Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal...

18
Experimental Poroviscoelasticity of Common Sedimentary Rocks Roman Y. Makhnenko 1 and Yury Y. Podladchikov 2 1 Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, IL, USA, 2 Institute of Earth Sciences, University of Lausanne, Lausanne, Switzerland Abstract The success of geoenergy applications such as petroleum recovery or geological storage of CO 2 depends on properly addressing the physical coupling between the pore uid diffusion and mechanical deformation of the subsurface rock. Constitutive models should include short-term hydromechanical interactions and long-term behavior and should incorporate the principles behind the mathematical models for poroelastic and poroviscoelastic responses. However, the viscous parameters in constitutive relationships still need to be validated and estimated. In this work, we experimentally quantify the time-dependent response of uid-lled sedimentary rocks at room temperature and isotropic stress states. Drained, undrained, and unjacketed geomechanical tests are performed to measure the poroelastic parameters for Berea sandstone, Apulian limestone, clay-rich material, and Opalinus clay (shale). A poroviscous model parameter, the bulk viscosity, is included in the constitutive relationships. The bulk viscosity is estimated under constant isotropic stress conditions from time-dependent deformation of rock in the drained regime for timescales ~10 5 s and from observations of the pore pressure growth under undrained conditions at timescales of ~10 4 s. The bulk viscosity is on the order of 10 15 10 16 Pa s for sandstone, limestone, and shale and ~10 13 Pa s for clay-rich material, and it decreases with an increase in pore pressure despite a corresponding decrease in the effective stress. In the long term, uid pressure can asymptotically approach minimum principal stress, which in natural reservoirs may lead to liquefaction or rock embrittlement, causing slip instabilities and earthquakes and creating high-permeability channels in low-permeable rock. 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore uid. When defor- mation is coupled with pore uid ow, time dependence is introduced into the response of an otherwise rate-independent solid, which can affect the deformation and failure of rock (Rice, 1975; Rice & Cleary, 1976). Physical coupling of deformation and uid ow for reservoir rock needs to be properly addressed because its effects are highly important for geoenergy applications such as petroleum recovery or geological storage of CO 2 and nuclear waste (e.g., de Waal et al., 2015; Rutqvist, 2012; Tsang et al., 2005). Biot (1941) and Biot and Willis (1957) rst introduced the theory of linear poroelasticity to describe the elastic deformation of uid-saturated geomaterials. Their theory of uid-saturated media has recently been extended based on rigorous thermodynamic constitutive modeling (e.g., Coussy, 2004; Lopatnikov & Cheng, 2004). Additionally, Biots macroscopic model has been conrmed by a number of upscaling techni- ques applied to microscale elasticity (Berryman, 2005; Gurevich, 2007; Lopatnikov & Cheng, 2004). Biots the- ory and its extensions are commonly used to describe the behavior of shallow reservoirs that are assumed to deform elastically. Time-dependent behavior (creep) is usually reported for dry rock subjected to high devia- toric stress (Brantut et al., 2013; Costin, 1987; Scholz, 1968). Creep is believed to be a property of a dry matrix, especially when low-viscosity components (clay or mica) are present (Chang et al., 1997). Indeed, mineral creep has been observed in laboratory experiments at high temperatures (~ 1001,000 °C) and pressures (~10 8 10 9 Pa) (e.g., Cox & Paterson, 1991; Hilairet et al., 2007; Kaboli et al., 2017; Proietti et al., 2016). Although, the extrapolation to natural conditions yields enormous timescales (tens to hundreds of years) in order to observe time-dependent deformation of minerals (Karato, 2010). However, some experiments on low-porosity rock (Elsworth & Yasuhara, 2006) have shown that reactions involving reactive uids (as in CO 2 storage) are fast (hours to weeks), and under relatively modest loads (~10 6 10 7 Pa) can lead to pressure solution and effectively viscous deformation. MAKHNENKO AND PODLADCHIKOV 1 Journal of Geophysical Research: Solid Earth RESEARCH ARTICLE 10.1029/2018JB015685 Key Points: Presence of aqueous uids enhances viscous deformation of sedimentary rock Poroviscoelastic behavior of sandstone, limestone, shale, and clay is quantied by laboratory tests Bulk viscosity of uid-saturated rock is ~10 15 10 16 Pa s and decreases with pore pressure increase Supporting Information: Supporting Information S1 Data Set S1 Correspondence to: R. Y. Makhnenko, [email protected] Citation: Makhnenko, R. Y., & Podladchikov, Y. Y. (2018). Experimental poroviscoelasticity of common sedimentary rocks. Journal of Geophysical Research: Solid Earth, 123. https://doi.org/10.1029/2018JB015685 Received 9 MAR 2018 Accepted 28 AUG 2018 Accepted article online 1 SEP 2018 ©2018. American Geophysical Union. All Rights Reserved.

Transcript of Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal...

Page 1: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

Experimental Poroviscoelasticity of CommonSedimentary RocksRoman Y. Makhnenko1 and Yury Y. Podladchikov2

1Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign, Urbana, IL, USA, 2Instituteof Earth Sciences, University of Lausanne, Lausanne, Switzerland

Abstract The success of geoenergy applications such as petroleum recovery or geological storage ofCO2 depends on properly addressing the physical coupling between the pore fluid diffusion andmechanical deformation of the subsurface rock. Constitutive models should include short-termhydromechanical interactions and long-term behavior and should incorporate the principles behind themathematical models for poroelastic and poroviscoelastic responses. However, the viscous parameters inconstitutive relationships still need to be validated and estimated. In this work, we experimentally quantifythe time-dependent response of fluid-filled sedimentary rocks at room temperature and isotropic stressstates. Drained, undrained, and unjacketed geomechanical tests are performed to measure the poroelasticparameters for Berea sandstone, Apulian limestone, clay-rich material, and Opalinus clay (shale). Aporoviscous model parameter, the bulk viscosity, is included in the constitutive relationships. The bulkviscosity is estimated under constant isotropic stress conditions from time-dependent deformation of rockin the drained regime for timescales ~105 s and from observations of the pore pressure growth underundrained conditions at timescales of ~104 s. The bulk viscosity is on the order of 1015–1016 Pa s forsandstone, limestone, and shale and ~1013 Pa s for clay-rich material, and it decreases with an increase inpore pressure despite a corresponding decrease in the effective stress. In the long term, fluid pressurecan asymptotically approach minimum principal stress, which in natural reservoirs may lead to liquefactionor rock embrittlement, causing slip instabilities and earthquakes and creating high-permeability channelsin low-permeable rock.

1. Introduction

Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation is coupled with pore fluid flow, time dependence is introduced into the response of an otherwiserate-independent solid, which can affect the deformation and failure of rock (Rice, 1975; Rice & Cleary,1976). Physical coupling of deformation and fluid flow for reservoir rock needs to be properly addressedbecause its effects are highly important for geoenergy applications such as petroleum recovery or geologicalstorage of CO2 and nuclear waste (e.g., de Waal et al., 2015; Rutqvist, 2012; Tsang et al., 2005).

Biot (1941) and Biot and Willis (1957) first introduced the theory of linear poroelasticity to describe the elasticdeformation of fluid-saturated geomaterials. Their theory of fluid-saturated media has recently beenextended based on rigorous thermodynamic constitutive modeling (e.g., Coussy, 2004; Lopatnikov &Cheng, 2004). Additionally, Biot’s macroscopic model has been confirmed by a number of upscaling techni-ques applied to microscale elasticity (Berryman, 2005; Gurevich, 2007; Lopatnikov & Cheng, 2004). Biot’s the-ory and its extensions are commonly used to describe the behavior of shallow reservoirs that are assumed todeform elastically. Time-dependent behavior (creep) is usually reported for dry rock subjected to high devia-toric stress (Brantut et al., 2013; Costin, 1987; Scholz, 1968). Creep is believed to be a property of a dry matrix,especially when low-viscosity components (clay or mica) are present (Chang et al., 1997). Indeed, mineralcreep has been observed in laboratory experiments at high temperatures (~ 100–1,000 °C) and pressures(~108–109 Pa) (e.g., Cox & Paterson, 1991; Hilairet et al., 2007; Kaboli et al., 2017; Proietti et al., 2016).Although, the extrapolation to natural conditions yields enormous timescales (tens to hundreds of years)in order to observe time-dependent deformation of minerals (Karato, 2010). However, some experimentson low-porosity rock (Elsworth & Yasuhara, 2006) have shown that reactions involving reactive fluids (as inCO2 storage) are fast (hours to weeks), and under relatively modest loads (~106–107 Pa) can lead to pressuresolution and effectively viscous deformation.

MAKHNENKO AND PODLADCHIKOV 1

Journal of Geophysical Research: Solid Earth

RESEARCH ARTICLE10.1029/2018JB015685

Key Points:• Presence of aqueous fluids enhances

viscous deformation of sedimentaryrock

• Poroviscoelastic behavior ofsandstone, limestone, shale, and clayis quantified by laboratory tests

• Bulk viscosity of fluid-saturated rockis ~10

15–10

16Pa s and decreases with

pore pressure increase

Supporting Information:• Supporting Information S1• Data Set S1

Correspondence to:R. Y. Makhnenko,[email protected]

Citation:Makhnenko, R. Y., & Podladchikov, Y. Y.(2018). Experimental poroviscoelasticityof common sedimentary rocks. Journalof Geophysical Research: Solid Earth, 123.https://doi.org/10.1029/2018JB015685

Received 9 MAR 2018Accepted 28 AUG 2018Accepted article online 1 SEP 2018

©2018. American Geophysical Union.All Rights Reserved.

Page 2: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

The viscous model formulated by McKenzie (1984) forms the basis for understanding sedimentary compac-tion and melt migration in partially molten rocks. A few models have been developed to describe the poro-viscoelastic behavior of rock (e.g., Abousleiman et al., 1993; Coussy, 2004), and the principles behind themhave recently been reviewed (Yarushina & Podladchikov, 2015). The influence of creep on fluid-saturated rockhas been shown to play an important role in the long-term storage of nuclear waste (Belmokhtar et al., 2017;Gasc-Barbier et al., 2004) and CO2 (Le Guen et al., 2007; Liteanu et al., 2012; Räss et al., 2018). In general, geo-materials have a nonlinear rheology even when only partially saturated with water, and the time-dependentbehavior of reservoir rock is a function of the combination of the mean stress, deviatoric stress, pore pressure,pore fluid chemistry, temperature, and microstructural properties of the rock, among others (e.g., Brantutet al., 2013; Bürgmann & Dresen, 2008; Sone & Zoback, 2014; Yang et al., 2011; Zhang et al., 2007).

Significant developments in laboratory techniques related to testing fluid-saturated rock at elevated porefluid pressures have now made it possible to properly characterize the poroelastic and poroviscoelasticresponse. Poroviscoelastic models require validation and an accurate estimation of model viscous para-meter(s). The majority of creep observations to date are related to applying a significant deviatoric loadingto rock specimens. This article describes hydrostatic compression experiments at the level of mean stress thatallows us to observe only the poroviscoelastic behavior. Laboratory tests are performed on four fluid-saturated geomaterials representative of quartz-rich, calcite-rich, and clay-rich formations. Here we describethe laboratory techniques used to measure the poroelastic response and creep. A linear model is used tointerpret the results and to project long-term deformation of the rock.

2. Background

Fluid flow in deforming porous media and the classical conservation of mass laws for solid and fluid phases(cf. Coussy, 2004; Frenkel, 1944) can be written as

∂ 1� φð Þρsð Þ∂t

þ ∇j 1� φð Þρsυsj� �

¼ 0 (1)

and∂ ϕρ f� �∂t

þ ∇j ϕρ fυ fj

� �¼ 0; (2)

where ϕ is the interconnected porosity, ∇j = ∂/∂xj, ρs and υs are solid density and velocity, and ρf and υf arefluid density and velocity, respectively. Under small strain approximation and the assumption of homoge-neous deformation (that ρ s and ϕ do not change in space), (1) and (2) can be simplified to

1� φρs

∂ρs

∂t� ∂φ

∂tþ 1� φð Þ∇jυsj ¼ 0 (3)

andϕρ f

∂ρ f

∂tþ ∂ϕ

∂tþ ϕ∇jυsj þ ∇j ϕ υ f

j � υsj� �� �

¼ 0 (4)

Fluid and solid pressures, pfand ps, respectively, can be expressed through the fluid bulk modulus Kf, theunjacketed bulk modulus Ks0, and the unjacketed pore modulus Ks″ (Detournay & Cheng, 1993):

dp f ¼ Kfdρ f

ρ f(5)

anddps ¼ Ks

’ dρs

ρs� ϕ

1� ϕð Þ 1� Ks’

Ks}

� ��dp f ; (6)

Equation (6) becomes a conventional definition of solid pressure (dps = Ks’dρs/ρs) for the case of dry deforma-

tion (pf = 0) and when the unjacketed moduli Ks0 and Ks″ are equal to each other, which is true for microsco-pically homogeneous monomineralic rock with no disconnected pores (Rice & Cleary, 1976). The total mean

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 2

Page 3: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

stress p can be introduced as a function of solid and fluid pressures: p ¼ 1� ϕð Þ�ps þ ϕ�p f (Lopatnikov &Cheng, 2002).

The main rheological assumption adopted here is the Maxwell bulk viscoelasticity. The gradient of solid velo-city can be expressed through Biot’s (1941) effective stress for elastic deformation, and the viscous part is

governed by the effective stress of Terzaghi (1943), pe ¼ p� p f ¼ 1� ϕð Þ ps � p f� �

∇kυsk ¼ � 1Kd

∂p∂t

þ 1Kd

� 1

Ks’

� �∂p f

∂t� p� p f

1� ϕð Þηϕ: (7)

Here Kd is the drained bulk modulus and ηϕ is the effective bulk viscosity, which reflects properties of porousrock such as its pore structure and the viscous and failure parameters of its mineral grains. Biot’s macroscopicmodel considers the effect of pore pressure on both bulk (through Kd) and solid (through Ks0) response of rock(second term on right-hand side of equation (7)). Terzaghi’s approach assumes incompressible elastic limit (1/Ks0 = 0) and simplifies equation (7), making solid velocity gradient being dependent only on Terzaghi’s effec-tive stress pe. Due to the analogy between slow incompressible viscous motion and elastic deformation(Goodier, 1936), the gradients of velocities and flow were shown to be the functions of Terzaghi’s effectivestress and bulk viscosity (Yarushina & Podladchikov, 2015).

Substituting equations (6) and (7) into the solid mass balance equation (1) provides an expression of the vis-cous and elastic evolution of porosity (cf. Detournay & Cheng, 1993; Yarushina & Podladchikov, 2015):

∂ϕ∂t

¼ � 1� ϕKd

� 1

Ks’

� �∂pe∂t

þ ϕ1

Ks’� 1

Ks}

� �∂p f

∂t� peηϕ

: (8)

Here the elastic part can be rewritten as the exact effective stress corresponding to the variation in porosity(Berryman, 1992).

The increment of fluid content is recovered by substituting equations (5) to (7) into the sum of mass balanceequations (1) and (2) to eliminate the time derivatives of densities and the porosity and divergence of solidvelocity (cf. Yarushina & Podladchikov, 2015):

∇j ϕ υ fk � υsk

� �� � ¼ 1Kd

� 1

Ks’

� �∂p∂t

� 1B∂p f

∂t

� �þ pe

1� ϕð Þηϕ; (9)

where B is Skempton’s (1954) coefficient:

B ¼1Kd� 1

Ks0

1Kd� 1

Ks0 þ ϕ 1

Kf� 1

Ks0 0

� � : (10)

The first term on the right side of equation (9) describes poroelastic deformation and is analogous to theincrement in fluid content (Detournay & Cheng, 1993). The second term is responsible for the viscousresponse governed by pe. Equations (8) and (9) present a poroviscoelastic constitutive model and are usedbelow to interpret the laboratory experiments.

3. Experimental Methods3.1. Materials

Three geomaterials are chosen as representative of sedimentary rock formed with different minerals: sand-stone, limestone, and shale. Additionally, a clay-rich soil-like material obtained from crushing the shale istested to compare its response with that of cohesive rock.

Berea sandstone (Ohio, United States) is a fine-grained (0.12 to 0.25 mm), well-compacted, well-sorted, cross-bedded sandstone with a dominant pore diameter of 20 μm. It consists of subrounded to rounded quartzgrains (~90% of solid volume), potassium feldspar (~7%), calcite (~1%), iron oxide (~0.5%), and traces of otherminerals, which makes it a good representative of quartz-rich reservoir material. It has an interconnected

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 3

Page 4: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

porosity ϕ of 0.23 and Darcy permeability k of 4 × 10�14 m2 (measured for steady state flow at pe = 5 MPa).The sandstone has slight (5% to 7%) elastic anisotropy caused by the presence of bedding planes. The uni-axial compression tests in a direction perpendicular to the bedding planes provide a Young’s modulus E of14–15 GPa, a Poisson’s ratio ν of 0.31, and an unconfined compression strength of 41–43 MPa, which is 5%larger than that along the beds (Makhnenko & Labuz, 2016).

Apulian limestone or calcarenite (Italy) is representative of calcite-rich rock. It is a pale orange to grayish glau-conitic fossiliferous limestone composed of calcite (95% to 98%) with traces of quartz, plagioclase, glauconite,and iron oxide. The dominant pore diameter of the rock is 1 to 2 μm, the interconnected porosity is 0.35, andthe permeability is measured as 3.0 × 10�15 m2 (at pe = 0.6 MPa). The elastic anisotropy of calcarenite is within2%, E is 7.3 GPa, ν is 0.25, and the unconfined compression strength is 15–16 MPa (R. Y. Makhnenko &Labuz, 2014).

Opalinus clay (shaly facies) is the Jurassic shale recovered at the Mont Terri underground rock laboratory (DR-A niche). The shale contains 50% clay (illite, kaolinite, chlorite, and smectite), 24% calcite, 20% quartz, 5%organic matter, and traces of other minerals. The interconnected porosity is measured by mercury intrusionporosimetry method to be 0.12, the dominant pore throat diameter is ~15 nm, and the intrinsic permeabilityis ~10�21 m2. Boreholes drilled perpendicular to the apparent bedding planes allow recovering 150 mm indimeter cores, and specimens are prepared from parts of the cores that come from more than 1 m depthin the gallery. They have no visible surface cracks and are immediately sealed after core extraction invacuum-evacuated aluminum foil before cylindrical specimens are prepared. These are tested directly afterthe machining and have 80% to 90% brine saturation. Natural brine that contains 6.13 g/L sodium chlorite(NaCl), 1.63 g/L sodium sulfate (Na2SO4), 1 g/L of CaCl2.2H2O, and MgCl2.6H2O and traces of other chemicalcomponents (Pearson, 2002) is used as the pore fluid to minimize the chemical effect. Its bulk modulus Kf ismeasured in a fluid pressure controller to be 2.0 GPa, slightly less than that of pure water (Makhnenko,Vilarrasa, et al., 2017).

The clay-rich material used is the isotropic remolded phase of Opalinus clay; hence, it has the same mineral-ogy as the shale. Remolded specimens are prepared by crushing the intact material in a grinder, sieving theparticles to a size smaller than 0.5 mm, mixing with brine corresponding to approximately 1.5 times the liquidlimit (or 60%), and consolidating the material at 350-kPa vertical stress for at least 72 hr under one-dimensional strain conditions. The cylindrical specimens obtained have a porosity of 0.33 (calculated fromsolid to bulk density ratio), and the degree of saturation is 0.85 to 0.90. The diameter of the pores in conso-lidated specimens is estimated from mercury intrusion porosimetry tests and ranges from 10 nm to 20 μm.The dominant grain size is 2 μm (for the clay particles), although some grains (quartz) reached 0.2 mm in dia-meter. The interconnected porosityϕ and permeability k are strongly stress dependent at low effective meanstresses and are calculated from volume strain and flow measurements to be 0.28 and 3 × 10�18 m2 atpe = 0.5 MPa, respectively (Makhnenko, Vilarrasa, et al., 2017).

3.2. Poroelastic Parameters

The parameters introduced in section 2 that govern the poroelastic behavior of isotropic fluid-saturated rockare the drained bulk modulus Kd, the unjacketed bulk modulus Ks0, and undrained Skempton’s (1954) B coef-ficient. They can be measured experimentally under three limiting conditions: drained, unjacketed, andundrained (Detournay & Cheng, 1993; Makhnenko & Labuz, 2016). The drained condition occurs when thepore fluid is allowed to leave the rock during loading, and the pore fluid pressure is maintained at a constantlevel: dpf = 0 (where d is the infinitesimal increment). For the undrained condition, the fluid content inside therock does not change, meaning that fluid does not flow through the boundaries of the considered element:

∇j φ υ fj � υsj

� �� �¼ 0. The third possible limiting condition for fluid-rock interaction, the unjacketed condition,

occurs when an increase in the total mean stress p ¼ σ1 þ σ2 þ σ3ð Þ=3 (where σ1, σ2, and σ3 are major, inter-

mediate, and minor principal stresses) is equal to the increase in fluid pressure: dp ¼ dp f .

Alternatively to Skempton’s B coefficient, the unjacketed pore bulk modulus Ks″ can be used. Oftentimes,Ks″ = Ks0 = Ksolid (the bulk modulus of the dominant mineral) is assumed (Detournay & Cheng, 1993;Yarushina & Podladchikov, 2015). However, in general, Ks0 and Ks″ differ: Ks0 is related to a change in the totalspecimen volume in unjacketed compression, whereas Ks″ is associated with a change only in the pore

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 4

Page 5: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

volume. An assumption of Ks″ = Ks0 may lead to improper calculation of other poroelastic properties, althoughKs″ is rarely reported because its direct measurements are associated with very small pore volume changesunder pe = const (Makhnenko, Tarokh, et al., 2017). Ks0, however, is measured under unjacketed boundaryconditions, which are achieved in hydrostatic (σ1 ¼ σ2 ¼ σ3 ¼ p) compression experiments (Makhnenko &Labuz, 2016). Prismatic specimens (50 × 35 × 35 mm) with strain gage rosettes on their sides are saturatedwith the confining fluid (hydraulic oil), which has no chemical effect on the tested rock in the shortterm. This is confirmed by the repeatability of the unjacketed tests results for the same sandstone,limestone, and shale specimens (the latter one only reloaded to 10 MPa) tested a few days after the firstloading—unloading cycles. After at least 24 hr of saturation at 60 MPa pore pressure, unjacketed unload-

ing (p ¼ p f and dp ¼ dp f ) down to 0 MPa is performed in a stepwise manner giving at least 10 min formeasured strains to equilibrate between the steps. Considering the dominant pore throat sizes for thesandstone (20 μm) and limestone (1–2 μm), their saturation with oil (acting as a wetting fluid in air-dryspecimens) can be guaranteed at pore pressures above 30 MPa. Saturation process of the shale (15 nmdominant pore throat size) is more involved, and the application of 60 MPa fluid pressure for 15 days isrequired to bring the shale specimen with ≈0.85 brine saturation to full saturation. It is assumed that atthis point, oil pushes all the in situ brine (wetting fluid) into the smallest pores and occupies the remainingpore space. The unjacketed unloading of the shale specimen is performed for 30 days, and each 1- to 3-MPa unloading step is considered to be complete when measured normal strains are equilibrated after thediffusion of pore fluid pressure inside the rock (Makhnenko, Tarokh, et al., 2017). Measurements of three-dimensional strains at different pressures provided data to determine the elastic (quasi-static) anisotropyof the materials and the unjacketed bulk moduli.

Drained and undrained compression experiments on sandstone and limestone are conducted with theUniversity of Minnesota plane strain apparatus (Labuz et al., 1996), specifically modified and calibrated fortesting fluid-saturated rock (Makhnenko & Labuz, 2016). Prismatic specimens (100 × 87 × 44mm) are coveredwith polyurethane and wedged inside the stiff frame that inhibits their deformation in one of the directions. Ifonly axial loading is applied, the increments in intermediate strain (aligned with the direction of passiverestraint) are 2 orders of magnitude smaller than strains in the maximum (axial strain) and minimum (lateralstrain) principal stress directions. The apparatus allows accurate measurements of all principal stressesand strains (within 5 kPa and 10�5, respectively) at different mean stress and pore pressure levels(R. Makhnenko & Labuz, 2014). If the stress induced by the passive restraint is preset in a way that is equalto the applied axial and lateral stresses, an approximately hydrostatic loading condition can be assumedfor the testing with plane strain apparatus.

Drained and undrained shale testing is performed on cylindrical specimens (105-mm high and 50 mm in dia-meter) in a high-pressure (70 MPa) triaxial cell that allows both upstream and downstream fluid injection andrecording as well as measurements of axial and lateral deformation (Figure 1). The volume strain, stresses, andpore pressures are measured with 1 × 10�5, 10 kPa, and 5 kPa accuracy, respectively. A similar triaxial cell witha 3.5 MPa capacity is utilized for sandstone, limestone, and clay saturation and accurate parameter measure-ments at low total stresses. The 3.5 MPa triaxial apparatus allows total and pore pressure control within 1 kPa,and volume strain measurements are accurate to approximately 1 × 10�5.

Skempton’s B coefficient characterizes the increase in pore pressure resulting from the applied loading underthe undrained condition,0 ≤ B ≤ 1, and is calculated from the following expression (Bishop, 1976):

B ¼ 1

dpdp f

� �measured� 1

1Kd� 1

Ks ’

� � VLVKf

þ CLMV

� � ; (11)

which takes into consideration the contribution of VL and CLM, the dead volume and the compressibility of thepore pressuremeasuring system, respectively. The latter one contributes to less than 5% of the correction fac-tor (the second term in equation (11)), because high-pressure stainless steel tubing and pressure transducersare used. The dead volume is equal to 14 ml in plane strain apparatus, 12 ml in 70 MPa triaxial cell, and 9.5 mlin 3.5 MPa triaxial cell. The bulk modulus of the pore fluid is assumed to be equal to the bulk modulus ofdeaired water, namely, Kf = Kwater = 2.24 GPa at full saturation. The correction factor is material dependent

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 5

Page 6: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

and is determined to be 0.1 to 0.2 for materials tested in the plane strain apparatus and 70 MPa triaxial celland not more than 0.1 for the 3.5 MPa triaxial cell (less than 0.01 for clay-rich material at low pressure).

The back pressure saturation technique is implemented for porous sandstone and limestone (Makhnenko &Labuz, 2016). After applying desired mean stress and flushing deaired water through the rock until the steadystate flow is achieved and a few pore volumes of water go through the specimen, increments of pore (orback) pressure are applied while keeping the effective mean stress pe approximately the same. The purposeof the back pressure technique is to achieve 100% saturation by forcing any gas into solution of the porewater. Increase of the pore pressure in a partially saturated specimen compresses the gas in the pores anddissolves it in the pore water in accordance with Henry’s law of solubility (Lowe & Johnson, 1960).Skempton’s B coefficient is measured at gradually increasing back pressures (at pe = constant), and thisprocedure is referred to as B check. The value of B is increasing with increasing pf because more air is forcedinto solution and compressibility of the pore fluid (1/Kf in equation (11)) decreases. The number of steps andtime between them (from one to several hours) depends on the desired final value of back pressure at whichspecimen is expected to be saturated. Short time between the steps (a few minutes) normally does not allowair bubbles to dissolve in pore fluid and corresponds to higher back pressures needed to be applied toachieve full saturation. A constant and independent of the magnitude of the back pressure value of B indi-cates full saturation, and the air-water mixture behaves as a fluid with a bulk modulus equal to that of purewater, so dissolved air in the water has no influence on Kf.

Low permeability of shale and clay-rich material do not allow flushing a few pore volumes through them,although consolidation of the partially saturated specimens even at low mean stresses brings their degreeof saturation to about 0.97. Consequently, the back pressure saturation technique is applied with pf beingincreased in 0.1 MPa steps for clay-rich material and 0.5 to 1.0 MPa steps for Opalinus clay with the waitingperiod between the steps being between 1 and 3 days.

Figure 1. Sketch of the conventional triaxial compression test on a cylindrical specimen (105 mm high and 50 mm in dia-meter). Total isotropic (all-around equal) stress p is controlled, and pore pressure pf is either controlled by upstream anddownstream controllers (drained test) or measured by upstream and downstream pressure transducers when the “in” and“drainage” valves are closed (undrained test). The direct method of measuring bulk viscosity implies the volume strainevolution is monitored with time over at least 5 × 104 s (14 hr) under the drained condition. The indirect method is basedon recording the evolution of pore pressure buildup under the undrained condition for the timescales ~104 s (3 hr).

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 6

Page 7: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

3.3. Bulk Viscosity

The material response under constant mean stress (dp=dt ¼ 0) and the drained boundary condition for porefluid pressure (dpf/dt = 0) are considered. The bulk viscosity is calculated following the formulation in equa-tion (7), hereafter referred as the direct method:

ηϕ ¼ � p� p f

1� ϕð Þ∇kυsk; (12)

where ∇kυsk is the divergence of solid velocity and is interpreted as the measured volume strain rate, �dε/dt.The volume strain is calculated to an accuracy of 1 × 10�5, and the effective mean stress value for a shallowsedimentary rock is on the order of a few megapascals. Therefore, for bulk viscosity values of ~1015 Pa s, thetimescale required to evaluate ηϕ to two significant digits is ~104 to 105 s (a few hours to a few days). To com-pensate for the effects of temperature and pressure perturbations, the timescales are even longer; hence, thebulk viscosity of shallow rock is rarely reported for isotropic loading conditions.

For porous materials, the degree of saturation is the critical parameter governing their viscous responseunder undrained conditions. If saturation is not complete, pore pressure does not increase significantly underthe undrained condition but rather decreases as the air bubbles dissolve in pore water. However, when fullsaturation is achieved and thematerial deformation can be considered poroviscoelastic, equation (9) predictsa pore pressure buildup with time under conditions of constant total stress. The simplified form of equa-

tions (8) and (9) for the undrained deformation (no-flow) condition, ∇j φ υ fj � υsj

� �� �¼ 0, at constant mean

stress (dp=dt ¼ 0) leads to another method of evaluating ηϕ, hereafter referred to as the indirect method:

ηφ ¼ B p� p f� �

1� φð Þ dp f

dt1Kd� 1

Ks0

� � : (13)

The evaluation of bulk viscosity then requires accurate measurements of the poroelastic parameters.Changes in pore pressure (measured with 1 to 5 kPa accuracy) needed to report bulk viscosity ηϕ~1015 Pa s with two significant figures can be recorded over the timescales of 103 to104 s (one to few hours).Hence, pore pressure can be used as the monitoring parameter for viscous behavior and for calculating thebulk viscosity. In the undrained tests, pore pressure is observed to increase monotonically, making this indir-ect method significantly less time consuming. Figure 1 illustrates both the direct and indirect methodsapplied in this study.

4. Experimental Results4.1. Poroelastic Response

The characteristic time for dissipation of the induced pore pressure in a specimen of length L drained at twoends is on the order of L2/4c, where c is the diffusion coefficient (Detournay & Cheng, 1993), which can beexpressed from the permeability k, fluid viscosity μ, and measured poroelastic parameters including shearmodulus G:

c ¼ kBKd

μ 1� Kd

Ks’

� �1� 1� Kd

Ks’

� �4BG

3Kdþ4G

� � : (14)

Values of c for the tested materials are reported in Table 3 and range from 0.4 m2/s for water-saturated Bereasandstone to ~10�8 m2/s for water-saturated clay-rich material and shale. The characteristic timescale forequilibration of the pore pressure inside the rock, at the range of stresses applied, varies from 0.01 s forwater-filled 100-mm long sandstone specimens to hundreds of seconds for oil-filled limestone and tens ofthousands of seconds for shale (35-mm wide specimens), which is still significantly smaller than the time-scales of the reported tests (hours to a few days).

Poroelastic parameters are determined from instantaneous (short-term) deformation before any viscouseffects occurred, with the exception of low-permeable shale and clay, for which the diffusion time for the

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 7

Page 8: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

drained deformation is on the order of tens of hours. The influence of time-dependent deformation on the measurements of poroelastic properties isdiscussed in sections 4.2 and 4.3. The results and corresponding data forthe unjacketed, drained, and undrained tests are presented inFigures 2a–2c and Table 3, along with the respective total and pore pres-sures at which they were measured. Instead of a drained test for calcare-nite, the results of a dry (pf= 0) hydrostatic compression test arereported. The unjacketed bulk modulus is measured for Berea sandstone,Apulian limestone, and Opalinus clay. Linear response of these materialsprovides stress-independent values for Ks’, which qualitatively confirmsthat full saturation has been achieved. The unjacketed bulk modulus ofclay-rich material is assumed to be the same as for the shale, because theyhave the same mineral composition and approximately the same porestructure. The assumption is appropriate because interconnected porosityof the shale (measured with mercury intrusion porosimetry on three speci-mens to be ≈0.124) is just 0.003 smaller than the total shale’s porosity(measured via pulverization method on three other specimens to be≈0.127), meaning that the nonconnected pore volume of the material isvery small.

Skempton’s B coefficient values are measured at each increment of porepressure, while the effective mean stress is held constant. The timescaleof the saturation process is known to depend on the geometry of the spe-cimens and the applied pressure gradients (Makhnenko & Labuz, 2016).The smaller the applied pressures, the longer it takes to force any gas intosolution of the pore water and achieve full saturation (Lowe & Johnson,1960). The Berea sandstone specimen can be fully saturated within 1 daywith the application of pf> 2.5 MPa (at pe = 7MPa), and the saturation pro-cess takes at least 1 week if pore pressures are below 1MPa. For the calcar-enite specimen tested at pe = 0.2 MPa, the maximum B coefficient isachieved for pf > 1.1 MPa after 7 days of back pressure saturation. At thesame time, it is possible to saturate the calcarenite if pore pressure is gra-dually increased to 2 MPa within a 2-day period. Moreover, it takes 3 weeksto fully saturate Opalinus clay at pf > 8 MPa and 4 weeks to fully saturatethe clay-rich material at pore pressures above 1.0 MPa and pe = 0.3 MPa. Inaddition, the clay-rich material and shale exhibited some swelling duringthe saturation process (at most, 0.5% and 0.05%, respectively), but theirbehavior is compactive thereafter and could be interpreted by means ofthe poroviscoelasticity.

4.2. Time-Dependent Deformation

The bulk viscosity values for Berea sandstone, calcarenite, clay-rich mate-rial, and Opalinus clay are calculated under drained condition from thevolume strain changes at a constant hydrostatic loading (equation (12))after achieving full saturation (Table 1). Note that each measurement istaken for at least 5 × 104 s to guarantee two significant figures accuracyin reporting ηϕ. The case of zero fluid pressure corresponds to dry creepunder the hydrostatic condition for the sandstone and limestone.

The diffusion times are on the order of a few seconds when drained load-ing is applied to sandstone and limestone specimens, but they are muchlonger for the shale and clay-rich material. In these cases, the viscousresponse starts playing an important role in evaluation of poroelastic prop-erties and, conversely, the diffusion processes affect the measurements ofviscous parameters at early times. Figure 3a demonstrates the results of

Figure 2. Poroelastic response of Berea sandstone (blue), calcarenite (green),clay-rich material (maroon), and Opalinus clay (orange): (a) mean (hydro-static) stress-volume strain data for the unjacketed compression; (b) effectivemean stress-volume strain data for the drained compression (a dry com-pression test, pf = 0, instead of a drained test is performed on calcarenite);and (c) Skempton’s B coefficients from the undrained compression as afunction of the pore pressure. Note that for the undrained tests, the porepressure at full saturation is also a function of time.

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 8

Page 9: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

drained test on Berea sandstone when the effective mean stress is mono-tonically increased for 1 hr from 6.0 to 8.0 MPa, while the pore pressure ispreserved constant at 2.2 MPa. Measurement of volume strain allows thecalculation of bulk modulus Kd. The consequent deformation of the speci-men for 20 h is interpreted as viscous response of rock and bulk viscosity iscalculated according to equation (12). The diffusion process takes approxi-mately 2 days in 100-mm long shale specimen loaded within 4 hr from 24to 26 MPa while preserving the pore pressure at 9 MPa. When the diffusionis over, the specimen keeps compacting and volume change with time isused to calculate the bulk viscosity from equation (12). The differentiationbetween diffusion and viscous processes in clays is extensively discussedin soil mechanics literature (cf. Mesri & Castro, 1987). In our case, approxi-mately linear volume increase with time is subtracted from the overallvolume strain and provides the calculation of the bulk modulus Kd.Similarly to soil mechanics terminology where viscous deformation iscalled secondary compression (Taylor & Merchant, 1940), we also presentthe volume strain vs time on a semi logarithmic plot where the end of dif-fusion process (primary compression) appears to be more obvious (inset inFigure 3a).

4.3. Undrained Pressure Buildup

The responses of Berea sandstone, calcarenite, clay-rich material, andOpalinus clay are monitored under the undrained condition at room tem-perature (24 °C) after each B check, while a constant isotropic total stress pis acting on the specimens. An example of B check performed on Bereasandstone for 14 min from p = 8 to 9 MPa with initial pore pressurepf = 4.0 MPa is shown in Figure 3b. Another curve in Figure 3b demon-strates the experiment on Opalinus clay where p is continuously increasedfrom 11 to 12 MPa for 2 hr with initial pf = 8.1 MPa. The undrained loadingis followed by the pore pressure buildup dpf/dt at p = 12 MPa = constant.The measurement of Skempton’s B coefficient is affected by the viscous

deformation of rock, so the viscous pressure buildup needs to be subtracted from the total pore pressure

increase during the application of loading. The obtained dp f=dp value needs to be corrected for the deadvolume in the system to get the final value of Skempton’s B coefficient (equation (11)). Knowledge of B, Kd,and Ks0 then allows for the calculation of bulk viscosity ηϕ from the viscous pressure buildup usingequation (13).

If the behavior of rock is purely poroelastic, application of constant p should not produce any perturbationsin the pore fluid pressure. However, pore pressure buildup occurred for fully saturated specimens at differentinitial pressures and at fixed values of the total stress for all four tested materials. Figures 4a to 4d indicateviscous and approximately linear behavior of the rock (in terms of pore pressure) to applied undrained load-ing after all elastic response is complete (see Figure 3b). These measurements are not related to any inertiaeffects and are reproducible for different experimental timescales. The observed behavior can be explainedas resulting from the time-dependent or poroviscoelastic response of water-saturated rock, where steepercurves indicate lower bulk viscosity values. Parameters corresponding to the pore pressure increase with timeare reported in Table 2, along with the goodness of linear fit (R2) values, which demonstrates that the linearityof pore pressure buildup is a good assumption for short-term (~104 s) viscous response of sedimentary rock.Some nonlinearity at the beginning of observation is, probably, related to uncertainty of determining theonset of purely viscous response. Note that the starting values for the pressure buildup are not material prop-erties but are related to the specific saturation procedure.

Themeasured poroelastic parameters and data on the pore pressure increase presented in Figures 4a–4d andTable 2 are used to calculate the bulk viscosity of the sedimentary rock. Measurements at the beginning ofmonitoring are very sensitive to pore pressure fluctuations but then become approximately constant. The

Table 1Parameters Corresponding to the Direct Method Measurements of BulkViscosity ηϕ for Berea Sandstone, Calcarenite, Clay-Rich Material, andOpalinus Clay (Shale)

Material p(MPa) pf(MPa) ηφ (Pa s)

Berea sandstone 1.7 1.4 3.0 · 1015

1.8 1.3 5.0 · 1015

6.0 0.0 2.1 · 1016

6.0 1.8 1.8 · 1015

8.0 2.2 3.4 · 1015

8.0 2.6 2.0 · 1015

8.8 2.1 4.3 · 1015

10.0 1.8 6.6 · 1015

11.0 3.0 4.6 · 1015

19.0 3.0 1.2 · 1016

Calcarenite 1.85 1.7 4.6 · 1014

2.2 0.0 7.7 · 1015

2.2 1.4 8.6 · 1014

2.7 1.0 1.6 · 1015

6.0 0.95 9.1 · 1014

7.1 1.9 1.3 · 1015

8.0 1.8 3.7 · 1015

Clay-rich material 1.72 1.42 4.0 · 1014

1.72 1.6 3.2 · 1013

1.72 1.66 3.9 · 1013

1.77 1.66 5.0 · 1013

2.7 1.7 3.0 · 1014

Opalinus clay 12.3 6.0 3.2 · 1015

13.4 7.0 3.1 · 1015

14.9 9.0 2.5 · 1015

26.0 9.0 1.5 · 1016

26.0 9.5 1.4 · 1016

26.0 12.0 7.8 · 1015

35.0 10.0 2.6 · 1016

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 9

Page 10: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

values of ηϕ calculated for each material at the end of the observation(~104 s) at different effective stresses are reported in Table 3.

Based on the accuracy of the experimental equipment, the absolute errorsare reported to be the following: strain (10�6–10�5), length (10�7–10�5 m,depending on the device), volume of fluid (1 mm3), time (1 s), load(0.025–0.25 kN), and pore pressure (1–50 kPa, depending on the device).Accuracy of the experimental measurements and repeatability of the testresults for the same specimens allow for calculation of the relative errorsfor the measured parameters reported in Table 3 as 1% for the porosity,3% for Ks0, and 5% for the permeability. Kd is measured within 3%, 4%,5%, and 10% accuracy for Berea sandstone, calcarenite, Opalinus clay,and the clay-rich material, respectively. Special attention is paid to precisemeasurements of Skempton’s coefficient B and calibration factors in equa-tion (11), so B is reported within ±0.01 accuracy. The diffusion coefficient cis calculated from equation (14), and consideration of the relative errors forthe parameters consisting it provides the accuracy of its determinationbeing 15% for the sandstone and limestone and 20% for the clay-richmaterial and shale. Considering less than 0.3% error in measuring pres-sures (0.1% of full scale) and time (a few seconds at most) allows for theevaluation of the calculated bulk viscosity accuracy ηϕ (equation (13)) as10% for Berea sandstone and calcarenite and 15% for the clay-rich materialand Opalinus clay.

4.4. Effective Stress for Bulk Viscosity

The experimental measurements of bulk viscosity as a function of theTerzaghi effective pressure pe obtained by the direct and indirect methodsfor each material are presented in Figure 5. In addition, the Mohr-Coulombfailure surfaces are plotted in pe, q plane, and q = (σ1� σ3) at failure can beexpressed from cohesion cc and friction angle φ

q ¼ 6 sinφ3� sinφ

pe þ6cc cosφ3� sinφ

: (15)

Cohesion cc and friction angle φ are taken from the previous work on Bereasandstone, cc = 7.2 MPa and φ = 44° (Makhnenko et al., 2015), Apulian lime-stone, cc = 5.3 MPa and φ = 20° (Meyer & Labuz, 2013), and Opalinus clay,cc = 2.2–5.0 MPa and φ = 23–25° (Bossart & Thury, 2008). The specimens inthose works are prepared from the same blocks of rock as in this studywith the exception of shaly facies of Opalinus clay, where the potentialvalues of cc and φ are reported for the range of porosities and mineralogi-cal compositions. Additionally, cohesion and friction are measured for

clay-rich material to be cc = 0.02 MPa and φ = 24°. In general, no conclusion can be drawn regarding the effec-tive stress dependency of the bulk viscosity except for Opalinus clay, where ηϕ seems to be pe dependent atthe range of effective mean stresses explored. These stresses are relatively low because of the elevated porepressures applied to promote shale saturation. More indirect tests would be helpful in judging if the trendwould stay approximately linear after including measurements at higher mean stresses.

4.5. Viscous Compaction

During undrained pressure buildup, specimens are compacting, and the corresponding change in porositydϕ is calculated from equation (8) by expressing Ks″ from Skempton’s B coefficient (equation (10)) as pre-sented in Figure 6. In the case of Berea sandstone, we observe no chemical effect of the pore fluid on poro-viscoelastic parameters of the rock even at fluid pressures above 8 MPa, which is confirmed by therepeatability of the results for the measurements performed on the same specimens at the same conditions.Measured volume strains are consistent with the predicted compaction of the material. For example, at p

Figure 3. Measurements of poroelastic and viscous parameters of Bereasandstone and Opalinus clay via (a) direct method when the drained load-ing is applied and bulk modulus Kd and bulk viscosity ηϕ are calculated fromvolume strains, and (b) indirect method when the undrained loading isapplied and Skempton’s B coefficient and bulk viscosity are calculated frompore pressure increase.

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 10

Page 11: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

= 7.0 MPa and pf = 4.1 MPa, the change in rock porosity (equation (8)) is calculated as�3 × 10�5 for 104 s, andthe measured volume strain is recorded as 5 × 10�5 (compacting). The experiments performed on calcareniteat p = 9.0 MPa and pf = 2.1 MPa provide a volume strain of 1 × 10�4, which is consistent with the predicteddecrease in the porosity of the tested specimen (�8 × 10�5). The clay-rich material is tested at pore pressuresbelow 2 MPa, and the observed volume changes are consistent with the predicted changes in porosity (e.g.,dϕ = �5.2 × 10�5 and volume strain = 7 × 10�5 at p = 1.7 MPa and pf = 1.2 MPa). Similarly, for Opalinus claytested at p = 11 MPa and pf = 8.5 MPa, the volume strain is recorded as 2.5 × 10�5 for 5 × 104 s, whereas thecalculated porosity change is�1.2 × 10�5. Increasing the pore pressure at a constant mean stress should leadto an increase in porosity if a geomaterial is deforming poroelastically. However, all the tested specimens areexperiencing time-dependent compaction, and the measured volume strain values are qualitativelyconsistent with porosity changes predicted by the poroviscoelastic relationships presented, meaning thatthe constitutive model predicts the direction of deformation and its order of magnitude.

5. Discussion

Accurate stress, pore pressure, and strain measurements allow the determination of the parameters govern-ing the poroviscoelastic response of Berea sandstone, Apulian limestone (calcarenite), clay-rich material, andOpalinus clay (shale). The poroelastic behavior of Berea sandstone and clay-rich material is observed to bestrongly stress dependent: parameters governing the material response in equations (8) and (9) change withthe effective mean stress applied. For calcarenite and Opalinus clay, the poroelastic parameters are found to

Figure 4. Pore pressure development with different initial values as a function of time in the undrained tests with (a) Bereasandstone (at p = 8 MPa), (b) calcarenite (at p = 9 MPa), (c) clay-rich material (at p = 1.5 MPa), and (d) Opalinus clay (at p= 12 MPa).

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 11

Page 12: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

be weakly stress dependent: the dry bulk modulus changed only slightlywith increasing effective mean stress, and Skempton’s B coefficientdecreased moderately.

The effective stress does not appear to be the only parameter governingthe bulk viscosity. We note that the bulk viscosity for all four testedmaterials exhibits smooth decreasing trends when plotted versus adimensionless parameter, the ratio between the pore pressure and total(isotropic) stress. Figure 7 illustrates the general trend of decreasing bulkviscosity with an increase in the pore fluid pressure ratio, and at high values

of p f=p, Berea sandstone, calcarenite, andOpalinus clay reach ηϕ ~ 1015 Pa s.Shale and clay-rich material are extremely difficult to saturate at low pore

pressures, so their bulk viscosities are reported only for p f=p > 0:3, with

most of the tests performed at p f=p > 0:5. The clay-rich material shows

a continuously decreasing trend of ηϕ down to ~1013 Pa s.

Bulk viscosities below 1015 Pa s measured for calcarenite at p f=p≈0:3(shown in Figures 5b and 7b) correspond to the tests in which the porefluid pressure exceeds 2.5 MPa. The decrease in material viscosity in thesecases can be explained by the chemical effect of pore water at elevatedpressures (pf > 2.5 MPa) on the calcite-rich rock matrix. This observationis confirmed by the decreased values of P wave velocity (R. Y.Makhnenko & Labuz, 2014), irreversible deformation, and the relativelylarge volume strain of ≈1 × 10�3 (compression positive) recorded at theend of the observation in these tests. In contrast, the predicted decreasein porosity attributable to the poroviscoelastic effect is an order of magni-tude smaller. At the level of utilized mean stresses and pore pressures, weobserve no chemical effect of the pore fluid (in terms of the repeatability ofthe results for the same specimens) on the poroviscoelastic materialresponse for distilled water-saturated Berea sandstone and calcarenite atpf < 2.5 MPa and for clay-rich material and Opalinus clay (shale) saturatedwith the natural brine for shale (Pearson, 2002).

Although the bulk viscosities calculated via the direct and indirect meth-ods show little difference, we suggest that the latter be used if measure-ments of poroelastic parameters are available. During undrained loading,the rock matrix is assumed to deform simultaneously everywhere; hence,the corresponding pore pressure increase also happens very quickly andcan be measured on pore pressure transducers attached to the pressurelines connected to the specimen. The indirect method does not requirechallenging volume strainmeasurements, and performing involved triaxialtests is not necessarily for evaluating bulk viscosity. In fact, core floodingdevices with pressure transducers attached to pore pressure lines are sui-table for performing undrained loading and evaluating viscous response

of rock at timescales of ~104 s (hours). Whereas strain measurements are less accurate and more sensitiveto environmental noise than are pore pressure measurements, so significantly longer waiting times(>5 × 104 s or tens of hours) are required for consistent measurements of bulk viscosity by the direct method.In addition, the drained loading requires achieving pore pressure equilibration, and the dissipation process istime consuming for low-permeable rock (on the order of days).

It is also interesting to note that the ratio of the bulk viscosity to the drained bulk modulus, the characteristicobservation time tchar = ηϕ/Kd (or Maxwell time at which the viscous strain becomes comparable to the elasticstrain) is on the same order of magnitude of ~105 s or days for each rock-like material (see Table 3) and is onthe order of 106 s (weeks) for the clay-rich material. This characteristic time can be considered as currentbecause the bulk viscosity it contains could change with time. Typical viscosity values for solid Earth

Table 2Parameters Corresponding to the Linear Fit of Pore Fluid Pressure pf BuildupWith Time t, pf = c0 + c1t, and the Resulting Indirect Calculations of BulkViscosity ηϕ for Berea Sandstone, Calcarenite, Clay-Rich Material, andOpalinus Clay

Material p (MPa) c0 (MPa)c1

(10�5 MPa/s) R2 ηϕ (Pa s)

Berea sandstone 3.2 1.48 1.3 0.97 4.0 · 1015

7.0 4.10 1.7 0.99 9.2 · 1014

8.0 2.32 1.5 0.99 4.4 · 1015

8.0 2.52 2.6 0.99 2.7 · 1015

9.0 4.03 1.5 0.98 3.4 · 1015

9.9 6.19 2.4 0.99 1.5 · 1015

10.0 1.80 1.2 0.99 6.6 · 1015

10.0 2.03 1.7 0.99 4.6 · 1015

10.1 4.83 3.1 0.98 1.8 · 1015

11.0 5.91 1.8 0.98 3.0 · 1015

20.0 3.02 2.0 0.99 8.4 · 1015

23.7 2.97 1.3 0.99 1.4 · 1016

Calcarenite 1.3 1.11 0.14 0.97 8.9 · 1014

1.4 0.79 0.24 0.99 1.5 · 1015

1.4 0.96 0.14 1.00 1.6 · 1015

1.8 0.98 0.16 0.98 3.3 · 1015

1.8 1.12 0.17 0.99 2.6 · 1015

8.0 2.45 1.5 0.99 2.1 · 1015

8.7 2.93 7.0 0.99 4.5 · 1014

9.0 2.09 0.70 0.97 4.9 · 1015

9.0 2.28 3.2 1.00 1.3 · 1015

9.0 2.70 6.5 0.99 6.3 · 1014

Clay-richmaterial

1.30 1.17 0.03 0.95 3.4 · 1013

1.50 1.41 0.006 0.96 4.0 · 1013

1.50 1.44 0.004 0.99 5.2 · 1013

1.50 1.47 0.002 0.98 5.0 · 1013

1.70 1.20 0.09 0.97 9.3 · 1013

1.72 1.01 0.13 0.98 1.4 · 1014

2.80 2.32 0.14 0.96 6.4 · 1013

Opalinus clay 10.0 3.43 0.39 0.95 5.4 · 1015

11.0 8.11 0.74 0.94 1.2 · 1015

12.0 8.07 0.81 0.99 1.9 · 1015

12.0 9.42 0.80 0.97 9.0 · 1014

12.0 9.70 0.50 0.97 1.9 · 1015

13.0 8.92 0.79 0.96 1.5 · 1015

13.5 6.33 0.53 0.98 3.5 · 1015

14.0 7.59 0.83 0.96 2.4 · 1015

14.0 10.40 0.76 0.97 1.4 · 1015

15.0 10.05 0.89 0.99 1.7 · 1015

15.0 12.91 0.62 0.99 9.5 · 1014

15.0 13.97 0.36 0.99 6.0 · 1014

16.5 13.61 0.58 0.97 1.4 · 1015

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 12

Page 13: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

materials measured under laboratory conditions at elevated temperatures and differential stresses aresimilar to the values reported here (~1014�1016 Pa s; e.g., Renshaw & Schulson, 2017), whereas theirextrapolation to natural conditions yields much higher values of 1019 to 1022 Pa s (Karato, 2010),consistent with indirect geophysical constraints such as the ellipticity of the Earth and postglacialrebound (Turcotte & Schubert, 2014).

Most often, creep is reported for high-pressure or high-temperature conditions, where it becomes observableabove half the yield stress and/or half the melting point (cf. Vanel et al., 2009). Time-dependent rupture, how-ever, is one factor and the yield strength of materials is reported to drop with time. A range of micromechan-isms including atomic diffusion, dissolution, ion exchange, microplasticity, and stress corrosion could beresponsible for subcritical crack growth and subsequent creep (Atkinson & Meredith, 1987; Brantut et al.,2013). Aharonov and Scholz (2018) argue that creep on asperities is the reason of frictional phenomena inrock. Laboratory experiments have shown that brittle creep occurs in all major rock types and that creep

Figure 5. Bulk viscosity as a function of the Terzaghi effective pressure pe for (a) Berea sandstone, (b) Apulian limestone(calcarenite), (c) clay-rich material, and (d) Opalinus clay measured by the direct (open symbols) and indirect (solid sym-bols) methods. Plotted on the right vertical axis is deviatoric stress q at failure as predicted by the Mohr-Coulomb criterion.

Table 3Measured and Calculated Poroelastic and Poroviscoelastic Material Parameters of Berea Sandstone, Apulian Limestone (Calcarenite), Clay-Rich Material, and OpalinusClay (Shale) at Different Levels of Total Mean Stress p and Pore Pressure pf

Material Test conditions

Measured Calculated

ϕ (�) k (m2) Kd (GPa) Ks0 (GPa) B (�) c (m2/s) ηϕ (Pa s) tchar (105 s)

Berea sandstone p = 7.0 MPa pf = 4.1 MPa 0.23 4.8 · 10�14 6.5 30 0.71 0.4 0.9 · 1015 1.4p = 8.0 MPa pf = 2.5 MPa 0.23 4.0 · 10�14 9.5 30 0.61 0.4 2.7 · 1015 2.8p = 20 MPa pf = 3.0 MPa 0.23 2.8 · 10�14 11.3 30 0.40 0.2 8.4 · 1015 7.4

Apulian limestone p = 1.3 MPa pf = 1.1 MPa 0.35 3.0 · 10�15 4.8 43 0.76 1.8 · 10�2 0.9 · 1015 1.9p = 1.4 MPapf = 0.8 MPa 0.35 3.0 · 10�15 4.9 43 0.74 1.7 · 10�2 1.5 · 1015 3.1p = 9.0 MPa pf = 2.3 MPa 0.35 2.7 · 10�15 5.1 43 0.70 1.5 · 10�2 1.3 · 1015 2.5

Clay-rich material p = 1.5 MPa pf = 1.4 MPa 0.30 9 · 10�18 0.01 8.9 0.99 1.3 · 10�7 4.0 · 1013 40p = 2.8 MPa pf = 2.3 MPa 0.28 3 · 10�18 0.043 8.9 0.98 1.9 · 10�7 6.4 · 1013 15p = 1.7 MPa pf = 1.0 MPa 0.27 2 · 10�18 0.072 8.9 0.97 2.1 · 10�7 1.4 · 1014 19

Opalinus clay (shale) p = 15.0 MPa pf = 14.0 MPa 0.12 9 · 10�21 1.85 8.9 0.92 2.6 · 10�8 6.0 · 1014 3.5p = 11.0 MPa pf = 8.1 MPa 0.12 7 · 10�21 2.15 8.9 0.90 2.3 · 10�8 1.2 · 1015 5.7p = 14.0 MPa pf = 7.6 MPa 0.12 6 · 10�21 2.45 8.9 0.88 2.3 · 10�8 2.4 · 1015 10

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 13

Page 14: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

strain rates are very sensitive to the environmental conditions, such as dif-ferential and mean stresses, temperature, and pore fluid composition (cf.Brantut et al., 2013). The effect is enhanced with increase of water satura-tion and pore pressure (e.g., Bernabe et al., 1994; Wawersik & Brown, 1973),which is confirmed by the presented experimental data. Increase in fluidpressure reduces effective stress globally and promotes stress concentra-tion on asperities locally, thus enhancing time-dependent solid deforma-tion. While the common approach dictates that fluid pressure growthtoward mean stress leads to dehydration embrittlement (cf. Paterson &Wong, 2005), this study shows the drop of effective viscosity leading tofluidization. So more ductile behavior should be expected instead of thebrittle one, although the mechanisms behind viscous response of rockneed further experimental verification including measurements atmicroscales.

The bulk viscosity of serpentine antigorite can be calculated from themea-

surements provided by Hilairet et al. (2007) and appears to be ~1017 Pa s at200 °C and 0.1 GPa of deviatoric stress (as shown in Figure 7a), which ismuch lower than that of the major mantle-forming minerals. This valueapproaches those for Berea sandstone at low pore pressure to total stressratios. Vu et al. (2012) reported a cement paste (similar in poroelastic prop-erties to calcarenite) as having a bulk viscosity of approximately

2.5 × 1014 Pa s, and its decrease by a factor of 2 with an increase in tem-perature from 20 to 90 °C.

An increase in the clay mineral fraction of geomaterials has been observedto make the viscous effects more pronounced. For example, Chang et al.(1997) measured a decrease in bulk viscosity (recalculated from creep)by a factor of 4 when 10% wetted montmorillonite is added to unconsoli-dated Ottawa sand (ηϕ ~ 1014 Pa s). Moreover, an increase in the relativehumidity or pore pressure appears to enhance time-dependent deforma-tions in clay-rich materials with ηϕ ~ 1014 to 1016 Pa s for both argillites(Armand et al., 2016; Fabre & Pellet, 2006; Yang et al., 2011; Zhang et al.,2007) and shales (Sone & Zoback, 2014). Gasc-Barbier et al. (2004) andBelmokhtar et al. (2017) also report an increase by a factor of 2 in creeprate with a temperature increase from 20–25 to 80–85 °C for fully saturatedBure clayey rock and Callovo-Oxfordian claystone, respectively. For bothmaterials, ηϕ ~ 1015–1016 Pa s and is reported even at isotropic stress(Belmokhtar et al., 2017) or at a very low deviatoric loading (10% strengthin Gasc-Barbier et al., 2004). However, the majority of creep tests on clay-richmaterials are aimed at studying an increase in creep rates with increas-ing deviatoric stress (>50% strength), and the bulk viscosity is notreported. The bulk viscosity values recalculated from the creep datareviewed here provide results that are in agreement with our observationsof time-dependent rock deformation at isotropic loading and elevatedpore pressures.

Additionally, our experiments show the bulk viscosity decrease with theincrease of pore fluid pressure under constant total pressure, thus withdecreasing effective pressure. This contradicts a commonly held belief thatthe effective viscosity of nonlinear-creeping materials must drop with therise of the deviatoric stresses and effective pressure (e.g., Atkinson &Davies, 2000; Karato, 2010). Interestingly, the long-term observation(4.5 × 106 s or 52 days) of pore pressure buildup in undrained test onclay-rich material shows that pore pressure increase slows down with

Figure 7. Bulk viscosity as a function of pore fluid pressure to the total isotro-pic stress ratio p f =p for Berea sandstone, Apulian limestone (calcarenite),clay-rich material, and Opalinus clay (shale) measured by (a) the directmethod and (b) the indirect method. Colors designate different levels ofapplied total stress. The bulk viscosity of dry serpentine is calculated fromHilairet et al. (2007).

Figure 6. Change (�) of porosity with time in the undrained experimentswith a constant mean stress on Berea sandstone at pe = 5.3 MPa (blue), cal-carenite at pe = 6.9 MPa (green), clay-rich material at pe = 0.5 MPa (maroon),and Opalinus clay at pe = 2.5 MPa (orange).

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 14

Page 15: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

time (Figure 8a). The simplest two-parameter linear dependence of bulkviscosity on fluid pressure:

ηϕ ¼ ηo δp f

p� 1

� �(16)

is calibrated by finding the best fit to the experimental data (Figure 8b).The best fit reveals the rise of bulk viscosity ηϕ in time (4–8 × 1013 Pa s).Thus, the decrease of bulk viscosity with rising fluid pressure under con-stant total pressure is reverted at fluid pressure values approaching thetotal pressure. This reverse is expected as the typical nonlinear powerlaw creep models predict singular rise of effective viscosity to infinity atdifferential stresses approaching zero (Karato, 2010; Renshaw &Schulson, 2017).

Longer-than-laboratory-term integration in time of our experimentalobservations would lead to a growth of the fluid pressure under constanttotal load (e.g., the weight of the overlying rocks), toward total pressure,even if the stress state is isotropic. Stabilization of the fluid pressure riseat approaching the total pressure is predicted. If this process is triggeredin natural reservoirs by a fluid pressure perturbation, it may lead to lique-faction or rock embrittlement causing slip instabilities and earthquakes(National Research Council, 2012; Scholz, 2002).

Furthermore, the poroelastic bulk moduli and viscosity are important para-meters that define the character of fluid flow through porous media.Viscous compaction of fluid-filled porous media allows a special type offluid flow instability to be generated that leads to the formation of high-porosity, high-permeability domains able to self-propagate upwardbecause of the interplay between buoyancy and viscous resistance ofthe deforming porous matrix. This instability is known as a porosity wave(Connolly & Podladchikov, 2015; Yarushina et al., 2015), and its formationis possible under conditions applicable to deep CO2 storage in reservoirs.It also explains the creation of high-porosity channels and chimneys(Räss et al., 2018), whose number and spatial distribution are controlledby the permeability, fluid viscosity, and bulk viscosity of the rock(McKenzie, 1987). Connolly and Podladchikov (2015) and Yarushina et al.(2015) have established the relationship between episodic behavior andtemporal focusing of fluid flow. Furthermore, Omlin et al. (2017) demon-strates that porosity waves may be generated not only by deformation

in the ductile domain but also by kinetic control on the fluid-producing reactions. However, experimental ver-ification of the expected behavior and resulting calibration of the relevant parameters has yet to beaddressed. Another issue is a scaling problem: porosity waves are expected to occur and evolve at scales(meters to kilometers) that do not lend themselves easily to high-pressure and elevated temperature labora-tory experimentation (scales of millimeters to centimeters) under conditions at which the rock and its consti-tuents deform in a viscous or plastic manner. The latter is a prerequisite for modeling (Chakraborty, 2017). Ourpreliminary results (Räss et al., 2017) and the data presented here show that high-permeabile pathways aremost likely to form in low-permeable clay-rich materials and that these pathways can be observed at labora-tory timescales (weeks to months). Accordingly, proper characterization of sedimentary rock behavior is cri-tical for understanding a long-term effect of hydrocarbon extraction or waste storage on the subsurface.

6. Conclusions

The poroelastic and time-dependent (creep) response of fluid-filled sedimentary rock, namely, quartz-richBerea sandstone, calcite-rich Apulian limestone, clay-rich material, and Opalinus clay (shale), is

Figure 8. (a) Undrained loading test and consequent long-term pore pres-sure buildup for clay-rich material at initial p = 1.30 MPa and pf = 1.21 MPa.Model prediction for pore pressure buildup based on equation (16) is pre-sented. (b) Best fitting model for bulk viscosity based on the average overtime absolute fluid pressure difference between the model and the experi-ment (in Pa). The best fitting parameters are indicated by the white diamondin the inclusion.

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 15

Page 16: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

experimentally quantified at hydrostatic loading and low temperatures. A poroviscous model parameter, thebulk viscosity, is included in the constitutive equations by analogy with Biot’s (1941) poroelastic relationshipsand is estimated from the time-dependent deformation of rock in the drained regime and by observation ofthe pore pressure growth in the undrained regime under constant isotropic stress conditions. Measurementsof the bulk viscosity through pore pressure growth appear to be more precise and significantly less time con-suming, with timescales on the order of 104 s (hours). Rapid fluid pressure growth corresponds to low bulkviscosity, which is on the order of 1015 to 1016 Pa s for sandstone, limestone, and shale, and is ~1013 to1014 Pa s for clay-rich material. This result is somewhat in agreement with previously reported creep testson saturated geomaterials at low deviatoric stresses. Moreover, the viscosity is observed to decrease withan increase in pore fluid pressure under constant total pressure, thus with a decrease in effective pressure.This result contradicts a commonly held belief that the effective viscosity of nonlinearly creeping materialsmust drop with an increase in effective pressure. Pore fluid pressure growth toward the total stress can resultin liquefaction or rock embrittlement, causing slip instabilities and earthquakes. In addition, viscous compac-tion of fluid-filled porous media allows the generation of high-permeable domains, or porosity waves, andtheir formation is most probable in low-permeable clay-rich materials.

ReferencesAbousleiman, Y., Cheng, A. H.-D., Jiang, C., & Roegiers, J.-C. (1993). A micromechanically consistent poroviscoelasticity theory for rock

mechanics applications. International Journal of Rock Mechanics and Mining Science and Geomechanics Abstracts, 30(7), 1177–1180. https://doi.org/10.1016/0148-9062(93)90090-Z

Aharonov, E., & Scholz, C. H. (2018). A physics-based rock friction constitutive law: Steady state friction. Journal of Geophysical Research: SolidEarth, 123, 1591–1614. https://doi.org/10.1002/2016JB013829

Armand, G., Conil, N., Talandier, J., & Seyedi, D. M. (2016). Fundamental aspects of the hydromechanical behaviour of Callovo-Oxfordianclaystone: From experimental studies to model calibration and validation. Computers and Geotechnics, 85, 277–286.

Atkinson, B. K., & Meredith, P. G. (1987). The theory of subcritical crack growth with applications to minerals and rocks. In B. K. Atkinson (Ed.),Fracture mechanics of rock (pp. 111–116). London: Academic Press. https://doi.org/10.1016/B978-0-12-066266-1.50009-0

Atkinson, H. V., & Davies, S. (2000). Fundamental aspects of hot isostatic pressing: An overview. Metallurgical and Materials Transactions A,31A, 2981–3000.

Belmokhtar, M., Delage, P., Ghabezloo, S., & Conil, N. (2017). Thermal volume changes and creep in Callovo-Oxfordian claystone. RockMechanics and Rock Engineering, 50(9), 2297–2309. https://doi.org/10.1007/s00603-017-1238-7

Bernabe, Y., Fryer, D. T., & Shively, R. M. (1994). Experimental observations of the elastic and inelastic behaviour of porous sandstones.Geophysical Journal International, 117(2), 403–418. https://doi.org/10.1111/j.1365-246X.1994.tb03940.x

Berryman, J. G. (1992). Exact effective-stress rules in rock mechanics. Physical Review A, 46(6), 3307–3311. https://doi.org/10.1103/PhysRevA.46.3307

Berryman, J. G. (2005). Comparison of upscaling methods in poroelasticity and its generalizations. Journal of Engineering Mechanics, 131(9),928–936. https://doi.org/10.1061/(ASCE)0733-9399(2005)131:9(928)

Biot, M. A. (1941). General theory of three-dimensional consolidation. Journal of Applied Physics, 12(2), 155–164. https://doi.org/10.1063/1.1712886

Biot, M. A., & Willis, D. G. (1957). The elastic coefficients of the theory of consolidation. Journal of Applied Mechanics: Transactions of the ASME,79, 594–601.

Bishop, A. W. (1976). Influence of system compressibility on observed pore pressure response to an undrained change in stress in saturatedrock. Géotechnique, 26(2), 371–375. https://doi.org/10.1680/geot.1976.26.2.371

Bossart, P., & Thury, M. (2008). Mont Terri Rock Laboratory—Project, programme 1996 to 2007 and results (Vol. 3). Wabern: SwissGeological Survey.

Brantut, N., Heap, M. J., Meredith, P. G., & Baud, P. (2013). Time-dependent cracking and brittle creep in crustal rocks: A review. Journal ofStructural Geology, 52, 17–43. https://doi.org/10.1016/j.jsg.2013.03.007

Bürgmann, R., & Dresen, G. (2008). Rheology of the lower crust and upper mantle: Evidence from rock mechanics, geodesy, andfield observations. Annual Review of Earth and Planetary Sciences, 36(1), 531–567. https://doi.org/10.1146/annurev.earth.36.031207.124326

Chakraborty, S. (2017). A new mechanism for upper crustal fluid flow driven by solitary porosity waves in rigid reactive media? GeophysicalResearch Letters, 44, 10,324–10,327. https://doi.org/10.1002/2017GL075798

Chang, C., Moos, D., & Zoback, M. D. (1997). Anelasticity and dispersion in dry unconsolidated sands. International Journal of Rock Mechanicsand Mining Sciences, 34(3–4), 48.e1–48.e12.

Connolly, J. A. D., & Podladchikov, Y. Y. (2015). An analytical solution for solitary porosity waves: Dynamic permeability and fluidization ofnonlinear viscous and viscoplastic rock. Geofluids, 15(1-2), 269–292. https://doi.org/10.1111/gfl.12110

Costin, L. S. (1987). Time-dependent deformation and failure. In B. K. Atkinson (Ed.), Fracture mechanics of rock (pp. 167–216). London, UK:Academic Press. https://doi.org/10.1016/B978-0-12-066266-1.50010-7

Coussy, O. (2004). Poromechanics. Chichester, UK: John Wiley.Cox, S. F., & Paterson, M. S. (1991). Experimental dissolution-precipitation creep in quartz aggregates at high temperatures. Geophysical

Research Letters, 18(8), 1401–1404. https://doi.org/10.1029/91GL01802de Waal, J. A., Muntendam-Bos, A. G., & Roest, J. P. A. (2015). Production induced subsidence and seismicity in the Groningen gas field—

Can it be managed? Proceedings of the International Association of Hydrological Sciences, 372, 129–139. https://doi.org/10.5194/piahs-372-129-2015

Detournay, E., & Cheng, A. (1993). Fundamentals of poroelasticity. In C. Fairhurst (Ed.), Comprehensive rock engineering: Principles, practice andprojects: Vol. II. Analysis and design methods (pp. 113–171). Oxford, UK: Pergamon Press.

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 16

AcknowledgmentsWe thank Joe Labuz for permission topublish the results of plane straincompression tests. ChristopheNussbaum and Swisstopo areacknowledged for providing Opalinusclay cores. Research by R.Y. Makhnenkowas supported as part of the Center forGeologic Storage of CO2, an EnergyFrontier Research Center funded by theU.S. Department of Energy (DOE), Officeof Science, Basic Energy Sciences (BES),under award DE-SC0C12504. SusanKrusemark edited the manuscript. Datadisplayed or used to generate figuresand plots are available as supportinginformation.

Page 17: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

Elsworth, D., & Yasuhara, H. (2006). Short-timescale chemo-mechanical effects and their influence on transport properties of fractured rock.Pure and Applied Geophysics, 163(10), 2051–2070. https://doi.org/10.1007/s00024-006-0113-3

Fabre, G., & Pellet, F. (2006). Creep and time-dependent damage in argillaceous rocks. International Journal of Rock Mechanics and MiningSciences, 43(6), 950–960. https://doi.org/10.1016/j.ijrmms.2006.02.004

Frenkel, Y. (1944). On the theory of seismic and seismoelectric phenomena in moist soil. Journal of Physics, 8, 230–241.Gasc-Barbier, M., Chanchole, S., & Bérest, P. (2004). Creep behavior of bure clayey rock. Applied Clay Science, 26(1-4), 449–458. https://doi.org/

10.1016/j.clay.2003.12.030Goodier, J. N. (1936). Slow viscous flow and elastic deformation. Philosophical Magazine, 22(149), 678–681.Gurevich, B. (2007). Comparison of the low-frequency predictions of Biot’s and de Boer’s poroelasticity theories with Gassmann’s equation.

Applied Physics Letters, 91(9), 091919. https://doi.org/10.1063/1.2778763Hilairet, N., Reynard, B., Wang, Y., Daniel, I., Merkel, S., Nishiyama, N., & Petitgirard, S. (2007). High-pressure creep of serpentine, interseismic

deformation, and initiation of subduction. Science, 318(5858), 1910–1913. https://doi.org/10.1126/science.1148494Kaboli, S., Burnley, P. C., Xia, G., & Green, H. W., II (2017). Pressure dependence of creep in forsterite olivine: Comparison of measurements

from the D-DIA and Griggs apparatus. Geophysical Research Letters, 44, 10,939–10,947. https://doi.org/10.1002/2017GL075177Karato, S. (2010). Rheology of the Earth’s mantle: A historical review. Gondwana Research, 18(1), 17–45. https://doi.org/10.1016/j.

gr.2010.03.004Labuz, J. F., Dai, S.-T., & Papamichos, E. (1996). Plane-strain compression of rock-like materials. International Journal of Rock Mechanics and

Mining Sciences, 33(6), 573–584. https://doi.org/10.1016/0148-9062(96)00012-5Le Guen, Y., Renard, F., Hellmann, R., Brosse, E., Collombet, M., Tisserand, D., & Gratier, J. P. (2007). Enhanced deformation of limestone and

sandstone in the presence of high fluids. Journal of Geophysical Research, 112, B05421. https://doi.org/10.1029/2006JB004637Liteanu, E., Niemeijer, A., Spiers, C. J., Peach, C. J., & De Bresser, J. H. P. (2012). The effect of CO2 on creep of wet calcite aggregates. Journal of

Geophysical Research: Solid Earth, 117, B03211. https://doi.org/10.1029/2011JB008789Lopatnikov, S. L., & Cheng, A. H.-D. (2002). Variational formulation of fluid infiltrated porous material in thermal and mechanical equilibrium.

Mechanics of Materials, 34(11), 685–704. https://doi.org/10.1016/S0167-6636(02)00168-0Lopatnikov, S. L., & Cheng, A. H.-D. (2004). Macroscopic Lagrangian formulation of poroelasticity with porosity dynamics. Journal of the

Mechanics and Physics of Solids, 52(12), 2801–2839. https://doi.org/10.1016/j.jmps.2004.05.005Lowe, J., & Johnson, T. C. (1960). Use of back pressure to increase degree of saturation of triaxial test specimen. In Proc. ASCE Research Conf. on

Shear Strength of Cohesive Soils, Boulder, CO, pp. June, 1960, 819–836.Makhnenko, R., & Labuz, J. (2014). Plane strain testing with passive restraint. Rock Mechanics and Rock Engineering, 47(6), 2021–2029. https://

doi.org/10.1007/s00603-013-0508-2Makhnenko, R., & Labuz, J. (2016). Elastic and inelastic deformation of fluid-saturated rock. Philosophical Transactions of the Royal Society A:

Mathematical, Physical and Engineering Sciences, 374(2078), 20150422. https://doi.org/10.1098/rsta.2015.0422Makhnenko, R. Y., Harvieux, J., & Labuz, J. F. (2015). Paul-Mohr-Coulomb failure surface of rock in the brittle regime. Geophysical Research

Letters, 42(17), 6975–6981. https://doi.org/10.1002/2015GL065457Makhnenko, R. Y., & Labuz, J. (2014). Calcarenite as a possible host rock for geologic CO2. In Proceedings of the 48th U.S. Rock

Mechanics/Geomechanics Symposium, Minneapolis, MN 1–4 June 2014, paper No. 7559.Makhnenko, R. Y., Tarokh, A., & Podladchikov, Y. (2017). On the unjacketed moduli of sedimentary rock. In M. Vandamme, P. Dangla,

J.-M. Pereira, & S. Ghabezloo (Eds.), Poromechanics VI—Proceedings of the 6th Biot Conference on Poromechanics, (pp. 897–904). AmericanSociety of Civil Engineers: Reston, VA.

Makhnenko, R. Y., Vilarrasa, V., Mylnikov, D., & Laloui, L. (2017). Hydromechanical aspects of CO2 breakthrough into clay-rich caprock. EnergyProcedia, 114, 3219–3228. https://doi.org/10.1016/j.egypro.2017.03.1453

McKenzie, D. (1984). The generation and compaction of partially molten rock. Journal of Petrology, 25(3), 713–765. https://doi.org/10.1093/petrology/25.3.713

McKenzie, D. (1987). The compaction of igneous and sedimentary rocks. Journal of the Geological Society (London), 144(2), 299–307. https://doi.org/10.1144/gsjgs.144.2.0299

Mesri, G., & Castro, A. (1987). Cα/Cc concept and K0 during secondary compression. Journal of Geotechnical Engineering, 113(3), 230–247.https://doi.org/10.1061/(ASCE)0733-9410(1987)113:3(230)

Meyer, J. P., & Labuz, J. F. (2013). Linear failure criteria with three principal stresses. International Journal of Rock Mechanics and MiningSciences, 60, 180–187. https://doi.org/10.1016/j.ijrmms.2012.12.040

National Research Council (2012). Induced seismicity potential in energy technologies. Washington, DC: National Academies Press.Omlin, S., Malvoisin, B., & Podladchikov, Y. Y. (2017). Pore fluid extraction by reactive solitary waves in 3-D. Geophysical Research Letters, 44,

9267–9275. https://doi.org/10.1002/2017GL074293Paterson, M. S., & Wong, T. F. (2005). Experimental rock deformation—The brittle field (2nd ed.). Berlin Heidelberg: Springer-Verlag.Pearson, F. J. (2002). PC experiment: Recipe for artificial pore water (Tech. Note 2002–17). St-Ursanne, Switzerland: Mont Terri Project.Proietti, A., Bystricky, M., Guignard, J., Bejina, F., & Crichton, W. (2016). Effect of pressure on the strength of olivine at room temperature.

Physics of the Earth and Planetary Interiors, 259, 34–44. https://doi.org/10.1016/j.pepi.2016.08.004Räss, L., Makhnenko, R. Y., Podladchikov, Y., & Laloui, L. (2017). Quantification of viscous creep influence on storage capacity of caprock.

Energy Procedia, 114, 3237–3246. https://doi.org/10.1016/j.egypro.2017.03.1455Räss, L., Simon, N. S. C., & Podladchikov, Y. Y. (2018). Spontaneous formation of fluid escape pipes from subsurface reservoirs. Nature Scientific

Reports, 8(1), 11116. https://doi.org/10.1038/s41598-018-29485-5Renshaw, C. E., & Schulson, E. M. (2017). Strength-limiting mechanisms in high-confinement brittle-like failure: Adiabatic transformational

faulting. Journal of Geophysical Research: Solid Earth, 122, 1088–1106. https://doi.org/10.1002/2016JB013407Rice, J. R. (1975). On the stability of dilatant hardening for saturated rock masses. Journal of Geophysical Research, 80(11), 1531–1536. https://

doi.org/10.1029/JB080i011p01531Rice, J. R., & Cleary, M. P. (1976). Some basic stress diffusion solutions for fluid-saturated elastic porous media with compressible constituents.

Reviews of Geophysics and Space Physics, 14(2), 227–241. https://doi.org/10.1029/RG014i002p00227Rutqvist, J. (2012). The geomechanics of CO2 storage in deep sedimentary formations. International Journal of Geotechnical and Geological

Engineering, 30(3), 525–551. https://doi.org/10.1007/s10706-011-9491-0Scholz, C. H. (1968). Mechanism of creep in brittle rock. Journal of Geophysical Research, 73(10), 3295–3302. https://doi.org/10.1029/

JB073i010p03295Scholz, C. H. (2002). The mechanics of earthquakes and faulting. Cambridge, UK: Cambridge University Press. https://doi.org/10.1017/

CBO9780511818516

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 17

Page 18: Journal of Geophysical Research: Solid Earth · 1. Introduction Under typical upper crustal conditions, rock is generally saturated with an aqueous pore fluid. When defor-mation

Skempton, A. W. (1954). The pore-pressure coefficients A and B. Geotechnique, 4(4), 143–147. https://doi.org/10.1680/geot.1954.4.4.143Sone, H., & Zoback, M. D. (2014). Time-dependent deformation of shale gas reservoir rocks and its long-term effect on in-situ state of stress.

International Journal of Rock Mechanics and Mining Sciences, 69, 120–132. https://doi.org/10.1016/j.ijrmms.2014.04.002Taylor, D. W., & Merchant, W. (1940). A theory of clay consolidation accounting for secondary compression. Studies in Applied Mathematics,

19(1–4), 167–185.Terzaghi, K. (1943). Theoretical soil mechanics. New York, NY: John Wiley. https://doi.org/10.1002/9780470172766Tsang, C.-F., Bernier, F., & Davies, C. (2005). Geohydromechanical processes in the Excavation Damaged Zone in crystalline rock, rock salt, and

indurated and plastic clays in the context of radioactive waste disposal. International Journal of Rock Mechanics and Mining Sciences, 42(1),109–125. https://doi.org/10.1016/j.ijrmms.2004.08.003

Turcotte, D. I., & Schubert, G. (2014). Geodynamics. Cambridge, UK: Cambridge University Press. https://doi.org/10.1017/CBO9780511843877Vanel, L., Ciliberto, S., Cortet, P.-P., & Santucci, S. (2009). Time-dependent rupture and slow crack growth: Elastic and viscoplastic dynamics.

Journal of Physics D: Applied Physics, 42(21), 214007. https://doi.org/10.1088/0022-3727/42/21/214007Vu, M.-H., Sulem, J., Ghabezloo, S., Laudet, J.-B., Garnier, A., & Guédon, S. (2012). Time-dependent behaviour of hardened cement paste under

isotropic loading. Cement and Concrete Research, 42(6), 789–797. https://doi.org/10.1016/j.cemconres.2012.03.002Wawersik, W. R., & Brown, W. S. (1973). Creep fracture in rock. Utah University Salt Lake City Department of Mechanical Engineering, Final report

UTEC-ME-73-197, 28 December – 28 July, Defense Technical Information Center, Vancouver, Canada.Yang, D. S., Bornert, M., & Chanchole, S. (2011). Experimental investigation of the delayed behaviour of unsaturated argillaceous rocks by

means of Digital Image Correlation techniques. Applied Clay Science, 54(1), 53–62. https://doi.org/10.1016/j.clay.2011.07.012Yarushina, V. M., & Podladchikov, Y. Y. (2015). (De) compaction of porous viscoelastoplastic media: Model formulation. Journal of Geophysical

Research: Solid Earth, 120, 4146–4170. https://doi.org/10.1002/2014JB011258Yarushina, V. M., Podladchikov, Y. Y., & Connolly, J. A. D. (2015). (De) compaction of porous viscoelastoplastic media: Solitary porosity waves.

Journal of Geophysical Research: Solid Earth, 120, 4843–4862. https://doi.org/10.1002/2014JB011260Zhang, C.-L., Rothfuchs, T., Su, K., & Hoteit, N. (2007). Experimental study of the thermo-hydro-mechanical behaviour of indurated clays.

Physics and Chemistry of the Earth, 32(8-14), 957–965. https://doi.org/10.1016/j.pce.2006.04.038

10.1029/2018JB015685Journal of Geophysical Research: Solid Earth

MAKHNENKO AND PODLADCHIKOV 18