Вы находитесь на странице: 1из 21

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elseviers archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright

Author's personal copy


Tectonophysics 566-567 (2012) 6786

Contents lists available at SciVerse ScienceDirect

Tectonophysics
journal homepage: www.elsevier.com/locate/tecto

Inuence of alteration on physical properties of volcanic rocks


Antonio Pola a,, Giovanni Crosta b, Nicoletta Fusi b, Valentina Barberini b, Gianluca Norini c
a
b
c

Instituto de Geofsica, Universidad Nacional Autnoma de Mxico, Campus Morelia-UNAM, Antigua carretera a Ptzcuaro 8701, 58190 Morelia, Michoacn, Mexico
Dip. di Scienze Geologiche e Geotecnologie, Universit degli Studi di Milano-Bicocca, Piazza della Scienza 4, 20126 Milano, Italy
Consiglio Nazionale delle Ricerche (IDPA), via Mario Bianco 9, 20133 Milano, Italy

a r t i c l e

i n f o

Article history:
Received 13 May 2011
Received in revised form 21 July 2012
Accepted 21 July 2012
Available online 1 August 2012
Keywords:
Volcanic rocks
Alteration
Physical properties
Porosity
Pulse velocity
X-ray tomography

a b s t r a c t
Physical properties of some weathered/altered volcanic rocks and their variation with the degree of alteration are described in detail. A series of tests was performed to identify and quantify the progressive degradation of the properties: 1) petrographycal and chemical studies; 2) effective (e) and total porosity (t)
measurements and 3D pore reconstruction 3) ultrasonic pulse velocity and spatial attenuation (s)
measurements.
Four different volcanic lithologies have been tested: i) trachytic lava with abundant crystals; ii) pyroclastic deposits,
with lava clasts and pumice elements; iii) green tuff, made prevalently of pumice clasts; and iv) non-welded ignimbrite deposits. Chemical indices of weathering (CIW) reveal large differences (42.73b Chemical Index of Alteration
[CIA]b 69.24) not only between lithotypes, but also between samples. These differences are reected by physical
properties, in particular t (6.0b t from X-ray tomography imagesb 49.8%), e (11.0b e from mercury
porosimeterb 65.0%) and shear wave velocity values (0.50b Vs b 2.90 km/s). Pore network evolution with alteration
for each lithology is well documented by fractal dimension (D) and s. Mean values of porosity are strictly related to
P and S wave velocity (Vp and Vs) and the degree of alteration.
Values of CIA are well correlated with the degradation trend exhibited by measured physical properties. The
combination of techniques to measure the t and e provides a good estimate of grain size and pore size distribution and rock structure. Defects and particular characteristics in the rock sample (e.g. micro-fractures,
voids, cavities and orientation and sizes of certain minerals and clasts) are revealed by s values: the smaller
the s the more homogeneous and less altered is the sample.
2012 Elsevier B.V. All rights reserved.

1. Introduction
Volcanic rocks exhibit complex behaviors ranging from hard to
extremely soft rocks, depending on mineralogy, cementation, porosity,
and degree of alteration. The study of these rocks is often problematic
because of the variability of the materials (e.g. individual lava ows, pyroclastic deposits, and interbedded units) and their heterogeneity and
the presence of abundant voids. These characteristics make difcult to
reach a representative characterization and understanding of the physical
and mechanical behavior. Volcanic rocks are frequently found in altered/
weathered conditions because of the highly active volcanic environment
and the presence of hydrothermal conditions. Generally, the strength,
deformability and stiffness of these rocks show a dependence on physical
and chemical changes due to alteration and weathering. In particular, porosity (e.g. size, shape, distribution and frequency) and ultrasonic parameter values could be related to the strength and deformability of the rocks
(Ceryan et al., 2008; Rotonda et al., 2010; Sousa et al., 2005; Vinciguerra
et al., 2009). Then the quantication of these physical parameters is

Corresponding author. Tel.: +39 2 6448 2029; fax: +39 2 64484273.


E-mail address: antonio_pola@yahoo.com.mx (A. Pola).
0040-1951/$ see front matter 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tecto.2012.07.017

fundamental to evaluate and interpret rock behavior observed in situ


and during laboratory tests.
Porosity can be formed by voids, between grains or minerals,
interconnected or disconnected, of different size and shape, with a
particular frequency distribution of sizes. Volcanic rocks often present
a brecciated, porous or vesicular texture characterized by abundant
vesicles and pores, of different sizes, sometimes lled with secondary
minerals (Al-Harthi et al., 1999; Hudyma et al., 2004; Tillerson and
Nimick, 1984). Several researchers investigated the physical and
mechanical behavior of rocks as a function of their porosity (Al-Harthi
et al., 1999; Avar and Hudyma, 2007; Avar et al., 2003; Aversa and
Evangelista, 1998; Hudyma et al., 2004; Luping, 1986; Nimick, 1988;
Price et al., 1994). Results suggest that the compressive strength is controlled by total porosity t, the abundance of macro-pores and its structure, size distribution and the type of forming particles.
Several techniques exist to evaluate and characterize porosity
(e.g. thin-section image analysis, pycnometer, mercury porosimeter,
computerized X-ray tomography). Conventional image analysis on
thin-sections does not retain the complete 3D information. Quantitative data obtained by 2D images can be transformed into 3D
petrophysical properties, using stereological corrections. These techniques showed satisfactory results in petrophysical studies (Higgins,

Author's personal copy


68

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

1994, 2000; Morgan and Jerram, 2006; Proussevitch et al., 2010;


Sahagian and Proussevitch, 1998), even if the shape of the grains/
voids has to be necessary assumed (Gualda and Rivers, 2006). A more
complete three dimensional reconstruction of rock heterogeneities
and porosity derives from computerized X-ray tomography. The quality
of the nal results depends on the mode of image acquisition and image
processing (e.g. noise reduction, ltering, thresholding, and particle separation steps) according to Ashbridge et al., 2003; Geiger et al., 2009;
Gualda and Rivers, 2006; Ketcham, 2005; Mess et al., 2003; Taud et al.,
2005. Mercury porosimetry has been used to characterize porosity
(Metz and Knfel, 1992; Roels et al., 2001; Wardlaw and McKellar,
1981) and some aspects of uid distribution.
Pore structure, rock texture, fracturing and mineralogical changes
and their relations to physical and mechanical properties have been
studied and quantied by measuring ultrasonic P and S wave velocity
(Vp, Vs, s and temporal attenuation [t]) (Martnez-Martnez et al.,
2007, 2011; Patonin, 2003; Sousa et al., 2005; Stanchits et al., 2003;
Vinciguerra et al., 2009; Zamora et al., 1994). Sousa et al. (2005)
describe the inuence of alteration on granitic rocks (e.g. effective
porosity e, uniaxial compressive strength [UCS], thermal shock, linear crack density, and Vp), concluding that the variations in Vp are
strictly correlated to the increment of the width of micro-fractures,
which in turn depends on the degree of alteration. Stanchits et al.
(2003) present laboratory measurements of changes in Vp and its
attenuation during the injection of water into a granite sample as it
was loaded to failure. They show that Vp and its attenuation are very sensitive to the process of opening of micro-cracks and their subsequent
resaturation. Recently, Martnez-Martnez et al. (2011) present an evaluation of carbonatic rock quality by comparing the sensitivity of ultrasonic
parameters (Vp, Vs, Velocity ratio [Vp/Vs], waveform energy [], t and s).
The authors demonstrate that among these parameter, s is the most
sensitive parameter to identify the rock defects (e.g. ssures and voids).
They enclose the values of s between two limits: 0 dB/cm, when the
wave is non-attenuated and 40 dB/cm, when the wave is completely
scattered. Alteration is represented by values of s higher than 12 dB/cm.
The evaluation of physical-mechanical characteristics is very important in the geological-geotechnical context. The progressive degradation
of the materials is an important factor at inducing volcanic ank collapses
(e.g. Capra and Macas, 2002, Finn et al., 2001, 2007; Reid et al., 2001).
Finn et al. (2001) describe that the evaluation of the hazards associated
with this factor is very difcult, principally because of the lack of knowledge with regard to typical physical-mechanical processes that could
generate volcanic collapse by progressive alteration of the materials.
The understanding of the behavior of altered volcanic rocks requires
a detailed geotechnical characterization, in which the physical and mechanical characteristics and their evolution during alteration have to be
included. The aim of this work is to provide a detailed description of the
evolution of rock alteration in three different volcanic environments.
Experimental campaign and laboratory tests for the mineralogical
petrographical and geomechanical characterization of the materials
are based on sampling of various altered volcanic rocks (e.g. lava, pyroclastic deposits, tuff, and non-welded ignimbrite).
Rock samples consist of trachytic lava and pyroclastic rock, characterized by different degrees of alteration. They were collected in three
different Italian volcanic environments: Solfatara crater, Ischia Island
and Bolsena volcanic zone. These zones present excellent settings to
study the evolution of physical-mechanical properties (e.g. geometry
and topology of pore network and grain size and shape) in altered
volcanic rocks (Fig. 1ad).
1.1. Geological description
1.1.1. Solfatara
Phlegraean Fields caldera is an active volcano affected by at least
two collapses. The youngest collapse is related to the eruption of
the Neapolitan Yellow Tuff (NYT) (12 ka) (Civetta et al., 1997; Orsi

et al., 1992, 1995). The volcanism younger than the NYT has been
characterized by several explosive events which generated a large number of cones and craters (Di Vito et al., 1999; Orsi et al., 1996), where the
crater of Solfatara was included (4.1 and 3.8 kyr BP) (Di Vito et al., 1999).
Solfatara crater is made up of phreatomagmatic breccia overlain by
pyroclastic-ow deposits. Cipriani et al. (2008) suggest that eruptive activity in the Solfatara changed from phreatomagmatic to magmatic repeatedly throughout the eruption. The juvenile fraction of pyroclastic
series is trachytic, similarly to the eruptions of Phlegraean Fields (Fig. 1a).

1.1.2. Ischia
The island of Ischia, located 35 km west of the Gulf of Naples
(Fig. 1b), is an active volcanic area belonging to the Neapolitan volcanic region (Cipriani et al., 2008; Orsi et al., 1992). Stratigraphic studies and radiometric dating indicate four phases of activity (e.g. Chiesa
et al., 1986; Civetta et al., 1991; Gillot et al., 1982; Orsi et al., 1992).
Sbrana et al. (2010) show that after the second phase (7356 ka) intense explosive volcanic activity occurred with numerous ignimbritic
eruptions that led to the formation of a caldera. The last calderaforming eruption originated the green tuff, a welded pyroclastic ow
deposit emplaced 5655 ka in the Mt Epomeo area. The green tuff
was followed by a third phase of explosive and effusive eruptions
(5520 ka) (Chiesa et al., 1986; Civetta et al., 1991; Gillot et al., 1982;
Orsi et al., 1992) and a last phase (10 ka1302 AD) which originated
lava and pyroclastic deposits (Orsi et al., 1991). The resurgence of
Mt Epomeo block occurred from 55 to 5 ka, exposing the Green
tuff deposit on the north-western ank of the Mt Epomeo (Sbrana
et al., 2010) (Fig. 1b). The Green tuff deposit is subjected to strong
hydrothermal alteration, visible along the Monte Epomeo fault
scarps.

1.1.3. Bolsena
Bolsena is part of the Vulsini volcanic District (VVD) (Beccaluva et
al., 1991) (Fig. 1c). The VVD is made up of four volcanic complexes
(Paleo-Bolsena, Bolsena, Monteascone and Latera) (Nappi et al., 1991),
characterized by one or more eruptive cycles. The Bolsena volcanic zone
(BVZ) is localized in the eastern part of VVD. Ten eruptive phases, characterized by a wide range of magma compositions, have been recognized in
this area (Nappi et al., 1998). The oldest products, attributed to Pliniantype activity, have been dated 5766.5 ka, but most of the BVZ volcanic
products are younger than about 400 ka (Nappi et al., 1995). The outcropping rocks of the study area are comprised in the last two eruptive phases.
The penultimate phase is characterized by several lava ows and scoria
cones (Nappi et al., 1998); the last phase is mostly represented by effusive
activity. Trachytic Plinian pumice fall and trachytic ignimbrite were
emplaced from a source located in NE sector of the Bolsena caldera
(Nappi et al., 1994).

2. Methods and theoretical background


We investigated the relationships between physical properties
(e.g. porosity, micro-structure and texture) and the degree of alteration
of volcanic rocks. Mineralogical and petrographical changes were examined by means of optical microscopy, X-ray powder diffraction (XRD)
and X-ray uorescence (XRF). The chemical indices of weathering (CIW)
were applied in order to identify chemical changes with alteration within
samples. Physical properties including unit weight (), apparent porosity
(), and ultrasonic pulse velocities (Vp, Vs and s) were determined, both
under saturated and dry conditions. t and e were obtained by bulk
weight and specic weight measurements, mercury porosimetry, thinsection images analysis, and computerized X-ray tomography image
analysis; pore distribution of each sample was analyzed by fractal dimension (D).

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

69

Fig 1. Localization map of eld study area of a) Solfatara. Some structural features are included in the map (modied after Isaia et al., 2004). b) Ischia. Green tuff deposit is indicated
by the green polygon; star indicates the sampling point c) Bolsena. Star indicates the sampling point (modied after Nappi et al., 1998). d) Illustrative map of Italy, where Fig. 1a, b
and c are located.

2.1. Sampling methodology and petrographical study


Fourteen rock blocks of different volcanic lithotypes (lava, pyroclastic rock deposits, tuff, and non-welded ignimbrite) were collected
during eld investigation. The quantity and size of blocks collected of
each lithotype depends on the accessibility and characteristics of the
respective outcrops. Tables 13 illustrate schematically the size and
the number of rock blocks collected. All samples were marked and
oriented in a vertical plane. Sampling of lava and Green Tuff units

was focused on covering the entire range of alteration. Five block


samples ( 50 50 15 cm in dimension) were collected in the lava
unit (SLA), one for each degree of alteration, while in the Green Tuff
unit four block samples (30 30 15 cm in dimension) were collected, two for fresh and two for highly altered lithotype. ; Three
block samples (606015 cm in dimension) were collected in the
pyroclastic unit, which presents the same degree of alteration, focusing on the grain size distribution (see Table 2). Finally, two
block samples ( 50 50 20 cm in dimension) were collected in

Author's personal copy


70

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

Table 1
Summary of eld description for Solfatara lava series.

Hand specimen
ID and alteration grade

Visual Effects
Identification description

Mineral composition

SLA1. Fresh lava: Grey coloured, fresh, and very


dense lava with porphyritic texture (euhedral
phenocrysts are plagioclase with an average size of
3 mm). Microcracks are not visible, some pores and
grains are bonded by oxidation. argilization is
present in the matrix and around minerals as blurred
stains.

Plg
+
+
+

Snd
+
+
+

Prx
+
+

Bio
+

Ox
+

arg

SLA2. Slightly altered lava: Grey coloured, slightly


stained, dense lava with trachytic texture (euhedral
phenocrysts are sanidine with an average size of 4
mm). Microcracks are visible, we can observe
argilization and oxidation along them and within the
boundary of minerals. Biotite is almost totally
replaced.

Plg
+
+

Snd
+
+
+

Prx
+

Bio
+

Ox
+

arg
+

SLA3. Moderately altered lava: Greenish,


yellowish, dense lava with preserved porphyritic
texture (major constituents are potassic feldspar
with =2 mm). Oxidation afects all crystals
boundaries whereas biotite and pyroxene are almost
completely replaced. Argilization is present into the
matrix and all around minerals as blurred stains.

Plg
+

Snd
+
+

Prx
+

Bio
+

Ox
+
+

arg
+
+

SLA4. Highly altered lava: reddish, yellowish lava.


Texture is not preserved. Although all minerals are
altered, sanidines prevail over pyroxenes and
plagioclases. Pores are presented along visible
cracks. Matrix is totally replaced by argilization.

Plg

Snd
+

Prx

Bio

Ox
+

arg
+
+
+

SLA5. Totally altered lava : White lava. texture is


not preserved. Complete alteration of all minerals
occurs, some pyroxenes could be recognized from
its partially-distorted geometry. Matrix is totally
replaced by argilization.

Plg
?

Snd
-

Prx
?

Bio

Ox

arg
+
+
+

Note. Plg = Plagioclase, Snd = Sanidine, Prx = Pyroxene, Bio = Biotite, Ox = Oxidation, Arg = Argilization. +++ = abundant, ++ = modest, + = few, + = rare, = absent, ? =
difcult or uncertain identication.

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

71

Table 2
Summary of eld description for Solfatara pyroclastic series.

Hand specimen
ID and alteration grade

Visual effects
Identification description

Grain sizes proportions (%)


visual estimation

SPRA1. Highly altered: White to yellowish


pyroclastic rock, composed prevalently by subangular glass fragments. Matrix is very dense, with
micro-fragments of glass and micro-crystals.
Oxidation appears all around the largest fragments
(pumice). cracks and Sub-rounded lithic fragments
of lava and pyroclastic rocks are common.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)

0
2
5
15
20
40

SPRA2. Highly altered: Gray or white discoloured.


It is friable, but contains relict texture from the
original unit. It is composed by subrounded
fragments of glass and very altered-lithics (trachytic
lavas >4 mm). Matrix is composed by medium sand
fragments of glass.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)

20
5
15
10
10
20

SPRA3. Highly altered: Gray or white discoloured


with yellow stains, wholly decomposed rock. It is
friable, but well sorted strata are easy to be
identified. Subrounded fragments of pumice and
very altered-lithics are common. In general, matrix
is composed by medium sand fragments of glass.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)

10
5
10
5
10
30

Note. Pe = Pebbles, Gr = Granules , CoS = Coarse sand, MeS = Medium sand, FiS = Fine sand, Si = silt. Visual estimation was performed following charts to aid the visual estimation
of modal proportions of grains and minerals (in Best, 2003).

non-welded ignimbrite, focusing on differences in the textural characteristics (e.g. grain-sizes distribution).
Initial description of the degree of alteration followed the BS standard methods (BS5930, 1999), which includes: fresh (F), slightly altered (SA), moderately altered (MA), highly altered (HA), completely
altered (CA) and residual soil (RS). Initial description of grain size was
performed by visual estimation of modal proportions of grains and minerals (Best, 2003).
Samples were described in terms of lithology, color, mineral composition, texture, structure and sampling site and they were grouped into
four different series: Solfatara lavas (SLA), with ve different degree of
alteration (Table 1); Solfatara pyroclastics (SPRA), with three different
pyroclastic facies and grain-sizes content (Table 2); Green Tuff ignimbrite (IGT) with a fresh and an altered facies (Table 3). Bolsena pyroclastics (BoPra) a non-welded ignimbrite deposit, with the same degree of
alteration (Table 4).

More than 100 cylindrical specimens were prepared (= 18 and


54 mm in diameter), by coring the sampled blocks. Coring of cylindrical
samples and all laboratory measurements (e.g. ultrasonic pulse velocity,
thin-section image analysis) were conducted on a vertically-oriented
plane, following the orientation plane marked in eldtrips. The number
of the cylindrical specimens for each lithotype depended on the size,
the geometry and the number of blocks collected. The information
concerning the number of specimens used in each test/study for each
lithotype are listed in Table 5.
Basic petrographical information, visual estimation of grain size,
sorting, porosity, mineral abundance and fabric were obtained by
the analyses of 22 thin-sections (2 for lithotype). Petrographical changes
among samples were studied in plane-polarized light (PPL) and in crosspolarized light (XPL) (Fig. 2). Observations were reinforced by 22 X-ray
powder diffraction (XRPD) (two analysis of each lithotype) and 11
X-ray uorescence (XRF) analyses (one analysis of each lithotype).

Author's personal copy


72

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

Table 3
Summary of eld description for Ischia green tuff series.

Hand specimen
ID and alteration grade

Visual effects
Identification description

Grain sizes proportions (%)


visual estimation

IGTF. Fresh green tuff: Greenish, composed


prevalently by glass. Curved and elongated pumice
fragments are very common. High percentage of
argilization and oxidation appears as small stains
into the largest fragments of pumice (friable). Subrounded lithic fragments of lava with 5 mm of
maximum size are very common.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)
ox
arg

15
10
20
15
5
30
+

IGTA. Highly altered: Reddish, composed


prevalently by glass. Matrix is very dense, large
pores seem filled. Sub-rounded lithic fragments of
lava are very altered, but crystals (sanidines and
biotites) are recognizable. High percentage of
argilization is present.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)
ox
arg

15
15
20
5
5
40
++
+++

Note. Pe = Pebbles, Gr = Granules , CoS = Coarse sand, MeS = Medium sand, FiS = Fine sand, Si = silt, Ox = Oxidation, Arg = Argilization, +++ = abundant, ++ = modest, + = few,
+ = rare, = absent. Visual estimation was performed by following charts to aid the visual estimation of modal proportions of grains and minerals (in Best, 2003).

2.2. Geochemistry
Geochemical characteristics of weathered/altered rocks provide
information about the degree of weathering and could be related to
the differences in physical-mechanical properties. The relationship
between chemical changes and weathering could be specied using
Chemical Weathering indices (CWI), which are principally based on
the assumption that distributions of chemical elements, as well as
loss on ignition content (LOI), are mainly regulated by the degree of
weathering (Duzgoren-Aydin et al., 2002).
Many different CWI, mostly expressed as molecular or weight percentage ratios of major elements, have been proposed for characterizing the weathering degree of different rocks (Duzgoren-Aydin and Aydin,
2003; Fedo et al., 1995; Harnois and Moore, 1988; Irfan, 1996; Kim and
Park, 2003; Nesbitt and Young, 1982; Ohta and Arai, 2007; Parker,
1970; Ruxton, 1968). Duzgoren-Aydin and Aydin (2006) suggest that
CWI should be used to evaluate chemical heterogeneity, rather than determining the weathering stage. The following CWI were adopted in our
study (Table 6):
a) Chemical index of alteration, CIA (Nesbitt and Young, 1982),
applied in the literature to identify weathering degree in granites
and basalts (Gupta and Rao, 2001) and to reconstruct climatic conditions (Bahlburg and Dobrzinski, 2011; Krissek and Kyle, 2001);
b) Alumina to calcium-sodium oxide ratio, ACN (Harnois and Moore,

1988), applied unsatisfactorily in pyroclastic rocks under subtropical


conditions (Duzgoren-Aydin et al., 2002); c) Silica to Alumina ratio or
Ruxton ratio, SA (Ruxton, 1968), applied satisfactorily to volcanic
and granitic rocks (Irfan, 1996, 1999), which assumes that Al2O3 remains immobile during weathering and that changes in the ratio reect the loss of silica; d) Plagioclase index of alteration, PIA (Fedo et
al., 1995), obtained by excluding K2O molecular concentration
from CIA formula and used when the weathering of low-K crystals
such as plagioclase needs to be monitored (Btard et al., 2009);
e) Chemical index of weathering, CIW (Harnois, 1988) identical
to the CIA index, except for the K2O content that is neglected. It
has been applied satisfactory to felsic metamorphic rock and to
hydrothermally altered faults (Price and Velbel, 2003). CIW together
with CIA and PIA are susceptible to subtle geochemical changes that
may occur in hydrothermal alterated environments. Authors suggest that these indices should be carefully applied to heterogeneous
proles, because they assume aluminium immobility.

Major oxides contents were used in the CIW formulae to evaluate


chemical changes along the rock series (Table 6). The major oxides
content (e.g. altered and non-altered) were obtained from whole-rock
XRF geochemical analyses, performed on one representative sample of
each degree of alteration for each lithology. The Oxides content
obtained in altered samples was normalized to those of unaltered

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

73

Table 4
Summary of eld description for Bolsena ignimbrite.

Hand specimen
ID and alteration grade

Visual effects
Identification description

Grain sizes proportions (%)


visual estimation

BoPRA. Slightly altered: Pinkish rock, composed


prevalently by small sub-angular silica-amorphous
minerals. Sub-rounded lithic fragments of lava and
juvenils are also common. The matrix is composed
by micro-fragments of glass and micro-crystals.
Oxidation is present along micro-cracks and within
the largest clasts. Argillization is presented within
the matrix.

Pe (>4 mm)
Gr (2-4 mm)
CoS (0.5-2 mm)
MeS (0.25-0.5 mm)
FiS (0.06-0.25 mm)
Si (0.06> mm)
ox
arg

0
5
30
5
20
30
++
+++

Note. Pe = Pebbles, Gr = Granules , CoS = Coarse sand, MeS = Medium sand, FiS = Fine sand, Si = silt, Ox = Oxidation, Arg = Argilization, +++ = abundant, ++ = modest, + = few,
+ = rare, = absent. Visual estimation was performed by following charts to aid the visual estimation of modal proportions of grains and minerals (in Best, 2003).

Total porosity (t) was obtained indirectly by pycnometer tests, following the standard test procedures (Germaine and Germaine, 2009).
t was also obtained by two-dimensional digital image analysis. Digital images were obtained by scanning 22 thin-sections at 600 dpi resolution (2 images for each lithotype) (Table 5), using a photogrammetric
scanner, with a color background. Image segmentation, which consists
in pores identication, denition and differentiation from minerals or
other clasts, was carried out. This step was supported and conrmed
by thin section analysis and petrographical description (mineral constituents, rock texture, crystal size range, and porosity types). Final
processing included a more accurate image segmentation and image
calibration (ImageJ code). Once each pore has been identied (Fig. 3),
porosity shape parameters (i.e. location, perimeter, surface area, circularity and aspect ratio) have been automatically extracted.
A complete three dimensional reconstruction of rock structure and
porosity was obtained from a set of contiguous two dimensional X-ray
computerized tomography images (CT) (Fig. 4a). 76 cylindrical samples
(Table 5) were scanned by means of a GE D-600 medical CT hybrid

samples (xAltered/xfresh). If this ratio is less or greater than 1, the element


is considered leached out or xated within the rock system, respectively
(Duzgoren-Aydin and Aydin, 2006).
2.3. Total and effective porosity
Effective porosity (e) was initially obtained following the procedure recommended by ISRM (1972), which consists in calculating
dry density (dry) and saturated density (w) of cylindrical core samples.
In order to measure the degree of saturation and water absorption, 82 cylindrical specimens ( 54 mm, h 130 mm) from 11 different lithotypes
(Table 5) were submerged in distilled water under a constant air vacuum
pressure.
e was also determined by mercury intrusion porosimetry. The effective porosity and related characteristics (e.g. size frequency, connected
porosity, pore volume, total pore area, and bulk density) were determinate by a Pascal 140/240 Thermo-Fisher mercury porosimeter on 33 cubical specimens (3 specimens for each lithotype) (Table 5).

Table 5
Summary of samples used in laboratory tests.
Cylindrical specimens

3D (XRT)

3D (MsI)

2D (T-S)

e (Hg)

Vp, Vs and s

(54 mm )

(18 mm )

(54 mm )

(54 mm )

(54 mm )

(20 40 mm)

(15 15 15 mm)

(54 mm )

11
5
7
8
7
8
9
7
4
3
13
82

6
10
9
12
10
13
12
10
6
3
14
105

11
5
7
8
7
8
9
7
4
3
13
82

3
3
3
3
3
3
3
3
3
3
3
33

5
5
7
8
7
8
9
7
4
3
13
76

2
2
2
2
2
2
2
2
2
2
2
22

3
3
3
3
3
3
3
3
3
3
3
33

11
5
7
8
7
8
9
7
4
3
13
82

Author's personal copy


74

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

Fig. 2. Polarized light microscope images of lava and pyroclastic rocks from Solfatara and Ischia: a) sample SLA1. Sanidine (Snd) as a major constituent and a minor amount of biotite
(Bio) in a matrix formed prevalently by micro-sanidine (Mi-Snd) and micro-biotite (Mi-Bio); b) sample SLA3. Altered pyroxene (Al-Pry) and argilization stains (Ar) and some pores
(P) are easily recognizable; c) sample SLA4. Altered pyroxene (Al-Pry), silica-amorphous (Si-am) minerals, and pores (P) are easily recognizable; d) sample SLA5. Totally argillized
pyroxene (Ar-Pry) is recognized just by its geometry. Micro-sanidine (Mi-Snd) and silica-amorphous (Si-am) are recognized; e) sample SPRA1. Micro-sanidine (Mi-Snd), quartz
(Qz), and silica-amorphous (Si-am) minerals in a matrix formed by hydrothermal processes; f) sample SPRA2. Fragments of glass (Gl-fr), silica-amorphous (Si-am),
micro-quartz (Mi-Qz), and lithic fragments of lava (Li-la); g) sample IGTF. Major constituents are plagioclase (Plg) and sanidine (Snd) with minor amounts of biotite (Bio);
sub-rounded fragments of lava (Li-La) and micro-biotite (Mi-Bio) are very common; h) sample IGTA. Composed prevalently by fragments of glass (Gl-fr), plagioclase (Plg) and
sanidine (Snd). Fragments of lava (Li-la) are very common.

scanner and a BIR Actis 130/150 Micro CT/DR system. The advantages of
medical scanner include high image acquisition velocity and conguration versatility. On the contrary, 33 cylindrical samples (3 samples for
each lithotype) (Table 5) were scanned by means of a Micro CT/DR,
which allows the acquisition of images at higher resolution (4060 m).

Image processing, data extraction and data analysis required the


following steps (Fig. 4a): 1) image rectication, which consists in dening image brightness and applying a series of lters to reduce noise
(Gualda and Rivers, 2006; Ketcham, 2005) (Fig. 4b); 2) identication and
isolation of elements of interest (e.g. porosity and voids distribution)

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

75

Table 6
Chemical weathering indices (CWI).
Index

Formula (molecular ratio)

References

Chemical index of alteration (CIA)


Alumina to calciumsodium oxide ratio (ACN)
Silica to alumina ratio (SA)
Plagioclase index of alteration (PIA)
Chemical index of weathering (CIW)

(Al2O3)(100)/(Al2O3 + CaO + Na2O + K2O)


Al2O3/Al2O3 + CaO + Na2O
SiO2/Al2O3
(Al2O3-K2O)(100)/(Al2O3 + CaO + Na2O-K2O)
(Al2O3)(100)/(Al2O3 + CaO + Na2O)

Nesbitt and Young (1982)


Harnois and Moore (1988)
Ruxton (1968)
Fedo et al. (1995)
Harnois (1988)

(Fig. 4c) 3) extraction of porosity shape parameters (location, perimeter,


surface area, circularity and aspect ratio) (Fig. 4d), resulting in a geometrical, morphological and topological description of pore volume features.
2.4. Pore size distribution
To assess the inuence of the degree of alteration on pore size distribution, we applied power law analysis to pore size distribution values
obtained by mercury porosimeter, X-ray tomographies obtained by
Micro CT/DR system, and thin section image analysis (Table 5). Fractal
distributions were identied using a bi-logarithmic diagram of the number of pores larger than r. The pore-size distribution obeys a power law
where its angular coefcient corresponds to the fractal dimension
(D) (Fig. 5) (see Turcotte, 1992 and Higgins, 2006 for a more complete review). In addition, fractal dimension values are simply related
to the overall properties of pores: the increase of D values corresponds
to more graded particle size distributions and a larger number of ne
pores.
In order to compare 2D pore size distribution from thin section
analysis with 3D pore size distribution from X-ray tomography, a
two- to three-dimensional conversion method based on following
relationship: V = [A(a + b)/2] was followed, where A is pore area, a
and b are the major and minor axes, respectively (Farmer et al., 1991).
Before applying the two to three dimensional conversion methods,
the two dimensional values of porosity obtained from thin-section and
X-ray images analysis are compared and the most important characteristics are presented in next paragraphs.

The problem of conversion of two-dimensional data to threedimensional is not simple for objects more geometrically complex
than a sphere (Higgins, 2000). This problem could be treated through
the utilization of more complex stereological 2D3D conversion
techniques, based on the observed number of particle cuts, on a randomly oriented cross-section through the volume (Sahagian and
Proussevitch, 1998). Stereological conversion techniques have been
applied to quantify textural features (e.g. pore and crystal sizes and
distribution) of a wide range of rock formations (Higgins, 2000;
Morgan and Jerram, 2006; Shea et al., 2010).
2.5. Ultrasonic pulse velocity
Compressional and shear wave velocities (Vp and Vs) depend on
the elastic modulus and mass density of the medium; whereas, the attenuation of a wave front depends on textural characters of the rock
(e.g. scattering of fractures and voids, weathering state, and uid saturation) and various energy absorbing mechanisms, such as saturation and permeability (Price, 2009). The degree of Vp anisotropy may
depend on the crack distribution, the effective pressure, and the frequency at which the measurements are made. On the contrary, Vs can increase
or decrease with saturation, and Vs anisotropy depends on rock microstructure; pore structure plays an important role in explaining the
wave attenuation and wave dispersion.
Pore structure, texture of the rock (geometric arrangement and
sizes of grains, crystals, pores, glass shards, uid saturation, preferred
orientation of crystals), fracturing and severe mineralogical changes

Fig. 3. Illustration of total Porosity (t) in black color, extracted from two-dimensional digital image analysis. The matrix and minerals are shown in white. a-f) total porosity for
Solfatara lava series (12.8%, 16.9%, 18.4%, 15%, 9.5%, respectively); g-h) total porosity for Solfatara pyroclastic rock (24.8%, 17.8%, respectively); i) total porosity for Ischia Green
tuff rock (15.3%).

Author's personal copy


76

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

Fig. 4. (a) example of X-ray micro CT original images (b) Example of a ltered image reconstructed from a tomographic slice through sample SL4; (c) pore distribution as obtained
from thresholding of ltered image (d) 3D pore-system reconstruction.

Fig. 5. Bi-logarithmic pore-size frequency diagram. The power law tting relationship
is illustrated by a line, the way this power law changes within each sample is illustrated
by the dash-lines. The exponent of each Power law, which corresponds to the fractal
dimension, is included in Fig. 9a.

in a rock have been commonly related to the variation and changes in


wave velocity (Al-Harthi et al., 1999; Andriani and Walsh, 2002;
Drrast and Siegesmund, 1999; Eberhardt et al., 1999; Kern et al.,
1997; Ketcham and Carlson, 2001; Ketcham and Iturrino, 2005; Kili
and Teymen, 2008; Martnez-Martnez et al., 2006; Taud et al.,
2005; Wulff et al., 2000; Xu et al., 2006). Vp and Vs generally decrease
with increasing porosity for a certain kind of rock. Vinciguerra et al.
(2009) found that the Vp of tuff, with 10% to 30% in porosity, ranges
between 3.1 and 4.1 km/s. The velocity-porosity relationship becomes
complicated when micro-cracks exist, because the elastic properties
are more affected by the micro-cracks than by open porosity (see
Martnez-Martnez et al., 2006; Sousa et al., 2005). Furthermore,
the aspect ratio (i.e. longer dimension to shorter dimension) of cavities has a pronounced effect on the elastic properties. For example,
Martnez-Martnez et al. (2007) explain the variation of Vp and Vs in
brecciated carbonate rocks by a series of petrographical parameters
(e.g. brecciation, ne-grained matrix content, clast size distribution
and preferred orientation of clasts).
Effective pressure on transducers is an external variable that
also modies the ultrasonic parameters measured. Benavente et
al. (2006) studied in depth the inuence of pressure on the value
of several ultrasonic parameters. They compare results obtained
from a pressure-controlled system and hand-held measurement
procedure. Conclusions show that the error in the manual measurements is negligible in the velocity quantication. Variations
in the energy and attenuation parameters are also very low in the

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

range of pressures typical for tests carried out with manual


procedures.
We measured Vp and Vs by using a Pulse Velocity Test Instrument,
following Naik et al. (2004). The ultrasonic pulse velocity was obtained
by direct transmission of a 1 MHz signal in 82 cylindrical samples,
54 mm in diameter and 130 mm high (Table 5). The adopted sampling
frequency was 1 MHz. Ultrasonic waveform was registered using Parametric transducers (model 5660B, gain 40/60 dB, bandwidth 0.02
2 MHz). Velocities (Vp and Vs), spatial (s) and temporal attenuation
(t) have been calculated. s, the gradual loss in energy or reduction
in amplitude of the wave as it propagates through the rock medium,
was computed according to the equation (Martnez-Martnez et al.,
2011):
s dB=cm 20LogAe =Amx =L
where Ae is the maximum amplitude emitted by the transmitter, Amx is
the maximum amplitude registered by the receiver and L is the length
of the cylindrical sample.
3. Sample description
A brief geological description of each outcrop (Solfatara, Ischia and
Bolsena) together with mineralogical and petrographycal study of all
series and the degree of alteration has been completed (Figs. 2 and 6,
Tables 13).

77

3.1. Solfatara lava (SL)


The lava from Solfatara is heavily fractured, with strongly altered
joints suggesting a widespread fumarolic and thermal springs-activity.
Discontinuities are often very small in size. Altered lava varies signicantly on a short distance especially when approaching to fumarolic activity
(Table 1). The rock fabric and texture at some places are completely lost
making the in situ identication of the original rock very difcult. Five different degree of alteration, from SLA1 to SLA5, have been recognized.
3.1.1. Fresh Lava (SLA1)
The sample has a porphyritic texture (P.I. = 30%) with euhedral
plagioclase phenocrysts, on average 3 mm size. Major constituents
are sodic plagioclase (oligoclase and andesine) and potassic feldspar
with minor amounts of pyroxene and biotite. The matrix presents a
sub-parallel arrangement of micro-pyroxenes (30%) and microplagioclases (60%) (pilotaxic texture), 200 and 100 m of maximum
size, respectively. Two main types of alteration are observed: oxidation
at the boundary of most minerals, affecting all biotite crystals, and
argilization in the matrix and around minerals as blurred stains
(Fig. 2a). According to XRD results (Fig. 6), SLA1 contains mainly
sanidine, nepheline, and pyroxene (augite). Small peaks of biotite
and albite are also identied.
3.1.2. Slightly altered lava (SLA2)
It presents a trachytic texture with predominant euhedral phenocrysts of sanidine (average size 4 mm), minor amounts of plagioclase

Fig. 6. XRPD results of fresh and weathered rocks: a) Lava from Solfatara; b) pyroclastic rock from Solfatara; c) Green tuff from Ischia; d) Bolsena ignimbrite.

Author's personal copy


78

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

(average size 1.5 mm) and a small amount of pyroxene (average size
0.8 mm) and biotite (average size 1.2 mm). The matrix is mainly
composed by micro-plagioclases (60%) (sanidine) and microcrystals
of biotite (20%), 200 and 70 m of maximum size, respectively. The
remaining part of the matrix is composed by 5% of pores, 35 m in
maximum size and 15% of argilization stains. Argilization and oxidation
are present along micro-fractures and boundary of minerals. Biotite is
almost totally replaced. According to XRPD results (Fig. 6) this sample
contains mainly sanidine. Peaks of biotite and pyroxene (augite) show
a slight reduction in intensity; low peaks of gypsum and hematite are
also present.
3.1.3. Moderately altered lava (SLA3)
The major constituents are potassic feldspar (sanidine) with minor
amounts of sodic plagioclase and pyroxene, 1.2 mm in maximum size; biotite is rare. Relatively large crystals of sanidine (2 mm in average size)
are surrounded by a matrix composed principally by micro-plagioclases
(60%) and micro-pyroxenes (10%), 90 and 30 m in maximum size, respectively. Oxidation affects all crystals boundaries; biotite and pyroxene
are almost completely replaced (Fig. 2b) Argilization is present into the
matrix and all around minerals as blurred stains. According to XRPD results (Fig. 6a) this sample contains mainly sanidine, with a slight reduction in intensity for peaks of biotite and pyroxene (augite). Low peaks of
gypsum and hematite are present.
3.1.4. Highly altered lava (SLA4)
Although all minerals are altered, sanidines prevail over pyroxenes and plagioclases. XRPD shows that alunite (derived from acidic
alteration of potassic feldspar) is abundant. Matrix is totally replaced
by argilization and a new process of silicication can be observed into
the potassic feldspar and within the matrix (Fig. 2c). Silica-amorphous
minerals, 40 m in maximum size, appear all over the sample. Sanidine
(Fig. 6a) peaks are strongly reduced, whereas those of biotite and pyroxene disappear, and those of alunite and amphibole can be identied.
3.1.5. Totally altered lava (SLA5)
Complete alteration of all minerals occurs at this degree of alteration,
but some pyroxenes can still be recognized from their partially-distorted
geometry. Matrix is totally replaced by argilization and silica-amorphous
minerals, 40 m in maximum size (Fig. 2d). In XRPD (Fig. 6a) all mineral
peaks disappear and a large content of amorphous silica can be identied.
3.2. Solfatara pyroclastic rocks (SPRA)
A stratigraphic section, representative of the Solfatara crater pyroclastic series, was selected. The fresh outcropping part of the section
is located outside the crater, about 500 m from its northern ank.
The series is composed of slightly stratied deposits with highly variable textures and grain size. The lithofacies consist of decimeter- to
meter-thick layers of breccia, lapilli, ne-coarse ash and sub-angular
pumice lapilli with lithic fragments (altered lavas). The stratigraphic
series inside the north-western ank of the Solfatara crater is chaotic
and affected by fumarolic activity, which obliterates the primary structure of the pyroclastic series. The three samples of the pyroclastic series
were classied as a highly altered, even if they present large differences
in grains size.
3.2.1. Highly altered pyroclastics (SPRA1)
It is composed prevalently by sub-angular glass fragments with an
average size of 0.5 mm (Fig. 2e). Oxidation appears as small stains all
around the largest glass fragments (0.4 mm in maximum size) replacing the phenocrystals of feldspars, which can be identied by their
preserved geometry. Cracks, sub-rounded lithic fragments of lava and
pyroclastic rock fragments are common. The largest glass fragments contain elongated-shape micro-crystals. XRPD results (Fig. 6b) show the

presence of alunite, jarosite and quartz and a large content of amorphous


silica.
3.2.2. Highly altered pyroclastics (SPRA2)
It is composed prevalently by sub-angular glass fragments (average size 4 mm). The matrix is formed by micro-fragments of glass
surrounded by micro-crystals of feldspars and quartz (Fig. 2f). Oxidation appears as stains in the matrix and in the largest glass fragments
(pumice, maximum size 5 mm). Very large sub-rounded lithic fragments
of lava and pyroclastic rocks are common (maximum size 7 mm). The
size of lithic fragments reaches 10 mm, in hand specimen. XRPD results
(Fig. 6b) show presence of sanidine, alunite, pyrite and quartz and a
large content in amorphous silica.
3.2.3. Highly altered pyroclastics (SPRA3)
It is composed prevalently by sub-angular glass fragments (average
size 6 mm), forming also the matrix and by very large sub-rounded
fragments of lava and pyroclastic rocks. Oxidation appears as diffuse
stains in the matrix. Lithic fragments are altered, but original trachytic
structure, with sanidine crystals, are easily identied. The size of lithic
fragments reaches 5 cm in hand specimen.
3.3. Green tuff from Ischia (IGT)
A stratigraphic section, outcropping along the northern ank of the
Mt Epomeo resurgent block and partly affected by hydrothermal processes, has been selected (Table 3). The hydrothermal alteration is induced by hot uids migrating along the main faults of the Mt Epomeo
resurgent block (Inguaggio et al., 2000). Samples are representative of
both fresh and hydrothermally altered portions of the Green tuff.
3.3.1. Fresh Ischia Green tuff (IGTF)
The fresh Green tuff unit consists of a pumice- and crystal-rich
pyroclastic density current deposit. This deposit is welded, massive, matrix supported, with sub-rounded pumices less than 10 cm in diameter
and angular lithics fragments less than 5 cm in diameter. The matrix
consists of a medium ash made by crystals, lithics and pumice fragments. In the studied outcrop, the deposit shows 0.11 cm secondary
vesicles formed by the collapse of the pumice structure in some juvenile
fragments.
This sample has a pyroclastic texture composed mainly by glass
(Fig. 2g) and curved and elongated pumice fragments. Major constituents are plagioclase and sanidine with minor amounts of pyroxene and
biotite (maximum size of 0.8 mm). The matrix is made of micro-crystals
of plagioclases feldspars (sanidine) (10%), biotite and pyroxenes (5%).
Oxidation appears as small stains within the largest fragments of pumice.
Sub-rounded lithic fragments of lava (maximum size 548 m) are very
common. Crystals of sanidine, plagioclase and biotite from lithic fragments of lava are easily recognizable, although a diffuse argilization is
present.
3.3.2. Highly altered Ischia Green tuff (IGTA)
The altered green-tuff is composed by hydrothermalized, red to
white, compacted soft rock. The pumice juvenile fragments of this altered portions of the deposit are not vesiculated, due to compaction
and lling eventually promoted by the hydrothermal processes.
It is composed prevalently by glass, whereas among the phenocrysts, plagioclase and sanidine are the most abundant, with small
amounts of biotite (Fig. 2h). The matrix has the same composition.
Micro-plagioclases and micro-sanidine with a maximum size of 70
and 50 m, respectively, form 5% of the matrix, while biotites represent just the 3%. Sub-rounded lithic fragments of lava with trachytic
texture are common; crystals of sanidine and biotite are easily recognizable, although argilization is diffuse.

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

79

(blurred stains) and silica-amorphous minerals along micro-cracks


are observed. Two slightly different facies are typical of BoPRA and they
are identied on the basis of the prevalence of ner or coarser particles.
This characteristic slightly controls the physical mechanical properties
as described by Pola et al. (in preparation).

3.4. Bolsena ignimbrite


A stratigraphic section, 3.5 km east of the Bolsena lake and 4 km
north of Monteascone, representing the remnants of the ninth and
tenth eruptive phases of Bolsena volcanic Zone (Nappi et al., 1998),
was selected. It is composed by a series of pyroclastic ows, pyroclastic falls and lava layers (Table 4). Below a 2 m thick lava, a series of
un-welded very low density pyroclastic ows outcrops. This series
consists of meter-thick layers made of sub-angular pumice lapilli with
scattered lithic fragments and crystals. These deposits are characterized
by decimetric yellow stains. Well-stratied surges, composed by coarse
ash and pumice fragments, are underlayered by a series of fall deposits,
separated by thin ash layers. These deposits are clast-supported, with
angular pumices ranging from 1 to 20 mm. Some of the fall levels are
compacted and slightly cemented.

4. Results and discussion


Physical properties are summarized in terms of their mean values
in Table 7. Bulk density varies in the range of 9.60 b b 24.52 KN/m 3,
while e and t, obtained by different techniques, range between
5.4 b e b 65% and 0.7 b t b 57.7%. The highest e and t values (65%
and 57.7%), obtained through Medical scanner images and pycnometer
analysis, respectively, characterize BoPRA samples, which also present
the lowest value of (9.60 KN/m3). The lowest value of e and t (6%
and 0.7%), obtained by bulk specic weight measurements and Medical
scanner image analysis, respectively, correspond to SLA1 sample, the
less altered sample. Fractal dimension is in the range between 1.24
and 2.14 and decreases progressively from intact fresh rocks to the
most altered samples (Table 7), suggesting an increase in large pore frequency. Vp and Vs vary from 1.14 to 4.39 km/s and from 0.42 to 2.91 km/s,
respectively. They are strongly connected with porosity and consequently
with the degree of alteration. It is observed (Table 7) that the lowest

3.4.1. Slightly altered Bolsena Ignimbrite (BoPRA)


This sample has a pyroclastic texture composed by small sub-angular
silica-amorphous minerals (average size 0.6 mm), sub-rounded lithic
fragments of lava and pyroclastic rocks. The matrix is composed by
well-dened micro-fragments of glass and micro-crystals of feldspars
and quartz. Micro-crystals are unaltered, but oxidation along microcracks and within the largest clasts, argillization within the matrix
Table 7
Summary of physical properties (mean values) of weathered/altered volcanic rocks.
Sample

(kN/m3)

(kg/m3)

t (%)

e (%)

T-S

XRT
2D
10.2
15.3
21.3
17.6
13.0

MsI

Pm

Hg

3D
6.0
6.4
25.6
30.7
31.5

0.7
2.1
7.7
19.9
22.5

14.6
15.1
28.4
33.1
30.3

11.0
15.0
18.6
32.0
26.8

6.0
5.4
23.0
21.0
19.2

Lava
SLA1
SLA2
SLA3
SLA4
SLA5

23.3
24.5
19.0
16.2
14.7

2375
2500
1938
1650
1500

2D
12.8
16.9
18.4
15.0
9.5

Pyroclastic
SPRA1
SPRA2
SPRA3

14.6
15.1
14.0

1483
1540
1425

24.8
17.8

19.5
10.0
20.0

20.3
34.9
42.9

23.9
31.4
30.2

39.3
38.3

41.5
44.8

24.7
22.4
30.2

Tuff
IGTF
IGTA

15.1
17.8

1540
1810

15.3

24.7
8.0

25.0
24.0

15.7
22.9

27.8
37.5

25.5
29.7

19.2
29.9

9.6

980

32.0

22.0

49.8

57.7

53.8

65.0

41.2

Ignimbrite
BoPRA

Data set is given as mean values. : unit weight, : density, t: total porosity, e: effective porosity, T-S: results from thin-section, XRT: results from x-ray tomography images,
MsI: results from Medical Scanner Image, Pm: results from pycnometer, Hg: results from mercury porosimetry, M: results from bulk-specic weight measurements.
Sample

Waves (km/s)

T-S

XRT

Hg

Vp

Vs

(dB/cm)

Lava
SLA1
SLA2
SLA3
SLA4
SLA5

0.06
0.06
0.3
0.27
0.24

1.59
1.44
1.47
1.35
1.44

1.92
1.84
1.80
1.63
1.52

2.13
2.06
2.07
2.03
2.08

4.39
4.14
3.16
3.11
2.79

2.13
2.91
2.00
1.00
1.48

1.98
4.06
2.76
4.66
2.53

Pyroclastic
SPRA1
SPRA2
SPRA3

0.33
0.29
0.43

1.34
1.40

1.59
1.42
1.45

2.15
1.95

2.18
2.06
1.65

1.02
0.79
0.51

2.54
2.91
3.23

Tuff
IGTF
IGTA

0.24
0.43

1.24

1.31
1.34

2.05
2.14

1.14
2.25

0.42
0.81

4.42
3.39

Ignimbrite
BoPRA

0.7

1.90

1.63

2.11

1.14

0.86

2.90

e: void ratio, D: fractal dimension, Vp: compressional wave velocity, Vs: shear wave velocity, s: spatial attenuation.

Author's personal copy


80

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

velocity in each rock series (SLA, SPRA, IGT) was determined in the most
altered sample or in those with highest porosity (SLA5 and SPRA). On the
contrary, the highest value was determined in low porosity rocks
(SLA1). IGT series represents an exception, since the most altered
sample showed lowest porosity.
Furthermore, samples with the highest s values are those in which
micro-fractures, interconnected pores, mineral degradation and large
fragments of pumices have been identied (Table 7). Values of s
range from 1.98 dB/cm for the most homogeneous sample (SLA1) to
4.66 dB/cm for the sample with a large value of interconnected-pores
(SLA4) (Fig. 3). The highest values of s in SLA series are determined
for SLA2 and SLA4 (4.06 and 4.66 dB/cm, respectively), where microfractures (perpendicular to the wave emission) and cavities are observed.
The most altered sample of SLA series (SLA5) shows a decrease in s
values, probably because of hydrothermal depositional processes. In
SPRA series, s values are mostly controlled by the groundmass and
the grain size: the smaller the grain size the smaller s (2.54 to
3.23 dB/cm). In addition, in IGT series, s values are mostly controlled
by the structure of the pumice fragments (IGTF, 4.42 dB/cm) and by
hydrothermal depositional processes (IGTA, 3.39 dB/cm) (Table 7).
4.1. Chemical analysis
Changes in element concentration within alteration series (SLA,
SPRA, IGT) are illustrated in Fig. 7. In SLA series, the amount of SiO2,
Al2O3, Na2O, K2O, and MnO remain relatively stable or decrease slightly
along the early stages of alteration (SLA2 and SLA3 samples), while the
amount of Fe2O3, CaO, MgO, Na2O, and, MnO decreases drastically within the most altered stages (SLA4 and SLA5 samples), because of the
leaching processes. Large amounts of SiO2 and TiO2 are found in
the most altered stage (SLA5), because of the recrystallization of
amorphous-silica (see SA ratio in Table 8). In all series, the content of
Fe2O3, CaO, MgO, Na2O and MnO presents a notable decrease, when the
altered samples are compared with fresh sample. In particular, these contents decrease from 1 to 0.08 in SLA4, SPRA2, and SPRA3 samples (Fig. 7).
In addition, all components in SPRA series present the same variation pattern. The components (CaO, MgO, MnO, and P2O5) of the most altered
sample in IGT series changes drastically, in particular MnO contents increase signicantly.
The chemical indices shown in Table 6 have been applied to our
samples to assess the degree of weathering and to relate these indices to some physical properties (e.g. , , Vp and Vs). The indices

Fig. 7. The unaltered rock-normalized diagram for rocks from Solfatara and Ischia.

Table 8
More representative chemical weathering indices for volcanic rocks from Solfatara Ischia
and Bolsena.
Site

Sample

CIA

ACN

SA

PIA

CIW

Solfatara

SLA1
SLA2
SLA3
SLA4
SLA5
SPRA1
SPRA2
IGTF
IGTA
BoPRA

42.73
45.98
46.33
70.93
58.80
70.41
67.04
47.09
55.13
69.24

0.55
0.60
0.60
0.93
0.67
0.97
0.91
0.61
0.69
0.76

5.63
5.70
5.50
3.93
238.10
6.94
6.29
5.64
5.65
383.52

37.13
42.30
43.15
90.46
61.50
94.60
85.60
44.76
58.69
73.68

54.63
60.39
60.37
93.50
66.63
96.61
90.68
60.51
69.35
76.40

Solfatara
Ischia
Bolsena

CIA: Chemical index of alteration (Nesbitt and Young, 1982). CAN: Alumina to calcium
sodium oxide ratio (Harnois and Moore, 1988). SA: Silica to alumina ratio (Ruxton,
1968). PIA: Plagioclase index of alteration (Fedo et al., 1995). CIW: Chemical index of
weathering (Harnois, 1988).

have been calculated using the molecular proportions of major element oxides (Fig. 7) and they were selected according to their capability of best describing chemical alteration by mineral degradation
pattern.
Patterns of chemical weathering indices indicate a progressive increase of weathering intensity (Table 8). The CIA, ACN and CIW indices range from 42.73 to 70.93, from 0.55 to 0.97, and from 54.63 to
96.61, respectively. In lava series, at the fourth degree of alteration
(SLA4), the high value of CIA and ACN are mainly due to an increase
in Al2O3 content, which could be associated to clay minerals deposition. The reduction in Ca and Na becomes evident in all the series,
suggesting that these elements are highly mobile and are preferentially
removed from rocks during weathering/altering processes. Values of
PIA, ranging from 37.13 to 94.60, suggest that the strong mobility of
Ca and Na is caused by the heavy dissolution of plagioclase (albite). Optical microscope observations (Fig. 2) reveal that the alteration of plagioclase results from dissolution along fracture planes. At second stage
of alteration, plagioclase from the SLA series shows clear evidence of
argilization and oxidation as fracture lling. At fourth and fth stages of
alteration, plagioclase from SLA series is severely affected by weathering
and completely replaced by clay minerals and recrystallization of amorphous silica. This pattern is very well described by the SA ratio, ranging
from 3.93 to 6.94 (Table 8). The large amount of amorphous silica is
reected in SLA and BoPRA samples, where the values of SA ratio are
very high (238.10 and 383.52, respectively).
In general, an acceptable correlation is observed between the degree of alteration and the value of the chemical indices (CIA, CIW,
ACN, SA and PIA). CIA, ACN, PIA, and CIW values increase with alteration, whereas SA decreases. The increase of CIW and CIA values can
be attributed to the loss of mobile cations (Table 6; Fig. 7) and alteration of the crystal structure (Haskins, 2006). Index values do not exhibit large changes among the rst three degrees of alteration, making
classication difcult. In particular, ACN and CIW in SLA series range
from 0.55 to 0.60 and from 54.63 to 60.39, respectively. Mobile oxides
(Al2O3, Fe2O3, CaO, Na2O, K2O) are highly removed from the rock at
fth degree of alteration (Fig. 7 and Table 8). As a result, drastic changes
in the common patterns of all chemical indices are observed.
Values of CIA obtained in this study, at the last stages of weathering
(third to fth) are in the range of values described in Arikan et al. (2007)
for unaltered/slightly altered acidic volcanic rocks (54 b CIAb 74).
Since CWI have limitation in identifying the degree of weathering
in a rock system, they should be reinforced and reinterpreted by additional information (e.g. petrographical, mineralogical, physical and
mechanical) (Kim and Park, 2003). Price and Velbel (2003) report that
high values of CIW, CIA and PIA may not always reect weathering,
because of their sensitivity to changing bulk chemistry associated

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

81

with hydrothermal alteration. Duzgoren-Aydin et al. (2002) suggest


that signatures of chemical weathering should be used to establish a
model for the interpretation of chemical heterogeneity rather than for
quick classication, mainly because the distribution of chemical elements is determined by local conditions at the site, which could vary
drastically in a short distance. It is clear that no single CWI would collect
all of the chemical characteristics of the samples and the most representative index might be chosen by evaluating the chemical patterns of
each geological environment. Summarizing, CIW and CIA indices seem
to be the most representative, even if some local alteration is identied
by values derived from SA index.
4.2. Characterization of porosity
Image analysis represents a rapid and precise method to obtain 3D
reconstruction of porosity and quantication of each pore (e.g. area,
volume, shape, frequency, and spatial distribution). We focused on
comparison of images of samples with different alteration degree,
which allows to reconstruct porosity changes with alteration.
In general, total porosity (t) values obtained from different techniques increase with the degree of alteration for almost all rock series
(SLA, SPRA, and BoPRA). Results from X-ray tomography image analysis and pycnometer tests reveal that t increases progressively with
the degree of alteration with some minor changes in lava series (values
of SLA3 increase drastically and values from SLA5 present a small reduction). Lava series show the highest increase in porosity with the degree
of alteration, from 0.69% to 22.48% in medical CT images and from 6% to
31.5% in microCT images. Differences in the results from microCT and
medical CT are caused by different resolution between the two techniques and the consequent processing (averaging, ltering and manual
thresholding). Furthermore, total porosity (Fig. 8a) is obtained from estimated density values, based on the specic weight of the solid phase.
In X-ray medical CT, density values are inuenced by the coarser resolution with respect to micro CT.
Effective porosity (e) was obtained from bulk-specic weight
measurements and by mercury intrusion porosimetry. Values from
bulk-specic weight measurements vary between 6% for fresh lava
and 41.2% for non-welded ignimbrite, while values from mercury
porosimetry varies between 11 for fresh lava and 65% for non-welded
ignimbrite.
Values of e increase with the degree of alteration of the samples
(Fig. 8b) for all lithology, except for IGT, where e from porosimetry
decreases from 29.7% for fresh sample (IGTF) to 25.5% for the altered sample (IGTA); this is thought to result from post-depositional hydrothermal
alteration and deposition processes.
e obtained from bulk-specic weight measurements seems to
have no clear relationship with the degree of alteration (Fig. 8b). The
reason could be the percentage and size of interconnected pores and
fractures contained in each sample. It means that porous system connectivity does not increase progressively with the degree of alteration.
For IGT series, the decrease in e from bulk-specic weight measurements could be the result of a decrease in size of the pores with the
increased alteration and then with a consequent difculty in their
saturation under low air vacuum conditions.
Pore evolution in all samples is described in terms of structure of
the groundmass, nature and degradation of the crystals and pumice
fragments and post-depositional alteration processes as follows:
a) In general, the evolution of pores in SLA series is related to oxidation
and argilization process and connected micro-cracks, as it could be
observed by optical microscopy (Fig. 2ad). Reduction of porosity
in SLA5 is related to hydrothermal processes, as large pores have
been lled by new minerals, like clay and amorphous silica.
b) The evolution of pore structure in SPRA series is related to grainsize content and degree of groundmass cementation. The dominant
pore types are secondary and developed during and after a selective

Fig. 8. a) Total porosity as a function of lithotype obtained from pycnometer test, thin
section analysis (3D) and X-ray tomography (Micro CT). b) Effective porosity as a function
of lithotype obtained from bulk specic weight measurements and mercury porosimeter.

dissolution of minerals.
c) The evolution of pore structure in IGT series is related to the high proportion of pumice fragments and their open structure with weak
walls. The fragmentation of these structures is common, enhancing
pore connectivity and consequently forming open structures. The
most altered sample (IGTA) presents a very dense structure and a
drastic reduction in porosity, lled by clays, amorphous minerals
and small fragments of other materials transported by hydrothermal
processes.
d) High porosity values in BoPRA series are related to the depositional
processes: glass shards, derived from the fragmentation of the vitric
bubble walls of pumice vesicles, are well-preserved. Reconstruction
of pore structure and microscope observations reveal a high percentage of interconnected pores, which is also promoted by degradation and fragmentation of pumice elements.
The combination of techniques described above gives a good estimate of grain size and pore size distribution. They provide qualitative
and quantitative evaluations of t and e and allow the quantication
of spatial pore structure and size distribution. The most relevant conclusions are as follows:
Signicant relationship exists between porosity, as obtained by different techniques, and the degree of alteration for all the samples. t increases with the degree of alteration for most of the series (Table 7
and Fig. 8).
Thin-section image analysis allows describing pore shape characteristics on a plane (e.g. area, perimeter, circularity, and roughness), even
though, preparation, orientation (perpendicular to the base of the cylindrical specimens) and size (3 2 cm), could inuence the nal
estimates.
e values obtained from mercury porosimeter are generally slightly higher than the values obtained by water immersion method

Author's personal copy


82

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

(Table 7). This could result by the forced mercury intrusion (0.1
400 kPa) damaging pore walls or opening of small fractures, and/or
because of air vacuum pressure does not reach to permeate all pores.
Porosity obtained from thin-section is usually bigger than porosity
obtained from 2D X-ray tomography. In general, variation in all samples is constant (3%), although, bigger variation in BoPRA sample
(from 22 to 32%) are observed. Variation in both methods depends
on image segmentation. On thin-sections, pores are easily identied
from other clasts or minerals by using a color background, while pores
identication on X-ray tomography, depends on the gray scale of the
image. Sometimes, pores perimeter presents the same gray scale of
some other clasts or minerals.
Final values of 3D porosity computed from 2D data analysis of thin sections images (see Farmer et al., 1991) are controlled by the adopted
transformation relationship.
X-ray tomography is the fastest and more precise technique to obtain
3D textural information (e.g. volume, pore geometry, distribution, circularity, barycentre, interconnection). The quality of the results is
closely related to equipment calibration and resolution and image
processing (ltering and thresholding).
t values obtained from medical X-ray CT are inuenced by the coarser resolution and the adopted specic weight value of the solid phase
used in the calculations. At the same time medical CT allows to

evaluate the total porosity on large cylindrical specimens (b 20 cm


height) so it is able to provide a porosity value at a larger scale.
The fractal nature of porosity in altered volcanic rocks is a function of
the lithology and the degree of alteration. Fractal dimension was determined from pore volume values derived from thin-sections and X-ray
tomography image analysis, which represent t and mercury porosimeter, which represent e. Once 2D to 3D conversion was made
following the approach proposed by Farmer et al. (1991), values collected from X-ray tomography image and thin sections image were
widely comparable.
Pore frequency distribution, obtained from X-ray tomography image
analysis suggests fractal behavior of porosity in all series: in lava series
(SLA) values range between 1.52 and 1.92, in pyroclastic series (SPRA)
values range between 1.45 and 1.59, whereas in IGT series values vary
from 1.31 to 1.34 (see also, Table 7). As Table 7 shows, D decreases
progressively with the degree of alteration in all series, suggesting a
relative increase in frequency of large pores (Fig. 9bf). The fractal dimension corresponding to the most altered sample of ignimbrite
(IGTA) and lava sequence (SL5), deviate from this trend, suggesting
an increment of the relative frequency of smaller pores, as well as
large pores lled by new minerals (amorphous silica and clay minerals). This hypothesis is supported by thin section observations and
XRD analyses.

Fig. 9. a) Fractal dimension and most representative pore volume distributions obtained from b) thin section analysis (SLA series), c) mercury porosimetry (IGT and BoPRA series),
d) X-ray micro CT analysis (SPRA series).

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

D values obtained from Mercury porosimeter, represents distribution


of interconnected pores. Variation in D values is small in SLA and IGT
series, with D varying from 2.03 to 2.13 and from 2.05 to 2.14, respectively. A larger variation is observed in SPRA series (1.95 to 2.15) where
large anisotropy in the matrix, associated to hydrothermal processes
(leaching and deposition), and grain sizes is observed.
Fig. 9a indicates that different values of D occur at different scales
of analysis. These results suggest that scaling relationships based on
fractal geometry may be very useful at describing real distribution of
pores. Scaling relationship in D is restricted principally by the mode
of data acquisition. For example, acquisition of pore values in thinsection depends principally on the area of the thin-section, particles
size and quality of the image. In turn, the quality of nal results
depends on a series of noise reduction, ltering, thresholding, and particle separation steps.

4.3. Evaluation of P and S waves velocity


Statistical evaluation of Vp and Vs values and their relationships
with lithotype and degree of alteration are presented in Fig. 10c and
d, respectively. Vp values follow a decreasing trend with increasing
the degree of alteration. Mean values of Vp decrease from 4.39 km/s,
for SLA1 sample, to 2.79 km/s, for SLA5 sample (totally altered). Values
of Vp in SPRA series vary from 1.65 to 2.18 km/s; variation in this series
depends principally on sample structure and grains arrangement. IGT series follows an increasing trend with increasing the degree of alteration.
Vp values increase drastically from 1.14 km/s for IGTF to 2.25 km/s for
IGTA (altered sample). Based on thin section and X-ray tomography images observations, IGT samples are composed by interconnected pores,

83

which in some cases form at and elongated cavities. In this way, significant variations in Vp observed in Fig. 10c are closely related to physical
changes, consisting in the development of new pores, clay minerals
and the increased and decreased width of interconnected cavities by
hydrothermal processes. These observations are conrmed also by s
values (Table 7). The most homogeneous sample is the less altered sample of lava series (SLA1), with s value of 1.98 dB/cm, while the most
heterogeneous samples are SLA2, SLA4 and IGTF with s values of 4.06,
4.66 and 4.42 dB/cm, respectively (Fig. 11). According to petrographical
and physical descriptions, these samples are characterized by long
micro-fractures with a large void content, cavities and pumice fragments with very open and very weak structures. s values in SPRA series
vary from 2.54 to 3.23 dB/cm, depending on the textural characteristics
(Table 2 and Fig. 3): the highest s values were obtained for SPRA3 samples, which contain a large amount of lithic and pumice fragments
(2 mmb b 4 mm). s measured in IGT series varies according to the
characteristics derived from hydrothermal processes (Table 7). Textural
characteristics of fresh sample (IGTF) change drastically due to clay minerals deposition within the cavities formed by fragmentation of pumices.
The dispersion of Vp and Vs data is represented by the standard deviation (Fig. 10c and d, respectively). Dispersion of Vp values, is small
in all samples (b 0.36 km/s), with larger values observed in samples
SLA3, SLA5, and SPRA3, where strong anisotropy (e.g. matrix, grain
sizes and cavities content) is identied. On the contrary, dispersion of
Vs values is large for most of the samples (b 0.61 km/s), due to the
error related to the difcult to pick the rst arrival of the S-wave when
it shows low amplitude. Vs values do not have clear relationship with
the degree of alteration in all series (Fig. 10d). Large differences are present for each degree of alteration and each sample. For example, large variations in Vs measurements are observed along SLA series; in particular

Fig. 10. Propagation wave velocity of: a) P-wave vs effective porosity from bulk specic weight measurements, b) S-wave vs effective porosity from bulk specic weight measurements, c) P-wave vs lithotype and degree of alteration, d) S-wave vs lithotype and degree of alteration. Values of standard deviation (S) are included.

Author's personal copy


84

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

mineralogical, petrographycal and visual characteristics was carried


out. Then, different laboratory tests were performed in order to nd
out the physical behavior of volcanic materials. Physical characteristics
were compared and relationships found giving reference values and
ranges for properties and their trends with weathering. Some new and
interesting results were obtained and rocks were characterized
completely. However, further research is still necessary in order to corroborate the obtained relationships, as well as to dene more precise
correlations between petrographic and petrophysic variables. This
could be done by increasing the number of tested samples, obtaining
3D orientated thin sections of each lithology, and analyzing different
weathered volcanic rocks.
The most important conclusions obtained in this paper are briey
summarized as follow:
Fig. 11. Spatial attenuation (s) vs lithotype and degree of alteration. Homogeneous
rock (SLA1) presents the lowest s value. Dashed circles enclose the most heterogeneous samples, (with micro-fractures, SLA2; voids and cavities, SLA3; pumice fragments with open structures). Dashed rectangle encloses SPRA series, where s values
present small variation due to textural and grain sizes content.

measurements on specimens in SLA1 and SLA5 samples vary from 1.40


to 3.9 km/s and from 0.68 to 2.40 km/s, respectively. The reason could
lay in the percentage and size of interconnected pores, severe mineralogical changes and fractures characteristics (e.g. width and orientation). A
smaller variation is observed for SPRA series, which has a more persistent
network of pores, micro-fractures, and stratication sub-perpendicular
with respect to the direction of the wave emission (Fig. 10d), It should
be remembered that all of the samples in SPRA sequence present same
degree of alteration and variation in all the physical properties is strongly
related with the differences in grain sizes composition.
The relationship between values of Vp and e, from bulk-specic
weight measurements, are of the exponential type (Fig. 10a). In general,
all values are well correlated, although, few dispersed values are observed for SLA3, SLA4, and SLA5 and BoPRA samples. e values obtained
from bulk-specic weight measurements are not clearly related to Vs
(Fig. 10b) and best tting exponential relationship has a very low coefcient of determination (0.80 > R2 >0.040). Vs values in SLA1, SLA2,
SLA5 and BoPRA show large variation that could be attributed to abundance and orientation of small cracks or micro-fractures, and to the difference in pore network arrangement. Despite of the considerably large
number of total samples (187, Table 5), some specic lithotypes are underrepresented (e.g. IGT series, Table 5) and the scattering observed in
both Vp and Vs values could be related to this. Further researches will be
focused on increasing the number of tested rocks, contributing to a
more accurate and reliable relationship.
Finally, all these measurements have been carried out using a couple
of transducers centered at 1 MHz. This frequency is located between the
commonly used values at laboratory studies (usually between 100 kHz
to 1 MHz) and the seismic frequencies reported at eld studies (10 to
100 Hz) (Vinciguerra et al., 2006). Differences in transmitted wave frequency imply variations on the measured propagation velocity. This deviation can be corrected applying the BiotGassman equation. However,
other studies carried out on similar volcanic rocks have concluded that
corrected velocities are, in general, similar to the uncorrected values,
with increases ranging from 1% to 10% (Vinciguerra et al., 2006;
Zamora et al., 1994).
5. Conclusions
Different procedures to quantify physical properties of altered/
weathered volcanic materials from Solfatara, Ischia and Bolsena volcano, have been implemented and compared. Results demonstrate that
physical properties of rock change with the degree of alteration. As the
degree of alteration increases, physical properties decrease whereas e
and t increase. As a rst step, classication based on chemical,

The most common controlling alteration process is the hydrothermal ow. Joints and micro-fractures also appear as an important
control factor, as they inuence the behavior of Vp and Vs.
The collected volcanic rocks have been classied into ve alteration
categories (fresh, slightly altered, moderately altered, highly altered,
and completely altered) and into four lithotype series (lava, pyroclastic, tuff and ignimbrite series) according to the variability of petrographical, chemical, and physical characteristics.
The relationship between chemical changes and weathering can be
highlighted by using Chemical Weathering Indices (CWI). In general,
an acceptable correlation is observed between the degree of alteration and the value of CWI. The CIA is well correlated with the degradation trend exhibited by physical properties. We used the CWI only
to interpret the chemical heterogeneity, which is correlated with petrographical, mineralogical, physical and mechanical information.
Most of the physical and mechanical properties of altered volcanic rocks
are inuenced by porosity (voids, cavities, and fractures). In general, t
and e increase with the degree of alteration for almost all lithologies. In
lava (SLA) and non welded ignimbrite (IGT) series percentage and size
of interconnected pores, micro-fractures and pores lling by hydrothermal processes reveal some minor deviation from the general trend.
Fractal dimension (D) obtained from t and e, decreases progressively
with the degree of alteration, suggesting a relative increase in frequency of large pores. The most altered rocks are characterized by
an increase in small pore frequency (Table 7 and Fig. 9) with respect
to large ones, resulting in increasing fractal dimension. This is the result of hydrothermal deposition of clay minerals and recrystallization
of amorphous silica within large pores.
Vs values, reveals large dispersion not only within each degree of alteration, but also within each lithology. On the contrary, Vp is well related to the degree of alteration (Fig. 10c and d).
Spatial attenuation value (s) can be important in order to dene
rock defects. The smaller s value, the fewer the defects in rocks:
the most homogeneous rock (SLA1) shows values of 1.98 dB/cm,
while the most heterogeneous show values ranging between 4.06
to 4.66 dB/cm (e.g. SLA2, SLA4 and IGTF).
The described property values could be used to classify rocks and
to assign properties to similarly altered rocks. Finally, relationships
with mechanical behavior for these rocks will be presented in a geotechnical paper.
Acknowledgements
We gratefully acknowledge CONACYT and the Italian government
for the nancial support. We also gratefully acknowledge the Instituto
italiano de Cultura Ciudad de Mxico, especially Luigi Pironti. We would
like to thank Dr. Federico Agliardi, for the continuous and tireless support
in performing laboratory tests and for unraveling the mysterious world of
fractal analysis. We thank Dr. Giovanni Orsi and Dr. Sergio Chiesa for their
advice and assistance in the eld. Thanks to Elena De Ponti (Azienda
Opsedaliera San Gerardo, Monza, Italy), who provided the medical

Author's personal copy


A. Pola et al. / Tectonophysics 566-567 (2012) 6786

scanner images. Thanks also to Dr. Javier Martnez-Martnez and Dr.


Alberto Villa for helpful recommendations on geophysical aspects. Finally,
we thank Dr. S. Ammaiyan, Dr. J. Piriyadarshini and Dr. S. Vinciguerra for
their comments and suggestions which greatly improved the manuscript.

References
Al-Harthi, A.A., Al-Amri, R.M., Shehata, W.M., 1999. The porosity and engineering properties of vesicular basalt in Saudi Arabia. Engineering Geology 54, 313320.
Andriani, G.F., Walsh, N., 2002. Physical properties and textural parameters of
calcarenitic rocks: qualitative and quantitative evaluations. Engineering Geology
67, 515.
Arikan, F., Ulusay, R., Aydin, N., 2007. Characterization of weathered acidic volcanic
rocks and weathering classication based on a rating system. Bulletin of Engineering Geology and the Environment 66, 415430.
Ashbridge, D.A., Thorne, M.S., Rivers, M.L., Muccino, J.C., O'Day, P.A., 2003. Image optimization and analysis of synchrotron X-ray computed microtomography (CT)
data. Computers and Geosciences 29, 823836.
Avar, B.B., Hudyma, N., 2007. Observations on the inuence of lithophysae on elastic
(Young's) modulus and uniaxial compressive strength of Topopah Spring Tuff at
Yucca Mountain, Nevada, USA. International Journal of Rock Mechanics and Mining
Sciences 44 (2), 266270.
Avar, B.B., Hudyma, N., Karakouzian, M., 2003. Porosity dependence of the elastic modulus of lithophysae-rich tuff: numerical and experimental investigations. International
Journal of Rock Mechanics and Mining Sciences 40, 919928.
Aversa, S., Evangelista, A., 1998. The mechanical behaviour of a pyroclastic rock: yield
strength and destructuration effects. Rock Mechanics and Rock Engineering 31
(1), 2542.
Bahlburg, H., Dobrzinski, N., 2011. A review of the Chemical Index of Alteration (CIA) and
its application to the study of Neoproterozoic glacial deposits and climate transitions.
In: Arnaud, E., Halverson, G.P., Shields-Zhou, G.A. (Eds.), The Geological Record of
Neoproterozoic Glaciations. Geological Society, London. Memoir 36, 8192.
Beccaluva, L., Di Girolamo, P., Serri, G., 1991. Petrogenesis and tectonic setting of the
Roman Volcanic Province, Italy. Lithos 26, 191221.
Benavente, D., Martnez-Martnez, J., Juregui, P., Rodrguez, M.A., Garca del Cura, M.A.,
2006. Assessment of the strength of building rocks using signal processing procedures. Construction and Building Materials 20, 562568.
Best, M.G., 2003. Igneous and Metamorphic petrology, Second edition. Blackwell
Publishing.
Btard, F., Caner, L., Gunnell, Y., Bourgeon, G., 2009. Illite neoformation in plagioclase
during weathering: evidence from semi-arid Northeast Brazil. Geoderma 152,
5362.
BS5930, 1999. Code of Practice for Site Investigations. British Standards Institution,
London.
Capra, L., Macas, J.L., 2002. The cohesive Naranjo debris-ow deposit (10 km3): a dam
breakout ow derived from the Pleistocene debris-avalanche deposit of Nevado de
Colima Volcano (Mxico). Journal of Volcanology and Geothermal Research 117,
213235.
Ceryan, S., Tudes, S., Ceryan, N., 2008. A new quantitative weathering classication for
igneous rocks. Environmental Geology 55, 13191336.
Chiesa, S., Poli, S., Vezzoli, L., 1986. Studio dell'ultima eruzione storica dell'isola
d'Ischia: la colata dell'Arso-1302. Dipartimento di Scienze della Terra, Universit
di Milano, Centro Alpi Centrali, CNR, Milano.
Cipriani, F., Marianelli, P., Sbrana, A., 2008. Studio di una sequenza piroclastica del
vulcano della Solfatara (Campi egrei). Considerazioni vulcanologiche e sul sistema
di alimentazione. Atti della Societa Toscana di Scienze Naturali. Memorie, Serie A
113, 3948.
Civetta, L., Gallo, G., Orsi, G., 1991. Sr- and Nd-isotope and trace element constrains on
the chimica evolution of the magmatic system of Ischia (Italy) in the last 55 Ka.
Journal of Volcanology and Geothermal Research 46, 213230.
Civetta, L., Pappalardo, L., Fisher, R.V., Heiken, G., Ort, M., 1997. Geochemical zoning,
mingling, eruptive dynamics and depositional processes. The Campanian Ignimbrite, Campi Flegrei caldera, Italy. Journal of Volcanology and Geothermal Research
75, 183219.
Di Vito, M.A., Isaia, R., Orsi, G., Southon, J., deVita, S., D'Antonio, M., Pappalardo, L.,
Piochi, M., 1999. Volcanic and deformational history of the Campi Flegrei caldera
in the past 12 ka. Journal of Volcanology and Geothermal Research 91, 221246.
Drrast, H., Siegesmund, S., 1999. Correlation between rock fabrics and physical properties
of carbonate reservoir rocks. International Journal of Earth Sciences 88, 392408.
Duzgoren-Aydin, N.S., Aydin, A., 2003. Chemical heterogeneities of weathered igneous
proles: implications for chemical indices. Environmental and Engineering Geoscience 9
(4), 363376.
Duzgoren-Aydin, N.S., Aydin, A., 2006. Chemical and mineralogical heterogeneities of
weathered igneous proles: implications for landslide investigation. Natural Hazards
and Earth System Science 6, 315322.
Duzgoren-Aydin, N.S., Aydin, A., Malpas, J., 2002. Re-assessment of chemical weathering
indices: case study on pryroclastic of Hong Kong. Engineering Geology 63, 99119.
Eberhardt, E., Stimpson, B., Stead, D., 1999. Effects of grain size on the initiation and
propagation thresholds of stress-induced brittle fractures. Rock Mechanics and
Rock Engineering 32, 8199.
Farmer, I.W., Kemeny, J.M., McDoniel, 1991. Analysis of rock fragmentation in bench
blasting using digital image processing. Proc. Int Cong. Rock Mech. 7th ISRM Congress,
Aachen, Germany, 2, pp. 10371042.

85

Fedo, C.M., Nesbitt, H.W., Young, G.M., 1995. Unrevelling the effects of potassium metasomatism in sedimentary rocks and paleosols, with implications for paleo-weathering
conditions and provenance. Geology 23, 921924.
Finn, C.A., Sisson, T.W., Deszcz-Pan, M., 2001. Aerophysical measurements of collapsephone hydrothermally altered zones at Mount Rainier volcano. Nature 409, 600603.
Finn, C.A., Deszcz-Pan, M., Anderson, E.D., John, D.A., 2007. Three-dimensional geophysical
mapping of rock alteration and water content at Mount Adams, Washington: implications for lahar hazards. Journal of Volcanology and Geothermal Research 112, B10204.
Geiger, J., Hunyadfalvi, Z., Bogner, P., 2009. Analysis of small-scale heterogeneity in
clastic rocks by using computerized X-ray tomography (CT). Engineering Geology
103, 112118.
Germaine, J.T., Germaine, A.V., 2009. Geotechnical Laboratory Measurements for Engineers.
John Wiley & Sons, Inc., Hoboken, New Jersey. 351 pp.
Gillot, P.Y., Chiesa, S., Pasquare, G., Vezzoli, L., 1982. b33000 yr K/Ar dating of the
volcano-tectonic horst of the Isle of Ischia, Gulf of Naples. Nature 299, 242245.
Gualda, G.A.R., Rivers, M., 2006. Quantitative 3D petrography using x-ray tomography:
application to Bishop Tuff pumice clasts. Journal of Volcanology and Geothermal
Research 154, 4862.
Gupta, A.S., Rao, S.K., 2001. Weathering indices and their applicability for crystalline
rocks. Bulletin of Engineering Geology and the Environment 60, 201221.
Harnois, I., 1988. The CIW index. Sedimentary Geology 55, 319322.
Harnois, L., Moore, J.M., 1988. Geochemistry and origin of the Ore Chimney Formation,
a transported paleoregolith in the Grenville Province of Southern Ontario, Canada.
Chemical Geology 69, 267289.
Haskins, D., 2006. Chemical and mineralogical weathering indices as applied to a granite
saprolite in South Africa. The Geological Society of London, 465. IAEG.
Higgins, M.D., 1994. Determination of crystal morphology and size from bulk measurements on thin sections: numerical modelling. American Mineralogist 79, 113119.
Higgins, M.D., 2000. Measurement of crystal size distributions. American Mineralogist
85, 11051116.
Higgins, M.D., 2006. Use of appropriate diagrams to determine if crystal size distributions
(CSD) are dominantly semi-logarithmic, lognormal or fractal (scale invariant). Journal of Volcanology and Geothermal Research 154, 816.
Hudyma, N., Burcin, B., Avar, B.B., Karakouzian, M., 2004. Compressive strength and
failure modes of lithophysae-rich Topopah Spring Tuff specimens and analog
models containing cavities. Engineering Geology 73, 179190.
Inguaggio, S., Pecoraino, G., D'Amore, F., 2000. Chemical and isotopical characterization
of uid manifestations of Ischia Island (Italy). Journal of Volcanology and Geothermal
Research 99, 151178.
Irfan, T.Y., 1996. Mineralogy, fabric properties and classication of weathered granites
in Hong Kong. The Quarterly Journal of Engineering Geology 29, 535.
Irfan, T.Y., 1999. Characterization of weathered volcanic rocks in Hong Kong. The Quarterly Journal of Engineering Geology 32, 317348.
Isaia, R., D'Antonio, M., Dell'Erba, F., Di Vito, M., Orsi, G., 2004. The Astroni volcano: the
only example of closely spaced eruptions in the same vent area during the recent
history of the Campi Flegrei caldera (Italy). Journal of Volcanology and Geothermal
Research 133, 171192.
ISRM Committee on Laboratory Tests, 1972. Suggested methods for determining water
content, porosity, density, absorption and related properties and swelling and
slake-durability index properties. Document no. 2, pp. 136.
Kern, H., Liu, L., Popp, T., 1997. Relationship between anisotropy of P and S wave velocities and anisotropy of attentuation in serpentinite and amphibolite. Journal of
Geophysical Research 102 (B2), 30513065.
Ketcham, R.A., 2005. Computational methods for quantitative analysis of three-dimensional
features in geological specimens. Geosphere 1, 3241.
Ketcham, R.A., Carlson, W.D., 2001. Acquisition, optimization and interpretation of
X-ray computed tomography imagery: application to geosciences. Computers and
Geosciences 27, 381400.
Ketcham, R.A., Iturrino, G.J., 2005. Nondestructive high-resolution visualization and
measurement of anisotropic effective porosity in complex lithologies using highresolution X-ray computed tomography. Journal of Hydrology 302, 92106.
Kili, A., Teymen, A., 2008. Determination of mechanical properties of rocks using simple methods. Bulletin of Engineering Geology and the Environment 67, 237244.
Kim, S.S., Park, H.D., 2003. The relationship between physical and chemical weathering
indices of granites around Seoul, Korea. Bulletin of Engineering Geology and the
Environment 62, 207212.
Krissek, L.A., Kyle, P.R., 2001. Geochemical indicators of weathering, Cenozoic
palaeoclimates and provenance in ne-grained sediments from CRP-3, Victoria
land Basin, Antartica. Terra Antartica 8, 561568.
Luping, T.N., 1986. A study of the quantitative relationship be-tween strength and pore
size distribution of porous materials. Cement and Concrete Research 16, 8796.
Martnez-Martnez, J., Benavente, D., Garca del Cura, M.A., Caaveras, J.C., 2006. Applications of ultrasonic to brecciated dolostones for assessing their mechanical properties. IAEG2006 Paper number 243.
Martnez-Martnez, J., Benavente, D., Garca del Cura, M.A., 2007. Petrographic quantication of brecciated rocks by image analysis. Application to the interpretation of
elastic wave velocities. Engineering Geology 90, 4154.
Martnez-Martnez, J., Benavente, D., Garca del Cura, M.A., 2011. Spatial attenuation:
the most sensitive ultrasonic parameter for detecting petrographic features and
decay processes in carbonate rocks. Engineering Geology 119 (34), 8495.
Mess, F., Swennen, R., Van Geet, M., Jacops, P., 2003. Applications of X-ray computed tomography in the geoscience. Geological Society, special publication, 215.
Metz, F., Knfel, D., 1992. Systematic mercury porosimetry investigations on sandstones. Materials and Structures 25, 127136.
Morgan, D.J., Jerram, D.A., 2006. On estimating crystal shape for crystal size distribution analysis. Journal of Hydraulic. Journal of Volcanology and Geothermal Research 154, 17.

Author's personal copy


86

A. Pola et al. / Tectonophysics 566-567 (2012) 6786

Naik, T.R., Malhotra, V.M., Popovics, J.S., 2004. The ultrasonic pulse velocity method.
Handbook on Non-Destructive Testing of Concrete. CRC Press. 170 pp.
Nappi, G., Renzulli, E., Santi, P., 1991. Evidence of incremental growth in the Vulsinian
calderas. Journal of Volcanology and Geothermal Research 47, 1331.
Nappi, G., Capaccioni, B., Mattioli, M., Mancini, E., Valentini, L., 1994. Plinian fall deposits from Vulsini Volcanic District. Central Italy. Bulletin of Volcanology 56,
502515.
Nappi, G., Renzulli, A., Santi, P., Gillot, P.Y., 1995. Geological evolution and geochronology of the Vulsini Volcanic District, Central Italy. Bollettino della Societ Geologica
Italiana 114, 599613.
Nappi, G., Antonelli, F., Coltorti, M., Milani, L., Renzulli, A., Siena, F., 1998. Volcanological
and petrological evolution of the Eastern Vulsini District, Central Italy. Journal of
Volcanology and Geothermal Research 87, 211232.
Nesbitt, H.W., Young, G.M., 1982. Early Proterozoic climates and plate motions inferred
from major element chemistry of lutites. Nature 199, 715717.
Nimick, F.B., 1988. Empirical relationships between porosity and the mechanical properties of tuff. In: Cudall, Sterling, Stareld (Eds.), Questions in Rock Mechanics.
Balkema, Rotterdam, pp. 741742.
Ohta, T., Arai, H., 2007. Statistical empirical index of chemical weathering in igneous
rocks: a new tool for evaluating the degree of weathering. Chemical Geology
240, 280297.
Orsi, G., Gallo, G., Zanchi, A., 1991. Simple shearing block-resurgence in caldera depressions. A model from Pantelleria and Ischia. Journal of Volcanology and Geothermal
Research 47, 111.
Orsi, G., D'Antonio, M., de Vita, S., Gallo, G., 1992. The Neapolitan Yellow Tuff, a largemagnitude trachytic phreatoplinian eruption: eruptive dynamics, magma withdrawal and caldera collapse. Journal of Volcanology and Geothermal Research 53,
275287.
Orsi, G., Civetta, L., D'Antonio, M., Di Girolamo, P., Piochi, M., 1995. Step-lling and development of a three-layers magma chamber: the Neapolitan Yellow Tuff case history. Journal of Volcanology and Geothermal Research 67, 291312.
Orsi, G., de Vita, S., Di Vito, M., 1996. The restless, resurgent Campi Flegrei Nested Caldera
(Italy): constraints on its evolution and conguration. Journal of Volcanology and
Geothermal Research 74, 179214.
Parker, A., 1970. An index of weathering for silicate rocks. Geological Magazine 107,
501504.
Patonin, A.V., 2003. Anisotropy of rock drill core elastic properties at uniaxial loading.
Russian Journal of Earth Sciences 5 (4), 299306.
Price D.G., 2009. Engineering Geology principles and practice. Edited and compiled by
M.H. de Freitas, Springer-Verlang Berlin Heidelberg. 450 pp.
Price, J.R., Velbel, M.A., 2003. Chemical weathering indices applied to weathering proles
developed on heterogeneous felsic metamorphic parent rock. Chemical Geology 202,
397416.
Price, R.H., Boyd, P.J., Noel, J.S., Martin, R.J., 1994. Relationship between static and dynamic
properties in welded and non-welded tuff. In: Nelson, P.P., Laubach, S.E. (Eds.), Rock
Mechanics: Models and Measurements Challenges from Industry. Proceedings of the
First North American Rock Mechanics Symposium. A.A. Balkema, pp. 505512.
Proussevitch, A.A., Sahagian, D.L., Jutzeler, M., 2010. Functional stereology for 3D particle size distributions from 2D Observations: a practical approach. AGU Fall Meeting
Abstracts, 11, p. 2332.
Reid, M.E., Sisson, T.W., Brien, D.L., 2001. Volcano Collapse promoted by hydrothermal
alteration and edice shape, Mount Rainier, Washington. Geological Society of
America 9 (32), 373376.

Roels, S., Elsen, J., Carmeliet, J., Hens, H., 2001. Characterisation of pore structure by
combining mercury porosimetry and micrography. Materials and Structures 34,
7682.
Rotonda, T., Tommasi, P., Boldini, D., 2010. Geomechanical characterization of the
volcaniclastic material involved in the 2002 landslide at Stromboli. Journal of Geotechnical and Geoenvironmental Engineering 136 (2), 389401.
Ruxton, B.P., 1968. Measures of the degree of chemical weathering of rocks. Journal of
Geology 76, 518527.
Sahagian, D.L., Proussevitch, A.A., 1998. 3D particle size distributions from 2D observations:
stereology for natural applications. Journal of Volcanology and Geothermal Research 84,
173196.
Sbrana, A., Fulignati, P., Marianelli, P., Boyce, A.J., Cecchetti, A., 2010. Exhumation of an
active magmatic hydrothermal system in a sesurgent caldera environment: the example of Ischia (Italy). Journal of the Geological Society 166, 10611073.
Shea, T., Houghton, B.F., Gurioli, L., Cashman, K.V., Hammer, J.E., Hobden, B.J., 2010. Textural studies of vesicles in volcanic rocks: an integrated methodology. Journal of
Volcanology and Geothermal Research 190 (34), 271289.
Sousa, L.M.O., del Rio, L.M.S., Calleja, L., Argandoa, V.G.R., Rey, A.R., 2005. Inuence of
microfractures and porosity on the physico-mechanical properties and weathering
of ornamental granites. Engineering Geology 77, 153168.
Stanchits, S.A., Lockner, D.A., Ponomareva, A.V., 2003. Anisotropic changes in P-wave
velocity and attenuation during deformation and uid inltration of granite. Bulletin of the Seismological Society of America 93 (4), 18031822.
Taud, H., Martinez-Angelesa, T.R., Parrot, J.F., Hernandez-Escobedo, L., 2005. Porosity
estimation method by X-ray computed tomography. Journal of Petroleum Science
and Engineering 47, 209217.
Tillerson, J.R., Nimick, F.B., 1984. Geoengineering properties of potential repository units
at Yucca Mountain, southern Nevada. Sandia National Labs Report, SAND84-0221.
Turcotte, D.L., 1992. Fractal and Chaos in Geology and Geophysics. Cambridge University Press, New York. 412 pp.
Vinciguerra, S., Trovato, C., Meredith, P.G., Benson, P.M., Troise, C., De Natale, G., 2006.
Understanding the seismic velocity structure of Campi Flegrei Caldera (Italy): from
laboratory to the eld scale. Pure and Applied Geophysics 163 (10), 22052221.
Vinciguerra, S., Del Gaudio, P., Mariucci, M.T., Marra, F., Meredith, P.G., Montone, P.,
Pierdominici, S., Scarlato, P., 2009. Physical properties of tuffs from a scientic
borehole at Alban hills volcanic. Tectonophysics 471, 161169.
Wardlaw, N.C., McKellar, M., 1981. Mercury porosimetry and the Interpretation of pore
geometry in sedimentary rocks and articial models. Powder Technology 29,
127143.
Wulff, A.M., Raab, S., Huenges, E., 2000. Alteration of seismic wave properties and uid
permeability in sandstones due to microfracturing. Physics and Chemistry of the
Earth, Part A: Solid Earth and Geodesy 25 (2), 141147.
Xu, X., Hofmann, R., Batzle, M., Tshering, T., 2006. Inuence of pore pressure on velocity in
low-porosity sandstone: implications for time-lapse feasibility and pore-pressure
study. Geophysical Prospecting 54, 565573.
Zamora, M., Sartori, G., Chelini, W., 1994. Laboratory measurements of ultrasonic wave
velocities in rocks from the Campi Flegrei volcanic system and their relation to
other eld data. Journal of Geophysical Research B7, 1355313562.

Вам также может понравиться