Vertical AMS Variation Within Basalt Ow Profiles From The Xitle Volcano (Mexico) As Indicator of Heterogeneous Strain in Lava Ows

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 21

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/290625362

Vertical AMS variation within basalt flow profiles from the Xitle volcano
(Mexico) as indicator of heterogeneous strain in lava flows

Article  in  Journal of Volcanology and Geothermal Research · January 2016


DOI: 10.1016/j.jvolgeores.2016.01.003

CITATIONS READS

5 268

6 authors, including:

Cecilia-Irene Caballero-Miranda Luis Alva Valdivia


Universidad Nacional Autónoma de México Universidad Nacional Autónoma de México
47 PUBLICATIONS   329 CITATIONS    155 PUBLICATIONS   921 CITATIONS   

SEE PROFILE SEE PROFILE

J. Urrutia Fucugauchi A. Kontny


Universidad Nacional Autónoma de México Karlsruhe Institute of Technology
562 PUBLICATIONS   19,256 CITATIONS    112 PUBLICATIONS   1,239 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Exp 364 View project

Influence of the Chicxulub impact and Deccan volcanism on climate, environment and biodiversity of the Cretaceous-Paleogene transition. PGC2018-093890-B-I00
(MCIU/AEI/FEDER, UE). View project

All content following this page was uploaded by A. Kontny on 16 July 2018.

The user has requested enhancement of the downloaded file.


Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Contents lists available at ScienceDirect

Journal of Volcanology and Geothermal Research

journal homepage: www.elsevier.com/locate/jvolgeores

Vertical AMS variation within basalt flow profiles from the Xitle volcano
(Mexico) as indicator of heterogeneous strain in lava flows
C.I. Caballero-Miranda a,⁎, L.M. Alva-Valdivia a, J.A. González-Rangel a, A. Gogitchaishvili b,
J. Urrutia-Fucugauchi a, A. Kontny c
a
Laboratorio de Paleomagnetismo, Instituto Geofísica, Universidad Nacional Autónoma de México, Ciudad Universitaria, 04510 México, D. F., Mexico
b
Laboratorio Interinstitucional de Magnetismo Natural, Instituto Geofísica, Universidad Nacional Autónoma de México, Morelia, Sede Michoacán, Mexico
c
Institute of Applied Geosciences, Karlsruhe Institute of Technology (KIT), Adenauerring 20a, D-76131 Karlsruhe, Germany

a r t i c l e i n f o a b s t r a c t

Article history: The within-flow vertical variation of anisotropy of the magnetic susceptibility (AMS) of three basaltic flow pro-
Received 10 September 2015 files from the Xitle volcano were investigated in relation to the lava flow-induced shear strain. Rock magnetic
Accepted 5 January 2016 properties and opaque microscopy studies have shown that the magnetic mineralogy is dominated by Ti-poor
Available online 15 January 2016
magnetite with subtle vertical variations in grain size distribution: PSD grains dominate in a thin bottommost
zone, and from base to top from PSD-MD to PSD-SD grains are found. The vertical variation of AMS principal di-
Keywords:
Anisotropy of magnetic susceptibility (AMS)
rection patterns permitted identification of two to three main lava zones, some subdivided into subzones. The
Heterogeneous strain models lower zone is very similar in all profiles with the magnetic foliation dipping toward the flow source, whereas
AMS lava flow zones the upper zone has magnetic foliation dipping toward the flow direction or alternates between dipping against
Magnetic fabric and toward the flow direction. The K1 (maximum AMS axis) directions tend to be mostly parallel to the flow di-
Xitle lava flows rection in both zones. The middle zone shows AMS axes diverging among profiles. We present heterogeneous
Trans-Mexican volcanic belt strain ellipse distribution models for different flow velocities assuming similar viscosity to explain the AMS direc-
tions and related parameters of each zone. Irregular vertical foliations and transverse to flow lineation of a few
samples at the bottommost and topmost part of profiles suggest SD inverse fabric, levels of intense friction, or
degassing effects in AMS orientations.
© 2016 Elsevier B.V. All rights reserved.

1. Introduction relationships between the magnetic fabric determined from AMS


measurements and the flow direction and/or other geological features
Since Graham (1954) proposed that magnetic anisotropy can be (e.g., flow structures, type of facies, tectonic tilting, position in the flow
used to characterize petrofabrics, many studies have investigated the unit) even if each kind of rock corresponds to a different kind of flow
anisotropy of magnetic susceptibility (AMS) for rocks in a wide range dynamics.
of geological and geophysical applications in order to assess their Such relationships are mainly found between the dip of the magnet-
primary or secondary origin (e.g., MacDonald and Ellwood, 1987; ic foliation plane and the plunge of magnetic lineation relative to the
Rochette et al., 1992; Tarling and Hrouda, 1993; Kodama, 1995; flow plane and flow direction. The magnetic foliation and lineation
Martín-Hernández et al., 2004). Magnetic fabric studies of volcanic may be graphically analyzed in stereographic projections where low-
rocks have been used to investigate flow direction and induced flow angle imbricate patterns in relation to the flow plane (nearly the
strain from the magnetic anisotropy relating AMS directions with flow paleo-horizontal or paleoslope plane) are frequently observed:
and stress directions and magnitude of AMS parameters with strain in- minimum K3 axes (magnetic foliation plane poles) show near vertical
tensity (e.g., Ort et al., 2015). Studies of basaltic (e.g., Kolofíková, 1976; positions and maximum K1 axes near horizontal positions (parallel, or-
Ellwood, 1978; Cañón-Tapia and Walker, 1998; Henry et al., 2003; thogonal, in favor or against flow direction). Sometimes when the mag-
Cañón-Tapia, 2004; Bascou et al., 2005; Boiron et al., 2013) and pyro- netic foliation plane is not well defined, some odd relationships have
clastic flows (e.g., Ellwood, 1982; Incoronato et al., 1983; Wolff et al., been reported, such as the distribution of K3 (and intermediate K2)
1989; MacDonald and Palmer, 1990; Palmer et al., 1996; Cagnoli and axes along nearly vertical planes. This arrangement has been rarely re-
Tarling, 1997; Le Pennec et al., 1998; Alva-Valdivia et al., 2005; ported in basaltic lava flows, and when it has, it has been attributed to
Caballero-Miranda et al., 2009; Gountié Dedzo et al., 2011) show clear diverse factors: to inverse fabric because of the magnetic mineralogy
(e.g., Cañón-Tapia et al., 1995), or related to a steep paleoslope (e.g.,
⁎ Corresponding author. Herrero-Bervera et al., 2002). In other types of flows, such as gravity-
E-mail address: cecilia@geofisica.unam.mx (C.I. Caballero-Miranda). driven mass-flows, this pattern has been related to the flow dynamics

http://dx.doi.org/10.1016/j.jvolgeores.2016.01.003
0377-0273/© 2016 Elsevier B.V. All rights reserved.
10 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

such as a traction fabric of a turbidite (Ellwood, 1980), or a rebounding of 40 m), and the Pedregal de San Angel batch (BPSA), unit VI (10 m av-
pyroclastic turbulent flow (Caballero-Miranda et al., 2009). In all cases, erage thickness). Unit V is considered the paroxysmal phase of the Xitle
the vertical plane of the described pattern appears to record the flow eruption (because it has the maximum lava volume) and the one with
direction. the lowest viscosity, since it has the greatest maximum extent, at
The relationship between the AMS fabric elements and the lava flow around 12.5 km from the vent (Delgado et al., 1998). Unit V and VI con-
plane and flow direction may be attributed to the flow dynamics (e.g., sist of several minor lava flows with individual thicknesses from 0.2 to
shear stress, flow viscosity, cooling rate) and/or to magnetic mineralogy 15 m.
(e.g., grain size, composition, oxidation degree). Both of these variables Lava flows are calc-alkaline olivine basalt with outcrop surfaces
may change not only in different parts of the same flow but also along a dominated by ropey pahoehoe structures, and internal complex lava
single vertical flow profile. tube structures that extend for more than 13 km (Espinasa, 1992;
Holocene–Pleistocene lava flows preserve indicators of flow direction, Martin del Pozzo et al., 1997). Lava rises, pressure ridges, tumuli, and
emplacement mechanism, and flow structure (Walker, 1991; Hon et al., hornitos are frequently identified in several sites, mainly around
1994; Peterson et al., 1994), useful for determining the relationships be- Cuicuilco pyramids for the case of unit V, and for the case of unit VI
tween these and the magnetic fabric (Bascou et al., 2005). Lava flows these features are present around the CU profile.
show strong viscosity variations within a flow profile that cause complex Previous magnetic fabric studies on the Xitle lavas have been carried
emplacement scenarios and multiple rheological layering, instead of rela- out on five irregularly sampled profiles, all probably from the same indi-
tively homogeneous flow emplacement modes (Loock et al., 2008 and ref- vidual topmost flow unit of the BPSA member and others apparently in
erences therein). Observations of the diverse elements of magnetic fabric the topmost flow unit from the BCU member (Cañón-Tapia et al., 1995).
along vertical profiles of Holocene–Pliocene, or occasionally some older Another magnetic fabric study is reported from the BCU member,
lava flows, are ideal to evaluate the relationships between magnetic fab- supporting the validity of the results derived from an extensive study
rics and flow mechanisms (Cañón-Tapia et al., 1995, 1996, 1997; Bascou aimed at evaluating the reliability of the paleomagnetic record of the
et al., 2005; Loock et al., 2008; Audunsson et al., 1992), since they preserve flows (Urrutia-Fucugauchi, 1996).
almost intact evidence of both. All the previous magnetic fabric studies report the K1 (maximum)
The purpose of this work is to study in detail vertical variations of the AMS axes parallel to the flow direction and in most cases plunging
AMS elements along three vertical lava flow profiles from the Xitle volca- upflow (Alva-Valdivia, 2005). Those fabrics are observed most frequent-
no, in central Mexico (1600 years old); each profile is located at a slightly ly in the lower part of flow profiles (Cañón-Tapia et al., 1995). Cañón-
different facies position. We focus first on the AMS direction variation and Tapia et al. (1995) showed clear vertical variation of the magnetic fabric
complement our analysis with their corresponding AMS parameter varia- directions and AMS parameters (anisotropy degree), and related the
tion. We use not only the principal component statistic but also a point AMS ellipsoid to the internal deformation of the lava flow, and the an-
density distribution (e.g., Cañón-Tapia, 2004, Caballero-Miranda et al., isotropy degree to the lava viscosity.
2009) in order to identify slightly different statistical populations. Our The sampling objective of this study was to obtain characteristic
purpose is to characterize the similarities and differences of AMS changes AMS and related magnetic mineralogy data on single lava flows. We
along these three profiles to test if magnetic fabrics can be related to flow focus on fresh and complete (not eroded) lava flows without outcrop
dynamics, particularly relative to vertical changes in flow velocity and rel- evidence of alteration; flows from similar facies, composition, and vis-
ative differences of maximum flow velocity between profiles. Our goal is cosity. Units V and VI were chosen for sampling because of their wide-
to provide some insights on how apparently irregular AMS axis orienta- spread occurrence (from aerial photographs), which suggests the
tions can be interpreted in young lava flows. lowest viscosity at the emplacement time (Delgado et al., 1998). Distal
facies were preferred in an attempt to minimize slope effects in AMS.
RM and PC profiles are 9 km and CU is 10 km away from the Xitle
2. The Xitle volcano: Geologic background and previous studies cone, CU is just at the border front of the flow; hence, all profiles are dis-
tal facies. Tumuli, hornitos, spatter cones, and other structures suggest-
The Xitle1 volcano is one of the youngest in the monogenetic ing disturbing flows were avoided along the chosen profiles. CU and RM
Chichinautzin volcanic field (Martin del Pozzo, 1982), which belongs profiles were selected because both belong to the topmost individual
to the Trans-Mexican Volcanic Belt (TMVB; see inner box in Fig. 1). flow of unit VI, showing apparent lateral continuity. The PC profile
Xitle is a 150-m-high cinder cone located at the southwestern part of was taken from an individual flow of unit V that does not naturally out-
the Basin of Mexico, Mexico City. The Xitle cinder cone has characteristic crop on the surface but is in the central part of a series of at least ten in-
steep slopes. Its activity started with a tephra emission followed by for- dividual flows exposed inside a former quarry with a well preserved
mation of a smaller parasitic lava cone, the Xicontle, to the west of Xitle internal structure and clear lower and upper flow boundaries.
and then by a series of several lava flows which spread downslope over Samples from profiles CU, RM, and others from nearby outcrops
an area of 70.2 km2, mostly to the northeastern part of Xitle (Fig. 1). The from the BPSA member were also collected for paleointensity surveys
Xitle lava flows have a particular archeological interest since they have (Morales-Contreras, 1995; Gonzalez et al., 1997, Böhnel et al., 1997);
covered the early human settlement of Cuicuilco. the samples studied by Morales et al. (2006) are the same used from
Earlier published 14C dating from sites with well controlled stratigra- the PC profile of this study. A comprehensive paleomagnetic study
phy indicate ages around 2000 yr. BP for the volcanic eruption (White (paleointensity, magnetic fabric, and rock magnetism) was also con-
et al., 1990; Ortega et al., 1993; Cordova et al., 1994; Urrutia-Fucugauchi ducted for the PC profile using a series of 10 individual lava flows that
and Martin del Pozzo, 1993, Urrutia-Fucugauchi, 1996; Delgado et al., outcrop in that place (Alva-Valdivia, 2005).
1998), but more recent works indicate ages around 1670 ± 35
(Gonzalez et al., 2000; Siebe, 2000).
Seven basaltic units from 2 to 40 m thick, each one corresponding to 3. Sampling description and local flow features
a different batch of magma, have been recognized and mapped accord-
ing to their morphology, petrography, and chemical analyses (units I to Flows were sampled with a gasoline portable drill using a mobile
VII from oldest to youngest, Fig. 1; Delgado et al., 1998). The two more scaffold or staircase in order to reach the different flow levels. Cores
widespread and thick lava members are the Ciudad Universitaria were obtained from different levels at close intervals from 10 to 20 cm
batch (BCU), labeled as unit V (25 m average thickness with a maximum and oriented with a magnetic and solar compass whenever possible.
We obtained one to three specimens from each core. Table 1 summa-
1
From nahuatl language, means navel. rizes sampling features (99 cores, 153 specimens were studied).
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 11

Fig. 1. Xitle flows (after Delgado et al., 1998) and profile sites. Profiles are on the outcropping flows mapped as unit VI (RM and CU profiles) and V (PC profile). General flow directions were
obtained from aerial photographs. Inner box shows location of the Chichinautzin VF (Volcanic Field), where the Xitle volcano belongs, in the Trans-Mexican Volcanic Belt (TMVB).

Two main zones are clearly visible within the three flows: (1) A by Pj, and ellipsoid shape by T (Jelínek, 1981). A Gaussian counting
lower compact part practically without vesicles, which includes (a) a point density distribution statistic (K = 1; Robin and Jowett, 1986)
narrow basal subzone with tiny vesicles, not always present, and was also calculated using SpheriStat software to check the agreement
(b) an upper subzone with very scarce to scarce vesicles. (2) An upper between the means of the principal directions and the maximum distri-
part with characteristic abundant vesicles of several sizes usually hori- bution areas. We performed a detailed comparison between AMS re-
zontally elongated. These parts are graphically represented in vertical sults from the Minisep and KLY2 instruments using a selection of
flow profiles in Fig. 2 (photographs with inserted cartoons). Flow profile exactly the same samples from the CU and RM profiles. Since the com-
features are detailed but schematically described in Table 2, including parison resulted in similar and consistent results, we employed the
vesicle size description, distribution, and corresponding aspect ratios whole collection of data obtained from all samples of these profiles
of a limited amount of vesicles; general vesicle distribution agrees (Figs. 1-3 and Table 1 in Supplementary Information).
with the model of Sahagian (1985). Thermomagnetic curves of samples from the CU and RM profiles
were obtained from measurements at ambient conditions using a
4. Methods laboratory-built horizontal balance (between room temperature and
700 °C) at the Geomagnetic Laboratory (University of Liverpool). The
AMS from all CU and RM samples were measured with a Molspin– low-temperature dependence of low field magnetic susceptibility of
Minisep anisotropy system and later the CU and half of the RM samples both profiles was determined between liquid nitrogen and room
were also measured with a Kappabridge KLY2; PC samples were all temperatures (77–290 K), using a Bartington AC bridge with a water-
measured with the KLY2. All data were processed using Anisoft software cooled sensor. PC thermomagnetic curves were obtained from measure-
to create the AMS data files. Means of AMS principal directions of ellip- ments at ambient conditions of crushed samples (~ 100 mg), using a
soids were obtained using Jelínek statistics (Jelínek, 1978), referred as Highmoore Susceptibility Bridge equipped with a furnace. Experiments
K1 for maximum, K2 for medium, and K3 for minimum directions. were completed between room temperature and 650 °C and were heat-
Mean susceptibility is indicated by Km, anisotropy degree parameter ed and cooled at a rate of 20 °C/min (Alva-Valdivia, 2005).

Table 1
Flow profile thickness and sampling features. Number of specimens measured with each instrument at each flow. Shaded cells correspond to measurements mainly employed in this
study.

Profile Thickness (cm) Core number Sampling interval Specimens per core (average) Instrument used
[specimens] (average, cm)
Minisep KLY2

CU 659 39 17.5 (1) 42 specimens 36 specimens (base to top)


[42] (base to top)
RM 495 38 6.3 (2) 75 specimens 38 specimens (base to 0.6 height)
[75] (base to top)
PC 480 22 13 (1.5) — 36 specimens
[36] (base to top)
12
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28
Fig. 2. Photograph and schematic structure of PC, RM, and CU flow profiles with a general ore microscopy summary. Vertical scale indicated is in centimeters (right) and in normalized height (left). Relative vesicle density and shape (ellipses) are
schematically represented; further details of vesicle size distribution and aspect ratios are in Table 2. Blue arrows along PC and RM profiles indicate the positions of samples of Figs. 4 and 5, respectively; stars are positions of samples of Fig. 3.
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 13

Flow profiles description: facies, stratigraphical relationships, profile subdivisions delineated by their main structure and their thickness, details about vesicle size and aspect ratio distribution, further details in Fig. 2. (Number of measurements for
Hysteresis parameters of small core-chips from all profiles were

Lower zone rich in small v = 1.6


Upper zone of scarce and lower
measured using a Princeton Instrument MicroMag employing maxi-

zone of rich small v = nearly

Upper zone of scarce v = 2.2


mum applied fields of 1–1.5 T.

Upper v = nearly 1–2.2


(Aspect ratio average) 5. Magnetic mineralogy
Vesicles (v) form

Middle v = 7.4

Middle v = 6.2

Middle v = 2.4
Average = 6.3

Average = 2.2
Small v = 5.1

Small v = 4.8

Average = 7
Big v = 10.8

Small v = 2
5.1. Ore microscopy
Big v = 6.1

1–2.5
Reflected light microscopy on polished sections from the three pro-
files was performed using the oxidation and texture classification

The upper half with scarce small v increasing density upward. The
30 cm (0.06), rich in tiny nearly round v and some basal large pipe
lowest 25–50 cm (0.04–0.08), rich in small nearly round v and
The upper half with scarce small v increasing density upward.

Abundant v increasing density upward. Mostly small (0.5–3) and


scheme suggested by Haggerty (1976) and Buddington and Lindsley

irregular tunnels, apparently produced by v interconnection.


distributed randomly along vertical. The v are elongated and

planes The v are elongated and aligned along preferable planes.


The upper third with scarce small v increasing density upward.
medium (5–18) sizes but also an important proportion of big v
The lower half practically w/out v. A thin red basal horizon:
Abundant v increasing density upward. More small (0.5–4)

The lower 2/3 practically w/out v. A thin basal horizon: lowest

Abundant v increasing density upward. Mostly small (1–4)


and medium (6–10) sizes and a couple of very big as small
(24–75). Big v sizes are randomly distributed along preferable
(1964), respectively. Oxidation classifications C1–C7 were applied for

lower half practically w/out v. Not enrichment of v at the base


titanomagnetite (Ti-Mag) and R1–R7 for titanohematite (Ti-Hem)
and medium (6–17) sizes than big (20–40). Size of v

v with its upper part slightly stretched in favor of flow.


(the prefix C for primary cubic phases and R for primary rombohedral
phases; 1–7 correspond to increasing oxidation states).
Typical characteristics of all profiles are titanomagnetite (Ti-Mag)
and large quantities of primary ilmenite (Ilm) with different tex-
tures, grain sizes, and oxidation states depending on their position
in the flow; outstanding along the central part of flows are large
Vesicle (v) distribution and size

aligned along preferable planes.

Ilm crystals (often N 100 μm) in comparison with Ti-Mag. Detailed


(maximum diameter in cm)

scattered basal large pipe v.

study of the RM profile indicates that the main oxide mineral is Ti-
Mag intergrown with Ilm; both display columnar shapes and skeletal
borders without any particular orientation; they are observed along
the borders of ferromagnesian anhedral crystals (pyroxene and
olivine). In minor proportion, there are also anhedral, subhedral,
and euhedral Ti-Mag crystals with Ilm lamellas, homogeneously
disseminated in the rock. A general summarized description of all
profiles follows (Fig. 2):
(normalized)

[272] (0.55)
[247] (0.37)

[412] (0.63)

[223] (0.45)

[150] (0.31)

[330] (0.69)
Thickness

5.1.1. Lowermost part 0–0.1 nh—normalized height (Fig. 3a–b)


[cm]

In this thin zone, there are small euhedral or subhedral Ti-Mag with
Ilm, 1–5 μm in size, showing either as disseminated or skeletal grain tex-
Upper Vesicular
Upper Vesicular

Upper Vesicular

tures. The inferred oxidation state in skeletal Ti-Mag is C2, whereas it is


Lower Compact

Lower Compact

Lower Compact
Main structure

C3 for subhedral grains and R1 for Ilm.


Profile part

5.1.2. Principal-central part 0.1–0.9 nh


In this main part, there are larger euhedral to subhedral Ti-Mag with
Ilm intergrowths showing light gray-brown color, trellis and sandwich
Upper part is slightly
covered by grass and
scarce vegetation and
Erosion surface with

10–30-cm-thick soil

textures that sometimes are fractured or brecciated. Ti-Mag crystals


individual flows.
Erosion surface
ropey structure.

vary from 1 to 40 μm in size with prevailing sizes around 25 μm; in the


thorn pioneer

Ropey to slab

A series of 4
vegetation.

most inner parts of the flow where all the grains are coarser, Ilm crystal
structures.

sizes are as large as 150 μm. The deuteric oxidation state here typically
Locally
Above

baked

varies between C3–C5 and R2–R5. Exsolution features are commonly ob-
served in the Ti-Mag, effectively sub-dividing the grains into smaller,
A baked flow and below it a series of 4
Laterally a lower flow unit with baked

magnetite-rich regions. Some crystals of the titanohematite (Ti-Hem)


Fine sediment layer with a few big

Soil or altered ash tuff with baked

series are as large as 80 μm and have inferred oxidations states of


subrounded clasts up to 15 cm in
diameter and ceramic fragments.

R2–R3.
Stratigraphical relationships

5.1.3. Uppermost part 0.9–1 nh


more individual flows

This thinner part is characterized by small euhedral Ti-Mag with Ilm


(sizes 1–5 μm) showing skeletal textures and oxidation states of C2–C3
and R1 for the Ti-Mag and Ti-Hem, respectively.
features.

Fig. 3 shows a detail of the RM profile based on samples taken from


features
Below

different positions above the base (nh): 0.06, 0.16, 0.28, 0.38, 0.43, 0.46,
0.54, 0.6, and 0.65. All samples show Ti-Mag partly or completely
aspect ratio average at each flow is ~50).

Intermediate and in internal


Distal in frontal lobe border.

altered to Ti-Hem (Fig. 3a–g), suggesting different alteration degrees.


profile there are tumuli and
Laterally and closely to the

Distal, but still in internal

thinning laterally closely

Ilmenite is homogeneously distributed in the Ti-Mag and has a trellis


individual flow sampled
part of flow lobe, but

shape texture; it has formed due to high-temperature oxidation from


earliest homogeneous Ti-Mag (Fig. 3b-f). Ti-Hem fills thin fractures
part of flow lobe

(15 μm thick) in the Ti-Mag and sometimes is altered to botryoidal


goethite (Fig. 3e). Titanomaghemite (Ti-Mgh) is rare in our samples
hornitos.

and occurs in very thin concave-shaped fractures in the Ti-Mag


Profile Facies

(Fig. 3c and 3h). Ilm generally is altered to rutile in granular shape


and fine grain size (b 10 μm).
Table 2

A systematic gradation was observed in the size and shape of crystals


RM
CU

PC

throughout the flow: (1) At the bottommost sample 0.06 and the
14 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 3. Opaque mineral photo-micrographs of representative samples from the RM profile. Sample numbers are their distance (nh) above the base (pointed as stars in Fig. 2). Images were
taken with a reflected light polarized microscope: a) Ti-Mag columnar intergrowth with very small Ilm homogeneously distributed without apparent orientation (some show skeletal
borders). Toward the center are anhedral crystals of Ti-Mag. b) Ti-Mag columnar crystal forming siebe texture with Ilm. Ti-Mag is cut by a fracture partly occupied by Ti-Hem. c) Ti-
Mag columnar crystal forming graphic texture with ilmenite. Ti-Mgh replaces the Ti-Mag through curved fractures. d) Columnar crystals of Ti-Mag with skeletal borders of Ti-Mag and
associated Ilm. These columns are larger than the other studied samples; some columns show veins of Ti-Hem. e) Fraction of an octahedral crystal pseudomorph of Ti-Hem, produced
by the Ti-Mag alteration, it partly contains Ilm relicts forming trellis texture with Ti-Hem. Goethite (Gth) is filling a thick fracture in the Ti-Hem pseudomorph, forming a botryoidal
texture vein. f) Ti-Mag columnar crystal showing skeletal borders replacing by Ti-Hem, some Ilm relicts are in siebe shape. g) Columnar Ti-Mag crystals partly altered to Ti-Hem and
intergrowth with Ilm, at the center is a Ti-Mag anhedral crystal. h) Ti-Mag euhedral crystal showing skeletal borders and some curved fractures were it is partly replaced by Ti-Mgh.

topmost samples 0.6 and 0.67, Ti-Mag crystals are anhedral to subhedral 5.2. Rock magnetism
and on average ~ 40 μm in size; (2) within the central part of the flow
(0.28–0.43 samples), in general the width and length size of crystals in- Magnetization, temperature-dependent magnetic susceptibility, and
creases up to 50 μm and 110 μm, respectively (Fig. 3a–h). The similarity hysteresis parameters (saturation magnetization, Ms; saturation rema-
between both parts of a faster cooling rate is obvious from these nence, Mrs; coercive force, Hc; and coercivity of remanence, Hcr) were
observations. measured from samples along the profiles. Results of the rock magnetic
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 15

properties were previously reported in Böhnel et al. (1997) for the CU


profile and some results from the PC profile are partly described in
Alva-Valdivia (2005) and Morales et al. (2006). A summary of results
from all profiles including those from previous and present studies (all
from RM and some of PC profiles) follows.

5.2.1. Thermomagnetic curves


K-T curves from CU indicate Curie temperatures around 575–580 °C
particularly at the flow base with slight inflections at around 200 °C
(Böhnel et al., 1997). PC results from selected samples representing all
zones (Fig. 4) indicate the presence of Ti-poor Ti-Mag at the lower
flow part (PC-62 and 67) and only one magnetic phase in all samples;
the Curie point is 580 °C. PC-62, 67, and 77 do not show perfect revers-
ibility probably due to the presence of some titanomaghemite, observed
in thin section. This Ti-Mgh might have transformed into magnetite
during experimental heating.
Low-temperature susceptibility curves from the RM profile (lower
and upper flow levels; Fig. 5) present two main features: a susceptibility
peak close to −150 °C and a rapid susceptibility increase at low temper-
Fig. 5. Representative low-temperature vs. susceptibility (K) curves, normalized with
atures; the same two features are also seen in the CU samples. The peak
respect to the value after warming to 5 °C. Position of samples is indicated in Fig. 2.
is an expression of the isotropic point of multidomain (MD) magnetite
(Dunlop and Özdemir, 1997) and is mostly observed in samples from
the lowermost part of the flow (0–0.1 nh) and from the upper part of viscous lava flow, is characterized by samples that vary strongly in
the flow (around 0.8–1 nh), but this magnetic behavior is not observed their Mrs/Hc ratio. Different symbols for samples from lower, central,
close to the lower flow margin, where it might indicate the conversion and upper flow zones in Fig. 6 clearly show that specimens from the
of Ti-Mag to Ti-Mgh. The increase in magnetic susceptibility at low tem- lower part of the flow tend to be nearer to the MD region and samples
peratures indicates a Fe3+-bearing ilmenite phase with a TC at about from the upper flow zone are nearer to the SD region. We observe two
−180 °C (Engelmann et al., 2010). Ilmenite was observed microscopi- exceptions of this pattern in the flow from CU; these exceptions corre-
cally in thin sections (see Fig. 3). A Curie temperature of a Ti-bearing he- spond to samples at the basal flow margin that are in the SD or near
matite phase has not been identified up to 650 °C. This behavior is the SD grain size region.
observed across the central parts of the flow and is most pronounced The within-flow variation of a selection of the measured magnetic
between 0.2 and 0.5 nh above flow-base where the ratio χ77K/χ278K oc- parameters within the RM profile is presented in Fig. 7. The trends in
casionally exceeds 3 (0.02–0.04 nh in the RM profile). Hcr/Hc, Mrs/Ms, and natural remanence (NRM) intensity in the upper-
most part of the flow, around 0.85–1 nh (the uppermost 60–70 cm
5.2.2. Hysteresis properties from RM), suggest a shift to SD-like grains and therefore to smaller
Ratios of hysteresis parameters (Fig. 6) suggest that almost all sam- grain sizes. Further down, between 0.05 and 0.15 nh above the flow
ples from the CU profile and most from the RM profile fall in the pseudo- base (10–70 cm), low Mrs/Ms and Hc values suggest a zone of coarse
single domain (PSD) grain size region (Day et al., 1977). This observa- magnetic grains. NRM is variable between 0.6 and 0.8 nh (300–
tion may indicate a mixture of MD and a significant amount of SD grains 400 cm) above the flow base. This interval is within the upper zone of
(Dunlop, 2002). The PC profile, which corresponds with the least the RM flow and we suggest that this may be due to small oxidation

Fig. 4. Susceptibility versus temperature curves. Position of samples above bottom is below samples name (nh = normalized height) and marked in Fig. 2.
16 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 6. Day plot (Day et al., 1977) shows the relationship between the hysteresis parameters of PC, RM, and CU profiles. As described in legend, larger dark circles are specimens from the
lower flow zones and smaller circles from the upper zones; open-light shadow circles are specimens from the middle zones (zones are defined in Figs. 9 and 10).

changes or to the amount of magnetic minerals. The within-flow varia- (Fig. 8a, b). In the PC profile, the lineation is better defined (smallest CE-
tions of magnetic parameters of the CU and PC profiles are similar. 95% of K1) than in the RM and the ellipsoids are more oblate (Table 3). In
the CU profile (top of lava flow, unit VI) the K1 mean is neither parallel
6. AMS results nor perpendicular to the general flow direction; it has the worst defined
lineation and the least oblate ellipsoids (Table 3).
The AMS principal axes for each site and their means are plotted on
stereographic projections (Fig. 8). Point density distribution curves con- 6.3. Point density analysis
tours (Gaussian counting) of the principal axes are also displayed. AMS
results are in Table 3: maximum (K1) and minimum (K3) means of sus- In all profiles, the K3 directions show two maximum peak areas, the
ceptibility directions; the minimum and maximum apical axes of their K3 means are between them (Fig. 8), suggesting two different statistical
corresponding 95% confidence ellipses (CE-95%); and the AMS parame- populations. Vertical K3 direction distribution along profiles revealed
ters Km, Pj, and T. that each statistical population corresponds to a different vertical posi-
tion within the profile. The same is also observed with the K1 directions,
6.1. Clustering particularly in the PC and RM profiles.

Samples from the PC and RM profiles show a better clustering of di- 6.4. Vertical AMS directional variation analysis
rectional data than the CU profile. In Fig. 8, the results from the PC and
RM profiles are represented each by only one diagram of all AMS prin- The vertical variation in direction of K1 and K3 within each profile is
cipal axes, whereas CU is represented by two diagrams in order to shown in Fig. 9. These graphs show different flow levels (zones) each
more clearly show the more scattered K1 and K2 axes distribution. The distinguished by a particular AMS pattern. In the PC profile, only a
better clustering in PC and RM compared to CU is also evident from lower and an upper flow level are identified, whereas in the RM and
the min and max CE values (Table 3). CU profiles, a central part is also recognized.
The lower zone has a normalized thickness of 0.43 in PC, 0.33 in RM,
6.2. Magnetic foliation and lineation and 0.30 in CU profiles. The AMS directions of this zone, in each profile,
show a very similar pattern, and the best clustering (CE-95% values in
In each of the three profiles AMS ellipsoids are mainly oblate (T N 0, Table 3). Both the well-defined magnetic lineation and the magnetic fo-
Table 3); K1 and K2 axes define a sub-horizontal magnetic foliation plane liation plunge and dip toward the SW. The magnetic lineation is parallel
that dips toward the SW, the direction of the flow source. In the PC and to the flow direction observed from aerial photographs and shown in
RM profiles, the mean magnetic lineation (K1 mean) is parallel to the Fig. 1. The low imbrication angle of the magnetic foliation (means
general flow direction, the K1 mean points upward to the flow direction from 5° to 23°) points to the NE, in the same direction as the flow
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 17

Fig. 7. Intra-flow variation of magnetic parameters for profile RM. (a) natural remanence, NRM; (b) Curie temperature, Tc; (c) saturation magnetization, Ms; (d) ratio of saturation
remanence to saturation magnetization, Mrs/Ms; and (e) ratio of coercivity of remanence to coercive force, Hcr/Hc.

direction. We describe this relationship as “positive” imbrication both central zones are similar and show a completely different AMS pat-
(lower equatorial diagrams of Fig. 9). Particularities of the lower tern not observed in other profiles: the K1 axes are perpendicular in re-
zone are as follows: In the RM and CU profiles, this lower flow level lation to the flow direction; the K2–K3 axes show a girdle distribution
has the particularity that some lowermost AMS ellipsoids (1–3 spec- that defines high dipping planes (nearly vertical in Middle-1 CU zone
imens) show irregular orientation (e.g., K1 directions perpendicular and dipping 40° toward SE in Middle-2 CU; Fig. 9c). Unique particulari-
to flow and/or K 2 directions switching position with K3 direction; ties observed in the central zone include thin levels at the base of the
Fig. 9b); these AMS ellipsoids correspond to a thin basal horizon central zone of the RM profile zones that are distinguished with AMS el-
with vesicular structure. lipsoids that show irregular orientations like those described also at the
The upper zone of the PC and RM profiles, from 0.43 nh to the top, and base of the lowermost part of this site (empty symbols in Fig. 9b). Sim-
from 0.55 nh to the top, respectively, show a very similar AMS pattern: ilar irregularities are also observed at the topmost thin section of the
their AMS ellipsoids plunge toward the NE; the K1 axes and mean mag- upper zone (filled crosses and stars in Fig. 9b). These particular AMS el-
netic lineation are parallel to the flow direction. The low imbrication lipsoid arrangements define a vertical foliation and a transverse (in re-
angle of magnetic foliation points to the SW in an opposite way to the lation to flow) lineation.
flow direction and as observed in the lower flow level (Fig. 9a, b). Sim-
ilar patterns have been noticed in other flows (e.g., Bascou et al., 2005). 6.5. Vertical AMS parameters variation analysis
Because the imbrication angle is in the opposite direction than the flow
direction, this pattern is referred as “opposed” imbrication (Cañón- Variation of the Km, Pj, and T parameters along each vertical flow
Tapia et al., 1997). In contrast, the upper zone of the CU profile shows profile are shown in Fig. 10 and Table 3. Fig. 10 show the relationship
an AMS pattern consisting of a mixture of “positive” and “opposed” im- between parameter variation and the vertical zones identified by
brications (toward and against flow source). Frequently, the K1 direc- means of directional AMS variations.
tions show very high inclinations (40°–50°). Additionally, the upper
zone in the CU profile has the worst clustering of all profiles (Fig. 9c). 6.5.1. Km parameter
The central zone is only distinguished in the RM and CU profiles, it The PC profile has the highest average values of the volumetric mean
range from 0.3 to 0.55 nh and from 0.33 to 0.62 nh, respectively, and it susceptibility; Km (around 6.6 × 10−3 SI units); the RM and CU profiles
is quite different in each profile. In the RM profile, the AMS of the central show lower average values, which are more similar between them
zone is similar to the one in the upper zone, both characterized by “op- (around 3.1 and 3.7 × 10−6 SI units, respectively). Higher PC Km values
posed” imbrication, but the central zone is less clustered, and with a pair are consistent with the fact that PC belongs to the unit V, which is con-
of specimens around 0.35 nh with positive imbrication (Fig. 9b). In the sidered having the lowest viscosity (Delgado et al., 1998) and hence a
CU profile, the central zone can be subdivided in two parts since their higher emplacement temperature, this might favor the formation of a
clustering shows slightly different directions in each part (Fig. 9c); higher amount of Ti-Mag crystals. Alternatively, the lower Km values
18 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 8. AMS principal axes distribution within the (a) PC, (b) RM, and (c) CU profiles: K1 (squares and crosses), K2 (triangles), and K3 (circles and stars), with their corresponding means and
95% confidence ellipses (Table 3). Different sizes and shapes of symbols are according with vertical position of samples as described in the legend; nh = normalized height. Contour curves
are point density distribution of AMS axes (Gaussian counting); notice the two maximum peaks for the K3 axes population in all sites. Arrows indicate local flow direction inferred. Dotted
perpendicular lines are vertical planes of symmetry of AMS principal directions.

in RM and CU may be due to a higher chilling velocity, which might 70% of the flow thickness, Fig. 2) and the position of the CU at the frontal
favor the formation of more volcanic glass and then less Ti-Mag crystals. lobe border seem to us to be in agreement with a higher chilling flow
The thick width of vesicle distribution observed in RM (the upper 55– velocity in both cases.

Table 3
AMS results from the 3 profiles: K1 and K3 principal directions, parameters and inferred flow direction. First row from each profile shows all specimens sampled; the shaded row describes
only specimens from the lower flow zone. Means are calculated using Jélinek statistics (K1 = maximum, K3 = minimum principal directions). CE 95% = angles (minimum and maximum)
of the 95% confidence ellipses for each direction. Km, mean susceptibility (in 10−6 SI units); Pj, anisotropy degree; and T, ellipsoid shape (T N 0 oblate, T b 0 prolate). The “prolate %” column
describes the percentage of specimens with prolate AMS ellipsoid shape (T b 0). Local flow directions inferred as described in the text.

Site n K1 CE K3 CE Km Pj T Prolate % Local flow


direction inferred
Dec Inc Min Max Dec Inc Min Max
(°) (°) (°) (°) (°) (°) (°) (°)

CU 42 172 20 33 62 13 68 31 39 3765 1.025 0.333 15% 38


CU 13 202 5 14 40 2 85 14 21 3289 1.029 0.579 0% 21
lower layer
RM 75 225 4 19 47 111 80 20 27 3136 1.044 0.329 12% 62
RM 34 252 20 8 30 88 69 7 18 3008 1.053 0.347 4% 72
lower layer
PC 36 237 12 21 27 37 77 20 27 6590 1.032 0.475 6% 54
PC 21 241 23 13 25 51 67 12 19 6494 1.037 0.581 0% 56
lower layer
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 19

The larger within-profile variations of Km are also within the PC pro- seems to be a general relationship between well-clustering AMS direc-
file (Fig. 10a), a right concave-shaped curve of these values is observed tions and higher Pj values.
along the flow profile. The higher values are at (a) the bottom portion of
the flow: 0–0.2 nh, and (b) the upper portion (the vesicular part of the 6.5.3. T parameter
flow): 0.7–1 nh. The lower values are at the central part along the sec- The average values of the shape parameter (T) at each profile, rang-
tion from 0.2 to 0.5 nh of the flow. Similar, but less distinct, concave ing from 0.329 to 0.475 (Table 3), indicate the dominance of oblate
curves are observed in the RM and CU profiles. Within the RM profile, shapes (T N 0 oblate, T b 0 prolate) with small fractions of prolate
some high Km values are also observed at the base of the central flow shape specimens (up to 6% in PC, 12% in RM, and 15% in CU). Fig. 10
level, showing for the specimens with these values irregular AMS shows that the highest positive T values (oblate shape) are within the
orientations. lower lava zones of all profiles. This zone also has the lowest fraction
of specimens with prolate AMS ellipsoids, and the strongest clustering
of AMS directions. In the PC and CU profiles, we observe no prolate
6.5.2. Pj parameter shapes within lower zones, and within the RM profile, the prolate
Along the three profiles, higher anisotropy degrees (Pj values) are shapes are restricted to a thin subzone near the bottom (Fig. 10b),
within the lower zone, and the lowest values are at the upper zone, where AMS ellipsoids show lineation or vertical foliation perpendicular
where the AMS directions show the lowest clustering (Fig. 10 and to flow. The PC profile has higher positive T values and the few prolate
Table 3; focus on the lowest maximum CE-95% values and higher Pj to nearly neutral shape specimens are only within the uppermost part
values of the shaded rows). The RM profile has the highest Pj values, of the flow (Fig. 10a).
and the CU profile shows the lowest Pj values; the latter has the weakest The CU profile has the highest proportion of specimens with prolate
clustering of AMS directions (Table 3). We note that in these flows there shape of all three studied profiles studied, including some prolate to

Fig. 9. Vertical flow declination and inclination variation of K1 (squares) and K3 (circles) within (a) PC, (b) RM, and (c) CU sites. Vertical scale is normalized, vertical axes are centered at
corresponding K1 and K3 mean values; a general vertical structure description is along with K1 inclination graph. Legend indicates size of symbols decreasing with height; alternate symbols
(e.g. crosses and stars) are ellipsoids with irregular orientations. Horizontal dotted lines mark boundaries between identified zones, different size and shadow of symbols are according
with these zones. Each zone is complemented with respective stereodiagrams, with corresponding means, confidence ellipses, vertical planes of symmetry (dotted lines), and flow
inference direction (arrow). Notice the AMS “positive imbrication” in all lower zones, “opposed imbrication” in PC and RM upper zones and transverse oblique imbrication
arrangement in CU middle zones (see text for relative explanation).
20 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 9 (continued).

nearly neutral shapes; these specimens are concentrated within the the deformed states of the square net, circle shapes change but areas are
central part of the flow (Middle 1 and Middle 2 zones in Fig 10c) preserved; and all circles inside squares of same row (and presumably
where Pj values are the lowest, although some come also from thin sub- AMS ellipsoids at same lava flow level) show homogenous deformation.
zones from the lowermost and topmost parts of the flow. All these zones If the deformation on the square were only due to the weight of lava
and subzones also have some specimens with transverse lineation (in flow (vertical shortening due to pure shear) the square net would be
relation to flow) or even sub-vertical foliation or lineation. The highest similar to Fig. 11f or g.
proportion of prolate to nearly neutral shape within the RM profile
also occurs within the central flow zone. We conclude that there is a 7.1. Flow-induced shear strain in models and vertical flow velocity
general relationship between (a) oblate shape specimens with the variation
lower flow layers and higher Pj values; (b) prolate and near-neutral
shape specimens with lower Pj values in the middle flow layers; (c) The pairs of half arrows in Figs. 11 and 12 indicate the sense of flow-
and prolate and near-neutral shapes in thin lowermost and uppermost induced shear strain for hypothetical states. In these figures, it can be
layers; and (d) prolate and near-neutral shape specimens and trans- seen that the vertical position of maximum flow velocity (Vmax) con-
verse to flow ellipsoid orientations. trols the sense of flow-induced shear strain and hence the symmetry
of AMS, ellipsoid orientations, and related parameters. The Vmax verti-
7. Heterogeneous deformation modeling and possible effects in AMS cal position should in turn be related with the flow viscosity and the dif-
ference between bottom velocity (Vb) and top velocity (Vt). The Main
In order to qualitatively understand the relations between AMS el- different cases of Vt vs. Vb relations are Vt N N Vb (Fig. 11b and c),
lipsoid imbrication and flow-induced shear strain, we have elaborated Vt = Vb (Figs. 11d and e), and Vt ≥ Vb (Fig. 12), each occurring in differ-
some sketches based on heterogeneous deformation on a net of minor ent scenarios. Different vertical rate of flow velocity change may also
squares obtained from a “toy” program used in structural geology lec- occur (Figs. 11d and e, 12a–c).
tures (CizHet program by Tolson, 2009), with which it is possible to If the Vmax were located at the top of a flow, as it may be the case for
model flow velocity changes. Sketches are in Figs. 11 and 12 (where a Newtonian fluid, or in the very moment when lava starts to flow, de-
the z axis is the vertical and the x axis represents the flow direction.); formation on the square net unit would be similar to the sketch of Fig.
they correspond to initial state of no deformation (Fig. 11a) and first 11b. If the weight of the flow is considered, some vertical shortening
stages of flow. In the initial state the square net shows perfect circles in- would be additionally applied and the deformation would be similar
side it, which would correspond hypothetically to an isotropic fabric. In to the sketch of Fig. 11c. In contrast, a viscous fluid such as a lava flow
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 21

Fig. 9 (continued).

would have the Vmax in a middle part of the flow, as has been described zone are more elongated than in any other part of the flow: Pj values
within dykes (Callot and Guichet, 2003), then the induced shear strain are mostly above the general average and T values are positive and
would have different sense below rather than above (Fig. 12d). Similar above the average. The thickness of this zone is different in each profile;
as in a dyke, there is a different sense of shear stress along both parts its height may be tentatively considered as the height of the Vmax plane
of the central flow axis (Callot and Guichet, 2003). If there were addi- of the flow. Considering inferred position of the Vmax of flow, and the Pj
tionally some shortening due to the lava flow weight, ellipses of the values (higher at RM, lower at PC, and lowest at CU), some simple dia-
square net would be slightly flattened (Fig. 11e). These sketches graph- grams were elaborated (Fig. 12) as an attempt to explain these features.
ically show how shortening strain becomes almost negligible, whereas Since flow velocity controls the sense and intensity of shear stress and
shear strain increases when flow velocity increases (as stated by induced strain, then the relative differences in ellipsoid eccentricity
Merle, 1998). The decrease of imbrication angle of the AMS ellipsoids should correspond to the flow velocity's vertical distribution.
due to shortening, if there is any, is more evident around the Vmax Models represent the final state of a progressive deformation in a re-
plane at the middle flow position (compare long axes of ellipses in Fig. gime of constant flow, and although they do not consider other vari-
11e vs. 11d). ables, such as those that affect mainly the upper part of flows; they
These qualitative sketches may then help to explain, from a graphi- may explain the main differences in imbrication and elongation of the
cal perspective, not only the variability of AMS imbrication along verti- AMS within the lower and middle zones. The other variables not consid-
cal lava flow profiles (Cañón-Tapia et al., 1996, 1997; Merle, 1998; ered that may give overimposed effects in AMS patterns represented by
Bascou et al., 2005) or within dykes (Callot and Guichet, 2003; these models are the rheological interactions between solidified crust
Eriksson et al., 2011), but also the variability of other AMS features. and fluid lava, and the possible flow sheeting due to drastic changes in
viscosity derived from differences in cooling rate.
7.2. Models explaining vertical AMS variations and other profile features In order to elaborate any inference about flow from models, we have
related to consider that new or continuous lava batch emissions may increase
or maintain flow velocity longer; that more fluid lava (lower viscosity)
The AMS patterns of the lower AMS zone in our profiles display pos- should favor higher flow velocity, and possible higher positions of
itive imbrication, suggesting that the Vmax of the flow that induced Vmax in the flow; and that lava cooling results in a decrease of velocity,
shear strain was above it. This zone fits with the lower part of profiles possibly lowering the position of Vmax in the flow and rheological ef-
with a compact structure lacking of vesicles. AMS ellipsoids from this fects due to viscosity changes derived from cooling.
22 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

8. Discussion this part the lava remained liquid and warmer for a longer time than
in the upper part where it cooled relatively quickly. The longer cooling
8.1. Magnetic mineralogy time and related slower cooling rate permitted the growing of crystals
while in the upper part the faster cooling is in agreement with smaller
Although magnetic properties and ore microscopy observations sug- grains. The small grains (SD or near SD) observed in the lowermost
gest that the main AMS carrier is Ti-poor magnetite within all the three lava flow may be explained by the chilling effect of the lava when it
profiles, meaningful grain size differences are also observed: a tendency flows over cooler paleo-landscape surfaces. The smaller oxidation states
for a higher proportion of large grain sizes (PSD-MD) along the lower- of the Fe\\Ti oxides suggested for the upper part of the flow, at least in
central flow zones and smaller grain sizes (PSD-SD) along the upper the RM profile, are probably associated with higher cooling rates and
flow zones (Fig. 6). Grain size also significantly decreases toward the subaerial exposure, which may agree with the greater thickness of the
lowermost flow margin. Ilmenite grains show similar size-increase to- upper vesicular structure in the RM profile in comparison with the CU
ward the flow center, where large grains are conspicuous; this may con- and PC profiles. In contrast, the higher oxidation states of the Fe\\Ti ox-
tribute to the paramagnetic signal in magnetic experiments (hysteresis ides, observed at the top most part of the PC profile, are probably related
curves). Additionally, mostly higher oxidation is present within the with the thinnest thickness of the upper vesicular structure (Fig. 2). The
upper flow zones. grain size distribution along each profile roughly correlates with the
The coarser grain size distribution along the lower flow zone might AMS pattern distribution observed. Both, oxidation degree and grain
be consistent with an inflated lava flow mechanism indicating that in size, are suggested to be a result of lava flow dynamics.

Fig. 10. Vertical flow variation of Km (in 10−6 SI units), Pj and T parameters within (a) PC, (b) RM, and (c) CU sites with general structure description. Vertical scale is normalized;vertical
axes (vertical dotted line at T graph) are located at corresponding average values. Horizontal dotted lines mark boundaries between identified zones, different sizes and shadow of symbols
are according with these zones; alternate symbols (e.g. crosses) are the same as Fig. 9. Average parameter at each zone is annotated. Notice how parameters change according with zones.
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 23

Fig. 10 (continued).

8.2. Flow dynamics The RM model suggests the highest flow velocity (highest Pj values
in lower zone and higher induced shear strain), compared to the PC
The vertical within-flow variation of the AMS directions marks flow and CU profiles if the viscosity for all flows is assumed to be similar
zones related to the flow-induced shear strain of lavas independently (Fig. 12). Because the RM profile has the thinnest lower zone (0.3 nh)
whether strain is carried on minerals from the silicate framework (e.g., the Vmax plane is at the lowest position. The large thickness of the vesic-
plagioclase) controlling the crystallization of titanomagnetite (Bascou ular structure suggests an intense dissipation process, which is in agree-
et al., 2005) or effects directly the magnetic mineral fraction (Cañón- ment with high oxidation states of Fe\\Ti oxides. We suggest that these
Tapia and Coe, 2002). In this section, we discuss the observed vertical features besides the highest Pj values observed in the lower zone of the
AMS variations and suggested models in relation to the flow dynamics RM profile are consistent with a more intense and continuous lava-in-
under consideration of the magnetic mineralogy and indicators of the flated process. Additionally, the upper vesicular zone thickness of the
volcanic flow structures as far as available. The elaborated models, RM profile agrees with the normalized thickness (0.4–0.6) proposed
which best explain the observed AMS variations (Fig. 12), show a by Cashman and Kauahikaua (1997) as indicative of flow inflation.
lower positive and an upper opposed imbrication separated by an inter- Within the CU (0.34 nh) and the PC (0.43 nh) profiles, the top of the
mediate zone of lowest degrees of anisotropy. The upper boundary of lower zone (Vmax plane), is at a higher position. A thicker lower zone is
the lower zone showing positive imbrication is considered as the posi- consistent with a slower cooling rate, high oxidation states at the top-
tion of the Vmax of the flow. The vertical distribution of strain ellipses most of the profile, and a relatively higher temperature required for a
shows the higher Pj values within the lower zones of the three basaltic higher Ti-Mag fraction, as can be inferred from the highest susceptibility
flow profiles. values observed in the PC profile (Fig. 10). A higher Ti-Mag fraction in
24 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 10 (continued).

the PC profile is also in agreement with the relatively lower viscosity parts of flow in semisolid phase. Since the CU profile is very near to
proposed for the lava flow unit V by Delgado et al. (1998), since lower the distal flow boundary, a lower velocity and a partial semisolid flow
viscosity lavas usually have compositions with higher Fe content. Lava are both plausible.
with a lower viscosity may induce less strain and lower Pj values may Irregular imbrication within middle zones of the RM and CU profiles
result as it is observed in the PC profile in comparison with the RM pro- (Fig. 11d and e) occurs around the Vmax and the line (sheet in 3D) of no
file. We suggest that these features are consistent with a less intense shear strain (Merle, 1998). It agrees with the lowest elongation of ellip-
and shorter lava-inflated process. ses and corresponds to the lower anisotropy degrees and frequently
Within the upper zones of all three flows opposed imbrication in re- changes in AMS shapes observed. These conditions might additionally
lation to the flows occur, although the K1 and K3 axes are better defined favor a lower accuracy in AMS measurements. Middle zone height is re-
in the PC and RM profiles (Fig. 9a and b, Table 3) suggesting similar flow lated to the Vmax plane but its thickness may be related to the rate of
conditions for both. The upper zone within the CU profile shows both change of the flow velocity above and below the Vmax along the profile
positive (7/16 specimens) and opposed (8/16) imbrication in an imper- (Fig. 12a, b, and c). For example, the lack of a middle zone in PC might be
fect mixed alternation without any vertical position preference (see ver- explained with sudden changes in flow velocity above and below the
tical K1 variation in Fig. 9c). Two possibilities can explain this magnetic Vmax position. The decreasing-up values of Pj within the upper AMS
fabric orientation. (a) Low flow velocities produce lower shear strain in- zone of the PC profile (Fig. 10a) may be explained by a sudden decreas-
tensities (Fig. 12c), which may decrease the accuracy in AMS measure- ing rate of flow velocity (Fig. 12a). The proposed model attempts to con-
ments. (b) Lava flow with higher resistance to flow due to the formation sider all of these variations (Fig. 12a). Within the RM and CU profiles,
of a solid or semisolid phase in this zone may favor inner rotations of the middle AMS zones may be explained by less sudden changes in
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 25

Fig. 11. Models of strain ellipses, presumably AMS ellipsoids projections along z–x and z–y planes (z is vertical, x is flow direction) during a flow-induced shear strain. (a) Starting point of
no flow and no deformation time = 0. In diagrams (b) to (e), maximum flow velocity (Vmax) is dotted red arrow, pairs of half arrows indicate sense of induced shear strain and bottom
velocity (Vb) is =0. (b) Is when flow starts, time = 1. (d) Is when a crust has formed and top flow velocity (Vt) is =0; Vmax is around the half of the flow moving column, reversing the
sense of shear; time = 2. In (c) and (e), an unspecific amount of pure shear has been added for times 1 and 2 which could correspond to flow weight, flattening ellipsoids and producing
horizontal maximum axes along sections are around line of maximum flow velocity. Diagrams (f) and (g) show AMS ellipsoids projected in an orthogonal to flow vertical plane with some
vertical shortening (see text for more explanation). (For interpretation of the references to color in this figure legend, the reader is referred to the web version of this article.)

the decreasing flow velocity next to the Vmax position and then a The thickest middle zone is within the CU profile where it is
thicker sheet of no induced shear strain to minimum induced shear subdivided in two subzones (Figs. 9c and 10c), both with perpendicular
strain evolve (Fig. 12b and c). to flow lineation and not well-defined magnetic foliation, features not
26 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Fig. 12. Models showing vertical flow velocity variation, along flow profiles and AMS ellipsoids imbrication and elongation distribution along a z–x plane (a–c) and a z–y plane (d); Vmax is
a dotted arrow and half arrows represent induced shear strain. (a) A model with a near half vertical position of Vmax, decreasing velocity along a narrow vertical transect (near concave
shape of velocity curve); possible the case of PC profile. (b) A model with a lower position of Vmax, a higher relative velocity which produces more elongate AMS ellipsoids in the lower
flow zone and decreasing velocity along a wider vertical section (convex shape of velocity curve); possible the case of RM profile. (c) A model with position of Vmax similar to (b) but a
relative lower velocity producing less elongated AMS ellipsoids in the lower flow zone, decreasing velocity along an even wider vertical section; possibly the case of CU profile. (d) A model
on a perpendicular to flow plane showing effects of shortening and of a slight spreading flow movement induced by an irregular paleoslope dipping perpendicular to the main flow
direction; both effects produce plunging on axes along middle zone section and imbrication angle toward the spreading direction. Models generated with CizHet program for
heterogeneous deformation (Tolson, 2009).

observed in the other profiles. We speculate that the transverse to flow (Henry, 1980; Henry et al., 2003; Alva-Valdivia et al., 2005). In the
lineation of ellipsoids may correspond to a component of the spreading AMS diagrams of the present study, the flow plane is the equatorial
direction induced by irregularities of the paleoslope in a low velocity plane (horizontal plane). Hence the planes of symmetry are vertical.
(Fig. 12d). The best plane may include the K3 mean azimuth (pole of magnetic fo-
At some specific levels within the CU and RM profiles in the low- liation), the K1 mean azimuth, or both. The more clustered the AMS di-
ermost and topmost part of them (crosses and stars in Fig. 9b and c), rections are, the easier it is to define these planes, whereas the more
zones where chilling processes may influence the AMS and size of scattered the AMS directions are, the more difficult it is to define
magnetic minerals, there are vertical and perpendicular to flow indi- them. A scattered AMS distribution may be attributed to a less efficient
vidual foliations. Those orientations may be attributed to inverse flow-induced shear strain process (e.g., lower flow velocity, higher vis-
fabrics due to SD. Since SD was not always documented, we propose cosity, lower lava temperature, higher chilling rates, and/or paleoslope
two alternative explanations. (a) Dragging flow mechanism because conditions).
of a higher friction along the thin basal subzone (see less oblate or The dominant AMS imbrication angle obtained from all profiles
even prolate AMS shapes, crosses in Fig. 10b and c). (b) Degassing points toward the NE and ranges from 7° to 21° (considering both K1
and elutriation process, similarly as Gountié Dedzo et al. (2011) and K3 means). Azimuths from local flow inferences are 38°, 62°, and
have suggested for vertical AMS fabrics observed in ignimbrites; 54° for the CU, RM, and PC profiles, respectively (Table 3, Fig. 9). This
the clear basal vesicular structure horizon that includes small pipes agrees with the general flow direction of distal facies of BCU and BPSA
observed in the RM and CU profiles and not in the PC (Fig. 2) may observed in aerial photographs (Fig. 1). This dominant imbrication
favor this possibility. agrees well with the “positive” imbrication of the lower AMS zone. Not-
withstanding positive imbrication is dominant, within the middle-
8.3. Flow directions upper part of profiles, imbrication is mostly opposed (Figs. 9 and 10)
as reported in other AMS studies on lava flows (Cañón-Tapia et al.,
General flow direction of the lava flow units from the Xitle volcano 1996, 1997; Bascou et al., 2005).
can be seen in the aerial photographs represented in Fig. 1 (Delgado Within the CU profile, the possible inferred component of spreading
et al., 1998). Flow directions inferred from our AMS study within each direction suggested from the AMS ellipsoid orientations in the middle
profile agree well with the observed general flow. zone, would correspond to the best vertical plane of symmetry: be-
Flow direction can be inferred from the AMS ellipsoid imbrication tween S21W and S67W (data taken from corresponding equatorial dia-
angles relative to the flow plane. The simplest way to define the imbri- grams on Fig. 9c), whereas the general flow direction in this unique case
cation angle from a collection of AMS directions is to find the best plane would correspond to the other vertical plane of symmetry and would be
of symmetry of the axes distribution perpendicular to the flow plane toward N21W and N67W.
C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28 27

9. Conclusions and partially eroded lava flows and inferring more reliable flow direc-
tions. Our study has shown that the lava zone recognition is important
The within-flow vertical variation of AMS for three basaltic flow pro- for elucidating flow dynamics and flow sources and can be helpful for
files from the Xitle volcano were investigated and discussed in relation selecting the part of the flow most advantageous for performing other
to heterogeneous deformation due to the lava flow-induced shear experimental paleomagnetic surveys such as paleointensity studies.
strain. A lower positive and an upper opposed imbrication zone separat- Supplementary data to this article can be found online at http://dx.
ed by an intermediate zone of lowest degrees of anisotropy were dis- doi.org/10.1016/j.jvolgeores.2016.01.003.
criminated. Opposed imbrication has been reported in numerous
works (e.g., Loock et al., 2008) and the zonation of both imbrications
Acknowledgments
has been noticed in earlier works where individual flows have been
clearly distinguished (e.g., Cañón-Tapia et al., 1995, Walker et al.,
This work was supported by DGAPA-UNAM grant PAPIIT-IN107114.
1999, Cañón-Tapia et al., 1996, Cañón-Tapia, 2004).
We thank Gustavo Tolson for letting us use his “toy” CizHet program for
We present strain ellipse distribution models for different flow ve-
heterogeneous deformation, v.1.0 created for his structural geology lec-
locities assuming similar viscosity, which represent good approxima-
tures since 2009, the program is available at his web page (http://www.
tions for the observed AMS fabrics within the basaltic flows. Proposed
geologia.unam.mx/igl/index.php/tolson-g) or can be requested to the
models explain that the vertical variation rate of flow velocity and the
corresponding author e-mail (cecilia@geofisica.unam.mx). Thanks to
location of Vmax are the main factors that control the formation of the
M. Espinosa and V. H. Macías for measurements of magnetic properties
identified AMS zones, since they control induced shear strain sense
and to M. L. Rivas-Sanchez for the microscopy observations. We ac-
and intensity. Models indicate that the lower AMS zone corresponds
knowledge the useful comments of E. Herrero-Bervera and of the two
to the zone below the plane of maximum flow velocity (Vmax), the
reviewers (B. Henry and anonymous). LA thanks DGAPA-UNAM grant
upper zone to the region above it, and the middle zone to the region
for his sabbatical stay in 2014 at KIT Germany.
around the Vmax plane. Identification of the AMS flow zones may be
useful to assess the vertical location of the maximum flow velocity.
We inferred that the higher Vmax corresponds to the PC and RM pro- References
files, that the higher vertical variation rate of flow velocity is within
Alva-Valdivia, L.M., 2005. Comprehensive paleomagnetic study on a succession of Holo-
the PC profile, that the lowest Vmax is within the CU profile where the
cene olivine-basalt flow: Xitle volcano (Mexico) revisited. Earth Planets Space 57,
vertical variation rate of flow velocity is also the lowest. 839–853.
The lower AMS zone is the most conspicuous and shows the most Alva-Valdivia, L.M., Rosas-Elguera, J., Bravo, T., Urrutia-Fucugauchi, J., Henry, B., Caballero,
C., Rivas, M., Goguitchaichvili, A., Lopez, H., 2005. Paleomagnetic and magnetic fabric
common features in all places. It shows positive imbrication, better-
studies of the San Gaspar ignimbrite, western Mexico- constraints on emplacement
clustered AMS directions, and higher anisotropy degree values along mode and source vents. J. Volcanol. Geotherm. Res. 147, 68–80.
each profile. This is the most reliable zone for inferring flow directions. Audunson, H., Levi, S., Hodges, F., 1992. Magnetic property zonation in a thick lava flow.
It may be broadly recognized because it fits almost perfectly with the J. Geophys. Res. 97 (B4), 4349–4360.
Bascou, J., Camps, P., Dautria, J.M., 2005. Magnetic versus crystallographic fabric in a basal-
compact structure zone of flows where almost no vesicles were formed. tic lava flow. J. Volcanol. Geotherm. Res. 145, 119–135.
Our observations suggest that the relative thickness of this zone is Böhnel, H., Morales, J., Caballero, C., Alva, L., McIntosh, G., González, S., Sherwood, G., 1997.
thicker in flows documented as less viscous (PC profile; Fig. 9a). Variation of rock magnetic parameters and paleointensities over a single Holocene
lava flow. J. Geomagn. Geoelectr. 49, 523–542.
The middle and upper zones show more variation in AMS directions Boiron, T., Bascou, J., Camps, P., Ferré, E.C., Maurice, C., Guy, B., Gerbe, M.-C., Launeau, P.,
and parameters in all profiles and between them. The most scattered 2013. Internal structure of basalt flows: insights from magnetic and crystallographic
AMS directions and lowest anisotropy degrees are observed within fabrics of the La palisse volcanics, French massif central. Geophys. J. Int. 193, 585–602.
http://dx.doi.org/10.1093/gji/ggs115.
the middle zone. This zone is the least reliable for inferring main flow di- Buddington, A.F., Lindsley, D.H., 1964. Iron-titanium oxide minerals and synthetic equiv-
rection. It corresponds to the compact structure that includes some alents. J. Petrol. 5, 310–357.
scattered vesicles and to the zone of nearly none shear strain. When Caballero-Miranda, C.I., Alva-Valdivia, L.M., Torres-Hernández, J.R., 2009. Anisotropy of
magnetic susceptibility of the cantera ignimbrite, San Luis Potosi, México: source de-
this zone is thinner, or absent, as documented in this work, it is consis-
lineation. Earth Planets Space 61, 173–182.
tent with a more fluid lava flow such as observed in the PC profile, with Cagnoli, B., Tarling, D.H., 1997. The reliability of anisotropy of magnetic susceptibility
a higher absolute maximum velocity and apparently a higher vertical (AMS) data as flow direction indicators in friable base surge and ignimbrite deposits:
Italian examples. J. Volcanol. Geotherm. Res. 75, 309–320.
decreasing velocity rate (from Vmax). The latter condition is consistent
Callot, J.P., Guichet, X., 2003. Rock texture and magnetic lineation in dykes a simple ana-
with a very narrow zone of minimum strain (Fig. 12a); this relationship lytical model. Tectonophysics 366, 207–222. http://dx.doi.org/10.1016/S0040-
would explain the absence of the middle AMS zone in the PC profile. Fol- 1951(03)00096-9.
lowing this idea, a wider middle AMS zone agrees with a wider zone of Cañón-Tapia, E., 2004. Flow direction and magnetic mineralogy of lava flows from the
central parts of the peninsula of Baja California, Mexico. Bull. Volcanol. 66, 431–442.
minimum induced strain. In our models, a wider middle zone is associ- Cañón-Tapia, E., Coe, R., 2002. Rock magnetic evidence of inflation of a flood basalt lava
ated to lower rate of change of flow velocity (Fig. 12c), and that seems to flow. Bull. Volcanol. 64, 289–302. http://dx.doi.org/10.1007/s00445-002-0203-8.
occur in the CU profile, which is located at the border of the flow lobe Cañón-Tapia, E., Walker, G.P.L., 1998. Mecanismo de emplazamiento de las mesetas
basálticas gigantes continentales estudiado a través de mediciones de anisotropía
where flow velocity should decrease at the time of emplacement. de susceptibilidad magnética. GEOS 18, 2–10.
The upper zone records mainly opposed imbrication; the AMS direc- Cañón-Tapia, E., Walker, G.P.L., Herrero-Bervera, E., 1995. Magnetic fabric and flow direc-
tions are better clustered compared with the middle zone. It may show tion in basaltic pahoehoe lava of xitle volcano, México. J. Volcanol. Geotherm. Res. 65,
249–263.
not only opposed but also positive imbrication; these may be due to Cañón-Tapia, E., Walker, G.P.L., Herrero-Bervera, E., 1996. The internal structure of lava
flow velocity fluctuations partly perturbed by formation of a semisolid flows—insights from AMS measurements I: near-vent a'a. J. Volcanol. Geotherm.
phase and formation of solidified crust, all of which may alter induced Res. 70, 21–36.
Cañón-Tapia, E., Walker, G.P.L., Herrero-Bervera, E., 1997. The internal structure of lava
shear strain sense or intensity. This zone may be useful for inferring
flows—insights from AMS measurements II: Hawaiian pahoehoe, toothpaste lava
flow direction if considering the opposed imbrication. It may be unam- and 'a'a. J. Volcanol. Geotherm. Res. 76, 19–46.
biguously identified in the field because it includes most of the vesicle- Cashman, K.V., Kauahikaua, J.P., 1997. Reevaluation of vesicle distributions in basaltic lava
flows. Geology 25 (5), 419–422.
structure in the upper part of the flows and exhibits the highest density
Cordova, C., Martin del Pozzo, A.L., López-Camacho, J., 1994. Palaeolandforms and volcanic
content of vesicles with the highest aspect ratios. However, their bound- impact of the environment of prehistoric cuicuilco, southern Mexico city. J. Archaeol.
aries do not exactly fit with the thickness of the vesicular structure of Sci. 21, 285–596.
the flow. Day, R., Fuller, M., Schmidt, V.A., 1977. Hysteresis properties of TM: grain-size and compo-
sitional dependence. Phys. Earth Planet. Inter. 13, 181–190.
Identification of AMS zones in a sequence of lava flows is a useful Delgado, H., Molinero, R., Cervantes, P., Nieto-Obregón, J., Lozano, R., Macías González, H.,
tool to identify single flows, particularly useful when studying older Mendoza-Rosales, C., Silva-Romo, G., 1998. Geology of xitle volcano in southern
28 C.I. Caballero-Miranda et al. / Journal of Volcanology and Geothermal Research 311 (2016) 9–28

Mexico city: a 2000 year old monogenetic volcano in an urban area. Revista Mexicana MacDonald, W.D., Ellwood, B.B., 1987. Anisotropy of magnetic susceptibility: sedimento-
de Ciencias Geológicas 15 (2), 115–131. logical, igneous and structural-tectonic applications. Rev. Geophys. 25, 905–909.
Dunlop, D.J., 2002. Theory and application of the Day plot (Mrs/Ms versus Hcr/Hc): 1. MacDonald, W.D., Palmer, H.C., 1990. Flow directions in ashflow tufts: a comparison of
Theoretical curves and tests using titanomagnetite data. J. Geophys. Res. 107 (B3), geological and magnetic susceptibility measurements, Tshirege Member (upper
2056. http://dx.doi.org/10.1029/2001JB000486. Bandelier tuff), Valles caldera New Mexico, USA. Bull. Volcanol. 53, 45–59.
Dunlop, D.J., Özdemir, Ö., 1997. Rock Magnetism: Fundamentals and Frontiers Cambridge Martin del Pozzo, A.L., 1982. Monogenetic vulcanism in the Sierra Chichinautzin. Mexico,
Univ. 1997. Press, Cambridge, UK (573p). Bulletin of Volcanology 41-9, pp. 9–23.
Ellwood, B.B., 1978. Flow and emplacement direction determined for selected basaltic Martin del Pozzo, A.L., Espinasa, R., Lugo, J., Barba, L., Lopez, J., Plunkett, P., Uruñuela, G.,
bodies using magnetic anisotropy measurements. Earth Planet. Sci. Lett. 41, 254–264. Manzanilla, L., 1997. Volcanic impact in Central Mexico. Excursion Guide, IAVCEI
Ellwood, B.B., 1980. Induced and remanent magnetic properties of marine sediments as Puerto Vallarta 1997 (31 pp).
indicators of depositional process. Mar. Geol. 38, 233–344. Martín-Hernández, F., Lüneburg, C.M., Aubourg, C., Jackson, M., Aubourg, C., 2004. Mag-
Ellwood, B.B., 1982. Estimates flow directions for calk-alkaline welded tuffs and paleo- netic fabric: methods and applications- an introduction, p. 1–7. In: Martín-
magnetic data reliability from anisotropy of magnetic susceptibility measurements: Hernández, F., Lüneburg, C.M., M., J. (Eds.), Magnetic Fabric: Methods and Applica-
Central San Juan mountains, southwest Colorado. Earth Planet. Sci. Lett. 59, 303–314. tions. Geological Society of London, Sp. Publ. 238 560 pp.
Engelmann, R., Kontny, A., Lattard, D., 2010. Low-temperature magnetism of synthetic Fe– Merle, O., 1998. Internal strain within lava flows from analogue modeling. J. Volcanol.
Ti oxide assemblage. J. Geophys. Res. 115, B12107. http://dx.doi.org/10.1029/ Geotherm. Res. 81, 189–206.
2010JB000865. Morales, J., Alva-Valdivia, L.M., Goguitchaichvili, A., Urrutia-Fucugauchi, J., 2006. Cooling
Eriksson, P.I., Riishuu, M.S., Sigmundsson, F., Elming, S.-Å., 2011. Magma flow directions rate corrected paleointensities from the xitle lava flow: evaluation of within-site scat-
inferred from field evidence and magnetic fabric studies of the Streitishvarf compos- ter for single spot-reading cooling units. Earth Planets Space 58, 1341–1347.
ite dike in east Iceland. J. Volcanol. Geotherm. Res. 206, 30–45. http://dx.doi.org/10. Morales-Contreras, J.J., 1995. Determinación de Paleointensidades del Campo
1016/j.jvolgeores.2011.05.009. Geomagnetico Para el Cuaternario en la Sierra Chichinautzin. Universidad Nacional
Espinasa, P.R., 1992. Tubos de lava y formas asociadas en el Pedregal del Xitle. Proceedings Autónoma de México, Mexico, M.S. thesis.
XII Geographical National Congress. Mexican Society of Geography and Statistics, Ort, M.H., Porreca, M., Geissman, J.W., 2015. The use of palaeomagnetism and rock mag-
Aguascalientes, I, pp. 21–30. netism to understand volcanic process: introduction, p. 1–11. In: Ort, M.H., Porreca,
Gonzalez, S., Sherwood, G., Böhnel, H., Schnepp, E., 1997. Palaeosecular variation in cen- M., Geissman, J.W. (Eds.), The Use of Palaeomagnetism and Rock Magnetism to Un-
tral Mexico over the last 30,000 years. Geophys. J. Int. 130, 201–219. derstand Volcanic Process. Geological Society of London, Sp. Publ. 396 281 pp.
Gonzalez, S., Pastrana, A., Siebe, C., Duller, G., 2000. Timing of the prehistoric eruption of Ortega, B., Urrutia-Fucugauchi, J., Nieto, J., 1993. Geología y Edades de C-14 del Derrame
Xitle volcano and the abandonment of Cuicuilco pyramid, southern basin of Mexico. del Pedregal de San Angel. Universidad Nacional Autónoma de México, Mexico City,
Geol. Soc. Lond., Spec. Publ. 171, 205–224. Technical Rep. Instituto de Geofísica.
Gountié Dedzo, M., Nédélec, A., Nono, A., Njanko, T., Font, E., Kamgang, P., Njonfang, E., Palmer, H.C., MacDonald, W.D., Gromme, C.S., Ellwood, B.B., 1996. Magnetic properties
Launeau, P., 2011. Magnetic fabrics of the Miocene ignimbrites from west- and emplacement of the bishop tuff, California. Bull. Volcanol. 58, 101–116.
Cameroon: implications for pyroclastic flow source and sedimentation. J. Volcanol. Peterson, D.W., Holcomb, R.T., Tilling, R.I., Christiansen, R.L., 1994. Developments of lava
Geotherm. Res. 203 (3–4), 113–132. tubes in the light of observations at Mauna Ulu, Kilauea volcano, Hawaii. Bull.
Graham, J.W., 1954. Magnetic anisotropy, an unexploited petrofabric element. Geol. Soc. Volcanol. 56, 343–360.
Am. Bull. 65, 1257–1258. Robin, P.-F., Jowett, E.C., 1986. Computerized density contouring and statistical evaluation
Haggerty, S. E., 1976. Oxidation of opaque mineral oxides in basalts. In Oxide Minerals Short of orientation data using counting circles and continuous weighting functions.
Course Notes, edited by D. Rumble III, 2nd ed., Mineral Society of America 3, 1–10. Tectonophysics 121, 207–223.
Henry, B., 1980. Contribution à l'étude des propriétés magnétiques de Roches Rochette, P., Jackson, M., Aubourg, C., 1992. Rock magnetism and the interpretation of an-
magmatiques des alpes: conséquences structurales, régionales et générales. Trav isotropy of magnetic susceptibility. Rev. Geophys. 30, 209–226.
lab tectonophysique Paris. CRE 80 (07), 1–528. Sahagian, D.L., 1985. Bubble migration and coalescence during the solidification of basal-
Henry, B., Plenier, G., Camps, P., 2003. Post-emplacement tilting of lava flows inferred tic lava flows. J. Geol. 93 (2), 205–211. http://dx.doi.org/10.1086/628942.
from magnetic fabric study: the example of Oligocene lavas in the Jeanne d'arc pen- Siebe, C., 2000. Age and archaeological implications of xitle volcano, southwestern basin
insula (Kerguelen islands). J. Volcanol. Geotherm. Res. 127, 153–164. of Mexico city. J. Volcanol. Geotherm. Res. 104, 45–64.
Herrero-Bervera, E., Cañón-Tapia, E., Walker, G.P.L., Tanaka, H., 2002. Magnetic fabrics Tarling, D.H., Hrouda, F., 1993. The Magnetic Anisotropy of Rocks. Chapman & Hall,
study and inferred flow directions of lavas of the old Pali road, OPahu, Hawaii. London (217 p).
J. Volcanol. Geotherm. Res. 118, 161–171. Tolson, G., 2009. Programa CizHet v. 1.0 (CizHet program for heteregenous deformation,
Hon, K., Kauahikaua, J., Denlinger, R., Mackay, K., 1994. Emplacement and inflation of v.1.0), Instituto de Geología. Universidad Nacional Autónoma de México, México.
pahoehoe sheet flows: observations and measurements of active lava flows on Kilau- Urrutia-Fucugauchi, J., 1996. Palaeomagnetic study of the Xitle–Pedregal de San Angel
ea volcano, Hawaii. Geol. Soc. Am. Bull. 106, 351–370. lava flow, southern basin of Mexico. Phys. Earth Planet. Inter. 97, 177–196.
Incoronato, A., Addison, F.T., Tarling, D.H., Nardi, G., Pescatore, T., 1983. Magnetic fabric in- Urrutia-Fucugauchi, J., Martin del Pozzo, A.L., 1993. Implicaciones de los datos
vestigations of pyroclastic deposits from phlegrean fields, Southern Italy. Nature 306, paleomagnéticos sobre la edad de la sierra de chichinautzin, Cuenca de México.
461–463. Geofisica Internacional 32, 523–533.
Jelínek, V., 1978. Statistical processing of anisotropy of magnetic susceptibility measured Walker, G.P.L., 1991. Structure and origin of vesicle by injection of lava under surface crust
on groups of specimens. Stud. Geophys. Geod. 22, 50–62. of tumuli, “lava rises”, lava-rise pits and lava inflation clefts in Hawaii. Bull. Volcanol.
Jelínek, V., 1981. Characterization of the magnetic fabric of rocks. Tectonophysics 79, 53, 546–558.
T63–T67. Walker, G.P.L., Cañón-Tapia, E., Herrero-Bervera, E., 1999. Origin of vesicle layering and
Kodama, K.P., 1995. Magnetic Fabrics. Reviews in Geophysics 33, Supplement, 1991– double imbrication by endogenous growth in the Birkett basalt flow (Columbia
1994. river plateau). J. Volcanol. Geotherm. Res. 88, 15–28.
Kolofíková, O., 1976. Geological interpretations of measurement of magnetic properties of White, S.E., Reyes, M., Ortega, J., Valastro, S., 1990. El ajusco: geomorfología volcánica y
basalts. An example of the Chribsky les lava flow of the velky roudny volcano (nízký acontecimientos glaciales Durante el pleistoceno superior y comparación con las se-
jeseník Mts). Cas. Miner. Geol. 21, 287–348. ries glaciales mexicanas y las montañas rocallosas. Colec. Cientif., INAH, México
Le Pennec, J.L., Yan, C., Diot, H., Froger, J.l., Gourgaud, A., 1998. Interpretation of anisotropy City. 212 pp. 1–77.
of magnetic susceptibility fabric of ignimbrites in terms of kinematic and sedimento- Wolff, J.A., Ellwood, B.B., Sachs, S.D., 1989. Anisotropy of magnetic susceptibility in welded
logical mechanisms: an Anatolian case-study. Earth Planet. Sci. Lett. 157, 105–127. tuffs: application to a welded-tuff dike in the tertiary trans-Pecos Texas volcanic
Loock, S., Diot, H., Van Wyk de Vries, B., Launeau, P., Merle, O., Vadeboin, F., Petronis, M.S., province, USA. Bull. Volcanol. 51, 299–310.
2008. Lava flow internal structure found from AMS and textural data: an example in
methodology from the Chaîne des Puys, France. J. Volcanol. Geotherm. Res. 177,
1092–1104.

View publication stats

You might also like