Complex conductivity of volcanic rocks and the geophysical mapping of alteration in volcanoes

https://doi.org/10.1016/j.jvolgeores.2018.04.014Get rights and content

Highlights

  • Induced polarization is sensitive to the alteration of volcanic rocks.

  • Induced polarization of volcanic rocks can be understood in terms of a dynamic Stern layer model.

  • The effect of temperature and pyrite (or magnetite) content can be understood in terms of a simple model.

Abstract

Induced polarization measurements can be used to image alteration at the scale of volcanic edifices to a depth of few kilometers. Such a goal cannot be achieved with electrical conductivity alone, because too many textural and environmental parameters influence the electrical conductivity of volcanic rocks. We investigate the spectral induced polarization measurements (complex conductivity) in the frequency band 10 mHz–45 kHz of 85 core samples from five volcanoes: Merapi and Papandayan in Indonesia (32 samples), Furnas in Portugal (5 samples), Yellowstone in the USA (26 samples), and Whakaari (White Island) in New Zealand (22 samples). This collection of samples covers not only different rock compositions (basaltic andesite, andesite, trachyte and rhyolite), but also various degrees of alteration. The specific surface area is found to be correlated to the cation exchange capacity (CEC) of the samples measured by the cobalthexamine method, both serving as rough proxies of the hydrothermal alteration experienced by these materials. The in-phase (real) conductivity of the samples is the sum of a bulk contribution associated with conduction in the pore network and a surface conductivity that increases with alteration. The quadrature conductivity and the normalized chargeability are two parameters related to the polarization of the electrical double layer coating the minerals of the volcanic rocks. Both parameters increase with the degree of alteration. The surface conductivity, the quadrature conductivity, and the normalized chargeability (defined as the difference between the in-phase conductivity at high and low frequencies) are linearly correlated to the CEC normalized by the bulk tortuosity of the pore space. The effects of temperature and pyrite-content are also investigated and can be understood in terms of a physics-based model. Finally, we performed a numerical study of the use of induced polarization to image the normalized chargeability of a volcanic edifice. Induced polarization tomography can be used to map alteration of volcanic edifices with applications to geohazard mapping.

Introduction

Geophysical methods and the methods developed in the framework of hydrogeophysics are increasingly used on volcanoes to determine the distribution of material properties (e.g., seismic velocities, bulk density, electrical properties) with application to volcanic eruption forecasting, geohazard mapping, and hydrogeology. Electrical conductivity tomography can be used to image the 3D distribution of the electrical conductivity of volcanoes (e.g., Johnson et al., 2010; Revil et al., 2010; Rosas-Carbajal et al., 2016; Gresse et al., 2017). Electrical conductivity of volcanic rocks depends on two contributions. The first is a bulk contribution controlled by the water content and pore water salinity. It corresponds to electrical conduction through the connected pore network of the material (e.g., Archie, 1942). The second contribution is an interfacial contribution called surface conductivity. This contribution is associated with conduction in the electrical double layer coating the mineral grains (e.g., Revil et al., 2002 for volcanic rocks and Waxman and Smits, 1968 for sedimentary materials). This electrical double layer consists of the Stern layer (with weakly or strongly adsorbed counterions depending on their affinity for the mineral surface) and the diffuse layer. Since electrical conductivity of volcanic rocks depends on these two contributions, electrical conductivity tomography is rather difficult to interpret and therefore cannot be used as a stand-alone geophysical method (see discussion in Bernard et al., 2007; Komori et al., 2010; Kemna et al., 2012; Usui et al., 2016; Soueid Ahmed et al., 2018a, Soueid Ahmed et al., 2018b). For instance, a volcano can be a highly conductive body because of the high salinity of the pore water or a high degree of alteration (or both). Induced polarization can be used to discriminate these effects as long as the content in metallic particles (e.g., pyrite and magnetite) is not too high (i.e., 1% in vol.).

In the context of the present paper, it may be useful to recall what we mean by alteration. Classically, the alteration of volcanic rocks is produced by the circulation of hydrothermal fluids and involves the replacement of primary igneous glass and minerals (e.g., plagioclase, pyroxene, amphibole) by secondary minerals that are stable at the thermodynamic conditions of alteration (e.g., Bonnet and Corriveau, 2007). We are especially interested by the case where these secondary minerals are clay minerals (kaolinite, chlorite, illite, and smectite, see Honnorez et al., 1998). Note that some volcanic rocks can also be altered through surface weathering.

Complex conductivity characterizes the reversible storage of electrical charges in rocks (Schlumberger, 1920; Bleil, 1953; Seigel, 1959), a process known as (induced) polarization. This “polarization” is a low frequency (<10 kHz) characteristic of rocks that is unrelated to the dielectric polarization phenomena observed at higher frequencies (>10 kHz) (e.g., Revil, 2013). The origin of the low-frequency polarization of rocks is generally associated with the polarization of the electrical double layer around the mineral grains and can be described with two interrelated parameters: the quadrature conductivity and the normalized chargeability (see Revil et al., 2017b). The quadrature conductivity corresponds to the imaginary component of the complex conductivity. The normalized chargeability measures the dispersion of the in-phase conductivity with the frequency. Since the polarization and surface conductivity are both controlled by the properties of the electrical double layer, it is therefore unsurprising that these parameters are also interrelated (Revil et al., 2017b). This relationship is very important to interpret electrical resistivity and induced polarization tomographies that can be carried out at the scale of volcanic structures. As a side note, measurements of induced polarization can be performed in the field with the same equipment used for electrical resistivity tomography (e.g., Kemna et al., 2012). Induced polarization may therefore appear as an attractive method to study volcanic and geothermal systems.

The key questions we want to address are (1) Is the relationship developed in a previous paper for basaltic rocks (Revil et al., 2017a) valid for all types of volcanic rocks? (2) Is the cation exchange capacity (CEC) a proxy of the alteration of the volcanic rocks? (3) How do the surface and quadrature conductivities of volcanic rocks depend on their cation exchange capacity (CEC) and specific surface area? (4) How can the quadrature conductivity and the normalized chargeability be related to each other? (5) How is the polarization affected by temperature and the volume content of metallic particles? (6) Can we image electrical conductivity and normalized chargeability of a volcanic edifice and offer a combined approach to interpret these data? Answers to these questions are important to use induced polarization tomography to image the alteration of volcanic rocks. Since alteration is responsible for the weakening of the mechanical properties of volcanic rocks (see Pola et al., 2012; Frolova et al., 2014; Wyering et al., 2014; Heap et al., 2015), our study has strong implications regarding the mapping of geohazards in volcanic environments (Day, 1996; Reid et al., 2001; Finn et al., 2001; Reid, 2004).

We investigate here the complex conductivity of a set of 85 new core samples from five active volcanoes in the world: Merapi and Papandayan in Indonesia, Furnas in Portugal, Yellowstone in the USA, and Whakaari (White Island) in New Zealand. The laboratory measurements were collected in the frequency range 10 mHz–45 kHz with a very sensitive impedance meter. The interpretation of the laboratory data will be based on the dynamic Stern layer polarization model developed in Revil (2013) and recently updated in Revil et al., 2017a, Revil et al., 2017b, Revil et al., 2018.

Section snippets

The dynamic stern layer model

The (in-phase) conductivity of a volcanic rock represents its ability to carry electrical current. A volcanic rock can also reversibly store electrical charges (Revil et al., 2017c), a phenomenon known as polarization. This polarization is responsible for a phase lag between the electrical current and the electrical field for frequency-domain measurements (Kemna et al., 2012) or a secondary voltage for time-domain measurements (Schlumberger, 1920). The polarization phenomenon is represented

Merapi and Papadayan volcanoes, Indonesia

The lavas from Papandayan (Fig. 3) are mainly basaltic andesite, pyroxene andesite, and pyroxene dacite (Asmoro et al., 1989). The Papandayan samples contain pyroclasts with clinopyroxene crystals with simple core-rim zoning (typically ~ 5–15 vol%), euhedral magnetite crystals, unresorbed amphiboles, and iron oxides. Petrological studies on Papandayan lavas show that the most abundant alteration minerals in the altered materials are polymorphs of silica, pyrite, pyrophyllite, natroalunite, and

Induced polarization experiments

We use a four electrodes technique to perform the complex conductivity measurements, i.e., we separate the current electrodes A and B from the voltage electrodes M and N (which is a classical notation in geophysics, see for instance Herman, 2001; Michot et al., 2016). The complex conductivity measurement was conducted with a high-precision impedance analyzer (Zimmermann et al., 2008) (Fig. 11a). Two different sample holders are used for the consolidated and unconsolidated samples, the same

Specific surface area versus cation exchange capacity

Fig. 13 shows the specific surface area data as a function of the cation exchange capacity for the volcanic rocks investigated in the present study. Fig. 13 also includes data from volcanic rocks from Hawaii and tight sandstones from Revil et al. (2018). For volcanic rocks, it is known that specific surface area measurements can be used as a proxy of alteration (see Nielsen and Fisk, 2008). The pertinent question is whether the CEC can also be considered as a proxy of alteration.

In our study,

Application to a synthetic tomographic test

In this section, we explain how induced polarization tomography can assist electrical conductivity tomography and show the type of images the method can produce for a volcano. As mentioned in Section 2, the electrical conductivity has two contributions (see Eq. (5) and Fig. 14): a bulk contribution associated with conduction in the pore network and surface contribution associated with the conduction path in the electrical double layer. Time-domain induced polarization can be used to image the

Conclusions

We investigated the complex conductivity of a set of 85 volcanic rock samples from five volcanoes located in Indonesia, Portugal, USA, and New Zealand in order to provide answers to 6 fundamental questions. We recall here these questions and we summarize their answers found in this work.

  • (1)

    Is the relationship developed in a previous paper for basaltic rocks valid for all types of volcanic rocks? The response is clearly yes as the new data agree with the results obtained on basalts (from Hawaii)

Acknowledgements

M. J. Heap thanks the Buttle family, Pee Jay Tours, GNS Science, and all those that helped collect the materials from Whakaari volcano (particularly B. Kennedy). We thank the USGS Core research Center in Denver, Colorado for allowing us to sample Yellowstone drill cores Y-2 and Y-8. The field samples at Yellowstone were collected under the research permit YELL-2016-SCI-7006. The research at Yellowstone was funded by Labex OSUG@2020 (ANR10 LABX56) and CNRS-INSU program SYSTER. We thank François

References (86)

  • G.M. Marmoni et al.

    Gravitational slope-deformation of a resurgent caldera: new insights from the mechanical behaviour of Mt. Nuovo tuffs (Ischia Island, Italy)

    J. Volcanol. Geotherm. Res.

    (2017)
  • K. Mayer et al.

    Experimental constraints on phreatic eruption processes at Whakaari (White Island volcano)

    J. Volcanol. Geotherm. Res.

    (2015)
  • V. Moon et al.

    Geotechnical characterisation of stratocone crater wall sequences, White Island volcano, New Zealand

    Eng. Geol.

    (2005)
  • H. Ohmoto

    Formation of volcanogenic massive sulfide deposits—the Kuroko perspective

  • A. Pola et al.

    Influence of alteration on physical properties of volcanic rocks

    Tectonophysics

    (2012)
  • A. Revil et al.

    Alteration of volcanic rocks: a new non-intrusive indicator based on induced polarization measurements

    J. Volcanol. Geotherm. Res.

    (2017)
  • T. Shea et al.

    Textural studies of vesicles in volcanic rocks: an integrated methodology

    J. Volcanol. Geotherm. Res.

    (2010)
  • B. Voight et al.

    Historical eruptions of Merapi volcano, central Java, Indonesia, 1768–1998

    J. Volcanol. Geotherm. Res.

    (2000)
  • L.D. Wyering et al.

    Mechanical and physical properties of hydrothermally altered rocks, Taupo Volcanic Zone, New Zealand

    J. Volcanol. Geotherm. Res.

    (2014)
  • G.E. Archie

    The electrical resistivity log as an aid in determining some reservoir characteristics. Petroleum

    Trans. AIME

    (1942)
  • P. Asmoro et al.

    Geological Map of Papandayan Volcano, Garut, West Java, Scale 1:25.000

    (1989)
  • J.L. Ball et al.

    The hydrothermal alteration of cooling lava domes

    Bull. Volcanol.

    (2015)
  • K.E. Bargar et al.

    Hydrothermal alteration in research drill hole Y-2, Lower Geyser Basin, Yellowstone National Park, Wyoming

    Am. Mineral.

    (1981)
  • M.H. Beeson et al.

    Major and Trace Element Analyses of Drill Cores From Thermal Areas in Yellowstone National Park, Wyoming (No. 84-373)

    (1984)
  • M.-L. Bernard et al.

    Transport properties of pyroclastic rocks from Montagne Pelée volcano (Martinique, Lesser Antilles)

    J. Geophys. Res.

    (2007)
  • D.F. Bleil

    Induced polarization, a method of geophysical prospecting

    Geophysics

    (1953)
  • A.-L. Bonnet et al.

    Alteration vectors to metamorphosed hydrothermal systems in gneissic terranes

  • F.D. Börner

    Complex conductivity measurements of reservoir properties

  • S. Brunauer et al.

    Adsorption of gasses in multimolecular layers

    J. Am. Chem. Soc.

    (1938)
  • R.L. Christiansen

    The Quaternary and Pliocene Yellowstone Plateau Volcanic Field of Wyoming, Idaho, and Montana (No. 729-G)

    (2001)
  • H. Ciesielski et al.

    Determination of exchange capacity and exchangeable cations in soils by means of cobalt hexamine trichloride. Effects of experimental conditions

    Agronomie

    (1997)
  • S.C. Cook et al.

    Engineering geology model of the Crater Lake outlet, Mt. Ruapehu, New Zealand, to inform rim breakout hazard

    J. Volcanol. Geotherm. Res.

    (2017)
  • S.J. Day

    Hydrothermal pore fluid pressure and the stability of porous, permeable volcanoes

    Geol. Soc. Lond., Spec. Publ.

    (1996)
  • S. Erdmann et al.

    Amphibole as an archivist of magmatic crystallization conditions: problems, potential, and implications for inferring magma storage prior to the paroxysmal 2010 eruption of Mount Merapi, Indonesia

    Contrib. Mineral. Petrol.

    (2014)
  • S. Erdmann et al.

    Constraints from phase equilibrium experiments on pre-eruptive storage conditions in mixed magma systems: a case study on crystal-rich basaltic andesites from Mount Merapi, Indonesia

    J. Petrol.

    (2016)
  • C.A. Finn et al.

    Aerogeophysical measurements of collapse-prone hydrothermally altered zones at Mount Rainier volcano

    Nature

    (2001)
  • J. Frolova et al.

    Effects of hydrothermal alterations on physical and mechanical properties of rocks in the Kuril–Kamchatka island arc

    Eng. Geol.

    (2014)
  • W. Giggenbach et al.

    Formation of acid volcanic brines through interaction of magmatic gases, sea-water, and rock within the White Island volcanic-hydrothermal system, New Zealand

  • M. Gresse et al.

    3-D resistivity tomography of the Solfatara crater (Italy): implication for the multiphase flow structure of the shallow hydrothermal system

    J. Geophys. Res.

    (2017)
  • J.E. Guest et al.

    The volcanic history of Furnas Volcano, São Miguel, Azores

  • M.J. Heap et al.

    Microstructural controls on the physical and mechanical properties of edifice-forming andesites at Volcán de Colima, Mexico

    J. Geophys. Res. Solid Earth

    (2014)
  • J.W. Hedenquist et al.

    White Island, New Zealand, volcanic-hydrothermal system represents the geochemical environment of high-sulfidation Cu and Au ore deposition

    Geology

    (1993)
  • R. Herman

    An introduction to electrical resistivity in geophysics

    Am. J. Phys.

    (2001)
  • Cited by (60)

    • Electromagnetic field in a conducting medium modeled by the fractional Ohm's law

      2022, Communications in Nonlinear Science and Numerical Simulation
    • Induced polarization images alteration in stratovolcanoes

      2022, Journal of Volcanology and Geothermal Research
      Citation Excerpt :

      All these andesitic core samples were collected on outcrops based on their distinct porosity and alteration levels (Table 1). The samples for Papandayan are the same than those used in Ghorbani et al. (2018) but were measured at 6 salinities to further check the linear character of the dependence between the conductivity of the samples and the conductivity of the pore water. Samples from la Soufrière are lavas and debris flow deposits (Table 1) from fresh states to samples exhibiting different degrees of alteration.

    View all citing articles on Scopus
    View full text