fault zone hydrogeology

Upload: camilo-matta-torres

Post on 10-Jan-2016

30 views

Category:

Documents


0 download

DESCRIPTION

hydrogeologi

TRANSCRIPT

  • Available online 10 October 2013

    Keywords:Fault zoneHydrogeologyStructural geology

    the shallow crust (b1 km) introduces permeability heterogeneity and anisotropy,

    1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 172

    . . . . . . . . . 180

    . . . . . . . . . 182

    . . . . . . . . . 182

    Earth-Science Reviews 127 (2013) 171192

    Contents lists available at ScienceDirect

    Earth-Science Reviews4.3. Volcanic rock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1824.4. Carbonate rock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 182

    5. Hydrogeological evidence of the impact of faults on uid ow. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1834. Conceptual geological models of fault zone hydrogeology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.1. Unlithied- and lithied siliciclastic rock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .4.2. Crystalline rock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .2. Approaches to fault zone hydrogeology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1723. Fault zone processes from surface-focused studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 173

    3.1. Terminology and fault zone architecture . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1733.2. Impact of fault zone processes on permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 175

    3.2.1. Fault zone processes that enhance permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1753.2.2. Fault zone processes that reduce permeability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1763.2.3. Interacting, secondary or larger-scale processes that either reduce or enhance fault zone permeability . . . . . . . . . . . . . . 1795.1. Unlithied- and lithied siliciclastic5.2. Crystalline and volcanic rock . . .5.3. Carbonate rock . . . . . . . . .

    Corresponding author.E-mail address: [email protected] (V.F. Bense).

    0012-8252/$ see front matter. Crown Copyright 2013http://dx.doi.org/10.1016/j.earscirev.2013.09.008Contentsdrothermal uid circulation. Fault zones have the capacity to be hydraulic conduits connecting shallow and deepgeological environments, but simultaneously the fault cores of many faults often form effective barriers to ow.The direct evaluation of the impact of faults to uid ow patterns remains a challenge and requires a multidisci-plinary research effort of structural geologists and hydrogeologists. However, we nd that these disciplines oftenuse different methods with little interaction between them. In this review, we document the current multi-disciplinary understanding of fault zone hydrogeology.We discuss surface- and subsurface observations fromdi-verse rock types from unlithied and lithied clastic sediments through to carbonate, crystalline, and volcanicrocks. For each rock type, we evaluate geological deformationmechanisms, hydrogeologic observations and con-ceptualmodels of fault zone hydrogeology. Outcrop observations indicate that fault zones commonly have a per-meability structure suggesting they should act as complex conduitbarrier systems in which along-fault ow isencouraged and across-fault ow is impeded. Hydrogeological observations of fault zones reported in the litera-ture show a broad qualitative agreement with outcrop-based conceptual models of fault zone hydrogeology.Nevertheless, the specic impact of a particular fault permeability structure on fault zone hydrogeology can onlybe assessed when the hydrogeological context of the fault zone is considered and not from outcrop observationsalone. To gain a more integrated, comprehensive understanding of fault zone hydrogeology, we foresee numeroussynergistic opportunities and challenges for the discipline of structural geology and hydrogeology to co-evolveand address remaining challenges by co-locating study areas, sharing approaches and fusing data, developingconceptual models from hydrogeologic data, numerical modeling, and training interdisciplinary scientists.

    Crown Copyright 2013 Published by Elsevier B.V. All rights reserved.Received 6 September 2011Accepted 27 September 2013

    which has an important impact on processes such as regional groundwaterow, hydrocarbonmigration, and hy-

    Article history: Deformation along faults ina b s t r a c ta r t i c l e i n f oFault zone hydrogeology

    V.F. Bense a,, T. Gleeson b, S.E. Loveless a, O. Bour c, J. Scibek d

    a School of Environmental Sciences, University of East Anglia, Norwich NR4 7TJ, England, United Kingdomb Civil Engineering, McGill University, Montral, QC H3A 2K6 Canadac Gosciences Rennes, UMR 6118 CNRS, Universit de Rennes 1, Campus de Beaulieu, 35042 Rennes, Franced SRK Consulting Inc., 22nd Floor, 1066 West Hastings Street, Vancouver, BC V6E 3X2, Canada

    j ourna l homepage: www.e lsev ie r .com/ locate /earsc i revrock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 184. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

    Published by Elsevier B.V. All rights reserved.

  • ..

    .

    .

    .

    .

    .

    We focus on fault processes and single-phase groundwater owpat- rect hydrogeological evidence of the impacts of particular faults on uid

    172 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192terns in the shallow crust after fault deformation. At greater depths frac-tured rocks will contain predominantly saline groundwaters and/orhydrocarbon accumulations and density-dependent and multi-phaseow is beyond the scope of this review. At greater depths, seismogenicdeformation and other deformation mechanisms beyond our scope arealsomore important. Althoughwe do not discuss the impact of faults onow in hydrocarbon reservoirs in much detail, in Section 5 we includehydrocarbon reservoir models which in many cases could be directlytransferable to shallow single-phase groundwater systems. By focusing

    ow to test and rene these outcrop-based fault models is often lacking.Subsurface-focused studies, often carried out by hydrogeologists,

    infer the hydrogeological behavior of fault zones from arrays of bore-holes or springs usually without observing fault structure directly inoutcrop (Fig. 1b). Often groundwater hydrologists need to be opportun-ists in the study of fault zones as groundwater monitoring networks aredesigned to characterize the hydrodynamics of aquifers as a wholewithout any specic focus on fault zones. Hence, closely spaced arraysof boreholes over fault zones are rare. However, the shape of the hy-6. Modeling uid ow in fault zones. . . . . . . . . . . . . . . . . .7. Towards interdisciplinary fault zone hydrogeology . . . . . . . . . .

    7.1. Controls on uid ow around fault zones . . . . . . . . . . .7.2. The tale of two disciplines . . . . . . . . . . . . . . . . . .7.3. The future tale of one inter-discipline? . . . . . . . . . . . .

    Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . .References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

    1. Introduction

    As the plumbing of the earth, fault zones in the shallow crust(b1 km) impact a suite of geological processes. Faults affect uid owpatterns in groundwater aquifers (e.g., Levens et al., 1994; Mayeret al., 2007; Bense et al., 2008; Burbey, 2008; Folch and Mas-Pla,2008), and hydrocarbon migration and entrapment in reservoir rocks(e.g., Aydin, 2000; Sorkhabi and Tsuji, 2005) as well as the safe storageof nuclear waste (Bredehoeft, 1997; Douglas et al., 2000; Mal'kovskiiand Pek, 2001; Ofoegbou et al., 2001), and CO2 sequestration (e.g.,Shipton et al., 2004; Agosta et al., 2008; Dockrill and Shipton, 2010;Kampman et al., 2012; Tueckmantel et al., 2012b). Fluid expulsion andlocalized mineralization along faults can lead to the formation of eco-nomic mineral deposits (e.g., Deming, 1992; Garven et al., 1999;Person et al., 2008), and provides evidence of enhanced past uid circu-lation along faults both on Earth (e.g., Mozley and Goodwin, 1995;Heynekamp et al., 1999; Caine and Minor, 2009; Balsamo et al., 2013)and on Mars (Treiman, 2008). Mineralization associated with uidow and waterrock interaction can impact the mechanical strengthof faults, potentially affecting the character of fault slip in response toearthquakes in the deeper crust (e.g., Moore and Rymer, 2007;Carpenter et al., 2011). The specic impact of faults on groundwaterow in different geological environments is complex, diverse andoften not well understood despite the relevance of understanding uidow around fault zones at shallow crustal depths which is important tonumerous societal concerns.

    The fundamental law of uid ow through porous media, Darcy'sLaw, indicates that subsurface uid ow is controlled by a combinationof rock permeability and the hydraulic gradient within the rock mass.Examining deformation processes as well as the internal structure andarchitecture of fault zone structure using tools from structural geologyis critical to obtain a primary understanding of the permeability struc-ture of fault zones. However, hydraulic gradients around fault zonesare strongly controlled by hydrogeologic processes such as rates ofgroundwater recharge forcing topography-driven ow, anthropogenicinuences such as extraction of groundwater, and deeper processeslike uid ow driven by sediment compaction. Hence, examining hy-draulic gradients present near fault zones and delineation of uid owpaths in fault zones using tools from hydrogeology should also be cen-tral in a development of a comprehensive understanding of the role offault zones in hydrogeology (Karasaki et al., 2008). Though there is aconsiderable body of research on faults and uid ow in both hydroge-ology and structural geology disciplines, there is limited evidence of anexchange of insights or integration between them.on fault zones post-deformation, we are excluding the role of fault. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 185

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 186

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 187

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

    . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 188

    evolution on uid ow in active fault zones. For simplicity, in thispaper we use the term rock for both lithied and unlithied materials.In the context of fault zone hydrogeology we consider a fault zone tobe the volume of rock where permeability has been altered by fault-related deformation. The protolith is the undeformed geological materi-al surrounding the fault zone.

    Our objective in this paper is to propel research on fault zone hydro-geology forward by providing a comprehensive overview of the study ofuid ow in and around fault zones at shallow depths (b1 km) wheresurface and subsurface data ismost abundant.We discuss fault zone hy-drogeology as studied from a structural geological and hydrogeologicalviewpoint in Section 2. In Section 3 we discuss the geological processesand hydrogeological characteristics of fault zones derived from thestudy of rock outcrops, in a range of geological settings. Section 4 dis-cusses the conceptual models of uid ow in fault zones as derivedfrom outcrop data. Section 5 reviews hydrogeological evidence of theimpact of faults in a similar range of geological settings. Studies thathave employed models as integrative tools to describe uid ow pat-terns around fault zones are discussed in Section 6. Finally, we discussdifferent strategies to develop a more comprehensive and integratedframework for the investigation of uid ow around fault zones in theshallow crust (Section 7). Throughout the text terminology is boldedwhen it is rst dened.

    2. Approaches to fault zone hydrogeology

    Hydrogeologists and structural geologists use differentmethods andapproaches to studying fault zone hydrogeology (Fig. 1). Explicitlyreviewing the purpose, data required, limitations and advantages ofeach individual technique is beyond the scope of this paper but hasbeen summarized in Table 1. Here we provide an overview of the differ-ent approaches and how these are impacted by data availability, eldareas and scale.

    Surface-focused studies, mostly by structural geologists, use tech-niques such as outcrop mapping of fault zone attributes, for examplethe length, orientation and aperture of fractures within the fault zone,the fault rock grain size and porosity, and permeametry to obtain thepermeability of fault rocks (Fig. 1a). Outcrop data can be used in numer-ical owmodels of fault zones (Section 6). From outcrop-based studiesa set of conceptual hydrogeological models have been developed in theliterature (outlined in Section 3.2) that suggest that faults act as barriershampering uid migration, as conduits propagating the movement ofuids or as more complex conduitbarrier systems (Caine et al., 1996;Aydin, 2000; Rawling et al., 2001; Bense and Person, 2006). However, di-draulic head prole that might show steps or inections in hydraulic

  • 173V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192gradient at fault zones can be used to infer the directions and rate of uidow at the fault zone (e.g. Haneberg, 1995; Bense et al., 2003a; Andersonand Bakker, 2008) and provide an indication of the hydrogeological be-havior of a fault zone as a barrier or a conduit (Fig. 2a). Groundwatertemperature, geochemistry and age (e.g., Bethke and Johnson, 2008;Leray et al., 2012) can constrain groundwater ow paths across faultzones in conjunction with hydraulic head observations (Fig. 2b and c).Boreholes can also be used for hydraulic testing with packer or pumpingtests to enhance groundwater ow rates around fault zones to elucidatethe fault zone hydrogeological structure (Anderson and Bakker, 2008).Sampling groundwater geochemistry, measuring hydraulic head, tem-perature and other parameters during pumping tests can result in de-tailed inferences on fault zone hydrogeological properties (Roqueset al., submitted for publication).

    Differences in themethodologies used by hydrogeologists and struc-tural geologists is partly due to contrasts in the type of data available ineld areas (Fig. 1). Dense networks of wells available to collect detailedhydrogeological evidence on fault behavior are typically found in devel-oped areas where outcrops that are necessary for structural geologicaloutcrop-based studies are usually sparse. Vice versa, where fault zonesare well exposed in outcrop, a dense network of wells is often not avail-able as these areas are generally less developed and/or the groundwateris not near the surface. Therefore, probably in part due to the differencesinmethods, data availability and eld areas, we nd that there is limitedevidence of an exchange of insights or integration between hydrogeol-ogy and structural geological disciplines.

    Eachmethod used by structural geologists and hydrogeologists has acharacteristic scale over which the method integrates observations(Fig. 3). Consequently, the use of the different methods results in faults

    Fig. 1. Investigation of fault zone hydrogeology has often focused on (a) characterization of surfainformation on geology anddirect hydrogeological evidence of the impact of faults on uidowtions and methods. Table 1 lists examples of studies employing the different modes of invesdimensional surface exposure (i.e. at mine sites or in mountainous topography) and some subbeing examined at various scales by the two disciplines. Generally, themethods used by structural geologists examine smaller integrationscales of b0.01 m (microstructure) to e.g. 100 m (outcrop studies)whereas hydrogeologists often also infer fault hydrogeological behaviorat larger integration scales of e.g. N10km. Most of the methods used byhydrogeologists integrate over larger scales than the typical width ofmany fault cores (Section 3), which suggests that hydrogeologicalmethods can only infer the effective hydrogeologic impact of faultzones rather than constrain the internal permeability structure of faultzones. Conversely, it may be difcult to elucidate a regional impact offaults on uid ow from the smaller scale studies conducted by struc-tural geologists.

    3. Fault zone processes from surface-focused studies

    3.1. Terminology and fault zone architecture

    Structural geologists have described fault zones in outcrop in diversesettings (Fig. 4) such as poorly lithied sedimentary rock, for examplebasin ll sediments in the Gulf of Corinth Rift basin (Fig. 4a; Lovelesset al., 2011), the Lower Rhine Embayment (Bense et al., 2003b), theRio Grande Rift basin (Fig. 4b; Heynekamp et al., 1999; Sigda et al.,1999; Rawling andGoodwin, 2006; Caine andMinor, 2009), in southernItaly (Balsamo and Storti, 2010) and Austria (Exner and Grasemann,2010), and glacial deposits in Denmark (Kristensen et al., 2008). Inlithied sedimentary rocks, the Colorado plateau is a classic area ofstudy where many researchers have focused efforts to determine faulthydraulic properties (Fig. 4c; Antonellini and Aydin, 1994, 1995;Jourde et al., 2002; Davatzes et al., 2003). Outcrops of regional scale

    ceoutcrops often by structural geologists or (b) by investigation of borehole data providing. The source of eld data leads to different conceptualmodels, numericalmodeling assump-tigation and eld methodologies. Surface investigations sometimes have a more three--surface investigations include tunnels or underground mines.

  • ydr

    men

    lims ther csamsen

    samseneabe uprotialeab

    ranhyserties to

    idesinea

    174 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192Table 1Various approaches used by structural geologists and hydrogeologists to study fault zone h

    Method Purpose Requirements and data Com

    Structural geologyOutcrop mapping Map products of

    deformation and faultarchitecture

    Field exposures; lithologicand structural data

    Datazonedeep

    Fault rockmineralogyand geochemistrystudies

    Examine waterrockinteractions

    Cores or outcrop samplesMineralogy and elementalcomposition of faultcomponents

    Rockrepre

    Laboratory core tests Permeabilitymeasurement

    Rock samples RockreprepermCan bwith

    In-situ permeametry(usually using miniair-permeameters)

    Permeabilitymeasurement

    Suitable (fresh, smooth)surfaces in outcrop

    Potenperm

    HydrogeologyDrilling and boreholegeophysics

    Direct subsurface samplingand measurement ofpetrophysical properties

    For example, geophysicallogging of boreholes and/orinter-borehole tomography

    Largegeopprop

    Flowmeter testing Determine permeabilityand storativity of faultcomponent or fracture(s)

    Pressure or ow responsemeasured in well

    Allow

    Pumping testing Effective permeabilitymeasurement of aquifer

    Pressure response measuredin well eld

    Prova delbasement faults in the Rocky Mountains (Fig. 4d; Caine and Tomusiak,2003) and Japan (Jefferies et al., 2006) have also been closely studiedfor their hydraulic properties. Cementation patterns caused by uid cir-culation along faults for example seen on Spitsbergen Fault (Fig. 4e)have even been suggested as analogous to linear features seen on thesurface of Mars (Fig. 4f; Treiman, 2008), thereby providing indirect ev-idence for past circulation of groundwater on another planet in oursolar system.

    Normal faults generally occur in extensional tectonic settings, re-verse faults in compressional settings, and strike-slip faults where slipis horizontal. Normal and reverse faults typically have a dip of ~4570. The rock above the fault zone is called the hanging wall, the rockbelow the fault forms the foot wall.When reverse faults have a relative-ly shallow dip (45) they are referred to as thrust faults. The reader isreferred to textbooks on structural geology for further description offault zone classication (e.g. Fossen, 2010). Fault zone permeabilitystructures may be classied in a number of ways (e.g. Karasaki et al.,2008), including based on the type of displacement (normal, thrust/reverse and strike-slip). However, here we classify the conceptualmodels of fault zone hydrogeology primarily based on rock type. Wechoose this justication because the structural architecture of faults,and their resulting permeability structure and aquifer development,are strongly controlled by rock type (Section 3.2).

    A rst-order description of fault zones commonly includes a faultcore (FC), which is surrounded by a damage zone (DZ) (Fig. 5a). The

    Mapping hydraulichead gradients

    Identify owdiscontinuities

    Water levels in well eld Direct evwells neeto resolv

    Use of heat tocharacterize ow

    Delineating of uid owpaths

    Distribution of temperaturein aquifers

    Direct evsignicangeotherm

    Articial andenvironmentaltracers

    Delineating uid owpaths

    Groundwater geochemistry Direct evinterpretchalleng

    Numerical modelingDiscrete fracturenetwork models

    Integrating borehole and/or outcrop data,parameter estimation

    3D distributions of fracturecharacteristics, hydraulictests

    Often onof meterfractures

    Continuum models Integrating borehole and/or outcrop data,parameter estimation

    Effective parameters Challenginto theogeology.

    ts Literature examples

    ited to near-surface character of faultat might however be exhumed fromrustal levels

    Lehner and Pilaar (1997), Caine and Forster(1999), Jourde et al. (2002), and Shipton et al.(2006a)

    ples taken at a small scale might not betative of larger scale processes

    Evans and Chester (1995), Caine and Minor(2009), and Koukouvelas and Paoulis (2009)

    ples small scale, not necessarilytative of larger scale averageility, but provide direct measurements.sed to study permeability developmentgressive strain

    Evans et al. (1997), Ogilvie et al. (2001),Ngwenya et al. (2003), and Faoro et al. (2009)

    of mapping spatial distribution ofility across fault zones in detail

    Antonellini and Aydin (1994), Sigda et al.(1999), Rawling et al. (2001), and Balsamoand Storti (2010)

    ge of techniques available. Translation ofical properties to hydrogeologicals can be problematic

    Moretti (1998), and Ellsworth et al. (2005)

    directly measure uid uxes Davison and Kozak (1988), Martin et al.(1990), Hsieh (2000), and Le Borgne et al.(2006),

    effective aquifer properties but withouttion of ow paths

    Shan et al. (1995), Marler and Ge (2003),Anderson and Bakker (2008), and Medeirosfault core, as the zone of the most intense strain, is generally found inthe center of the fault zone, and accommodates the majority of the dis-placementwithin the fault zone. The damage zone has secondary struc-tures such as fractures, and minor faults extending into the foot-walland hanging-wall, which take up the remainder of strain withinthe fault zone. In unlithied sediments, in which mixing of sedimentscan occur in the fault zone (discussed in Section 3.1), an additionalzone, called the mixed zone exists in between the damage zone andthe fault core (Heynekamp et al., 1999). However, the use of a mixedzone, as a genetic term in the context of fault zone architecture, in theeld description of faults has led to confusion (e.g., Evans andBradbury, 2004) sowe do not use the term in this paper but only distin-guish a fault zone as to consist of a fault core and a damage zone.

    In the FCDZ framework ameasurable fault thickness is dened per-pendicular to the fault strikewhich is the sumof the fault core and dam-age zone thicknesses (Fig. 5b). The thickness of fault zones increaseswith fault throw or displacement (Knott et al., 1996; Sperrevik et al.,2002; Childs et al., 2007). This trend is relatively consistent for differenttypes of faults (normal, reverse and strike-slip) in a range of rock types(Fig. 6). However, Evans (1990) and Faulkner et al. (2010) cautionagainst over-interpretation of these trends as fault thicknessthrow re-lationships are commonly plotted on loglog scales and show scatterover multiple orders of magnitude. The fundamental reasons underly-ing these observed relationships are still not well understood (e.g.Shipton et al., 2006b). While some modes of strain localization may

    et al. (2009)idence of impact on hydraulic head butd to be very closely spaced at fault zonese gradient in sufcient detail

    Haneberg (1995), and Bense and Van Balen(2004)

    idence ofowpaths butowneeds to bet enough to disturb the backgroundal gradient

    Fairley and Hinds (2004), Anderson andFairley (2008), Bense et al. (2008), and Readet al. (2013)

    idence of uid ow but theation of geochemical data can be ae

    Whiteman (1979), Abelin et al. (1991),Gascoyne et al. (1993), Rugh and Burbey(2008), and Leray et al. (2012)

    ly feasible at limited spatial scale (e.g. 10ss) and realistic representation ofrequires large amount of data

    Long et al. (1982), Caine and Forster (1999),Jourde et al. (2002), Caine and Tomusiak(2003), and Surrette and Allen (2007)

    e of a realistic inclusion of fault zonesmodeling routine

    Oda (1986), Forster and Smith (1989), Lopezand Smith (1996), Bense and Person (2006),and Micarelli et al. (2006)

  • 175V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192be responsible for increasing fault thickness with increasing throw(Childs et al., 2009), inconsistencies in the denition of fault thicknessin the eld is also probably responsible for some of the scatter in datashown in Fig. 6 (Shipton et al., 2006b).

    Fig. 2.Hydrogeologic tools to examine uid ow around faults. (a) Hydraulic headmaps canbe derived fromwater levels measured in wells. Contours of mapped hydraulic head gener-ally slope uniformly in unfaulted aquifers with homogeneous permeability. Faults acting ashydraulic barriers or conduits can be inferred from non-uniform gradients. The downward(b) or upward (c) direction and/or path of groundwaterow in fault zones can be further de-lineated using tracers such as temperature and other environmental tracers. Fluidmovementcan be driven by natural hydraulic gradients or groundwater abstraction.Assuming enhanced permeability due to fracturing in the damagezone and a low permeability fault core, a conceptual permeabilitymodel of a fault zone can be derived (Fig. 5d). In a widely-cited reviewpaper by Caine et al. (1996), the geological FCDZmodel of fault zones istranslated to a number of possible fault permeability structures as afunction of the relative contributions of a low-permeability fault coreand a high-permeability fractured damage zone (Fig. 5f). Field data(e.g. Savage and Brodsky, 2011) discuss how the proportion of faultcore volume to that of the damage zone, evolves with increasing faultthrow. Beyond the scale of an individual fault strand, the larger scalestructure of faults can be segmented or otherwise complex (Fig. 5a;Childs et al., 1996; Walsh et al., 2003; Rotevatn et al., 2007; Bonsonet al., 2007; Faulkner et al., 2010). As a result, fault zones in outcropcan display multiple fault cores, with overlapping damage zones(Fig. 5c). The density of fracturing can be plotted against distancealong a scan line to identify the presence of a damage zone (Fig. 5d;Faulkner et al., 2010; Savage and Brodsky, 2011; Schueller et al.,2013). High fracture densities in these composite fault zones can beinterpreted as a superposition of that of individual damage zones, assuggested by eld observations (e.g. Savage and Brodsky, 2011).Where fault damage zones overlap, in cases of closely spaced faultzones, the permeability structure can be derived assuming superposi-tion of individual fault zones (Fig. 5e). The structural heterogeneity pro-vided by fault segmentation has a direct impact on the continuity anddistribution of fault rocks. In this paper, however, we will consider thepermeability distributions in individual fault zones and we will usethe fault core and damage zone concepts as descriptors of a fault zonewhile also recognizing that these fault zones can be segments of a largerscale fault structure.

    3.2. Impact of fault zone processes on permeability

    The products of deformation processes accommodating strain in afault zone can enhance or reduce the permeability within fault zones.We consider a rst-order classication of such processes relevant forfault zone hydrogeology (Table 2) based upon whether a fault processincreases (Section 3.2.1) or decreases permeability (Section 3.2.2) andindicate how the occurrence and impact of these varies for differentrock types. Subsequently, in Section 3.2.3 we discuss a range ofinteracting, secondary or larger-scale processes that can reduce or en-hance permeability such as compaction, tectonic stress and geochemicalprocesses. These processes are highly relevant to understanding thepermeability distribution and evolution in fault zones but are notconstrained to occur solely within fault zones.

    3.2.1. Fault zone processes that enhance permeability

    3.2.1.1. Particulate ow in unlithied rock. In rocks with low cohesivestrength, such as unlithied rock, the rst stage of accommodation oftensile stress is via the formation of dilation bands (e.g. Du Bernardet al., 2002). Along dilation bands the grain fabric is disaggregated(Fig. 7a) but individual grains have notmoved past each other to accom-modate any shear offset along the band. Once offset occurs along dila-tion bands they are termed shear bands (Fig. 7b). Both shear anddilation bands are types of deformation bands (e.g. Fossen et al.,2007). The process of grains rolling and sliding past one another is usu-ally referred to as particulate or granular ow (cf. Borradaile, 1981).Particulate ow will lead to a rearrangement of the pore network andthus impact permeability. The initial dilation could lead to a slight en-hancement of permeability potentially up to one order of magnitude(Sperrevik et al., 2002; Bense et al., 2003b). Subsequent shearing and as-sociated rotation of grains along deformation bands can introduce hy-draulic anisotropy as observed in laboratory experiments producingshear bands in loose sediments (Fig. 7b; Arch and Maltman, 1990).Under greater conning pressures (greater burial depths) the breakage

    of grains (cataclasis) will increasingly occur along deformation bands.

  • ) e

    176 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192Fig. 3. The methods used in structural geology (black line) and hydrogeology (gray lineapproximately.Hence, only at shallow depths (e.g. b100m) permeability in unlithiedrock is usually not strongly impacted by fault zone processes, a distinctreduction in permeability may be possible at greater depths, which isdiscussed in Section 3.2.2.

    Relatively little attention has been paid to the hydrogeological im-portance of permeable sands and gravels that are incorporated via par-ticulateow in fault zones instead of, or togetherwith, lowpermeabilitybeds such as clay and shale. However, the incorporation of sand alongfault zones, together with or instead of clay beds, is well documentedin the eld (Lewis et al., 2002), and interpreted on seismic images(Koledoye et al., 2003). This process will often lead to an enhancementof permeability of a fault zone since beds of permeable material alongthe fault zone will form conduits for the migration of uids along thefault zone, the efciency of which depends on the continuity and thick-ness of this material (Bense and Person, 2006).

    3.2.1.2. Fracturing and brecciation in lithied siliciclastic, crystalline andcarbonate rock. In response to stress imposed on the rock, dilational frac-tures (Fig. 7a) with no shear displacement (called joints) occur in andoutside fault zones in relatively strong rocks such as partially lithiedlow-porosity clastic rocks (e.g.; Balsamo et al., 2010), crystalline rocksand carbonate rocks (Berkowitz, 2002; Neuman, 2005). The mechanicsof fracture initiation and propagation are beyond the scope of this re-view but are well described elsewhere (Pollard and Aydin, 1988;Aydin, 2000). For low-porosity (b10%) and permeability (1015 m2)rock, bulk porosity and permeability are primarily controlled by thefracture network properties. The network connectivity, that controlspartly the permeability, depends on fracture density, fracture orienta-tions and fracture length distribution (e.g. Hestir and Long, 1990; Bourand Davy, 1997). Permeability of fracture networks depends both onconnectivity of fracture networks and the fracture aperture distribution(Long et al., 1982; Brown and Bruhn, 1998; Faybishenko et al., 2000; deDreuzy et al., 2001, 2002). Using a simple parallel plate model toach have a characteristic length scale that they spatially integrate over, as shown hererepresent fractures, it was shown that the cube of the aperture is agood approximation of their permeability (Snow, 1969: Kiraly, 1971;Witherspoon et al., 1980; Zimmerman and Bodvarsson, 1996). Thus, in-dividual fractures with a moderate (~100250 m) aperture can stillcontrol local permeability. Near fault zones the fracture density andconnectivity typically increases signicantly (Fig. 5). Therefore thebulk rock permeability near fault zones is generally larger than that ofthe protolith. For example, permeability measurements indicate an in-crease in effective permeability by 23 orders of magnitude relative togranite and gneiss protolith in the East Fork thrust faults in Wyoming,USA (Evans et al., 1997).

    Brecciation is a prominent fault process in low-porosity settingssuch as within crystalline rocks, and carbonates (Sibson, 1977; Sibson,1986). Breccia is fault rock composed of angular, coarse-grained frag-ments that formed via high-density fracturing (brecciation), which areembedded in a ne-grained matrix, and where the matrix does notmake up more than 30% of the total rock volume (Sibson, 1977;Woodcock and Mort, 2008). Breccia is often uncohesive when formedin the upper few kilometers of the crust. As with the development ofjoints, fracturing leading to breccia formation in fault zones will, atleast initially, cause an enhancement of permeability. This can amountto up to four to ve orders of magnitude (Walker et al., 2013). Cemen-tation of the pore space created by brecciation, or a grinding of the brec-ciated rock with further displacement along the fault zone, will in turnlead to a strong reduction in permeability.

    3.2.2. Fault zone processes that reduce permeability

    3.2.2.1. Particulate ow, sedimentmixing and clay smear in unlithied rock.Particulate ow was discussed above as a process to potentially lead tosubtle enhancements of permeability along shear bands. However,when fault displacements are higher than bed thickness, particulateow can result in the mixing of unlithied rock of different grain sizes,

  • 177V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192at the grain-scale as well as at the scale of sedimentary beds. Tectonicsediment mixing in fault zones generally leads to a reduction of perme-ability (Heynekamp et al., 1999; Faerseth, 2006; Rawling and Goodwin,2006; Caine and Minor, 2009; Balsamo and Storti, 2011). The physicalmixing of sedimentswith contrasting grain-size distributions can be ex-pected to result in a more poorly sorted sediment mixture than any ofthe source beds, and for this reason sediment mixing leads to the effec-tive reduction of pore space and permeability in the fault zone. The de-gree to which permeability is reduced as a result of sediment mixing inunlithied rock will depend on the contrast in permeability betweenthe end-member beds, and whether particulate ow is to any extentfacilitated by grain breakage, which adds an additional reduction in av-erage grain-size of themixture. Heynekamp et al. (1999) report a reduc-tion in permeability of up to six orders of magnitude as compared to theoriginal sand bed, where sandy clay form in the fault zone as a result ofmixing between sand and clay beds along the Sand Hill fault zone in

    Fig. 4. Fault zones in a variety of geologic settings. HW=hanging wall; FW=foot wall; arrowsgravels in theGulf of Corinth, Greece. (b) San Ysidro fault zone cross-cuts unlithied, siliciclasticentrainment along minor faults in sandstone sequences, in the damage zone of the Moab Faustones and shales. (e)Quartz-cemented,mineralised fault of the Billefjorden Fault Zone, Austfjorwhich have been interpreted (Treiman, 2008) as cemented fault zones analogous to the one shNew Mexico, USA. The latter study further illustrates that mixing,from relatively homogeneous source beds, causes strong permeabilityheterogeneity in the fault zone because of incomplete mixing. In addi-tion to permeability heterogeneity, an anisotropy of permeability canbe expected to be present in fault zones as a result of rotation of bladedsediment grains. Grains aligning preferentially with the main fault diphave been observed in both lab-experiments on loose sands (Arch andMaltman, 1990), and in naturally faulted sediments ranging from sand(Bense et al., 2003a,b; Fossen et al., 2007) to gravels (Loveless et al.,2011). At the grain scale, the increased tortuosity of uid ow pathsacross the fault as a result of the re-alignment of oblate grains in the di-rection of the fault dip results in permeability anisotropy so that per-pendicular to the shear zone, permeability can be up to two orders ofmagnitude lower than along it (Arch and Maltman, 1990).

    Where clay minerals are present in the rock matrix phyllo-silicateframework bands will develop (Fig. 7c) alongwhich platy clayminerals

    in (ad) indicate the direction of movement along the fault. (a) Fault offsetting unlithiedsediments in the RioGrande rift, NewMexico, USA (fromCaine andMinor, 2009). (c) Shalelt, southern Utah, USA; (d) Fault 6, Traill , East Greenland truncating interbedded sand-den, Spitsbergen. (f) High-albedo features on the surface ofMars (indicated by the arrows),own in (e).

  • some point. However, at intermediate depths particulate ow can be ac-companied by cataclasis in which case the deformation process has

    Fig. 5. (a) Block diagram showing aspects of the three-dimensional normal fault zone suchas a relay zone fromChilds et al. (2009). Fault zonewidth has beendened (b) as the com-bined width of the fault core and damage zone for faults with a single core (Chester andLogan, 1986; Caine et al., 1996) or (c) as the distance between the outer slip surfaces be-longing to the same fault structure where multiple strands of high-strainmaterial encloselenses of fractured protolith (Faulkner et al., 2004, 2010). Fracture density and permeabil-ity for a crystalline bedrock lithology for a single fault zone strand (d) and multiple faultzone strands (e) (for other models see Fig. 9 and Section 3.2), modied from (Faulkneret al., 2010). (e) A description of fault behavior based on the proportion of fault core todamage zone, from Caine et al. (1996).

    178 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192been called facilitated- or dependent particulate ow (Borradaile,1981; Rawling and Goodwin, 2003). Facilitated particulate ow charac-terized by grain aking has been observed in unlithied sedimentaryrock only buried to a few hundreds of meters (Heynekamp et al.,1999; Sigda et al., 1999; Rawling and Goodwin, 2003; Balsamo andStorti, 2010; Balsamo et al., 2010, 2013). Moreover, the effectivenessof cataclasis occurring in unlithied sedimentary rock, variesas function of grain composition; relatively weaker grains such as feld-spars can be entirely crushed while stronger quartz grains show low-intensity cataclasis characterized by the aking of grains rather thantheir entire disintegration by crushing (Heynekamp et al., 1999; Exnerand Grasemann; Loveless et al., 2011; Exner and Tschegg, 2012). Sedi-ment grain-size distributions allowed Rawling and Goodwin (2003) toshow how the efciency of cataclasis along deformation bands in par-tially lithied quartz-dominated sediments is greater for relativelylarge grain-sizes than for ner sediment. Major near-surface rupturingassociated with earthquakes at depth may result in the occurrence ofcataclasis under relatively low conning pressure in loose sedimentswhere normally particulate owwould dominate as the mode of defor-mation (Cashman et al., 2007; Doan and Gary, 2009; Balsamo and Storti,2011). However, occurrence of uid overpressure in rapidly subsidingsedimentary rocks can potentially lead to substantial reduction in rockstrength as a result of which particulate ow might dominate overcataclasis as the deformation mechanism in granular rock to depthsorient in the direction of the fault zone andwill so facilitate the sliding ofgrains past one another possibly reducing grain breakage (e.g. Fossenet al., 2007). Clay smears often develop along fault zones cuttingthrough clay beds (Figs. 4c and 7c). The mechanisms and parameterscontrolling the continuity of clay smears (Lehner and Pilaar, 1997;Aydin and Eyal, 2002; Eichhubl et al., 2005) and shale smears (Lindsayet al., 1993; Davatzes and Aydin, 2005) entrained along fault zoneshave been the focus of a substantial body of research. The focus onclay smear exists mainly because of their potential to form effectiveseals for hydrocarbon migration (e.g., Yielding et al., 1997; Sorkhabiand Tsuji, 2005), and to effectively block across-fault uid ow (Benseand Van Balen, 2004), compartmentalizing reservoirs and aquifers.Clay smears have been described in stratigraphies characterized byunlithied rock consisting of sandclay alternations (Yielding et al.,1997) as well as in predominantly carbonate rocks (Ferrill et al., 2004;Bonson et al., 2007; Bourouis and Cornet, 2009). Experimental workusing ring-shear apparatus (e.g. Sperrevik et al., 2002), sand-box(Schmatz et al., 2006) and numerical modeling using discrete elementapproaches (Egholm et al., 2008), has attempted to elucidate themech-anisms and geological parameters controlling the thickness and conti-nuity of clay-smear. Ring-shear experiments (Sperrevik et al., 2000)suggest that the competence contrasts arising between clay and sandunits during sediment burial, exert an important inuence on the conti-nuity of clay-smears that develop.

    3.2.2.2. Cataclasis in unlithied and lithied rock. Cataclasis is the perva-sive brittle fracturing and comminution of grains (Engelder, 1974;Chester and Logan, 1986; Blenkinsop, 1991; Davis and Reynolds,1996). In lithied rock with relatively high primary porosity (N15%)such as sandstone, displacement along deformation bands (shearjoints) is commonly accommodated through cataclasis, replacing par-ticulate ow that usually dominates deformation in soft-sediments(Fig. 7b). Cataclastic deformation band networks in lithied sandstoneshave been the focus of a very substantial body of literature (e.g. Fowlesand Burley, 1994; Shipton and Cowie, 2001; Schultz and Siddharthan,2005; see Fossen et al., 2007 for a review).While particulate ow occursnear the ground surface in unlithifed rock, at greater depths rock isstronger and cataclasis becomes the dominant deformation process atmuch larger than 1km.

  • diffeDiag

    179V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192Where particulate ow is accompanied by cataclasis a permeabilityreduction along slip surfaces of 13 orders of magnitude has been re-ported (Fig. 8; Sigda et al., 1999; Rawling et al., 2001; Balsamo andStorti, 2010; Ballas et al., 2013; Balsamo et al., 2013). Permeabilityalong cataclastic deformation bands in lithied sandstones with lowclay-content is typically reducedmore strongly by 45 orders ofmagni-tude, as compared to undeformed sandstone, which is demonstrated bymany eld- and laboratory permeability tests (e.g. Antonellini andAydin, 1994; Fisher and Knipe, 2001). Permeability along cataclastic de-formation bands is often anisotropic with the largest reduction inpermeability perpendicular to the band (Antonellini and Aydin,1994; Sigda et al., 1999). Currently, the controls on the magnitudeand spatial variability of the permeability reduction along deforma-tion bands is still the focus of a substantial research effort (e.g.Torabi and Fossen, 2009).

    3.2.3. Interacting, secondary or larger-scale processes that either reduce orenhance fault zone permeability

    In addition to the fault processes that impact the permeability offault rocks described above, a suite of other geological processesand conditions can have a signicant impact on the developmentof fault permeability and architecture. Here we focus on the mainprocesses relevant to the shallow crust, (1) the role of compactionand cementation of unlithied rock, (2) dissolution and cementa-tion in fracture networks, and (3) regional stress eld impactingfracture permeability.

    3.2.3.1. Compaction and cementation of unlithied rock. Compaction and

    Fig. 6. The relationship between fault thickness (width) and fault throw (displacement) forshows similar trends. Lines and shading covers range of observed data for each fault type.cementation of sediments in subsiding sedimentary basins generallyleads to porosity loss (e.g. Bethke and Corbet, 1988) and a reductionin permeability (Bense and Person, 2006). Rate of sedimentation, sedi-ment compressibility, and permeability will control rates and magni-tudes of mechanical compaction during burial of the rock (Bethke andCorbet, 1988). The inuence of compaction and cementation on fault-related permeability structures can be two-fold. Firstly, permeabilitypatterns formed by deformation at shallow depth will be overprintedby permeability changes as a result of compaction and cementation atlarger depths (e.g. Milliken et al., 2005). The compressibility of thesediment, which is governed in part by rock strength, is likely to be in-uenced by shallow diagenetic and tectonic processes where for exam-ple disaggregation leads to a loss of cohesion. Consequently, the impactof compaction on permeability of faulted sediments can be variable:permeability contrasts can either become homogenised when porosityloss is greater in the protolith than along deformed rock, or enhancedwhen compaction and cementation preferentially reduces permeabilityalong shear bands (Fossen et al., 2007). Secondly, compaction and ce-mentation strengthens the rock which causes, as discussed above, thedominant deformation mechanism to shift from particulate ow tocataclasis or jointing with increasing depth. Fieldmeasurements of per-meability of fault rocks formed in compressible clastic sediments thusreect the effects on permeability of both compaction and deformation(Fisher and Knipe, 2001). Bense and Person (2006) described perme-ability loss due to faulting in combination with compaction (but notcementation) as a function of volumetric clay content mimicking theobserved range in permeabilities of faulted unlithied and lithiedrocks associated with various deformation mechanisms as reported inthe literature (Fisher and Knipe, 2001; Sperrevik et al., 2002).

    3.2.3.2. Dissolution and cementation in fracture networks. Fluids carryingreactive solutes circulating through fault zones potentially enhance orreduce permeability as a result of waterrock interaction (e.g. Zhanget al., 2008). Whether precipitation of minerals into pore space occurs(cementation), or rock dissolution occurs depends on the specic geo-chemical conditions such as rock composition, solute concentrations,temperature, and rates of advective solute transport via uid ow (e.g.Appelo and Postma, 2005). Moreover, the enhancement or reductionof pore space provides a strong feedback mechanism on the conditionsfor waterrock interaction to progress or slow down. For example, if arequirement for cementation is that sufcient solutes are provided viaadvective uid and permeability is reduced via cementation, a negativefeedback mechanism will be in operation. Given the correct conditions,fracture networks surrounding fault zones can become fully cemented,but this is by no means always the case (Micarelli et al., 2006; Kimand Sanderson, 2009). Regardless of whether such cementation hap-pens, new generations of fractures can develop with further stress im-

    rent fault types (a) and between lithied and unlithied sediments and crystalline rock (b)rams are based upon data presented in Childs et al. (2009).posed on the rock (Roberts and Stewart, 1994; Geraud et al., 2006;Benedicto et al., 2008). This cyclic process leads to the formation of char-acteristic crack-seal breccias (cf. Ramsay, 1980; Woodcock et al., 2007;Kampman et al., 2012). Enhanced permeabilities may prevail along ac-tive faults if the opening and formation of fractures is more rapid thanthe sealing mechanism via cementation (Riggs et al., 1994; Constantinet al., 2004). Although uidrock interaction as described above is a typ-ical feature of carbonates which are relatively easy to dissolve, geo-chemical, geomechanical and outcrop observations suggest that uidow and chemical reactions are often important to fault developmentand growth in crystalline rock as well (Chester and Logan, 1986;Evans, 1988; Bruhn et al., 1994; Chester, 1995; Hickman et al., 1995;Wintsch et al., 1995). Laboratory and eld evidence show that quartzis commonly dissolved and re-precipitated in fault cores which can re-duce their permeability by several orders of magnitude (Chester andLogan, 1986; Evans, 1988). Feldspars often react with slightly acidicpore water to produce phyllo-silicate which are volumeloss reactionsthat reduce the rheological strength and permeability of the centralfault core (Evans, 1988; Goddard and Evans, 1995; Wintsch et al.,

  • d cite. Re

    sticnta

    bl et

    (19bl et

    180 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192Table 2Processes that impact the permeability of fault zones in different rock types. Cells with boldouble asterisk (**) implies process is uncommon but possible; and blank cells implies rar

    Process Product Unlithied to poorly lithedrock

    Siliciclasedime

    Primary fault processes that enhance permeability (section 3.2.1)Particulate ow Disaggregation bands,

    dilation bands, sand smearDuBernard et al. (2002), andBense et al. (2003a,b), andExner and Graseman (2010)

    Fracturing Shear fractures, joints Balsamo et al. (2010) Eichhu

    Brecciation Breccias Caine and Minor (2009) HipplerEichhu1995). Phyllo-silicate formation could transiently enhance or redistrib-ute uid ow.

    3.2.3.3. Regional stress eld impacting fracture permeability. The apertureof fractures and therefore fracture transmissivity, is very sensitiveto pressure and to the stress conditions in the rock (Guguen andPalciauskas, 1992). This is one of the reasons, with a possibledecrease of fracture density with depth, that explains the decrease ofpermeability with depth often observed in the rst hundreds of metersin crystalline rocks (Ingebritsen and Manning, 1999; Gleeson andNovakowski, 2009). Field evidence over the past decades (e.g. Bartonet al., 1995), shows how in active, or recently active areas, faults net-works appear to be more permeable, as evidenced by higher hydrocar-bon production rates at those locations. The most permeable fracturesare generally assumed to be those which are most closely oriented in

    Primary fault processes that reduce permeability (section 3.2.2)Particulate ow Sediment mixing; dilation Heynekamp et al. (1999),

    and Rawling and Goodwin(2006)

    Phylosilicatesmearing

    Phylosilicate band Fulljames et al. (1997) Antonellin(1994); Kn

    Clay smear Lehner and Pilaar (1997),and Bense and Van Balen(2004)

    Lindsay etand Eghol(2008)

    Cataclasis Cataclastic deformationbands

    Sigda et al. (1999), Sigda andWilson (2003), and Cashmanand Cashman (2000)

    Aydin and(1978), anAntonellin(1994, 199

    Cataclasite, ultracataclasite Labaume eand Tueck(2012a)

    Brecciation/cataclasis

    Fault gouge Engelder (Gibson (19

    Secondary and other geological processes that can either reduce or enhance permeability (sectioFluid owcontrolleddissolution &cementation

    Solution cavities orincomplete precipitation(e.g., vugs)

    Crack-seal veins or breccia Veins, concretions andlocalised precipitation

    Mozley and Goodwin (1995),and Balsamo et al. (2013)

    Eichhubl e

    Pressuresolution

    Solution band Knipe (199Gibson (19

    Stylolites

    Sedimentinlling

    Sediment inll

    Fault growth Relay ramp, segmentboundaries, asperities,juxtaposition of lithologies atthe fault zone

    Rawling and Goodwin(2006), and Loveless et al.(2011)

    Childs et aWatterson(1998);M(1999)ations or single asterisk (*) implies process is common for the lithology; italic citations orpresentative examples of studies that describe the geologic structure are included

    ry rockCarbonate sedimentary rock Crystalline rock Volcanic rock

    al. (2009) Roberts and Stewart(1994), Agosta andKirschner (2003), andFerrill et al. (2004)

    Martel, 1990, andCaine and Tomusiak(2003)

    Gray et al., 2005

    93);al. (2009)

    Roberts and Stewart(1994), and Billi (2005)

    Bruhn et al. (1994),and Caine et al. (2010)

    Gray et al.(2005), andKaven andMartel, (2007)the direction of the largest horizontal stress. However, Barton et al.(1995) nuances this view by proposing that it is those fractures that un-dergo a high ratio of shear to normal stress which are most likely to behydraulically conductive. The latter model is conrmed by Henningset al. (2012), who show that the number of critically stressed fracturesis the best indicator of hydrocarbon reservoir performance, as a proxyfor the fracture network permeability.

    4. Conceptual geological models of fault zone hydrogeology

    In fault zo\nes inwhich a fault core and a damage zone arewell devel-oped, the FC/DC concept provides a convenient framework to describethe hydrogeological characteristics of fault zones (Fig. 5f; Caine et al.,1996). The permeability structure of a fault zone can bederived fromout-crops by combining a map, cross-section or 3D model of architectural

    i and Aydinipe (1997)

    al. (1993),m et al.

    Ferrill et al. (2004); Bonson etal. (2007)

    Johnsondi and Aydin5)

    Storti et al. (2003); Rath et al.(2011)

    Wilson et al.(2003)

    t al. (2001),mantel et al.

    Agosta and Kirschner(2003);Micarelli et al.(2006)

    Chester and Logan(1987)

    Walker et al.(2013)

    1974), and98)

    Ferrill and Morris (2003),and Benedicto et al. (2008)

    Chester and Logan(1987), and Schulzand Evans (2000)

    Gray et al.(2005)

    n 3.2.3)Micarelli et al., (2006); Kimand Sanderson (2009)

    Roberts and Stewart (1994) Walker et al. (2012)t al. (2009) Bastesen and Braathen

    (2010);Matonti et al.(2012)

    Bruhn et al. (1994);Schulz and Evans(2000); Person et al.(2008)

    Person et al.(2008)

    3), and98)

    Peacock et al. (1998)

    Tondi et al. (2006), andAgosta et al. (2012)

    Roberts and Stewart(1994), and Ferril andMorris (2003)

    Walker et al.(2012)

    l., (1996);et al.,

    ailloux et al

    Benedicto et al. (2008), andBastesen et al. (2009)

    Bruhn et al. (1994)

  • 181V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192components with the permeability of each fault zone element. Idealizedmodels of fault permeability structures have been proposed from eldobservations of faults, over a range of magnitudes in fault throw and in

    Fig. 7.Deformationmechanisms in clastic rock. (a) Formation of dilation bands in unlithied rocas low-intensity cataclasis, and cataclastic deformation bands forming gouge. (c) Clay within tsilicate framework bands develop at the grain scale while at the scale of sedimentary beds clay

    Fig. 8. (a) Example of observed permeability contrasts between architectural elements makinget al., 2001). Gray shading within each box represents the spread in observed values where thdamage zones in crystalline rocks represents the possibility of the presence of open fractures irepresented as a distribution in the data set underlying this gure. (b) Permeability data co(Figure from Balsamo and Storti, 2010) in the Crotone Basin in southern Italy (HW=Hangingability distribution. In the damage zone (DZ), cataclastic deformation bands occur alonghydrogeological heterogeneity. In the fault core, more intense cataclasis leading to the formatioas compared to the protolith.different geological settings (Caine et al., 1996; Aydin, 2000; Rawlinget al., 2001). The promise of such generic fault zone hydrogeologymodelsis that they have generic applicability in areas with similar geological

    k, and joints in lithied rock (from du Bernard et al., 2002). (b) Facilitated particulateow,he rock matrix reduces the occurrence of cataclasis along deformation bands, and phyllo-units will fold along the fault plane to form clay smears.

    up fault zones in crystalline rock, and unlithied siliciclastic rock (modied from Rawlinge black hue represents the median value. The arrow towards high permeability values forn such settings. Measurements of permeability on protolith clay units were too few to bellected across a small fault zone (throw ~20 m) in unlithied rock dominated by sandWall; FW=Foot Wall). The undeformed sediment has a relatively homogeneous perme-which permeability is reduced by one to two orders of magnitude leading to strongn of gouge causes permeability to be reduced further by almost four orders of magnitude

  • 182 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192parameters such as the lithology and geology of the protolith, and effec-tive stress (the total overburden pressure reduced by the uid pressure).

    4.1. Unlithied- and lithied siliciclastic rock

    In unlithied siliciclastic rock the smearing and drag of sand- andclay-beds along the main fault plane results in relatively consistentfault architectures as apparent when different outcrop studies are com-pared (e.g. Lehner and Pilaar, 1997; Eichhubl et al., 2005; Rawling andGoodwin, 2006).With increasing fault displacement a layered structureof clay and sand smears often develops (Fig. 9a). Bense and Person(2006) argue that in faults where cataclasis does not play an importantrole in reducing permeability at near surface depths in sandy sediments,the smearing of both clay and sand along the main fault plane is animportant mechanism to generate a strongly anisotropic fault core inwhich the sand seams will act as along-fault conduits while the claysmear will strongly hamper across fault ow. Any damage zone inthis kind of scenario will not have a strong impact on the faulthydrogeological structure as permeability along deformation bands inunlithied rock will be hardly affected. Based upon these consider-ations, faults in unlithied siliciclastic rock, in which cataclasis is mini-mal are likely to behave as combined conduitbarrier systems as aresult of the strongly anisotropic nature of the fault core (Fig. 9a),which will be further enhanced by grain rotation (Loveless et al., 2011).

    When cataclasis does play an important role, however, any sand in-corporated in the coremight be impacted by cataclasis and, hence, havea signicantly reduced permeability. Empirical relationships in faultedhigh-porosity sediments suggest that permeability in the fault core rap-idly decreases up to 23 orders of magnitude in the rst 510m of dis-placement, and then decreases at a slower rate in relation to furtherdisplacement (Balsamo and Storti, 2010). Also, cataclastic deformationbands developing in the damage zone will reduce permeability there(Fig. 8b). A range of faults along which shale is entrained in the faultcore, and complex deformation bandnetworks have formed in the dam-age zone and the fault core, have been described in detail on the Colora-do plateau (Davatzes and Aydin, 2005; Shipton et al., 2006a) and the RioGrande Rift Basin (Heynekamp et al., 1999). In the absence of open-fracture networks, this type of fault is thus likely to behave as an ef-cient barrier to uid migration (Fig. 9b; Rawling et al., 2001).

    4.2. Crystalline rock

    Fault cores in crystalline rocks are commonly zones of fault gougeand breccia (Evans and Chester, 1995; Caine et al., 1996; Evans et al.,1997). Fault cores often display anastomosing slip surfaces and occa-sionally deformation bands (Hancock, 1985; Evans, 1988). Flankingthe fault core is typically a damage zone, characterized by a high densityof fractures (Fig. 4d; Caine et al., 1996; Chester and Logan, 1986). Thefault core and damage zone are surrounded by protolith of relatively un-deformed rock that may contain joints not primarily related to the faultzone, but exists as a background deformation pattern.While laboratoryand in situ permeability experiments suggest that often the damagezone has a larger permeability than both the fault core and the protolith,the permeability in each domain is internally anisotropic (Forster andEvans, 1991; Caine and Forster, 1999). However, the permeability con-trast between components is much greater than the anisotropy withincomponents suggesting that inter-component contrast controls thebulk anisotropy of the system (Caine et al., 1996). Laboratory perme-ability tests of orientated samples show that the damage zone is usuallya high permeability conduit parallel to the fault plane while the core isa low permeability barrier parallel and perpendicular to the faultplane (Chester and Logan, 1986; Caine et al., 1996). Each of four end-member architectural styles (Fig. 9e) is thus associated with a charac-teristic permeability structure which varies with the percentage ofthe fault thickness taken up by the relatively low permeability FC and

    the higher permeability DZ. In this scheme, the ratio of damage zonewidth to total fault thickness quanties the description of fault architec-ture. In principle faults in these settings are thus thought of as toa degree behaving along continuums between four end members:conduitbarriers, distributed conduits, localized barriers or localizedconduits (Fig. 5f; Caine et al., 1996).

    4.3. Volcanic rock

    As a result of the wide variety of volcanic rocks ranging fromunwelded tuffs, towelded tuffs and solidied lava, the fault zone hydro-geology models that will be applicable to volcanic rocks will thus bewide-ranging.

    For fault zones in unwelded, stratied tuffs a fault zone hydrogeologymodel similar to that for unlithied siliciclastic rocks has been described(Evans and Bradbury, 2004). Inwelded, crystallized tuffs other conceptu-al models have been proposed (Wilson et al., 2003; Gray et al., 2005;Riley et al., 2010), which show that generally fault zone hydrogeologicalarchitecture in this geological setting is comparable to that of faults de-scribed in crystalline rocks. Hence, the main fault processes are breccia-tion in the fault core, and fracturing in the damage zone which can beaccompanied bymineralization. In particular, Gray et al. (2005) classiesthe architecture of fault zones found inwelded tuffs aroundYuccamoun-tain, Nevada into four groups based upon the shift, with increasing faultthrow, in the relative thickness of the damage zone and fault core mak-ing up the entire thickness of the fault zone, similar to the model de-scribed by Caine et al. (1996) for crystalline rock (Fig. 9d).

    A conceptualmodel for the development of fault permeability archi-tectures in basalt sequenceswas proposed byWalker et al. (2013) large-ly based upon outcrop evidence from the Faroe Islands. Thiswork showsthat faults in basalt are likely to behave as conduitbarrier systems in ananalogous way to many faults in crystalline rock. Walker et al. (2013)report how fault permeability along well-developed fault cores isstrongly reduced through cataclasis and the growth of authigenic clay-minerals, while in the damage zone fracture networks can, if well con-nected, maintain a relatively high permeability promoting along-faultuid ow. However, asWalker et al. (2013) point out, during the initialstages of fault zone development the conduit behavior of the fault zoneis likely to dominate. Overall, however, the fault zone permeabilitymodel as proposed for these faults is very similar to that of crystallinerocks as discussed above (Fig. 9c).

    4.4. Carbonate rock

    While dissolutional weathering often dominates carbonate rocks,primary fault deformation mechanisms are largely similar to those incrystalline rocks and siliciclastic rocks inwhich fracturing and cataclasisplay important roles impacting permeability (Stewart and Hancock,1991; Roberts and Stewart, 1994; Cello et al., 2000; Ferrill and Morris,2003; Storti and Billi, 2003; Doan and Cornet, 2007; Benedicto et al.,2008; Bastesen et al., 2009; Matonti et al., 2012). Fine grained sedimentmay also be washed into the fault core through cavities (Ferrill andMorris, 2003). Because of the importance of waterrock interaction incarbonate systems the permeability structures of faults in carbonaterocks can be expected to be transient, heterogeneous and anisotropic(Agosta et al., 2007; Agosta, 2008). The presence of uncemented frac-ture networks, breccia and dilatational jogs provides conduits whereasthe formation of ne-grained cataclastic fault rocks, the smearing ofclay into the fault core and the mineralization of fracture networks re-duces the permeability of the fault zone. The permeability structuresof fault zones in carbonate rocks are thus likely to be similar to those de-scribed above for crystalline rock and form to varying degrees conduitbarrier systems through the potential of uid-ow conduits through afracture dominated damage zone anking a low permeability faultcore (Fig. 9d). However complex the permeability structure along faultsin carbonates might be as a result of deformation, when dissolutional

    weathering creates signicant high-permeability pathways these may

  • 183V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192well dominate over the secondary permeability patterns induced by de-formation related to fault zone processes.

    5. Hydrogeological evidence of the impact of faults on uid ow

    The variety of fault models discussed in the previous sections,are primarily inferred from outcrop descriptions and permeability mea-surements from the eld and laboratory (Fig. 1a). The testing of thesemodels for their hydrogeological use needs eld evidence of thehydrogeological impacts of these fault zones (Fig. 1b). Here we also di-vide faults into distinct lithological categories despite the observationthat, often, faults crosscut and/or juxtapose multiple lithologies whichmakes describing and studying these fault zones more complex.

    5.1. Unlithied- and lithied siliciclastic rock

    Clastic sedimentary rocks often have relatively high bulk rock per-meability (N1014m2) and porosity (N15%), and are important loci of

    Fig. 9. Fault architectures are commonly described in terms of a fault core (FC) and a damage zonassociated conceptualized permeability structures in faults dominated by soft-sediment deformlithied sedimentary rocks in which cataclastic deformation plays a dominant role. Fault archi(d) carbonate rock.groundwater and hydrocarbon reservoirs. Often fault zones in sedimen-tary aquifer systems are apparent by signicant differences in hydraulichead in groundwater aquifers (Bense and vanBalen, 2004) oruid pres-sure as observed in hydrocarbon reservoirs (Sverdrup et al., 2003) on ei-ther side of a fault zone, indicative of anomalously high hydraulicgradients across the fault zone itself.

    The impact of faults as barriersmay become apparentwhen ground-water abstraction takes place either for mine-dewatering or water sup-ply purposes. In the loosely unlithied sediments of the Lower RhineEmbayment (LRE), Germany, extensive groundwater withdrawal takesplace and hydraulic head differentials across faults occur of up to~120 m. In addition, data from this area illustrate the potential effectsof discontinuities such as relay zones to strongly impact uid ow pat-terns in faulted aquifers. The differences in water table elevation on ei-ther side of faults in unlithied sediments often result in contrastingvegetation characteristicswhen shallowwater tables are foundupstreamof the fault zone while downstream water tables are deeper (Bense andKooi, 2004; Mayer et al., 2007). Outcrop studies in the LRE indicate the

    e (DZ) anking the fault core into the footwall andhangingwall. (a) Fault architecture andation at relatively shallow depths (partially after Heynekamp et al., 1999), and (b) withintecture and associated conceptualized permeability structures in (c) crystalline rocks and

  • 184 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192presence of clay/shale along fault planes, possibly in combination withcataclastic deformation (Lehner and Pilaar, 1997), which suggests thatthese faults would impede lateral groundwatermovement. Anomalouslyhigh hydraulic gradients across fault zones are indeed commonlyinterpreted as being indicative of a low fault permeability so that thefault acts as a bafe to groundwater ow (Haneberg, 1995).

    While in the LRE the presence of clay smearing, as observed in out-crop, seems to provide a good explanation for the barrier properties offaults as inferred from hydraulic head data, often such direct links be-tween structural geological and hydrogeological fault properties arenot so directly made. For example, hydraulic head data collected byMedeiros et al. (2009) around a fault zone dominated by cataclastic de-formation bands in a sandstone aquifer show that, although permeabilitymeasurements ondeformation bands show that the permeability of indi-vidual deformation bands is reduced by almost ve orders of magnitudecompared to the host rock, there seems to exist a mostly uninterruptedhydraulic communication between either side of the fault zone as ob-served during a pumping test. Structures in the fault zone cause openuid ow pathways that by-pass the low-permeability deformationbands. Mayer et al. (2007) on the other hand, reports water table dropsof ~80m across the Mission Creek Fault, California, cutting through allu-vium dominated by gravel in which clay-smearing along the fault planeis unlikely to occur. In the latter case, rotation of oblate pebbles along thefault plane might introduce sufcient reduction in permeability perpen-dicular to the fault to explain the high hydraulic head gradients.

    Although an analysis of hydraulic heads can give an indication towhat extent a fault might form an effective barrier to lateral groundwa-ter ow, natural tracers for groundwater ow such as salinity, temper-ature and other geochemical parameters are essential to provideadditional information on the presence and nature of uid ow pathsacross fault zones (Fig. 2). For example, in the LRE thermal anomaliesoccur in the aquifers anking the Rurrand Fault which can only be ex-plained by signicant along-fault up- and downwelling of groundwaterand subsequent sidewaysmigration into the aquifer (Bense et al., 2008).Presumably, the transport of groundwater along the fault plane occursthrough sand lenses and possibly continuous drags of sand in the faultcore in combination with the presence of clay smear contributing tostrongly anisotropic ow.

    Along the coast of the Gulf of Mexico, hydraulic head observationsseem to indicate that the Baton Rouge Fault in unlithied Quaternarysiliciclastic sediments hampers lateral uid ow while hydrogeochemi-cal analyses suggest that the brackish to salinewater occurring along theBaton Rouge Fault in shallow aquifers has migrated upward along thefault plane by several hundreds of meters (Stoessel and Prochaska,2005; Bense andPerson, 2006). Along the BatonRouge Fault the abstrac-tion of water in shallow aquifers provides the driving hydraulic headgradient needed for the up coning of deeper groundwater along faultzone. In the deeper parts of the Gulf of Mexico basin offshore upwarduid ow is driven by ongoing sediment compaction (Harrison andSumma, 1991) and, possibly analogous to the system along the BatonRouge Fault is the upward ow along faults of hydrocarbons resultingin oil-seeps found at the seabed offshore in the Gulf of Mexico(MacDonald et al., 2000). Uniquely, in the Gulf of Mexico, the migrationof pockets of over pressurized uids along faults has beenmonitored di-rectly using time-lapse seismic tomography (Haney et al., 2005).

    Further evidence indicating the importance of along-fault ow insiliciclastic aquifers is illustrated by paleo-uid ow patterns and infer-ences of its evolution through geological time as function of basin devel-opment derived from, for example, oxygen-isotopic analysis of calcitecements along block-bounding faults in California (Boles et al., 2004),diagenetic elongate carbonate concretions associated with faults inpoorly lithied sediments in south Italy (Balsamo et al., 2013), andmin-eralogical analysis of cements along faults in New Mexico (Caine andMinor, 2009). Other examples are the distribution and geochemistryof clay and iron precipitates along faults in sandstones in Nevada that

    allowed to reconstruct uid ow patterns and permeability structuresof deformation band networks and fault zones (Lansing Taylor andPollard, 2000; Eichhubl et al., 2004), and in poorly lithied sandstonesin the Paraiba Basin in Brazil the distribution of iron-oxide concretionsallowed to infer the dynamics of groundwater ow as controlled byfault structure and permeability (Balsamo et al., 2013).

    5.2. Crystalline and volcanic rock

    Crystalline rocks such as igneous and/or metamorphic rocks (here,we do not focus on other types of crystalline rocks such as evaporites)generally have low porosity and permeability and are therefore notclassed, generally, as good aquifers or reservoirs. Volcanic rocks canhave high primary porosity and permeability (e.g. basalts), but oftendo not (e.g. tuffs). However, crystalline and volcanic rocks are importantto siting of nuclear waste disposal sites and locally important for waterresources, in particular where secondary processes such as faulting andfracturing play a role in enhancing the rock's permeability. Whilecrystalline and volcanic rocks may deform by different fault processesresulting in different structures (see Sections 4.2 and 4.3) they oftenoutcrop together or are within close proximity, so we describe thehydrogeological evidence together.

    In mountainous settings, faults in the basement metamorphic andigneous rocks often are the loci of spring systemswith elevated temper-atures which, often in combinationwith evidence from silica-chemistry(Andrews et al., 1982; Ferguson et al., 2009), are indicative of relativelydeep circulation along fault zones. Examples of such systems in crystal-line rock are found in the CoastalMountains of British Columbia (Forsterand Smith, 1988; Forster and Evans, 1991; Grasby andHutcheon, 2001).The well-documented distribution and variation of temperature ofsprings along fault zones in the Northwestern Great Basin, USA provideperhaps the clearest evidence that faults in crystalline and volcanic rockcan be conduits allowing uid circulation (Fairley and Hinds, 2004;Anderson and Fairley, 2008). The discharge of water at or near the boil-ing point in springs at this site is indicative of rapid, upward advectivetransport along preferential ow pathways of enhanced permeability(Curewitz and Karson, 1997; Anderson and Fairley, 2008).

    While such thermal anomalies indicate that faults behave as along-fault conduits, other hydrogeologic, hydrologic and remote sensingdata suggest that faults in crystalline terrain can be effective hydraulicbarriers (Marler and Ge, 2003; Gleeson and Novakowski, 2009; Illmanet al., 2009). Lineaments in the Canadian Shield were interpreted byGleeson and Novakowski (2009) to be low permeability faults actingas hydraulic barriers to across-fault ow using well yield data, remotesensing and numerical modeling, but it is uncertain whether the struc-tures are also vertical barriers. Numerical modeling suggests a reversefault in crystalline rock in the southern Rocky Mountains of Colorado,USA is a barrier to lateral groundwater ow (Marler and Ge, 2003). Re-cent geophysical data from the same fault is interpreted to indicateconduitbarrier behavior at the meter to tens-of-meters scale (Ballet al., 2010). Detailed articial tracer and hydraulic testing data from athrust fault in Virginia, USA also suggests that faults in crystalline rockcan be combined conduits and barriers (Seaton and Burbey, 2005;Rugh and Burbey, 2008). Articial dye tracers applied at the surfacetraveled N100 m along the fault within weeks (Rugh and Burbey,2008), yet, hydraulic test data suggest that the fault is a low permeabil-ity barrier. A new hydraulic tomography cross-hole testing method at aproposed nuclear depository site in Japan showed low permeabilityzone and compartmentalization effects (Illman et al., 2009).

    There are other eld examples that show that faults may be highlypermeable and may represent valuable groundwater resources, particu-larly in granitic basement. Some eld studies have shown that gentlydipping fault zones may be highly productive (Seaton and Burbey,2005; Le Borgne et al., 2006; Liou et al., 2010; Ruelleu et al., 2010;Leray et al., 2013). A multi-disciplinary eld experiment (Roques et al.,submitted for publication) recently investigated in detail the hydro-

    geological character of a sub-vertical fault zone in crystalline bedrock

  • 185V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192in Brittany (France). The eld site was selected because of high wellyields during drilling (100 m3/h) related to a main permeable faultzone, encountered below 100m. The fault zone was characterized bothin natural conditions and during a 9-week large scale pumping test car-ried out at a pumping rate of 45m3/h. High-resolution ow logs allowedidentication of fracture connectivity and preferential ow paths whiledye tracer tests at different depths allowed estimation of transfer timesand groundwater uxes between the main compartments. Results sug-gest that the sub-vertical fault acts as the primary hydrogeological linkbetween aquifers (Roques et al., submitted for publication). In naturalconditions, deep groundwater ows along the fault zone towardsaquifers in shallow weathered bedrock and alluvial sediments. Duringpumping, sub-vertical groundwater uxes are reversed. Field-based in-sight on the hydraulic functioning of this kind of fault zone has importantimplications for groundwater resources and protection.

    Regional-scale faults in crystalline and volcanic rock have been ex-amined in detail by nuclear waste repository research and the risk offaults forming uid ow pathways in otherwise low-permeabilityrock. For example, multiple low-angle thrust fault zones cross-cut mas-sive and sparsely fractured crystalline rock of the Lac du Bonnet batho-lith in the Canadian Shield. Hydraulic tests and geochemical tracersindicate that faults are connected kilometer-scale conduits that controlregional groundwater ow and discharge (Gascoyne, 2004). However,the permeability distribution within the fault zones is highly heteroge-neous at a meter- to kilometer-scale (Davison and Kozak, 1988). Highangle normal faults cut the volcanic rocks of Yucca Mountain, a pro-posed geologic repository for high-level nuclear waste located in south-ern Nevada (Bredehoeft, 1997; Fabryka-Martin et al., 1997; Flint et al.,2001; Painter et al., 2003). Bomb pulse isotopes, which were highly el-evated in the atmosphere during above ground nuclearweapons testingin the Pacic Ocean in the early 1960s, were found in fault zones (Flintet al., 2001). The presence of these isotopes indicates groundwaterwater that is b40 years old and supports the conceptual model ofdeep, rapid and preferential inltration along these faults (Flint et al.,2001).

    5.3. Carbonate rock

    Fine-grained carbonate rocks often have low primary porosity butfracturing and subsequent dissolution can cause large enhancements ofpermeability resulting in carbonate rocks often being considered aqui-fers or reservoirs.More than in any other lithology, faults cutting throughcarbonate rocks will often be dominated by secondary dissolution andprecipitation processes altering the permeability structure almost con-tinuously. As a result, in the literature faults in carbonate rocks havebeen reported to display a verywide range of hydrogeological behaviors;from signicant barriers where secondary cementation or smearing oflowpermeabilitymaterial occurs, to effective conduitswhere dissolutionalong fault and fracture planes dominates. For example, high hy-draulic head gradients across a fault in a carbonate aquifer in South-ern Italy provide evidence for the barrier-like behavior of faults(Celico et al., 2006), analogous to observations in siliciclastic aqui-fers. Similar borehole data across the Aigion fault zone, Greece,show a pressure differential of 0.5 MPa across the fault plane, indi-cating that a major hydraulic barrier exists between the carbonateaquifer on either side of the fault plane (Giurgea et al., 2004;Micarelli et al., 2006; Doan and Cornet, 2007).

    Springs and outows often occur along faults in carbonate rocks(Bredehoeft, 1997; Giurgea et al., 2004; Celico et al., 2006; Billi et al.,2007; Magri et al., 2010). One noteworthy example of a fault relatedthermal spring emerging from carbonate rocks is the Bath hot springin southwestern England (Andrews et al., 1982). These springs havebeen interpreted to represent emerging ow from within the faultzone itself, implying that the fault zone acts as a conduit (Billi et al.,2007) or below the fault zone implying that the fault zone acts as a bar-

    rier (Giurgea et al., 2004; Celico et al., 2006). An analysis of pumpingtests in a fractured carbonate aquifer carried out by Allen and Michel(1999) show a mixed inuence of both permeable fracture networksand barriers along faults cutting through a carbonate aquifer in Ottawa,Canada. A similar conduitbarrier behavior of faults in carbonates hasbeen inferred by Breesch et al. (2009) who reconstructed a paleo-uidow history along reverse faults, based upon detailed petrographicand geochemical analyses of calcite lled fracture networks in Triassicand Cretaceous carbonate strata of Northern Oman. In this study, deter-minations of oxygen isotope ratios on calcite cements and circulationdepth constraints provided by uid inclusion thermometry suggestthat saline formation waters migrated syn-tectonically during contrac-tion of the basin along a major fault acting as a conduit, but at thesame time the fault hydraulically compartmentalized the basin.

    Geochemical data have indicated that fracture networks parallel tofault planes divert uid ow along the fault zone rather than across it(Maslia and Prowell, 1990; Ferrill and Morris, 2003; Ferrill et al., 2004;Geraud et al., 2006; Bushman et al., 2010). On the other hand, Ferrillet al. (2004) found uid ow paths in the faulted Edwards Aquifer,Texas, crossing from the foot wall to the hanging wall and back to thefootwall.

    6. Modeling uid ow in fault zones

    In fault zone hydrogeology numericalmodels are used to evaluate thepotential impact of fault zones on uid ow patterns. The primary chal-lenge is how to represent and parameterise the complex hydrogeologicalstructure of fault zones into such uid ow simulations (Garven et al.,1999). The approach taken to model uid ow in fault zones, dependson the amount of detail required in the calculated ow paths and hy-draulic head distributions which relates to the practical aim of themodeling exercise (e.g. water resources management, design of nuclearwaste repositories), but also on the type of aquifer in which the faultingoccurs and the type of deformation occurring along the fault zone.The spatial scale considered by a model evaluation is an importantdecision when modeling. A fundamental concept in hydrogeologyrelating to scale, is the representative elementary volume (REV)which can be dened as the smallest volume over which a measure-ment can be made that will yield a value representative of thewhole (Bear, 1972). Below the REV, the spatial variability of param-eters such as porosity and permeability is too large for the materialto be satisfactorily represented by a continuum with an averagedproperty assigned to it. If ow paths and uid uxes at a scalebelow that of the representative elementary volume need to beknown, individual features, such as fractures, should be representedin models (Fig. 1a), which is often accomplished using a discretefracture network (DFN) modeling approach, whereas above therepresentative elementary volume, the material can be treated asa continuum (Fig. 1b) for which an implementation of effectiveproperties over larger volumes is reasonable.

    Hydrogeological models on a scale larger than a few tens of metersare usually continuummodels using effective properties for discrete do-main features dened, often based upon the fault core/damage zoneschematisation of fault zones. This approach is often taken for modelsof uidow along fault zones in sedimentary basins considering region-al scale (~10100 km, Forster and Evans, 1991; Person et al., 1996;Micarelli et al., 2006; Magri et al., 2010) and more local (~15 km)groundwater ow systems (e.g. Folch and Mas-Pla, 2008). In suchmodels, the thickness of a fault zone, outlining the fault zone as ahydrogeological unit, is a crucial parameter which is directly controllingthe rate of uid mass and associated solute and heat uxes in thefault zone (e.g. Haneberg, 1995; Roberts et al., 1996; Bense andPerson, 2006). Continuum models can explicitly describe variations infault width, the juxtaposition pattern of aquitards and aquifers at thefault zone (Mailloux et al., 1999; Bense and Person, 2006), and thethree dimensional structure of fault zones (e.g. Micarelli et al., 2006).

    Other examples of continuum models of fault zone hydrogeology

  • 186 V.F. Bense et al. / Earth-Science Reviews 127 (2013) 171192include nely-griddedmodels of a fault in a sandshale sequence to an-alyze the effect of varying fault permeabilities on uid owpatterns andpressure compartmentalization in hydrocarbon reservoirs in the Gulf ofMexico (Mattha and Roberts, 1996), and models of hydrothermal owin carbonate rocks along the west-coast of Turkey (Magri et al., 2010).Leray et al. (2013) studied the productivity of permeable fault zonesfor water supply purposes. Using a simple numerical approach, theyshowed how the productivity of fault zone for groundwater resourcesdepends on the fault hydraulic properties as well as the geometricalcharacteristics of fault zones, and in particular the dip of fault zones.Continuummodels at the outcrop scale have also been used to calculateeffective fault zone properties like permeability anisotropy and the im-pact of small scale internal structure of fault zones on such properties(e.g. Jourde et al., 2002; Lunn et al., 2008).

    In contrast to continuummodels that represent the bulk permeabil-ity of a fault zone or fault zone components, discrete fracture network(DFN) models, originally developed to simulate ow and transport infractured bedrock (see reviews by Berkowitz, 2002; Neuman, 2005),represent the shape, orientation, length and aperture of fractures(which is related to permeability), and connectivity of fracture net-works. DFN models can directly incorporate detailed and realistic geo-metric and hydraulic data of fracture networks measured at theoutcrop and/or in boreholes, but often fracture networks and propertiesare generated in the DFN based upon statistical properties equivalent tothose observed in the eld. DFN models have been used to quantify theeffective permeability of individual fault componentswithin outcrops ofbasement rocks dominated by fracturing (Caine and Forster, 1999;Caine and Tomusiak, 2003). Where appropriate, geostatistical evalua-tions are used both in DFN models as well as in continuum models tocondition fault zone attributes such as fault rock permeability (e.g.Lopez and Smith, 1996; Fairley, 2009), and distribution, shape, and con-tinuity of fault zone elements (e.g. Lunnet al., 2008). Continuummodel-ing and DFN models can be combined in the simulation of thepermeability structure of fault zones. In this approach, key faults andfractures can be located explicitly in the model domain, w