.

3.2.  Segmented Olivines

Several rounded and coarse-grained olivine fragments (100–200  μm) entrained in the SV show formation of
cellular walls resulting in segmentation texture with individual segments up to ∼5 μm in size (Figures 2b and 2c).
Two discrete lamellar zones are visible in the middle portion of the segmented olivine grain and are in light gray
contrast while the rest of the grain is in dark gray contrast as observed in BSE image. EPMA analyses of these
two domains show that the light gray lamellar zones are Fe-rich olivine (Fo67–70), whereas the dark gray zones are
relatively Fe-poor (Fo73–76) (Table 2).

TIWARI ET AL. 3 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Table 1
Average Chemical Composition of Different Silicate Phases Present in the Host Rock (HR) and Shock Vein (SV) in the
Kamargaon L6 Chondrite Analyzed by Electron Microprobe
Location HR SV

Phases Ol 1σ LPx 1σ HPx 1σ Plag 1σ En 1σ Am. Plag 1σ


n 21 13 6 4 17 2
SiO2 37.48 0.47 54.41 0.65 53.70 0.21 69.58 0.25 53.39 1.34 69.16 0.36
TiO2 n.d. 0.00 0.21 0.03 0.52 0.04 n.d. 0.00 0.21 0.04 n.d. 0.00
Cr2O3 n.d. 0.00 0.13 0.04 0.98 0.06 n.d. 0.00 0.15 0.12 n.d. 0.00
Al2O3 n.d. 0.00 0.14 0.04 0.51 0.08 23.07 0.13 0.16 0.04 23.04 0.14
FeO  a
24.44 0.24 15.55 0.55 5.73 0.14 0.71 0.24 17.13 1.67 1.13 0.01
MnO 0.45 0.04 0.50 0.04 0.27 0.06 0.00 0.00 0.59 0.13 0.00 0.00
MgO 38.01 0.16 28.43 0.35 16.28 0.11 0.05 0.02 27.56 0.62 0.145 0.01
CaO 0.04 0.03 0.77 0.11 22.28 0.33 2.08 0.06 0.82 0.18 2.365 0.07
Na2O n.d. 0.00 n.d. 0.00 n.d. 0.00 3.56 0.39 n.d. 0.00 3.77 0.02
K2O n.d. 0.00 n.d. 0.00 n.d. 0.00 1.13 0.02 n.d. 0.00 0.915 0.09
Total 100.42 100.14 100.27 100.17 100.01 100.53
Cation
 Si 0.978 1.953 1.977 3.194 1.915 3.187
 Ti – 0.006 0.014 – 0.005 –
 Cr – 0.000 0.029 – 0.003 –
 Al – 0.006 0.022 1.254 0.006 1.252
 Fe 
2+
0.533 0.467 0.177 0.023 0.536 0.040
 Mn 0.010 0.015 0.008 0.000 0.019 0.000
 Mg 1.478 1.521 0.894 0.003 1.485 0.010
 Ca 0.001 0.030 0.879 0.105 0.029 0.117
 Na – – – 0.357 – 0.337
 K – – – 0.065 – 0.054
Total 3.000 4.000 4.000 5.000 4.000 5.000
 Oxygen 4 6 6 8 6 8
 Fo 74
 En 78 46 78
 Fs 20 09 21
 Wo 02 45 01
 Ab 65 66
 An 21 23
 Or 14 11
Note. Ol  =  olivine, LPx  =  Low-Ca pyroxene, HPx  =  High-Ca pyroxene, Plag  =  plagioclase, Am. Plag  =  amorphized
plagioclase, n = number of analyses; 1σ = standard deviation, n.d. = not detected.
 aAll iron is assumed to be ferrous.

3.3.  Composite Olivines

Numerous olivine grains entrained in the SV occur as composite grains display the combination of different
textures. Such composite olivine grain present near the vein edge shows heterogeneous texture where the core
displays segmentation texture (Se-Ol), while the remaining part of the grain shows vesicular texture (Ve-Ol)
(Figures 2d and 2e). BSE image shows that the segmented core portion of the grain has a variable contrast within
individual segments consist of light gray cores and dark gray rims indicate their heterogeneous composition.

TIWARI ET AL. 4 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

The core segmented part of the olivine grain seems to be embayed and
engulfed by the vesicular portion. Such heterogeneous texture and composi-
tion are conspicuously common in the grains that are relatively coarser and/or
occur near the vein edge. The Raman spectra of the compositionally zoned,
segmented core part yield strong peaks at ∼821 and ∼853 cm −1, respectively
(Figure 3) which corresponds to olivine whereas for the vesicular part of the
grain, two main Raman shifts appeared at ∼821 and ∼851 cm −1 (Figure 3)
corresponds to olivine as well.

The portions correspond to vesicular and segmented regions of the composite


olivine grain were excavated using FIB for TEM analyses (Figure 4). TEM
images of the vesicular portion show the presence of equigranular subhe-
dral to anhedral grains about ∼100–200 nm across (Figure 4a). Bulk chem-
ical composition measured by STEM EDS gives the forsterite content of
Figure 1.  Backscattered electron (BSE) image of the Kamargaon L6 chondrite
∼76 mol% (Table 3) which is similar to the forsterite content in the host rock
shows ∼1.6 mm thick shock vein passing through the middle of the sample. olivine (∼Fo74). TEM images of the segmented core part show a polycrystal-
The scale bar is 1 mm. line assemblage of euhedral to subhedral grains with the presence of triple
junctions (equilibrated grain boundaries with interfacial angle of 120°) and
ranging ∼200–300 nm in size measured across (Figure 4b). STEM EDS anal-
yses of the segmented core part give the composition in two distinct clusters
of Fo68–73 and Fo84–86 (Table 3). Individual segments are chemically zoned as indicated by BSE images with light
gray cores and dark gray rims (Figure 2e). Although individual core and rim parts of each segment are too fine to
be analyzed distinctly and precisely, we expect that the composition of Fo68–73 belongs to the brighter (as seen in
BSE image) iron-rich core as it has higher average atomic number and Fo84-86 to the darker (as seen in BSE image)
iron-poor rim as it has lower average atomic number (Table 3).

Another type of composite grains of olivine present in the SV of the Kamargaon chondrite show dissociation
texture as well as vesicular texture. The core part of such grains shows vesicular texture whereas the outer rim
part of the grain exhibits dissociation texture and spherulitic texture is present in between them (see Figure 1;
Tiwari et al., 2021).

3.4.  Vesicular Pyroxenes

Many pyroxene grains show the presence of a large number of vesicles. A grain of such vesicular pyroxene
(Ve-Px) is illustrated in Figure 2f with the small veinlet offshoots branching from it into the neighboring grain.
The Raman spectra of the vesicular pyroxene grains yield strong peaks at ∼334, ∼658, ∼676, and ∼1,006 cm −1,
and weak peaks at ∼394 and ∼936 cm −1, corresponding to enstatite (Figure 3). The vesicular pyroxenes entrained
in the SV are compositionally more variable (En73–81) (Table  S1) than the parent low-Ca pyroxene (En77-80)
(Table 1). Nevertheless, the average composition of these vesicular pyroxenes (En78Fs21Wo1) is similar to the
low-Ca pyroxene (En78Fs20Wo2) present in the host rock (Table 1).

4. Discussion
4.1.  Formation of Shock Features
4.1.1.  Segmented Olivine

Numerous olivine grains exhibit cellular-like texture as a result of segmentation which represents subgrain
boundaries (Figure  2c). Such texture resembles closely the cleavable olivine common in the terrestrial shear
zone rocks, for example, mylonites (Nozaka & Ito, 2011). These cleavable olivines are formed during the release
process subsequent to the deformation of olivine grains subjected to localized shear stress. The cellular walls in
segmented texture are previously formed parting planes in the form of the subgrain boundary with high dislocation
density. Such subgrain boundaries form via diffusion of the dislocations common in the crystal and are thermally
activated. The SVs occurring in the shocked meteorites form in a similar manner. Collisions between asteroids in
outer space generate shock waves and numerous fractures. Impact-induced shearing along these fracture planes
produces frictional heating and subsequently causes melting. Therefore, the SVs are the result of impact-induced

TIWARI ET AL. 5 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Figure 2.  Backscattered electron (BSE) image of the Kamargaon L6 chondrite: (a) Low magnification image shows Fe-Ni
metal-troilite eutectic intergrowth. (b) Portion of the sample shows shock vein (SV) and the host rock demarcated by a dashed
line. Many host rock fragments are entrained in the SV which show different textures and compositional heterogeneity.
(c) High magnification image of the boxed area in Figure 1b shows a coarse olivine (∼100 × 150 μm 2) grain displaying
segmentation texture. Two discrete fayalite rich (Fo67–70) lamellar zones are present at the center of the grain. Rounded
metal-sulfide blebs present inside the segmented olivine grain are indicated by arrows. (d) An elongated, and relatively coarse
grain of olivine present near the vein edge shows heterogeneous texture and composition with segmented core and vesicular
rim. (e) High magnification image of the boxed in Figure 1e revealing that the individual segments in the segmented core
parts are heterogeneous in composition with relatively iron-rich core and magnesium-rich rim. (Ve-Ol = vesicular olivine;
Se-Ol = segmented olivine). The portions marked as TEM_1 and TEM_2 show the locations of TEM lamella prepared by
FIB. (f) High-magnification image shows the presence of vesicular pyroxene (Ve-Px). The arrows indicate the offshoots
branching from the pyroxene grain into the surrounding material.

localized shear zones at high pressure and temperature conditions. Thus, the SVs found in meteorites are believed
to be analogous to S-type (shock type) pseudotachylites observed in the terrestrial impact sites as they are consid-
ered to form during the shock compression stage of the impact (Kenkmann et al., 2000; Spray, 1998; Van der
Bogert et al., 2003). Therefore, as they form when the material is experiencing high-pressure due to shock, they
may contain high-pressure polymorphs. The characteristic properties of S-type pseudotachylites which can be

TIWARI ET AL. 6 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Table 2
Major Element Composition of Segmented Olivine Grain With Two Discrete Lamellar Zones at the Center Analyzed by Electron Microprobe
Lamellar zones (light gray) Dark gray zone

SiO2 35.55 36.01 36.95 38.02 37.71 37.86 37.68 37.38 37.33 37.48 37.78 37.2
FeO a 31.9 29.56 27.67 22.67 23.27 23.49 23.00 23.37 23.58 24.57 22.38 23.81
MnO 0.40 0.47 0.46 0.56 0.52 0.5 0.56 0.47 0.46 0.52 0.53 0.48
MgO 32.35 33.13 35.66 39.69 38.72 38.73 39.03 39.22 38.2 37.94 39.77 38.46
CaO 0.04 0.02 0.05 0.04 0.04 0.03 0.05 0.09 0.05 0.03 0.02 0.03
Total 100.24 99.18 100.79 100.98 100.26 100.61 100.32 100.53 99.62 100.54 100.48 99.98
Cation
 Si 0.923 0.977 0.975 0.978 0.981 0.982 0.978 0.968 0.979 0.977 0.975 0.972
 Fe  2+
0.692 0.671 0.611 0.488 0.506 0.509 0.499 0.506 0.517 0.536 0.483 0.520
 Mn 0.009 0.011 0.010 0.012 0.011 0.011 0.012 0.010 0.010 0.011 0.012 0.011
 Mg 1.375 1.340 1.403 1.521 1.501 1.497 1.509 1.513 1.493 1.475 1.530 1.497
 Ca 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.002 0.001 0.001 0.001 0.001
Total 3.00 3.00 3.00 3.00 3.00 3.00 3.00 3.00 3.00 3.00 3.00 3.00
 Oxygen 4 4 4 4 4 4 4 4 4 4 4 4
 Fo 66 67 70 76 75 75 75 75 74 73 76 74
 aAll iron is assumed to be ferrous.

used to differentiate them from other impact induced melts are (Kenkmann et al., 2000; Spray, 1998): (a) the
boundary between SV and the host rock is irregular but sharply defined as can be seen in Figures 2b and 2d and
Figures S1a and S1b in Supporting Information S1 (Kenkmann et al., 2000; Stöffler et al., 1991) (b) they consist
of angular to subrounded fragments of the host rock entrained in microcrystalline matrix formed by fictional
melting (Spray, 1998; Van der Bogert et al., 2003) similar to the texture observed in the SV of the Kamargaon
L6 chondrite (Figure 2d, Figures S1b–S1d in Supporting Information S1); (c) intense fracturing of the host rock
material in the vicinity of the SV attributed to shock deformation (Walton, 2013) and emplacement of melt from
SV along these fractures into the host rock (Kenkmann et al., 2000) which is present in the Kamargaon L6 chon-
drite (Figures S1a and S1b in Supporting Information S1); (d) the mechanical breakdown of host rock fragments
along the shear zone by comminution which ultimately leads to frictional melting, results in grains consisting of
fractures (Kenkmann et al., 2000) (Figure S1c in Supporting Information S1); (e) evidences of shear displace-
ment of grains across the SV (Kenkmann et al., 2000). These kinds of shear displacements are mostly visible
when the width of the SV is lower than the average grain size which is ˂1 mm for ordinary chondrites (Kuebler
et al., 1999; Simon et al., 2018). The main SV in the Kamargaon L6 chondrite is too thick (1.6 mm) for any shear
displacement to be observable across the SV. However, some thin veins that probably postdate the main SV (in
some cases they fill the fractures in the host rock fragments) that show shear displacement across them (Figure
S1c in Supporting Information S1). (f) Presence of high-pressure polymorphs which were also present in the
Kamargaon L6 chondrite but later got back-transformed to the low-pressure phase as discussed in Section 4.1.1
(also see Tiwari et al., 2021).

Furthermore, the olivine grains in and around the SV experience high temperature but the shock compres-
sion suppresses the thermal expansion which results in high thermal stress in the olivine grains. Consequently,
high-strain energy gets stored in the grains and they become unstable. To ease off this stored strain, as a recovery
process, dislocations from different parts of a grain migrate to form a dislocation wall and divide one strained
grain into two, relaxed grains with a subgrain boundary between them. The core of the grain preserving the
segmented texture in Figures 2d and 2e shows heterogeneous composition within each segment, where the core
of each segment is light gray (Fe-rich) and the rim of each segment is dark gray (Mg-rich). Such compositional
zoning may be due to the interdiffusion of Fe-Mg during the segmentation process as reported in previous studies
(Miyahara et al., 2016, 2019).

TIWARI ET AL. 7 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

The whole, relatively coarser (∼100 × 150 μm 2), grain in Figure 2c shows the


development of segmentation with two discrete lamellar zones of relatively
Fe-rich composition. Solid-state polymorphic transformation of olivine to
ringwoodite occurs through different mechanisms. The mechanism in which
ringwoodite nucleates along the grain boundary of parent olivine grain and
grow inwards via interface-controlled growth, is mainly dominant in smaller
grains (<100 μm) (El Goresy et al., 2005; Gillet et al., 2007). In such cases,
the parent olivine grain is eventually replaced with polycrystalline ringwood-
ite with triple junctions. In coarser grains (>100  μm), such as segmented
olivine grain in Figure 2c, intracrystalline nucleation and diffusion-controlled
growth of ringwoodite along deformation-induced planar defects like stack-
ing faults and fractures are more dominant (Chen et al., 2006, 2007; Ohtani
et al., 2004). It involves the initial nucleation of ringwoodite platelet. along
with planar defects via a coherent mechanism and is subsequently followed
by incoherent nucleation (Kerschhofer et al., 1998, 2000).
Moreover, during such polymorphic transformation of olivine to wadsley-
ite or ringwoodite, iron preferentially partitioned into the high-pressure
phases (Katsura & Ito,  1989). Therefore, in the shocked meteorites, when
olivine-wadsleyite or olivine-ringwoodite coexist together, they have a
compositional gap such that the olivine is more Mg-rich than its coexist-
ing high-pressure polymorph (e.g., Feng et al., 2011; Miyahara et al., 2008).
Furthermore, solid-state transformation of olivine to its high-pressure poly-
morphs is a reconstructive type polymorphic transformation involving the
breaking of bonds accompanied by structural rearrangement and thus are
energetically sluggish (Sharp & DeCarli,  2006). Therefore, metastable
olivine can persist up to the stability field of the ringwoodite/wadsleyite and
pressure overstepping is required to overcome the activation energy of nucle-
ation (Kerschhofer et al., 1998). It is important to note that at any point in
space and time, the phase having the least molar Gibbs free energy is the
most stable one. Hence, the larger the extent of overstepping the higher is the
difference between molar Gibbs free energy (ΔGm) of the stable (ringwoodite/
Figure 3.  Representative Raman spectra of olivine present in the host rock as wadsleyite) and metastable phase (olivine) (Riedel & Karato, 1997). Conse-
well as vesicular and segmented olivine and vesicular pyroxene; HR-Ol = host quently, it becomes kinetically easier to nucleate the new phase which leads
rock olivine; Ve-Ol = vesicular olivine; Se-Ol = segmented olivine;
to immediate site saturation resulting from high nucleation rate (reflected as
Ve-Px = vesicular pyroxene.
the value of n closer to 1 in Avrami equation, e.g., Kubo et al., 2022). More-
over, in such transformations where overstepping of equilibrium boundaries

Figure 4.  (a) Bright-field transmitted electron microscopy (TEM_ image of vesicular olivine in the SV of the Kamargaon
meteorite, from the rectangular box of Figure 2e, shows inequigranular subhedral and anhedral texture. The selected
area electron diffraction (SAED) pattern of olivine along the [011] zone axis is shown in inset. Scale bars, 10 1/nm. (b)
Bright-field TEM image of segmented olivine, from the rectangular box of Figure 2f, shows polycrystalline assemblage of
euhedral to subhedral grains. The SAED pattern of the olivine along the [011] zone axis from the segmented olivine core part
is shown in inset. (c) High magnification TEM image of the boxed area in (b) shows the presence of triple junctions.

TIWARI ET AL. 8 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Table 3 is required, the grain size of the product phase (ringwoodite/wadsleyite) is


Major Element Composition of Different Types of Olivine Present in the inversely related to ΔGm and nucleation rate (Riedel & Karato, 1997). Thus,
Shock Vein of Kamargaon L6 Chondrite Determined by STEM-EDS reconstructive type transformation of olivine results in grain size reduction,
Phases Ve-Ol Se-Ol and the parent olivine grain is replaced by a polycrystalline assemblage
because of the nucleation rate of ringwoodite is relatively higher than its
Location Bulk Core Rim
grain growth rate (Kerschhofer et al., 2000; Riedel & Karato, 1997). When
SiO2 39.68 36.54 36.99 43.06 41.29 42.31 polycrystalline ringwoodite assemblage experiences high temperature under
FeO  a
21.66 29.03 24.94 12.65 15.21 13.65 ambient pressure, it back-transforms to polycrystalline olivine assemblage
MgO 38.66 34.42 38.07 44.29 43.50 44.04 (Fukimoto et al., 2020; Ming et al., 1991). TEM images of the segmented
part reveal that it is a polycrystalline assemblage with triple junctions
Cation
(Figures 4b and 4c). Therefore, based on observed textures and composition,
 Si 1.031 0.978 0.968 1.080 1.042 1.063
we suggest that the two discrete Fe-rich lamellar zones in the segmented
 Fe 0.471 0.650 0.546 0.265 0.321 0.287 olivine grain in Figure 2c were lamellar ringwoodite/wadsleyite once formed
 Mg 1.498 1.373 1.486 1.655 1.637 1.650 by intracrystalline nucleation and diffusion-controlled growth. Later they
Total 3.00 3.00 3.00 3.00 3.00 3.00 got back-transformed to olivine post-decompression due to high post-shock
 Oxygen 4 4 4 4 4 4 temperature.
 XMg 0.76 0.68 0.73 0.86 0.84 0.85
4.1.2.  Vesicular Olivine and Pyroxene
Note. XMg  =  Molar Mg/(Mg  +  Fetot), Ve-Ol  =  vesicular olivine,
Se-Ol = segmented olivine, n = number of analyses. During a collisional event, the shock wave produced passes through the
 aAll iron is assumed to be ferrous. projectile and the target, both consisting of grains of different minerals
having varied shock impedance (defined as the product of shock wave veloc-
ity in a material and density of the material). Detailed microscopic study
and computational modeling of shock front behavior in shock experiments
suggest that pressure heterogeneity and localized spikes in temperature may occur when shock waves inter-
act with mineral grains of different shock impedance (Baer, 2000; Bischoff & Stöffler, 1992; DeCarli, 1979).
Furthermore, Shock experiments on volatile bearing minerals have revealed that high-pressure conditions impede
the degassing process however after decompression degassing can take place massively if the post-shock temper-
ature is sufficiently high (Langenhorst, 2002).

Degassing of volatiles and gas bubble formation commonly result in the formation of terrestrial vesicular basalt
(Oppenheimer et al., 2003). Vesicle formation is common in the terrestrial rocks, however, it is rarely seen in
meteorites mainly due to differences in the parent body size, presence or absence of atmosphere, type, and abun-
dance of volatiles (McCoy et al., 2006). The larger size of the Earth as compared to the asteroids and the presence
of atmosphere results in a high value of gravity and limits the expansion and subsequent escape of volatiles
during magma extrusion. These factors allow even thin terrestrial lava flows to retain the gas bubbles and thus
the consequent formation of vesicles. By contrast, smaller size and absence of atmosphere impede the formation
of vesicles in lava flows on asteroids (McCoy et al., 2006; Wilson & Keil, 1997). Although the base of lava flows
on the asteroids may retain gas bubbles due to the overlying weight of lava, bubble retention would require the
formation of much thicker lava flows which is unlikely on asteroids (McCoy et al., 2006; Wilson & Keil, 1997).

4.1.2.1.  Shock-Induced Localized Melting

Vesicle formation in meteorites has been mainly reported in basaltic achondrites (e.g., eucrites) with CO and CO2
as dominant vesicle-forming gases (McCoy et al., 2006; Wilkening & Anders, 1975). In addition, the formation
of vesicular olivine has been reported in Kobe and Karoonda CK4 carbonaceous chondrites (Ohnishi et al., 2007;
Tomeoka et al., 2001). Rubin and Moore (2011) found vesicles in the impact melt breccia in the Larkman Nuna-
tak 06299 LL chondrite, however, individual vesiculated olivine or pyroxene grains were absent in their sample.
The rare formation of vesicles in ordinary chondrites may be due to several reasons. First, the ordinary chondrites
consist of a much lower modal proportion of matrices. Second, they are relatively volatile-poor and less porous
as compared to the carbonaceous chondrites (Tomeoka et al., 2001).

Ohnishi et al. (2007) proposed that the vesicular olivines in the Karoonda CK4 chondrite are formed from a high
temperature and low-pressure shock event that caused a local spike in temperature which was sufficiently high to
melt the olivine (∼2,073K for Fo76). This led to shock-induced instantaneous heating followed by partial melting
of the part containing olivine. Then subsequent rapid cooling resulted in situ crystallization. They also found the
presence of a high density of dislocations and formation of subgrain boundaries in the vesicular olivine. They

TIWARI ET AL. 9 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

suggested that these dislocations and subgrain boundaries are the results of deformation from a different shock
event which must have occurred after the shock event that formed the vesicular olivine. Their rationale was that
such deformation features as dislocations and subgrain boundaries could not survive the high-temperature event
responsible for the vesicle formation.

Small veinlet offshoots are found to be branching off from vesicular grains into the neighboring chondritic
portion (Figure 2f) which serves as additional evidence that these grains were melted once. In Figures 2e and 2f,
the segmented core part (Se-Ol) of olivine is preserved surrounded by the vesicular portion (Ve-Ol). The core
part seems to be embayed and engulfed by the vesicular portion. Such observations indicate that the shock event
responsible for the formation of segmented olivine preceded the shock event that formed vesicular pyroxenes
and olivines. We propose that the vesicular gains were formed due to local melting caused by localized temper-
ature spike during a shock event and subsequent degassing after decompression when the post-shock tempera-
ture was still high. Furthermore, we suggest that the elevated post-shock temperature could have been retained
for a few  seconds by two mechanisms: (a) long slippage of fault walls along the SVs as suggested previously
(Miyahara et al., 2017) and, (b) conduction of heat from adjacent SVs and melt pockets that may have formed
overlapping thermal haloes (Shaw & Walton, 2013).

4.1.2.2.  Mechanism for Vesicle Formation

We evaluate the viability of three possible mechanisms which may be responsible for the formation of vesicular
grains in the Kamargaon L6 chondrite: (a) liberation of S2 vapor, (b) evaporation of moderately volatile elements
(MVEs), and (c) vaporization of superheated (above liquidus) olivine and pyroxene. These possible mechanisms
can be tested by calculating the maximum amount of a chemical species that can be released from our sample
during vaporization and the amount that is needed for the formation of the observed volume of vesicles. The
volume fraction of vesicles in BSE images of our sample is estimated to be 8%–12% by using ImageJ software.
The mass fraction of gas (mg) required to create the estimated volume fraction of vesicles (vg) can be calculated
using Equation 1 (Benedix et al., 2008):

𝑚𝑔= 𝑃 𝑤𝑣𝑔 ∕(1 − 𝑣𝑔 ) 𝜌𝑛𝑔 𝑅𝑇


(1)

where P is the pressure of vesicle formation (Pa), w is the molecular weight of the gas, ρng is the density of the
non-gas phase (solid or liquid) (kg/m 3), R is the universal gas constant, and T is the temperature of vesicle forma-
tion (K).

1. Liberation of S2 vapor: The presence of troilite melt has been previously proposed to be the reason for the dark-
ening of silicate grains (e.g., olivine) (Petrova & Grokhovsky, 2019; Rubin, 2015). Furthermore, troilite blebs
are present in the silicate grains in the Kamargaon chondrite (Figure 2c). Therefore, we propose that vesicles in
the Kamargaon meteorite may have formed due to the release of S2 vapor from the sulfide present in the silicate
grains. In the case of olivine, if the chemical species is S, then, w = 64 (for S2), vg = 0.12, ρng = 3,200 kg/m 3
(density of forsterite) (Tschauner, 2019), and T = 2,073K. If we consider the size of the parent body is ∼6.4 km
(estimated in Section 4.2), then the pressure at the center can be calculated by:

(2)
𝑃𝑃 = 𝐺𝐺𝐺𝐺𝐺𝐺∕2𝑅𝑅

where G is the universal gravitational constant (6.67 × 10 −11 m 3 kg −1s −2), M (4πρR 3/3) is the mass of the body
(∼4.66 × 10 14 kg), ρ is the density of the body (3,400 kg/m 3), and R is the radius (km). If ρ has the value of
3,400 kg/m 3 (Wilkison & Robinson, 2000), and R = 3,200 m, the calculated value of P comes to be ∼1.6 × 10 4 Pa.
Putting the values of all parameters in Equation 1 gives the value of mg as 2.62 × 10 −6. This means that the release
of ∼3 ppm of S is sufficient for the formation of the observed volume of vesicles in our sample. In case the
vesicles were formed in the primary parent body of the L chondrite before the separation of the ∼6.4 km sized
fragment, P = 5 × 10 5 Pa is adopted for the depth of a few km in the parent body of 150 km in size (Schmitz
et al., 2019). This value of pressure suggests that ∼80 ppm of S is required for the vesicle formation in the Kamar-
gaon meteorite. The average abundance of S in the ordinary chondrites is similar with a mean concentration of
∼2.3 ± 0.2% (∼23,000 ppm) (Dreibus et al., 1995; Wasson & Kallemeyn, 1988) and evaporation of only troilite
can liberate ∼210 ppm of S in the L chondrites (Benedix et al., 2008). These estimations suggest that the vapori-
zation of S2 vapor is a feasible mechanism for vesicle formation in the Kamargaon meteorite.

TIWARI ET AL. 10 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

2. Evaporation of MVEs: Although carbonaceous chondrites have a higher abundance of highly volatile elements
(e.g., N, O, C, and H), L chondrites are relatively enriched in MVEs (e.g., Na, Mn, and K) (Scott & Krot, 2003;
Wasson & Kallemeyn, 1988). Therefore, it is possible that the vesiculation resulted from the outgassing of such
volatiles. The average elemental abundance of Na in the L chondrites is ∼7,000 ppm (Wasson & Kallemeyn, 1988).
High-pressure experimental studies using ultramafic compositions on the partitioning of Na between olivine and
melt gives the value of the average partition coefficient 𝐷𝐷𝑁𝑁𝑁𝑁 as ∼0.0031 (Borisov et al., 2008). This indicates
( 𝑂𝑂𝑂𝑂−𝐿𝐿 )
𝐴𝐴
that the olivine while crystallizing from the primary melt, can incorporate up to ∼22 ppm of Na in the L chon-
drites. The amount of Na required for the vesicle formation in our sample can be estimated using Equation 1.
Putting w = 23 and keeping the values of all the other parameters the same as above, we get mg = ∼1 ppm and
∼33 ppm for the parent body size of 6.4 and 150 km, respectively. These values are either less or comparable to
the total amount of Na that can be present in the olivine (∼22 ppm).

The partition coefficient of Na for orthopyroxene and𝐴𝐴melt (𝐷𝐷𝑁𝑁𝑁𝑁


𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
) depends largely on pressure (GPa) and to a
lesser degree, temperature (K) and can be estimated using experimental expression (Blundy et al., 1995; McDade
et al., 2003) which yields:
[( ) ]
(3)
𝑂𝑝𝑥−𝐿
𝐷𝑁𝑎 = 10367 + 2100 − 165𝑃 2 ∕𝑇 − 10.27 + 0.358𝑃 − 0.0184𝑃 2

At T  =  1,800K (liquidus temperature of En80Fs20), estimated value𝐴𝐴 of 𝐴𝐴𝑁𝑁𝑁𝑁


𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
 is ∼0.011 for both the pressure
values of ∼1.6 × 10   Pa and 5 × 10   Pa. This indicates that ∼77 ppm (0.011 × 7,000) Na can be incorporated in
4 5

orthopyroxene in the L chondrites whereas the required concentration of Na in orthopyroxene for vesicle forma-
tion, as estimated using Equation 1, is ∼32 ppm.

The concentration of Mn in olivine present in the host rock of the Kamargaon meteorite is ∼3,500 ppm
(0.45  ×  0.774457) (Table  1). Furthermore, the average concentration of Mn in olivine in the L chondrites
can be estimated using the average abundance of Mn in the L chondrites which is ∼2,570 ppm (Wasson &
Kallemeyn, 1988), and the partition coefficient of Mn which can be calculated by (Beattie et al., 1991):

(4)
𝐷𝐷𝑖𝑖𝛼𝛼−𝐿𝐿 = 𝐴𝐴𝑖𝑖𝛼𝛼−𝐿𝐿 𝐷𝐷𝑀𝑀𝑀𝑀
𝛼𝛼−𝐿𝐿
+ 𝐵𝐵𝑖𝑖𝛼𝛼−𝐿𝐿

𝜓𝜓 𝑂𝑂𝑂𝑂 − 𝐵𝐵𝐹𝐹𝛼𝛼−𝐿𝐿
𝐹𝐹
𝑋𝑋𝐹𝐹𝐿𝐿𝐹𝐹𝐹𝐹

𝛼𝛼−𝐿𝐿
𝐷𝐷𝑀𝑀𝑀𝑀 =
𝐴𝐴𝛼𝛼−𝐿𝐿
𝐹𝐹 𝐹𝐹
𝑋𝑋𝐹𝐹𝐿𝐿𝐹𝐹𝐹𝐹 + 𝑋𝑋𝑀𝑀𝑀𝑀𝑀𝑀
𝐿𝐿

where
𝐴𝐴 𝐴𝐴 and 𝐴𝐴𝑀𝑀𝑀𝑀
𝐴𝐴𝑖𝑖𝛼𝛼−𝐿𝐿 𝛼𝛼−𝐿𝐿
are partition coefficients of element i and Mg for phase α and𝐴𝐴melt, 𝐴𝐴
𝐴𝐴𝑖𝑖𝛼𝛼−𝐿𝐿 , 𝐴𝐴
𝐴𝐴𝑖𝑖𝛼𝛼−𝐿𝐿 , 𝐴𝐴𝐹𝐹𝛼𝛼−𝐿𝐿
𝐹𝐹
 ,
𝐴𝐴 and 𝐴𝐴𝐹𝐹𝛼𝛼−𝐿𝐿
𝐹𝐹
are constants whose values depend on the element and the phase in consideration, 𝐴𝐴 𝐴𝐴 and 𝐴𝐴𝑀𝑀𝑀𝑀𝑀𝑀
𝐴𝐴𝐹𝐹𝐿𝐿𝐹𝐹𝐹𝐹 𝐿𝐿
are
mole fractions of FeO and MgO in the melt and ψ  is the molar proportion of element occupying the octahe-
ol

dral site in olivine. Values of constants are taken from Beattie et al. (1991) 𝐴𝐴 as 𝐴𝐴𝑀𝑀𝑀𝑀𝑂𝑂𝑂𝑂−𝐿𝐿
   = 0.290,
𝐴𝐴 𝑂𝑂𝑂𝑂−𝐿𝐿
𝐴𝐴𝑀𝑀𝑀𝑀    = −0.049,
𝐴𝐴 𝐴𝐴𝐹𝐹 𝐹𝐹    = 0.279,
𝑂𝑂𝑂𝑂−𝐿𝐿
𝐴𝐴 and 𝐴𝐴𝐹𝐹 𝐹𝐹    = 0.031 and the molar fraction is calculated using the bulk oxide composition of the
𝑂𝑂𝑂𝑂−𝐿𝐿

L chondrite (Jarosewich, 1990) which comes𝐴𝐴to be 𝐴𝐴𝐹𝐹𝐿𝐿𝐹𝐹𝐹𝐹   = 0.172 𝐴𝐴 and 𝐴𝐴𝑀𝑀𝑀𝑀𝑀𝑀


𝐿𝐿
   = 0.346. Feeding these values in
Equation 4 gives D = 5.063 and D = 1.262. We estimated the amount of Mn that can be incorporated in olivine
present in the L chondrites as: D× mean abundance of Mn in L chondrite (∼2,570 ppm) = ∼3,244 ppm which is
similar to the observed value in our sample, whereas the requisite concentration for vesicle formation given by
Equation 1 is ∼78 ppm.

Similarly, Mn partition coefficient for orthopyroxene and melt can be calculated using Equation 4 with constant
values
𝐴𝐴 of 𝐴𝐴𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
𝑀𝑀𝑀𝑀
  = 0.352,
𝐴𝐴 𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
𝐴𝐴𝑀𝑀𝑀𝑀   = −0.025,
𝐴𝐴 𝐴𝐴𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
𝐹𝐹 𝐹𝐹
  = 0.286,
𝐴𝐴 𝐴𝐴𝐹𝐹𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿
𝐹𝐹
  = 0.059 (Beattie et al., 1991) which gives
the value
𝐴𝐴 of 𝐴𝐴𝑀𝑀𝑀𝑀   = ∼1.747. This indicates that ∼4,500 ppm Mn can get partitioned into orthopyroxene in the L
𝑂𝑂𝑂𝑂𝑂𝑂−𝐿𝐿

chondrites, whereas ∼76 ppm of Mn is needed for the vesicle formation in the Kamargaon chondrite as estimated
using Equation 1. Therefore, evaporation of MVEs can be a viable mechanism for the formation of observed
vesicles in the Kamargaon meteorite.

3. Vaporization of superheated olivine and pyroxene: Laboratory experiments on the vaporization of Mg-rich olivine
(Fo95) suggest the fractional vaporization of Fe relative to Mg with the volatility sequence Fe > Si > O2 > Mg
(Costa et al., 2017). This result indicates that the total vapor pressure above the vaporizing olivine will mostly
consist of partial pressure of Fe. The reported vapor pressure of Fe over olivine is ∼9 Pa (Costa et al., 2017). The
corresponding concentration of Fe in vapor phase can be calculated using the following formula: (vapor pressure/
atmospheric pressure) × 10 6, which gives the value ∼90 ppm. This estimation indicates that the forsteritic olivine

TIWARI ET AL. 11 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

can liberate up to ∼90 ppm of Fe upon vaporization. Furthermore, the required Fe concentration for the vesicle
formation in the Kamargaon meteorite is estimated using Equation 1 and is ∼3 ppm and ∼80 ppm for 6.4 and
150 km large parent bodies, respectively.

The expected products of congruent vaporization of pyroxene can be estimated by the following equation
(Cremonese et al., 2005):

(5)
(𝐶𝐶𝐶𝐶1−𝑥𝑥−𝑦𝑦 𝑀𝑀𝑀𝑀𝑥𝑥 𝐹𝐹 𝐹𝐹𝑦𝑦 ) (𝑀𝑀𝑀𝑀1−𝑦𝑦 𝐹𝐹 𝐹𝐹𝑦𝑦 ) 𝑆𝑆𝑆𝑆2 𝑂𝑂6 ⇔(1 − 𝑥𝑥 − 𝑦𝑦)𝐶𝐶𝐶𝐶 + (1 + 𝑥𝑥 − 𝑦𝑦)𝑀𝑀𝑀𝑀 + 2𝑦𝑦𝑦𝑦𝑦𝑦 + 2𝑆𝑆𝑆𝑆 + 3𝑂𝑂2

where x and y are molar proportions of enstatite and ferrosilite. Taking the composition of low-Ca pyroxene in the
Kamargaon L6 chondrite as En80Fs20, the vaporization products will be:

 (𝐶𝐶𝐶𝐶0.02 𝑀𝑀𝑀𝑀0.78 𝐹𝐹 𝐹𝐹0.2 ) (𝑀𝑀𝑀𝑀0.8 𝐹𝐹 𝐹𝐹0.2 ) 𝑆𝑆𝑆𝑆2 𝑂𝑂6 ⇔0.02𝐶𝐶𝐶𝐶 + 1.58𝑀𝑀𝑀𝑀 + 0.4𝐹𝐹 𝐹𝐹 + 2𝑆𝑆𝑆𝑆 + 3𝑂𝑂2

Fe gas constitutes ∼6% of all the vapor products and being the most volatile of all the products will play the most
important role in the vesicle formation. Total equilibrium vapor pressure above vaporizing pyroxene at 1,800K is
∼536 Pa (Perez-Becker & Chiang, 2013). The partial pressure of Fe will be ∼32 Pa (0.06 × 536) which converts
to a concentration of ∼317 ppm of Fe. Thus, vaporization of pyroxene (En80Fs20) can liberate up to ∼317 ppm
of Fe which is much higher than ∼78 ppm, the concentration required to form the vesicles as estimated by Equa-
tion 1. Therefore, vaporization of the olivine and pyroxene could likely be the reason for the vesicle formation
in the Kamargaon meteorite. These estimated values of required and available abundance of different elements
suggest that all of the three possible mechanisms or any combination of them could have operated simultaneously
to result in the formation of observed vesicles in the Kamargaon L6 chondrite.

4.1.2.3.  Chronology of the Composite Textures

The composite grains have a vesicular core and dissociated rim. These two textures may have formed either from
a single shock event or from the different shock events. The shock event which resulted in the formation of the
dissociated olivine would be a high temperature-pressure event that probably formed the high-pressure polymorphs
(now back-transformed ringwoodite/wadsleyite, amorphous plagioclase, and dissociated olivine) and the SV in the
Kamargaon L6 chondrite. The vesicular grains require local melting due to localized temperature spike during the
shock event. As the vesicular grains are entrained in the SV (Figure 2d), which was molten during the  shock event,
an efficient localized temperature spike would not have been possible. Therefore, we suggest that these two textures
were formed as a result of different shock events where vesicular grains were formed before  the dissociated olivine.

4.2.  Estimation of Shock Conditions

Pressure-temperature conditions: Based on the textures and phases existing in our sample, the average shock
pressure and temperature conditions experienced by the Kamargaon L6 chondrite are estimated to be 24–25 GPa
and ∼2,310–2,633K (Text S1 in Supporting Information S1), respectively. Additionally, the calculated values of
post-shock and shock temperature are ∼1,226 and ∼1,386K, respectively (Text S2 in Supporting Information S1).

Thermal modeling: The crystallization of the melt starts at the contact of the SV and the host rock due to the
loss of heat via conduction from hot melt to cold host rock (Sharp & DeCarli, 2006). The crystallization then
proceeds toward the center of the SV, and the crystallization is completed when the temperature at the center of
the SV drops below the solidus temperature. The temperature (T)-time (t) history of the SV is estimated using a
one-dimensional non-steady heat conduction model and is described as:

(6)
𝜕𝜕T∕𝜕𝜕t = 𝛼𝛼(𝜕𝜕 2 T∕𝜕𝜕x2 ),


( )
𝛼𝛼 = 𝜅𝜅∕𝜌𝜌𝜌𝜌p

where T is the temperature (K), t is time (s), α is thermal diffusivity (m 2/s), κ is thermal conductivity (J
s −1m −1 K −1), ρ is density (kg m −3), and Cp is specific heat (JK −1 kg −1). Equation 6 can be discretized for mathe-
matical operations and modeling using the finite difference method. It will yield the following equation:
( 𝑎𝑎
𝑇𝑇𝑏𝑏+1 − 2𝑇𝑇𝑏𝑏𝑎𝑎 + 𝑇𝑇𝑏𝑏−1
𝑎𝑎 )
(7) 𝑇𝑇𝑏𝑏𝑎𝑎+1 = 𝑇𝑇𝑏𝑏𝑎𝑎 + 𝛼𝛼∆𝑡𝑡
∆𝑥𝑥2

TIWARI ET AL. 12 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Figure 5.  Calculated cooling history of the shock vein (SV) at its center and near vein edge, based on one-dimensional heat conductivity model. (a) The initial
temperature was 2,310K, corresponds to liquidus temperature of the Allende CV3 chondrite at 25 GPa. The shock vein reaches the temperature of 1,226K in ∼2 s.
Inset shows the quenching time along the vein center and vein edge calculated as time required for the temperature of shock vein to drop below the solidus temperature
(2,243K) for the Allende carbonaceous chondrite. The quenching time for vein center is ∼20 and 5 ms for vein edge. (b) The initial temperature was 2,633K,
corresponds to the liquidus temperature of KLB1 peridotite at 25 GPa. Inset shows the quenching time of 50 and 12 ms along the vein center and vein edge.

where a and b are integers, a is the known value of temperature at the current time step, and a+1 is the unknown
temperature value at the future point of time whereas b denotes location points. The thermal diffusivity in this
model is taken as 10 −6 m 2/s (Ohtani et al., 2004) whereas the starting temperatures for the model are taken as
2,310K (based on Allende data) (Figure  5a) and 2,633K (based on KLB-1 peridotite data) (Figure  5b). The
quenching time for both the data is calculated at the center of the vein and near the vein edge (Figures 5a and 5b,
inset). The post-shock temperature of 1,226K is taken as the final temperature for the model. The high-pressure
pulse duration is constrained by the quenching time as the crystallization of high-pressure minerals from liquid
requires that the high-pressure pulse duration must persist for at least when quenching lasts. The estimated quench
time at the center of the vein is ∼50 ms based on the KLB1 peridotite data (Figure 5b, inset) and ∼20 ms based on
the Allende chondrite data (Figure 5a, inset). The actual quench time for the Kamargaon L6 chondrite should lie
in between ∼20 and ∼50 ms but should be closer to ∼50 ms. According to the model, it takes the SV ∼2 s to cool
down to 1,226K. Most of the high-pressure minerals are stable at low pressure if the temperature is below 1,200K
or decompose sluggishly (Hu & Sharp, 2017). Thus, for the ringwoodite and
wadsleyite to back transform to olivine, the high-pressure pulse duration must
have been less than ∼2 s. This estimation of pulse duration is similar to previ-
ous studies (Miyahara et al., 2017; Ohtani et al., 2004; Xie et al., 2006). To
further investigate the ringwoodite/wadsleyite to olivine back-transformation
mechanism, we deciphered the T-t history of olivine grains by Equation  8
which gives the variation of T with t at a distance (x) from the vein center:
{ √ √ }
(8)
𝑇 = 0.5(𝑇𝑚 − 𝑇𝑠 ) erf [(𝑤 − 𝑥)∕(2 𝛼𝑡)] + erf [(𝑤 + 𝑥)∕ (2 𝛼𝑡)] +𝑇𝑠

where Tm, Ts, 2w, and α are melt temperature, shock temperature, vein width,
and thermal diffusivity, respectively. We further calculated the T-t history of
olivine grains present near the vein wall (w ≈ x) while adopting the values of
Tm = ∼2,471K (average of 2,310 and 2,633K), Ts = 1,386K, and w = 0.0008 m
which gives the maximum temperature (Tmax) of ∼1,929K attained by the
olivine grains (Figure  6). It should be noted that we do not observe any
evidence of melting in the segmented and host rock olivines. Therefore, the
maximum temperature attained by these olivines cannot be more than the
solidus temperature of olivine (Fo74–76) present in the Kamargaon L6 chon-
Figure 6.  Temperature-time history of the olivines present near or along the
drite, that is, ∼1,903K (Bowen & Schairer, 1935). This condition can only be
shock vein (SV) edge of the Kamargaon chondrite.

TIWARI ET AL. 13 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

satisfied if the Tm ≤ ∼2,420K at which Tmax ≤ ∼1,903K (Figure 6). Furthermore, the ringwoodite has been exper-
imentally shown to have a thermal stability limit of 1,273K (Wang et al., 1997) and above which if decompression
occurs, it back-transforms to olivine. The T-t history for Tm = ∼2,420K (Figure 6) shows that the temperature of
olivine grains reaches 1,684K after 2 s (estimated pulse duration) which is far above the stability limit of 1,273K
and thus the pressure release at 2 s will cause instantaneous back-transformation of ringwoodite to olivine.

Impact velocity and parent body size: Particle velocity (Up), shock wave velocity (Us), and impact velocity (Ui)
are estimated to be ∼1.15, ∼5.56, and ∼2.30 km/s, respectively, by using the Rankine-Hugoniot equations (Text
S3 in Supporting Information S1). In addition, we computed the size of the parent body to be > 6.4 km based on
estimated shock pressure and pulse duration (Text S3 in Supporting Information S1).

4.3.  Impact Events Recorded in Kamargaon Chondrite

Based on shock-induced textures and mode of occurrence of different phases, we attempt to establish the chron-
ological order of the formation of different shock-induced textures in the Kamargaon chondrite. As we have
already discussed (Section 4.1.2.1 and 4.1.2.3) that segmented olivine was formed first, followed by the forma-
tion of vesicular olivine and pyroxene grains, while the dissociated olivine and high-pressure minerals were the
last to formed. Coexistence of textures denoting distinct shock stages in a sample has previously been studied
and attributed to heterogeneous shock distribution in the target rock (e.g., Kato et al., 2017; Stöffler et al., 1991).
Friedrich et al. (2014) studied the NWA 7298 H chondrite breccia and observed that the presence of different
lithologies in their sample demarcated by distinct foliation patterns. These authors discounted the idea of heterog-
enous shock distribution and proposed that these distinct foliation domains were formed due to multiple impacts
and later brought together by material transport. Although heterogeneous shock distribution may have played
a role in the formation of the Kamargaon meteorite, it cannot describe the coexistence of different textures of
contrasting magnitudes within the same grain (e.g., Figure 2d). These observations suggest that these phases of
different magnitudes are formed by multiple impact events that occurred on the parent body of the Kamargaon
L6 chondrite. U-Th- 4He and K- 40Ar dating of the Kamargaon chondrite yielded radiometric ages of 170 ± 25 and
684 ± 93 Ma, respectively while Kamargaon L6 chondrite has cosmic ray exposure age of ∼7.0 ± 1.6 Ma (Ray
et al., 2017). These radiometric ages derived from U-Th- 4He and K- 40Ar are much younger as compared to the age
of the solar system (∼4.56 Ga) due to resetting of the K-Ar chronometer during a heating event caused by impact
and outgassing of He, respectively. Based on these observations, we envisage that the highest impact pulse prob-
ably corresponds to a ∼684 Ma event or similar catastrophic event that caused the major mineralogical-textural,
shock-induced manifestation in the Kamargaon L6 chondrite, for example, the formation of dissociated olivine
and high-pressure minerals. The vesicular grains and the segmented olivine formed probably prior to this main
reheating event. However, more impact age data from highly shocked L-chondrite data are necessary to better
constrain the impact histories in the asteroidal parent body. In addition, we suggest that this chronological order
of shock events may also have implications for the vesicle formation such that the high pre-shock temperature and
the shock-induced enhanced porosity (the result of the impact event that produced segmented olivine before the
vesicle formation) may have facilitated the vesiculation in the Kamargaon meteorite.

5. Conclusions
In this study, we investigated the textures, major element compositions, and microstructures present in the SV of
the Kamargaon L6 chondrite in detail. We also estimated the average shock pressure and temperature conditions,
shock pulse duration, impact velocity, and parent body size of the Kamargaon chondrite. Our main conclusions
are as follows:

1. T
 he host rock portion of the studied sample mainly consists of olivine (Fo73–74), low Ca pyroxene
(En77–80Fs19–22Wo1–2), high-Ca pyroxene (En45–46Fs9–10Wo44–46), albitic plagioclase (Ab62–70An18–23Or12–15),
Fe-Ni metal alloy (kamacite and taenite), troilite, with minor amounts of phosphate and chromite.
2. Plagioclase in the SV has been transformed into amorphized plagioclase. Many olivine and pyroxene grains
exhibit vesicular texture whereas olivines exclusively display dissociation and segmentation texture.
3. We report for the first time, the presence of vesicular olivines and pyroxenes in an ordinary chondrite.
These vesicular grains possibly formed due to local melting caused by localized temperature spike during a
shock event and subsequent degassing after decompression when the post-shock temperature was still high.

TIWARI ET AL. 14 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Segmentation in olivine grains is subgrain boundaries formed as a recovery process subsequent to localized
shear stress.
4. Three possible mechanisms have been discussed for the vesicle formation in the Kamargaon chondrite as
follows: (a) liberation of S2 vapor; (b) liberation of volatiles such as Na, Mn, etc.; (c) vaporization of olivine
and pyroxene. Estimated values of available and required abundance volatiles suggest that all three or any
combination of these mechanisms could be responsible for the vesicle formation in the Kamargaon L6
chondrite.
5. Based on the phase assemblages present in SV in the Kamargaon L6 chondrite and comparison with experi-
mental phase relation studies on KLB-1 peridotite and Allende carbonaceous chondrite, the shock pressure and
temperature conditions recorded in the SV of the Kamargaon chondrite are ∼24–25 GPa and ∼2,310–2,633K,
respectively.
6. Based on shock pressure estimation, calculated impact velocity for the Kamargaon meteorite is ∼2.3 km/s.
The calculated crystallization time at the center of the SV with a thickness of 1.6 mm is ∼50 ms. The shock
pulse duration required for back transformation of high-pressure polymorphs is estimated to be ∼2 s which
implies that the parent body of Kamargaon L6 chondrite is ∼6.4 km across in size at least.

Data Availability Statement


All the data used in this study, including all the figures (Figures 1–6) and tables (Tables 1–3) in this manuscript,
and Table S1, Table S2, and Figure S1 in Supporting Information S1 can be accessed through the Figshare archive
(Tiwari et al., 2022).

Acknowledgments References
S.G. acknowledges the financial
support by the Council of Scientific Agee, C. B., Li, J., Shannon, M. C., & Circone, S. (1995). Pressure temperature phase diagram for the Allende meteorite. Journal of Geophysical
and Industrial Research (CSIR) Grant Research: Solid Earth, 100(B9), 17725–17740. https://doi.org/10.1029/95jb00049
24(0362)/20/EMR-II. K.T. acknowledges Baer, M. R. (2000). Computational modeling of heterogeneous materials at the mesoscale. In M. D. Furnish, L. C. Chabildas, & R. S. Hixon
the support for his Ph.D. fellowship (Eds.), CP505, shock compression of condensed matter-1999 (pp. 27–33). AIP Conference Proceedings. https://doi.org/10.1063/1.1303415
from the University Grants Commis- Beattie, P., Ford, C., & Russell, D. (1991). Partition coefficients for olivine-melt and orthopyroxene-melt systems. Contributions to Mineralogy
sion, Government of India and teaching and Petrology, 109(2), 212–224. https://doi.org/10.1007/BF00306480
assistantship from the Indian Institute of Benedix, G. K., Ketcham, R. A., Wilson, L., McCoy, T. J., Bogard, D. D., Garrison, D. H., et al. (2008). The formation and chronology of the
Technology (IIT), Kharagpur. K.T. thanks PAT 91501 impact-melt L chondrite with vesicle–metal–sulfide assemblages. Geochimica et Cosmochimica Acta, 72(9), 2417–2428. https://
Ashvini K. Tiwari for stimulating discus- doi.org/10.1016/j.gca.2008.02.010
sions regarding the thermal modeling. Bischoff, A., & Stöffler, D. (1992). Shock metamorphism as a fundamental process in the evolution of planetary bodies: Information from mete-
The authors thank G. D. Mukherjee for orites. European Journal of Mineralogy, 4, 707–755. https://doi.org/10.1127/ejm/4/4/0707
providing a laser micro-Raman facility Blundy, J. D., Falloon, T. J., Wood, B. J., & Dalton, J. A. (1995). Sodium partitioning between clinopyroxene and silicate melts. Journal of
at IISER Kolkata. The authors greatly Geophysical Research: Solid Earth, 100(B8), 15501–15515. https://doi.org/10.1029/95JB00954
appreciate the careful review by two Borisov, A., Pack, A., Kropf, A., & Palme, H. (2008). Partitioning of Na between olivine and melt: An experimental study with application to the
anonymous reviewers and the editorial formation of meteoritic Na2O-rich chondrule glass and refractory forsterite grains. Geochimica et Cosmochimica Acta, 72(22), 5558–5573.
handling by Bradley Thomson and https://doi.org/10.1016/j.gca.2008.08.009
Mariek Schmidt. Bowen, N. L., & Schairer, J. F. (1935). The system MgO-FeO-SiO2. American Journal of Science, 5(170), 151–217. https://doi.org/10.2475/ajs.
s5-29.170.151
Chen, M., Chen, J., Xie, X., & Xu, J. (2007). A microstructural investigation of natural lamellar ringwoodite in olivine of the shocked Sixiangkou
chondrite. Earth and Planetary Science Letters, 264(1–2), 277–283. https://doi.org/10.1016/j.epsl.2007.09.037
Chen, M., Goresy, A. E., & Gillet, P. (2004). Ringwoodite lamellae in olivine: Clues to olivine–ringwoodite phase transition mechanisms in
shocked meteorites and subducting slabs. Proceedings of the National Academy of Sciences of the United States of America, 101(42), 15033–
15037. https://doi.org/10.1073/pnas.0405048101
Chen, M., Li, H., Goresy, A. E., Liu, J., & Xie, X. (2006). Fracture-related intracrystalline transformation of olivine to ringwoodite in the shocked
Sixiangkou meteorite. Meteoritics & Planetary Science, 41(5), 731–737. https://doi.org/10.1111/j.1945-5100.2006.tb00988.x
Chen, M., Sharp, T. G., El Goresy, A., Wopenka, B., & Xie, X. (1996). The majorite-pyrope + magnesiowüstite assemblage: Constraints on the
history of shock veins in chondrites. Science, 271(5255), 1570–1573. https://doi.org/10.1126/science.271.5255.1570
Chen, M., Shu, J., Mao, H. K., Xie, X., & Hemley, R. J. (2003). Natural occurrence and synthesis of two new postspinel polymorphs of chro-
mite. Proceedings of the National Academy of Sciences of the United States of America, 100(25), 14651–14654. https://doi.org/10.1073/
pnas.2136599100
Cordier, P., Amodeo, J., & Carrez, P. (2012). Modelling the rheology of MgO under Earth’s mantle pressure, temperature and strain rates. Nature,
481(7380), 177–180. https://doi.org/10.1038/nature10687
Costa, G. C., Jacobson, N. S., & Fegley, B., Jr. (2017). Vaporization and thermodynamics of forsterite-rich olivine and some implications for
silicate atmospheres of hot rocky exoplanets. Icarus, 289, 42–55. https://doi.org/10.1016/j.icarus.2017.02.006
Cremonese, G., Bruno, M., Mangano, V., Marchi, S., & Milillo, A. (2005). Release of neutral sodium atoms from the surface of Mercury induced
by meteoroid impacts. Icarus, 177(1), 122–128. https://doi.org/10.1016/j.icarus.2005.03.022
Dann, J. C. (2001). Vesicular komatiites, 3.5-Ga Komati Formation, Barberton Greenstone Belt, South Africa: Inflation of submarine lavas and
origin of spinifex zones. Bulletin of Volcanology, 63(7), 462–481. https://doi.org/10.1007/s004450100164
DeCarli, P. S. (1979). Nucleation and growth of diamond in shock wave experiments. In K. D. Timmerhaus, & M. S. Barber (Eds.), High pressure
science and technology, (Vol. 1, pp. 940–943). Plenum. https://doi.org/10.1007/978-1-4684-7470-1_124

TIWARI ET AL. 15 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Dreibus, G., Palme, H., Spettel, B., Zipfel, J., & Wänke, H. (1995). Sulfur and selenium in chondritic meteorites. Meteoritics, 30(4), 439–445.
https://doi.org/10.1111/j.1945-5100.1995.tb01150.x
El Goresy, A., Chen, M., Gillet, P., & Dubrovinsky, L. (2005). Two distinct olivine-ringwoodite phase transition mechanisms in shocked
L-chondrites: Genetic implications. Meteoritics & Planetary Science, 40, 5010.
Feng, L., Lin, Y., Hu, S., Xu, L., & Miao, B. (2011). Estimating compositions of natural ringwoodite in the heavily shocked Grove Mountains
052049 meteorite from Raman spectra. American Mineralogist, 96(10), 1480–1489. https://doi.org/10.2138/am.2011.3679
Friedrich, J. M., Weisberg, M. K., & Rivers, M. L. (2014). Multiple impact events recorded in the NWA 7298 H chondrite breccia and the dynam-
ical evolution of an ordinary chondrite asteroid. Earth and Planetary Science Letters, 394, 13–19. https://doi.org/10.1016/j.epsl.2014.03.016
Fukimoto, K., Miyahara, M., Sakai, T., Ohfuji, H., Tomioka, N., Kodama, Y., et al. (2020). Back-transformation mechanisms of ringwoodite and
majorite in an ordinary chondrite. Meteoritics & Planetary Science, 55(8). https://doi.org/10.1111/maps.13543
Ghosh, S., Tiwari, K., Miyahara, M., Rohrbach, A., Vollmer, C., Stagno, V., et  al. (2021). Natural Fe-bearing aluminous bridgmanite in the
Katol L6 chondrite. Proceedings of the National Academy of Sciences of the United States of America, 118(40), e2108736118. https://doi.
org/10.1073/pnas.2108736118
Gillet, P., El Goresy, A., Beck, P., & Chen, M. (2007). High-pressure mineral assemblages in shocked meteorites and shocked terrestrial rocks:
Mechanisms of phase transformations and constraints to pressure and temperature histories. Geological Society of America, 421, 57. https://
doi.org/10.1130/2007.2421(05
Goswami, T. K., Ray, D., Sarmah, R. K., Goswami, U., Bhattacharyya, P., Majumdar, D., et al. (2016). Meteorite fall at Komargaon, Assam, India.
Current Science, 110(10), 1894.
He, Q., Xiao, L., Balta, J. B., Baziotis, I. P., Hsu, W., & Guan, Y. (2015). Petrography and geochemistry of the enriched basaltic shergottite North-
west Africa 2975. Meteoritics & Planetary Science, 50(12), 2024–2044. https://doi.org/10.1111/maps.12571
Hu, J., & Sharp, T. G. (2017). Back-transformation of high-pressure minerals in shocked chondrites: Low-pressure mineral evidence for strong
shock. Geochimica et Cosmochimica Acta, 215, 277–294. https://doi.org/10.1016/j.gca.2017.07.018
Jarosewich, E. (1990). Chemical analyses of meteorites: A compilation of stony and iron meteorite analyses. Meteoritics, 25(4), 323–337. https://
doi.org/10.1111/j.1945-5100.1990.tb00717.x
Kato, Y., Sekine, T., Kayama, M., Miyahara, M., & Yamaguchi, A. (2017). High-pressure polymorphs in Yamato-790729 L6 chondrite and their
significance for collisional conditions. Meteoritics & Planetary Science, 52(12), 2570–2585. https://doi.org/10.1111/maps.12957
Katsura, T., & Ito, E. (1989). The system Mg2SiO4-Fe2SiO4 at high pressures and temperatures: Precise determination of stabilities of olivine,
modified spinel, and spinel. Journal of Geophysical Research: Solid Earth, 94(B11), 15663–15670. https://doi.org/10.1029/jb094ib11p15663
Kenkmann, T., Hornemann, U., & Stöffler, D. (2000). Experimental generation of shock-induced pseudotachylites along lithological interfaces.
Meteoritics & Planetary Science, 35(6), 1275–1290. https://doi.org/10.1111/j.1945-5100.2000.tb01516.x
Kerschhofer, L., Dupas, C., Liu, M., Sharp, T. G., Durham, W. B., & Rubie, D. C. (1998). Polymorphic transformations between olivine, wadsley-
ite and ringwoodite: Mechanisms of intracrystalline nucleation and the role of elastic strain. Mineralogical Magazine, 62(5), 617–638. https://
doi.org/10.1180/002646198548016
Kerschhofer, L., Rubie, D. C., Sharp, T. G., McConnell, J. D. C., & Dupas-Bruzek, C. (2000). Kinetics of intracrystalline olivine-ringwoodite
transformation. Physics of the Earth and Planetary Interiors, 121(1–2), 59–76. https://doi.org/10.1016/S0031-9201(00)00160-6
Kubo, T., Kamura, K., Imamura, M., Tange, Y., Higo, Y., & Miyahara, M. (2022). Back-transformation processes in high-pressure minerals:
Implications for planetary collisions and diamond transportation from the deep Earth. Progress in Earth and Planetary Science, 9(1), 1–16.
https://doi.org/10.1186/s40645-022-00480-9
Kuebler, K. E., McSween, H. Y., Carlson, W. D., & Hirsch, D. (1999). Sizes and masses of chondrules and metal–troilite grains in ordinary chon-
drites: Possible implications for nebular sorting. Icarus, 141(1), 96–106. https://doi.org/10.1006/icar.1999.616
Langenhorst, F. (2002). Shock metamorphism of some minerals: Basic introduction and microstructural observations. Bulletin of the Czech
Geological Survey, 77(4), 265–282.
Lunning, N. G., Corrigan, C. M., McSween, H. Y., Jr., Tenner, T. J., Kita, N. T., & Bodnar, R. J. (2016). CV and CM chondrite impact melts.
Geochimica et Cosmochimica Acta, 189, 338–358. https://doi.org/10.1016/j.gca.2016.05.038
McCoy, T. J., Ketcham, R. A., Wilson, L., Benedix, G. K., Wadhwa, M., & Davis, A. M. (2006). Formation of vesicles in asteroidal basaltic
meteorites. Earth and Planetary Science Letters, 246(1–2), 102–108. https://doi.org/10.1016/j.epsl.2006.04.002
McDade, P., Blundy, J. D., & Wood, B. J. (2003). Trace element partitioning on the Tinaquillo Lherzolite solidus at 1.5 GPa. Physics of the Earth
and Planetary Interiors, 139(1–2), 129–147. https://doi.org/10.1016/S0031-9201(03)00149-3
Ming, L. C., Kim, Y. H., Manghnani, M. H., Usha-Devi, S., Ito, E., & Xie, H. S. (1991). Back transformation and oxidation of (Mg, Fe)2SiO4
spinels at high temperatures. Physics and Chemistry of Minerals, 18(3), 171–179. https://doi.org/10.1007/BF00234000
Miyahara, M., El Goresy, A., Ohtani, E., Nagase, T., Nishijima, M., Vashaei, Z., et al. (2008). Evidence for fractional crystallization of wadsley-
ite and ringwoodite from olivine melts in chondrules entrained in shock-melt veins. Proceedings of the National Academy of Sciences of the
United States of America, 105(25), 8542–8547. https://doi.org/10.1073/pnas.0801518105
Miyahara, M., Ohtani, E., El Goresy, A., Ozawa, S., & Gillet, P. (2016). Phase transition processes of olivine in the shocked Martian meteorite
Tissint: Clues to origin of ringwoodite-bridgmanite- and magnesiowüstite-bearing assemblages. Physics of the Earth and Planetary Interiors,
259, 18–28. https://doi.org/10.1016/j.pepi.2016.08.006
Miyahara, M., Ohtani, E., Nishijima, M., & El Goresy, A. (2019). Olivine melting at high pressure condition in the chassignite Northwest Africa
2737. Physics of the Earth and Planetary Interiors, 291, 1–11. https://doi.org/10.1016/j.pepi.2019.04.001
Miyahara, M., Ohtani, E., & Yamaguchi, A. (2017). Albite dissociation reaction in the Northwest Africa 8275 shocked LL chondrite and implica-
tions for its impact history. Geochimica et Cosmochimica Acta, 217, 320–333. https://doi.org/10.1016/j.gca.2017.08.034
Miyahara, M., Tomioka, N., & Bindi, L. (2021). Natural and experimental high-pressure, shock-produced terrestrial and extraterrestrial materials.
Progress in Earth and Planetary Science, 8(1), 59. https://doi.org/10.1186/s40645-021-00451-6
Nakamura, T., Noguchi, T., Tanaka, M., Zolensky, M. E., Kimura, M., Tsuchiyama, A., et al. (2011). Itokawa dust particles: A direct link between
S-type asteroids and ordinary chondrites. Science, 333(6046), 1113–1116. https://doi.org/10.1126/science.1207758
Nozaka, T., & Ito, Y. (2011). Cleavable olivine in serpentinite mylonites from the Oeyama ophiolite. Journal of Mineralogical and Petrological
Sciences, 106(1), 36–50. https://doi.org/10.2465/jmps.100408
Ohnishi, I., Tomeoka, K., & Ishizaki, N. (2007). Microinclusion-rich vesicular olivine in the Karoonda CK4 chondrite: Transmission electron
microscopy. Journal of Mineralogical and Petrological Sciences, 102(6), 346–351. https://doi.org/10.2465/jmps.070323
Ohtani, E., Kimura, Y., Kimura, M., Takata, T., Kondo, T., & Kubo, T. (2004). Formation of high-pressure minerals in shocked L6 chon-
drite Yamato 791384: Constraints on shock conditions and parent body size. Earth and Planetary Science Letters, 227(3–4), 505–515.
https://doi.org/10.1016/j.epsl.2004.08.018

TIWARI ET AL. 16 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Oppenheimer, C., Pyle, D. M., & Barclay, J. (Eds.), (2003). Volcanic degassing. Geological Society of London. https://doi.org/10.1016/
B0-08-043751-6/03020-6
Perez-Becker, D., & Chiang, E. (2013). Catastrophic evaporation of rocky planets. Monthly Notices of the Royal Astronomical Society, 433(3),
2294–2309. https://doi.org/10.1093/mnras/stt895
Petrova, E. V., & Grokhovsky, V. I. (2019). High pressure impacts on meteorites. Pure and Applied Chemistry, 91(11), 1857–1867. https://doi.
org/10.1515/pac-2018-1119
Ray, D., Mahajan, R. R., Shukla, A. D., Goswami, T. K., & Chakraborty, S. (2017). Petrography, classification, oxygen isotopes, noble gases,
and cosmogenic records of Kamargaon (L6) meteorite: The latest fall in India. Meteoritics & Planetary Science, 52(8), 1744–1753. https://
doi.org/10.1111/maps.12875
Riedel, M. R., & Karato, S. I. (1997). Grain-size evolution in subducted oceanic lithosphere associated with the olivine-spinel transformation and
its effects on rheology. Earth and Planetary Science Letters, 148(1–2), 27–43. https://doi.org/10.1016/S0012-821X(97)00016-2
Rubin, A. E. (2015). Maskelynite in asteroidal, lunar and planetary basaltic meteorites: An indicator of shock pressure during impact ejection
from their parent bodies. Icarus, 257, 221–229. https://doi.org/10.1016/j.icarus.2015.05.010
Rubin, A. E., & Moore, W. B. (2011). What’s up? Preservation of gravitational direction in the Larkman Nunatak 06299 LL impact melt breccia.
Meteoritics & Planetary Science, 46(5), 737–747. https://doi.org/10.1111/j.1945-5100.2011.01187.x
Schmitz, B., Farley, K. A., Goderis, S., Heck, P. R., Bergström, S. M., Boschi, S., et al. (2019). An extraterrestrial trigger for the mid-Ordovician
ice age: Dust from the breakup of the L-chondrite parent body. Science Advances, 5(9), eaax4184. https://doi.org/10.1126/sciadv.aax4184
Scott, E. R. D., & Krot, A. N. (2003). Chondrites and their components. Treatise on Geochemistry, 1, 711–772. https://doi.org/10.1016/
B0-08-043751-6/01145-2
Sharp, T. G., & DeCarli, P. S. (2006). Shock effects in meteorites. In D. S. Lauretta, & H. Y. McSween (Eds.), Meteorites and the early solar
system II (pp. 653–677). The University of Arizona Press.
Shaw, C. S., & Walton, E. (2013). Thermal modeling of shock melts in Martian meteorites: Implications for preserving Martian atmospheric
signatures and crystallization of high-pressure minerals from shock melts. Meteoritics & Planetary Science, 48(5), 758–770. https://doi.
org/10.1111/maps.12100
Simon, J., Cuzzi, J., McCain, K., Cato, M., Christoffersen, P., Fisher, K., et  al. (2018). Particle size distributions in chondritic meteorites:
Evidence for pre-planetesimal histories. Earth and Planetary Science Letters, 494, 69–82. https://doi.org/10.1016/j.epsl.2018.04.021
Spray, J. G. (1998). Localized shock-and friction-induced melting in response to hypervelocity impact. Geological Society, London, Special
Publications, 140(1), 195–204. https://doi.org/10.1144/GSL.SP.1998.140.01.14
Stöffler, D., Hamann, C., & Metzler, K. (2018). Shock metamorphism of planetary silicate rocks and sediments: Proposal for an updated classi-
fication system. Meteoritics & Planetary Science, 53(1), 5–49. https://doi.org/10.1111/maps
Stöffler, D., Keil, K., & Scott, E. R. D. (1991). Shock metamorphism of ordinary chondrites. Geochimica et Cosmochimica Acta, 55(12), 3845–
3867. https://doi.org/10.1016/0016-7037(91)90078-J
Tiwari, K., Ghosh, S., Miyahara, M., & Ray, D. (2021). Shock-induced incongruent melting of olivine in Kamargaon L6 chondrite. Geophysical
Research Letters, 48(12), e2021GL093592. https://doi.org/10.1029/2021gl093592
Tiwari, K., Ghosh, S., Miyahara, M., & Ray, D. (2022). Vesicular olivines and pyroxenes in shocked Kamargaon L6 Chondrite: Implications for
primary volatiles and its multiple impacts history. Figshare. Journal contribution, v9. https://doi.org/10.6084/m9.figshare.19179236
Tomeoka, K., Ohnishi, I., & Nakamura, N. (2001). Silicate darkening in the Kobe CK chondrite: Evidence for shock metamorphism at high
temperature. Meteoritics & Planetary Science, 36(11), 1535–1545. https://doi.org/10.1111/j.1945-5100.2001.tb01844.x
Tomioka, N., & Miyahara, M. (2017). High-pressure minerals in shocked meteorites. Meteoritics & Planetary Science, 52(9), 2017–2039. https://
doi.org/10.1111/maps.12902
Tschauner, O. (2019). High-pressure minerals. American Mineralogist, 104(12), 1701–1731. https://doi.org/10.2138/am-2019-6594
Van der Bogert, C. H., Schultz, P. H., & Spray, J. G. (2003). Impact-induced frictional melting in ordinary chondrites: A mechanism for deforma-
tion, darkening, and vein formation. Meteoritics & Planetary Science, 38(10), 1521–1531. https://doi.org/10.1111/j.1945-5100.2003.tb00255.x
Walton, E. L. (2013). Shock metamorphism of Elephant Moraine A79001: Implications for olivine–ringwoodite transformation and the
complex thermal history of heavily shocked Martian meteorites. Geochimica et Cosmochimica Acta, 107, 299–315. https://doi.org/10.1016/j.
gca.2012.12.021
Wang, Y., Martinez, I., Guyot, F., & Liebermann, R. C. (1997). The breakdown of olivine to perovskite and magnesiowüstite. Science, 275(5299),
510–513. https://doi.org/10.1126/science.275.5299.510
Wasson, J. T., & Kallemeyn, G. W. (1988). Compositions of chondrites. Philosophical Transactions of the Royal Society of London. Series A,
Mathematical and Physical Sciences, 325, 535–544. https://doi.org/10.1098/rsta.1988.0066
Weisberg, M. K., McCoy, T. J., & Krot, A. N. (2006). Systematic and evolution of meteorite classification. In D. S. Lauretta, & H. Y. McSween
(Eds.), Meteorites and the early solar system II (pp. 19–52). The University of Arizona Press.
Wilkening, L. L., & Anders, E. (1975). Some studies of an unusual eucrite: Ibitira. Geochimica et Cosmochimica Acta, 39(9), 1205–1210. https://
doi.org/10.1016/0016-7037(75)90127-1
Wilkison, S. L., & Robinson, M. S. (2000). Bulk density of ordinary chondrite meteorites and implications for asteroidal internal structure. Mete-
oritics & Planetary Science, 35(6), 1203–1213. https://doi.org/10.1111/j.1945-5100.2000.tb01509.x
Wilson, L., & Keil, K. (1997). The fate of pyroclasts produced in explosive eruptions on the asteroid 4 Vesta. Meteoritics & Planetary Science,
32(6), 813–823. https://doi.org/10.1111/j.1945-5100.1997.tb01572.x
Xie, Z., Sharp, T. G., & DeCarli, P. S. (2006). High-pressure phases in a shock-induced melt vein of the Tenham L6 chondrite: Constraints on
shock pressure and duration. Geochimica et Cosmochimica Acta, 70(2), 504–515. https://doi.org/10.1016/j.gca.2005.09.003
Zhang, J., & Herzberg, C. (1994). Melting experiments on anhydrous peridotite KLB-1 from 5.0 to 22.5 GPa. Journal of Geophysical Research,
99(B9), 17.729–17.742. https://doi.org/10.1029/94jb01406

References From the Supporting Information


Bottke, J. W. F., Nolan, M. C., Greenberg, R., & Kolvoord, R. A. (1994). Velocity distributions among colliding asteroids. Icarus, 107(2),
255–268. https://doi.org/10.1006/icar.1994.1021

TIWARI ET AL. 17 of 18
Journal of Geophysical Research: Planets 10.1029/2022JE007420

Fritz, J., Wünnemann, K., Greshake, A., Fernandes, V. A. S. M., Boettger, U., & Hornemann, U. (2011). Shock pressure calibration for lunar
plagioclase. Lunar and Planetary Science Conference, 42. Abstract #1196.
Ozawa, S., Miyahara, M., Ohtani, E., Koroleva, O. N., Ito, Y., Litasov, K. D., & Pokhilenko, N. P. (2014). Jadeite in Chelyabinsk meteorite and
the nature of an impact event on its parent body. Scientific Reports, 4(1), 5033. https://doi.org/10.1038/srep05033
Zhang, F., & Sekine, T. (2007). Impact-shock behavior of Mg and Ca-sulfates and their hydrates. Geochimica et Cosmochimica Acta, 71(16),
4125–4133. https://doi.org/10.1016/j.gca.2007.06.037

TIWARI ET AL. 18 of 18

You might also like