Tag Archives: brittle failure

Analogue structure models: Scaling the materials

Facebooktwitterlinkedininstagram
Scaled sand-box experiments are an ideal medium to observe rock deformation that, in this example, involves synkinematic deposition during rift-like crustal extension. The choice of model materials, in addition to imposed boundary conditions such as strain rates, will determine the outcome of the experiment. Dry sand was chosen for this model because its brittle behaviour under the model conditions is a good representation of natural rock failure. Diagram modified slightly from Eisenstadt and Sims, 2005, Figure 3a.

Scaled sand-box experiments are an ideal medium to observe rock deformation that, in this example, involves synkinematic deposition during rift-like crustal extension. The choice of model materials, in addition to imposed boundary conditions such as strain rates, will determine the outcome of the experiment. Dry sand was chosen for this model because its brittle behaviour under the model conditions is a good representation of natural rock failure. Diagram modified slightly from Eisenstadt and Sims, 2005, Figure 3a.

This post deals with the materials, their rheological behaviours, and scaling as they apply to analogue structure models.

Rock deformation

Two of the most common structures generated by rock deformation are faults (and fractures) and folds. The rheological conditions for each are fundamentally different: faults and fractures result from brittle behaviour, whereas folds require materials to act more like plastics with ductile behaviour. Both styles represent deformation conditions beyond the elastic limits of the material involved. For example, hard granite at near surface conditions will exhibit brittle behaviour at high strain rates (confining pressures will be low), but at depth may act in a ductile manner at lower strain rates and high confining pressures (and elevated temperatures). How rocks and sediments behave under stress depends on:

  • Lithology and grain/crystal composition and size. The compressive strength of limestone is commonly less than that for quartz arenites because of the propensity for calcite-dolomite crystal cleavage.
  • Stress magnitude and orientation. For example, is the principal stress due to lithostatic load, or are there deviatoric and differential stress components?
  • Confining pressures – compressive rock strength increases with depth and confining pressure.
  • Temperature.
  • Viscosity, that depends on lithology, temperature, and strain rate.
  • Preexisting anisotropies (fabrics or structures), such as older fracture networks, crystal or grain alignment, metamorphic foliation, and sedimentary stratification.

Rock strength is one measure that accounts for some of these variables; it is a measure of the maximum uniaxial stress at the point of material failure. Failure in this context is usually expressed as a function of the internal cohesion, or cohesive strength, and the internal angle of friction (φ)that for many Earth materials is about 30o to 35o. The SI unit of rock strength is the Pascal (Pa), or megapascal (MPa – dimensions ML-1T-2). Rock strength is defined for tensile and compressional conditions, but for most geological systems where deformation occurs at depth within the crust, compressive strength exceeds tensile strength. Furthermore, the dominant mode of failure is by shear.

Fracture sets in Old Red Sandstone, Portskerra (N Scotland) – the outcrop face is almost vertical. Red dashed lines follow conjugate fracture sets at a reasonably consistent 60o separation. The dashed yellow lines indicate bedding plane parting that is approximately parallel to σ3 – bedding dips 50o. The point of rock failure in generating these fractures corresponds to the rock peak friction, with a friction angle of 30o (see discussion notes below).

Fracture sets in Old Red Sandstone, Portskerra (north Scotland) – the outcrop face is almost vertical. Red dashed lines follow conjugate fracture sets at a reasonably consistent 60o separation. The dashed yellow lines indicate bedding plane parting that is approximately parallel to σ3 – bedding dips 50o. The point of rock failure in generating these fractures corresponds to the rock peak friction, with a friction angle of 30o (see discussion notes below).

For almost two centuries, analogue modeling of fault (and fold) systems has focused on rifts and inverted rifts, fold – thrust belts and orogenic wedges at convergent plate margins, and strike-slip fault arrays at transcurrent or oblique transcurrent margins. Here are some excellent reviews of analogue models for geodynamic systems (Shellart and Strak, 2016),  rifting and salt mobilization (Zwaan and Schreurs, 2022), and orogenic wedges (Graveleau et al., 2012).

 

What are these structure models attempting to do?  

Fundamentally, the models allow us to glimpse processes that under normal geological conditions of time and space we cannot observe. If we scale the model variables correctly, there is a good chance we will answer some of the questions about how these systems operate. For example, the master extensional faults in rift systems are usually accompanied by networks of synthetic and antithetic fault arrays that form at different stages of rifting and sedimentation in fault-bound basins. Analogue models can help us visualize the overall distribution of strain by teasing out the relative timing, orientation, and magnitude of the fault arrays and the displacement of syn-rift stratigraphy. To do this, the model materials need to be appropriately scaled to natural examples.

 

Material rheology

Structure model construction requires materials that are scaled to represent brittle and ductile rheological behaviour, plus behaviours in between these two end-member types. Analogue models built to examine fault initiation and evolution are usually layered – layering can reflect broad stratigraphic ordering of a sedimentary basin, or at larger scales, the layering of lithosphere crust and mantle. The order in which layers of different rheology and thickness are laid down in a model will influence the style of deformation; ductile layers that are prone to folding or flow may affect the distribution of strain in underlying and overlying brittle layers.

The chosen materials must also suit the modeled deformation mechanism. Models that represent diapirism of salt or igneous melts will require viscous materials that flow and have buoyancy contrasts with surrounding rock-material. The reverse buoyancy problem exists for lithosphere-scale models that represent subduction. The energy required to drive these experiments comes from within the models and is manifested as positive or negative buoyancy.

Materials having fundamentally different rheologies are required to model rifted upper crust and synrift sedimentation, where compressive or tensile stresses dominate, and where buoyancy and inertial forces are negligible (although this generalization is complicated where synkinematic salt is deposited). In these experiments, the energy to drive compression-extension originates outside the model, for example, from a worm-screw driven plate at one end of the model container.

 

Common model materials

One of James Hall’s first experiments on folding of stratified rock used cloth layers (Hall, 1815). One can imagine Hall at his dinner table, watching the tablecloth deform as plates were moved from diner to diner, and intuiting the analogy with folded rock he had observed in the field. His later experiments used layers of granular and viscous materials, as did many subsequent experimentalists (Caddell, 1889 (see also Butler et al., 2020 ; Willis, 1894;   Hubbert, 1934).

Except for the silicon polymers, the materials used by recent experimentalists have changed little from these early pioneering days – the difference is that we now have a good understanding of material properties in relation to their rheological behaviour.  Some commonly used materials and their properties include:

  • Well sorted, dry, quartz and feldspar sands (brittle behaviour, low cohesion). Bulk densities range from 1.56 g.cm-3 (quartz sand) to 1.3 g.cm-3 (feldspar) (bulk density is less than the mineral density). Mean peak friction coefficients are 0.5 – 0.6 and friction angles are 31o-36o. Some studies have assumed that dry sand is cohesionless, but repeated lab analyses show that mean peak cohesion ranges from 10-140 Pa.
  • Corundum/magnetite sands, commonly used as marker layers (brittle behaviour). Bulk density about 3.3 g.cm-3.
  • Wet clay – also used for brittle failure and has cohesive strength slightly higher than dry sand (Eisenstadt and Sims, 2005). Density depends on water content – commonly ~1.6 – 1.8 g.cm-3. Peak cohesion up to 100 Pa.
  • Kinetic sand (adding a touch of viscosity with 1% -2% silicon polymers). Either quartz or corundum mixtures. Viscosity varies with density (% polymer) and ranges from about 4 x 104 to 105 Pa.s
  • Gelatins (visco-elastic to brittle rheology in the gel state depending on strain rates and confining pressures) (Kavanagh et al., 2013)
  • Silicone putty (a viscous polymer representing ductile behaviour). Viscosity 104 to 105s.
  • PDMS (polydimethylsiloxane) – a more fluid silicone polymer. Density 0.98 cm-3 and a viscosity around 1.6 x 104 Pa.s.
  • Syrup (sugar, honey – low viscosity ductile flow). Viscosities <1 to 20 Pa.s.

Material properties

Faulting in the brittle upper crust generally obeys the Coulomb criterion for failure and frictional sliding, and is written as:

τ = C0  + Tanφ σN    where τ is the shear stress at the point of failure, C0 is the rock cohesive strength, φ is the internal friction angle, and σN  is the normal stress.

The same empirical relationship should also hold for models and model materials (to be consistent with the rules of similarity).  The Coulomb function emphasizes the variables we need to scale such that the ratios between model variables and natural variables are similar – for example, C0 model/ C0 real Lengthmodel/ Lengthnatural.

 

Model variables that map brittle behaviour

Tan φ: This is the friction factor, or friction coefficient – it is dimensionless, usually written as

                                   τ = C0  + γσN  where, for rock materials, γ is the coefficient of internal friction and is the ratio of the frictional force to normal force – it depends primarily on surface roughness. Thus, γ is zero for a frictionless, smooth, or lubricated surface. For most rock materials γ is 0.6 – 0.7. The presence of clays, or mica-lined schistose foliation will tend to lower the values of γ.

C0 – Cohesion (Cohesive strength): (Dimensions ML-1T-2). C0 is a function of material density, the acceleration due to gravity, and a length dimension. Bulk densities are well within an order of magnitude of real rock densities, the gravitation constant is the same for model and the real world, and lengths commonly scale from 10-4 to 10-6 (at 10-6 one cm scales to 10 km). If the Coulomb criterion is to apply to the models, then the cohesive strength must also scale to 10-4 to 10-6.

φ – friction angle: (dimensionless). A good analogy for representation of the internal friction angle (φ) is the angle of repose for dry, well-sorted sand; the actual value is about 34o. If the slope increases it becomes gravitationally unstable and grains will slide or tumble downslope until the repose angle is re-established – in other words, the granular material shears. This basically is the model applied to failure of harder rock where shear is caused by differential normal stresses.

Friction angles for different rock types and rock strengths, from Wyllie and Norrish, 1996. Table 14-1.

Friction angles for different rock types and rock strengths, from Wyllie and Norrish, 1996. Table 14-1.

Viscosity – mapping ductile behaviour

 This is one of the fundamental measures for materials where ductile behaviour is important. Viscosity is dependent on temperature and strain rate, and testing is done at specified values of these variables. Silicone putty is commonly used in fault models to represent shale or salt that are inter layered with brittle lithologies. It is also used as a basal layer of models that represent salt flow or more ductile behaviour deeper in the crust and mantle lithosphere.

Kinetic sand is a mixture of either quartz or corundum/magnetite sand and about 2% silicone polymer (polydimethylsiloxane, or PDMS). The polymer adds a degree of cohesion and viscosity to the mix, that represents rheological behaviour somewhere between brittle and ductile. PDMS has a density of 0.98 g/cm3, its viscosity averages 3×104  Pa.s at 23 °C (Konstantinovskaya et al., 2007). PDMS can also be used on its own to represent salt layers (e.g, Ferrer et al., 2023) or ductile mantle lithosphere.  An example of the PDMS viscosity- shear rate relationship is illustrated below.

Lab test plots of viscosity – shear rate for corundum sand – PDMS mixtures at different densities – pure PDMS density is 0.98 g/cm3 (top curve). From Zwaan et al., 2018, Figure 1.

Lab test plots of viscosity – shear rate for corundum sand – PDMS mixtures at different densities – pure PDMS density is 0.98 g/cm3 (top curve). From Zwaan et al., 2018, Figure 1.

Silicone putty (a kind of Silly Putty) is also a viscoelastic silicon polymer that has unique properties in that it reacts elastically if the strain rate is high (e.g., bouncing it off the floor), and as a viscous fluid at much lower strain rates. The latter property makes it ideal for models of deformation involving ductile behaviour. Some of the strain in the deformation models is taken up by folding of the ductile layers. It has a viscosity of about 4×105  Pa.s at 23 °C depending on composition (a bit higher than PDMS). At the experimental strain rates of most models, silicone putty behaves as a Newtonian fluid (i.e., it has no inherent strength and begins to deform immediately stress is applied). Newtonian properties are important in analogue modeling because the behaviour of the materials doesn’t change during the experiment.

Sugar syrups are also used in lithosphere-scale models to represents the asthenosphere-lithosphere boundary. Syrup is also used in experiments where isostatic compensation is modeled (Schmid et al., 2022). Average measured viscosity, depending on composition, ranges from  about 7 Pa to 64 Pa.s

 

Material scaling

The average values for C0, m, and γ for granular materials used in the models have been determined from repeated lab determinations. Both the friction coefficient and cohesion have measured values that record peak stress conditions at the point of failure – these are recorded as peak cohesion (peak cohesive strength) and peak friction respectively. Dynamic cohesion and dynamic friction represent the plateau on the shear stress curves, that reflect the forces opposing continued motion along a fracture plane; their values are generally lower than those for peak conditions. The example below of shear stress-strain illustrates this relationship.

The average lab determined peak friction coefficients (dimensionless) for quartz sands are about 0.6 – 0.7, which is consistent with measured rock values.

Internal friction angles for the sands are commonly 27o-35o, also consistent with natural rock values.

Lab test results for determining cohesive strength and friction coefficient for quartz sand, at different normal stress values (Pa). The peak and dynamic domains for each quantity are well defined. Note that shear displacement has the same dimensions as strain (L). Modified from Zwaan et al., 2018, Figure 2.

Lab test results for determining cohesive strength and friction coefficient for quartz sand, at different normal stress values (Pa). The peak and dynamic domains for each quantity are well defined. Note that shear displacement has the same dimensions as strain (L). Modified from Zwaan et al., 2018, Figure 2.

Cohesive strengths for most granular materials used in these models range from about 50 Pa to 200 Pa, depending on mean grain size and standard deviation (sorting), and to some extent on the variability of grain shape in natural sands – the use of microspheres can reduce this type of uncertainty. In contrast, the cohesive strength for Earth materials ranges widely from less than 0.1 MPa for soft soils, to >200 MPa for rocks like isotropic granite, marble, and basalt. The range for sandstones is about 50-100 MPa, and < 25 MPa for chalk and salt. If we assume an average value of 60 MPa for sandstones, and 60 Pa for sands used in the models, then the scale ratio is 10-6, which is consistent with the upper end of our length scale.

Viscosity scaling is less obvious, because the scaling factor between ductile materials like silicone putty or PDMS (viscosities in the range 104 to 105 Pa.s) and the average crustal rocks (1018 – 1020 Pa.s) is 10-14 to 10-16. This is hugely different to the 10-5 – 10-6 commonly used for the other scaled variables.

[Scaling factor is the ratio between model variable and natural variable, and hence is dimensionless). It is usually indicated by an asterisk. Dimensionless quantities established for natural systems can be applied directly to analogue models.]

We can solve this apparent dilemma by going back to basic Newtonian mechanics. Newton was the first to demonstrate the empirical relationship between shear stress and the rate of shear deformation, or strain rate, in a fluid. This is expressed as:

                                τ = μ (v/d) where

τ is shear stress, μ is dynamic viscosity, v is velocity and d a length such that v/d is the velocity gradient for shear, also called the strain rate (dimensions of T-1). If the relationship between stress and strain rate is linear, then the fluid is Newtonian. Under the experimental conditions for most analogue models, materials like silicone polymers behave as Newtonian fluids.

We can rearrange this function for viscosity (μ = τ/ (v/d)) and write dimensionless quantities for its variables using scaling factors – a model/natural stress scaling factor (τ *) and a model/natural strain-rate scaling factor ((v/d)*)  and compare the ratio between these two factors and a viscosity scaling factor (i.e., viscositymodel/viscositynatural, or μ*). If the stress factor – strain rate factor ratio equals the viscosity factor, and d is scaled according to other variables in the model, then viscosity is considered to be scaled appropriately (i.e., μ* = τ */(v/d)*). Most modellers who employ ductile materials in their experiments use this method to check the scaling for viscosity (e.g., Davy and Cobbold, 1991; Caniven and Dominguez, 2021; Zwaan et al., 2022.

 

Afterword

Accounting for variable scaling is an essential part of analogue modeling – get the scaling correct increase the chances of producing model results that mean something, that answer fundamental questions about processes. However, the choice and scaling of model materials doesn’t necessarily ensure model success. Other imposed boundary conditions influence the result:

  • The geometry of the model container, particularly its moving parts that drive extension, compression, translation.
  • The rates of processes like deformation, sedimentation, or base-level change imposed on the model; should rates be linear or variable?
  • Accounting for temperature and geothermal gradients in analogue models is notoriously difficult, and yet for many crustal processes these variables are fundamental. For example, oceanic and continental rifts are commonly accompanied by volcanism and high heat flow. Can we find model materials that are compatible with such conditions where rock viscosity and cohesive strength are scaled appropriately?

Is there value in combining analogue modeling with numerical modeling to help solve some of these problems? At its heart, modeling is a creative scientific exercise that can lead us in unforeseen directions of discovery. Charles Lyell’s dictum The present is the key to the past is an invitation to recreate that past using all the tools at our disposal including using analogue, numerical, and conceptual models. We should continue to RSVP and accept the invitation.

Other posts in the series on geological models

Geological models: An introduction

Model dimensions and dimensional analysis

Analogue models

Strike-slip analogue models

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

The rheology of the lithosphere

Facebooktwitterlinkedininstagram

 

The mechanical behaviour, or rheology of the lithosphere.

Sedimentary basins are regions of long-term subsidence of, in most cases, the entire lithosphere (the crust and mantle lithosphere). This truism rolls easily off the tongue, but its implications are important – subsidence involves deformation that effects the whole lithosphere; the mechanics of that deformation determine the kind of basin that will form.

We can think of rheology in terms of the relationship between stress (force) and strain (deformation). Deformation occurs if the stress applied is greater than the strength of the rock body. The most familiar expressions of rock and sediment deformation are those we visualize in outcrop, mountain sides, and satellite images of entire mountain belts: Faults and fractures, tectonic and soft-sediment folds, translation of rock bodies from one place to another, cleavage, even compaction. The conditions for these deformations are reasonably well known; some can even be reproduced in the laboratory. Several factors influence this stress-strain relationship:

  • The magnitude of the stress.
  • The strength of the rock body, influenced by its composition and the presence of internal inhomogeneities or discontinuities (also called anisotropy), such as pre-existing fractures or fabrics like mineral alignment.
  • Temperature; as a general rule, ductility, or the ability to flow increases with temperature in concert with a decrease in yield strength (i.e. the point at which it deforms). Temperature is an important determinant of the transition from brittle to ductile behaviour.
  • Confining pressure; the yield strength of rock tends to increase with confining pressure.
  • Strain rate; Most rocks will fracture at very high rates of deformation (e.g. during earthquakes), but the same rocks may deform by ductile flow at geologically extended strain rates.

We can describe the behaviour of rock and sediment using three basic mechanical, or rheological models: elastic, plastic, and viscous behaviour. These rheological models can be applied to the lithosphere and asthenosphere in much the same way that we apply them to the deformation we see in outcrop and mountain belts.

 

Elasticity of the lithosphere

Flying through turbulence can be disconcerting, particularly if you have a window seat where you can see the aircraft wings moving up and down. This is not a flaw in wing design; it is quite deliberate.  The wings are responding to stresses developed during the violent changes in aircraft trajectory and air pressure.  If the wings were more rigid, they would be at greater risk of breaking off. Happily, there is no permanent deformation in the wing framework; they have responded elastically to the applied stress.

Most Earth materials respond to stress elastically where deformation up to some yield strength (the elastic limit) is non-permanent; the rock or sediment recovers its original shape. Deformation is permanent beyond this limit in rock strength. How this deformation occurs depends on the conditions noted above (e.g. confining pressure, temperature etc.). Deformation (extension or compression) at relatively shallow crustal levels tends to be brittle; at greater depths there is a transition from elastic to plastic behaviour (ductile flow).

The stress-strain relationships representing the three rheological models is shown in the diagram.  For elastic strain (deformation), stress is proportional to strain until the point of failure. Elastic deformation begins immediately a stress is applied; there is no yield stress (unlike plastic behaviour). In elastic bodies, this means that stress is stored until either it is released during recovery, or at the point of failure.

 

Basic stress-strain relationships for elastic and plastic behaviour (left), and viscous behaviour (right). Note that strength in viscous materials is represented as a strain rate. From multiple sources.

Basic stress-strain relationships for elastic and plastic behaviour (left), and viscous behaviour (right). Note that strength in viscous materials is represented as a strain rate. From multiple sources.

Lithospheric elasticity is one of the more important determinants of sedimentary basin formation; it allows the lithosphere to flex in response to loads. The term “load” applies to stresses that act:

  • Vertically; this includes physically emplaced sediment, volcanic or tectonic loads, plus the loading caused by temperature changes (such as cooling and density increase of oceanic crust), and
  • Horizontally, for example far-field horizontally-oriented stresses adjacent to convergent margins.

One of the more obvious manifestations of lithospheric flexure is the rebound of landmasses following retreat of large ice sheets. Post-glacial rebound of Belcher Islands in Hudson Bay, close to the centre of the former Laurentide Ice sheet, was a whopping 9-10 m/100 years about 8000 years ago, decreasing to its present rate of about 1 m/100 years. In this case rebound is recorded by spectacular flights of raised beaches, each one abandoned as the landmass rose above sea level.

 

The staircase of raised beach ridges, over an altitude gain of about 100 m from present sea level, has formed in response to lithospheric rebound following melting of the Laurentide Icesheet. The present rate of uplift is about 1 m/100 years. Tukarak Island, Hudson Bay.

The staircase of raised beach ridges, over an altitude gain of about 100 m from present sea level, has formed in response to lithospheric rebound following melting of the Laurentide Ice sheet. The present rate of uplift is about 1 m/100 years. Tukarak Island, Hudson Bay.

Lithospheric flexure is also the dominant mode of subsidence in foreland and forearc basins where the crust is tectonically loaded by thrust sheets. The amount of flexure, and therefore subsidence is controlled to a large degree by the elastic thickness of the lithosphere – thinner lithosphere will tend to bend more than thicker.

 

Flexure of an elastic beam resulting from progressive tectonic emplacement of loads (from the right). Deformation can be reversed by removal of the load.

Flexure of an elastic beam resulting from progressive tectonic emplacement of loads (from the right). Deformation can be reversed by removal of the load.

Plastic behaviour

Materials that resist deformation up to a certain yield stress, or yield strength, exhibit plastic behaviour (this is the mechanical context of the term plastic, rather than the more parochial term for things like plastic bags). Deformation beyond the yield strength is permanent.  As is shown on the stress-strain diagram, for an ideal plastic there is no deformation until the critical stress is reached.

In this model, deformation can occur in two ways (Ershov and Stephenson, 2006):

  • Instantaneously and discontinuously as brittle failure, or
  • Continuously as in ductile flow.

 

Viscous behaviour

Viscosity measures resistance to deformation, specifically that caused by flow. Thus, the dimensional units of measure are Force (in this case shear stress) multiplied by Time, all divided by Area. The standard unit is the poise, or in SI notation, Pascal-seconds (Pa.s).

In common language we usually apply the term viscosity to fluids or liquids, like paint or syrup. We can also apply the term to rocks, but we need to think of rock viscosity in a geological time frame, rather than the time it takes to apply shear stress (i.e. pouring) maple syrup on your pancakes. Some commonly used values of viscosity are listed below: the differences are measured in orders of magnitude (units of Pascal-seconds):

  • Water at 20oC 10-3
  • Maple syrup 10-1
  • Basalt lava 102
  • Granite 1020
  • Mantle 1023
  • Average crust 1025

Viscous deformation is also known as creep. During viscous behaviour, creep begins at the point stress is applied such that strain rate (rate of deformation, or in this case the rate of shear) is a function of stress; i.e. there is no yield strength. Viscous deformation is permanent.

 

Strength envelopes

The strength of the lithosphere, and therefore its rheological behaviour in response to stress (brittle, ductile, or viscous) is determined primarily by its composition and temperature, both of which change with depth. These variables distinguish crust from upper mantle; for temperature, this is a function of geothermal gradient. This means that, with depth (and location) there will be transitions from one kind of behaviour to another – from brittle to plastic (e.g. ductile), and from ductile to viscous.

Lithosphere strength is commonly represented diagrammatically as a yield strength envelope (YSE). YSEs can be constructed for oceanic and continental lithosphere, to show how strength varies according to temperature and composition of the crust and mantle lithosphere. The boundaries of each domain represent the point of failure; the rheology within each domain is elastic (Ershov and Stephenson, 2006). These diagrams are an excellent way to portray the rheological changes across the MOHO. They also demonstrate the changes in relative strength when comparing the degree of hydration of crust and uppermost mantle; wet conditions tend to weaken crust and uppermost mantle layers.

The diagram shows the strength envelopes for the upper part of continental lithosphere with a layered crust (modified from Allen and Allen, 2013, Fig. 2.38). The panels show two extremes – one with ‘dry’ lower crust and uppermost lithosphere mantle, the other with both layers hydrated. As noted by Allen x 2 in their commentary, under ‘wet’ conditions both the lower crust and upper mantle lithosphere are very weak, such that lithosphere strength is maintained almost entirely by a strong upper crust.

 

Typical yield strength envelopes for two sets of conditions in the upper 60 km of continental lithosphere: Left – strong, dry lower crust and mantle lithosphere, where strength is distributed with depth; Right – weak and wet lower crust and mantle lithosphere, where most of the strength is in the upper, brittle crust. Conditions within each envelope promote an elastic response. Beyond the envelopes the response to deformation is ductile. Strength increases to the right. Modified from Allen and Allen, 2013, Fig 2.38.

Typical yield strength envelopes for two sets of conditions in the upper 60 km of continental lithosphere: Left – strong, dry lower crust and mantle lithosphere, where strength is distributed with depth; Right – weak and wet lower crust and mantle lithosphere, where most of the strength is in the upper, brittle crust. Conditions within each envelope promote an elastic response. Beyond the envelopes the response to deformation is ductile. Strength increases to the right. Modified from Allen and Allen, 2013, Fig 2.38.

Some generalisations

It is reasonably straight forward to define the three rheological models but applying them to the lithosphere-asthenosphere adds a different level of complexity. As a general rule we can think of the upper crust as responding elastically to the point of brittle failure, the lower crust and upper mantle as a transition from elastic to plastic behaviour (e.g. ductile flow), and the asthenosphere as viscous (which permits convective flow). However, complications with this simple story arise if, for example, the crust is layered, and again if the lower crust is dry (generally stronger) or wet (weaker). Through any section of lithosphere all three processes will operate simultaneously. The depth at which each process acts also varies laterally, depending on factors such as geothermal gradients and changes in composition.

 

Topics in this series

Sedimentary basins: Regions of prolonged subsidence

Defining the lithosphere

Isostasy: A lithospheric balancing act

Classification of sedimentary basins

Stretching the lithosphere: Rift basins

Nascent conjugate, passive margins

Basins formed by lithospheric flexure

Accretionary prisms and forearc basins

Basins formed by strike-slip tectonics

Allochthonous terranes – suspect and exotic

Source to sink: Sediment routing systems

Geohistory 1: Accounting for basin subsidence

Geohistory 2: Backstripping tectonic subsidence

 

Related topics

Crème brûlée, jelly sandwich, and banana split; the manger a trois of layered earth models

The sea level equation

Sea level change: busting a few myths

The thermal structure of the lithosphere

 

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Crème brûlée, jelly sandwich, and banana split; the manger a trois of layered earth models

Facebooktwitterlinkedininstagram

Some things in science are just too difficult to comprehend: the temperature at the center of the sun (15,000,000oC), the age of the earth (4.6 billion years), the size of a nano-particle (1-100 nanometres, or billionths of a metre). We can include in this list of imponderables, the skinny outer layers of the earth: the one we are in daily contact with (the crust), and other layers beneath it. Our familiarity with the crust is usually in terms of the dirt, rock, and water we work with. But what is it like 30km down? And, beneath the crust, the upper mantle is beyond reach of our senses. What does this layer look like? How does it respond to being pushed around?

Some scientists (geologists, geophysicists) spend a great deal of time pondering questions like these. The crust and upper mantle layers are collectively referred to as lithosphere. Beneath the continents it averages 150 km thick; beneath the oceans, it is as thin as 10 km below mid-ocean ridges. Given we spend our entire lives on the uppermost veneer, a reasonable person might ask ‘why is it important?’.

A few common answers include: Most earthquakes are generated in the lithosphere; Magmas erupted at volcanoes melt at these depths. But the overarching reason is that all tectonic plates are born and destroyed as lithosphere. Plate tectonics governs pretty well everything that happens on earth over geologically short and long time-scales. So, what appears arcane at first sight, does have practical applications.

Enter the dessert trolley. There are three choices: a crème brûlée, a jelly sandwich, and a banana split. Proposed as models of the layered earth, they serve a dual purpose: they provide visual descriptions of how the lithosphere might be structured and, after evaluating the merits of each, they can be consumed.

The crème brûlée is a two-layered model.  A viscous fluid base (custard) is capped by a thin crust of caramelized sugar. The crust behaves in two ways. Poke it gently in the centre, and it will bend slightly – release the pressure and it will return to its original shape.  This represents elastic behaviour (think also of wire springs, or rubber bands). Press it too hard and it will break into several ragged pieces; in this instance, you have exceeded the elastic limit, or strength, and induced brittle failure. Earthquakes represent brittle failure where earth’s crust fractures, is displaced, and in the process causes mayhem. The crème brûlée model is probably the simplest of the dessert trio in terms of its relevance to the lithosphere.

The jelly sandwich is potentially the more variable of the three analogues. It is a three-layered model where two pieces of bread are separated by a layer of jelly.  Here, the upper bread layer represents a strong upper crust, and the jelly a weak lower crust. The bottom bread layer is compared with a strong upper mantle – in contrast to the weak custard (mantle) layer in the crème brûlée. The upper and lower bread layers are both quite bendy (unless you have toasted the bread). If you use plain white bread, then bending will be uniform. But if you prefer whole-grain slices there will be lots of lumps and greater heterogeneity, and hence a less predictable response to the application of pressure, or stress. The jelly is much less fluid than custard. It can behave elastically – witness the wobbling, that represents deformation from which it recovers, but at a certain point it too will fail.  Bread is less rigid than a crème brûlée crust; any kind of twisting or bending will probably result in some permanent deformation (i.e. it doesn’t bounce back to its original shape). Unlike the crème brûlée crust, bread is less prone to brittle failure.

The banana split adds another level of complication to models of the lithosphere. The rationale for this model is that the lithosphere contains zones of weakness, particularly near the boundaries of tectonic plates – imagine these plates colliding or sliding past one another, where the forces are large enough to create mountain belts and consume oceans. Here, scoops of ice-cream represent blocks of crust and mantle that are separated by large, very deep faults. This is a very temperature-dependant model. As the ice-cream melts there is a zone of weakness between it and the adjacent scoop (block). The presence of fluid, particularly water, exacerbates this weakness. In this dessert, we need to translate the fluid boundary between scoops of ice-cream, to structures 10s of kilometres deep. Modern examples include the Alpine fault in New Zealand, and San Andreas Fault in California. Some of these large structures can last for very long periods of geological time (100s of millions of years), and potentially influence events in the crust-upper mantle long after they first formed.

All models in science are simplifications of the things we try to explain. It may be the case that some consider the dessert trio to be trivial, even silly, providing little useful scientific information for the representation of the crust and mantle.  But the utility of models and analogues is not only in scientific explanation, but to present a complex world in visually interesting, and yes even amusing ways. Models and analogues need to stir the imagination of folk who are not directly involved in this kind of research but have a vested interest in it. In this regard, the dessert trio works, even if folk can relate to them only via our taste buds.

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

It only takes a moment; the ups and downs of earthquakes

Facebooktwitterlinkedininstagram

 

Destruction wrought by the Kaikoura 7.8M earthquake, 2016.

Seismic metaphors, or seismic as metaphor? Seismic, a word that geologists and geophysicists traditionally thought was reserved for their use, has been purloined by politicians and social scientists to describe momentous shifts in things like public attitudes and voter propensities.  Seismic is the anglicized derivative of the Greek verb seien, (shake) and the word seismos (earthquake).  Apparently, it first appeared in English language writing about the mid-19th century. Personally, I find it satisfying that the social milieu sees fit to apply the scientific word in such a useful, metaphorical sense.

The word seismic is also a nice descriptor of our restless, physical world, especially the bits we live on.  Most of the earth’s crust is under stress, some parts more than others.  Most stresses are generated by the movement and jostling of tectonic plates, particularly at boundaries where plates converge, collide, or slide past one another.  Different parts of the crust respond to stress by bending; this includes seemingly hard, immovable rock. If the stress is removed then the deformed rocks return to their original states; this is referred to as an elastic response. However, if the earth materials are bent too far or too fast, they will break. An interesting analogue for this process, and a historical one, is the collapse of a Tacoma suspension bridge in 1940.

Here, steel girders and tarmac bent and twisted under stress until the deformation reached a limit (called the elastic limit), at which point the bridge failed. When the earth’s crust fails, the seismic event, or earthquake, can be devastating. Earthquakes are caused by the sudden brittle failure of rock under stress; the failure takes place along a fault, across which land (or sea floor) is moved up, down, or sideways.  The rapid displacement of rock masses produces pulses of energy, or seismic waves.  It is these waves that do the damage.

Finding an earthquake focus and epicenter

There are two main kinds of seismic wave; body waves that propagate through the earth’s interior, and surface waves that move along the earth’s surface. Body waves include a primary, or P-wave, and a secondary or S-wave.  P-waves travel fastest; they are also called compressional waves because they tend to push and pull materials as they propagate. Slower S-waves, or shear waves, produce a side-ways motion. Shear waves are not transmitted Schematic illustration of P and S seismic waves through fluids such as sea water, or the molten interior of the earth. The cartoon below illustrates how the earth reacts to these two wave types.  Surface waves are the slowest to propagate but they are also the seismic pulses that do most of the damage during an earthquake. The animation shows the earth motion for one kind of surface wave;

Rayleigh Waves. Rayleigh Waves produce a circular, or orbital motion of earth materials at the surface, a bit like particle motion beneath sea waves.

 

Each type of seismic wave is identified by the speed at which it moves, and the kind of movement that sediment and rock are subjected to.  These differences are expressed on seismograph recorders.  On a typical seismogram (below),  P-waves arrive first, followed by S-waves.   S-waves tend to have lower frequencies than P-waves (more spread out on the graph), but higher amplitude.  Surface waves usually have the highest amplitude.

First arrivals of P and S waves during an earthquake

Earthquake magnitude (M) reflects the severity of ground roll and shaking, and on seismograms corresponds to the amplitude of the signal (usually of surface waves).  M is expressed as a number (M1.8, M4.6, M7.8) up to a maximum of 10 (10 might be caused by a large meteorite impact – when there isn’t much left).  The numbers, and therefore the magnitude scale, are logarithmic, such that a magnitude of 4 (104) is 10 times smaller than M5 (105), and 100 times smaller than M6 (106).  We can also think of the changes in magnitude in terms of the energy released during a quake. It has been determined, (as an approximate empirical relationship) that for every unit increase in magnitude, there is a 27.5-fold increase in energy. The difference in energy is also logarithmic, such that an M8 event releases 571,914 times more energy than an M4 event (the magnitude is 10,000 times greater).  These kinds of numbers demand a degree of respect for the twists and turns our earth can throw at us.

Seismograms also help seismologists determine the Epicenter of an earthquake; the epicenter is the map location rather than the actual location, or focus, at depth. The calculation makes use of the fact that P-waves are faster than S-waves, so that the distance to an epicentre is based on the difference in arrival times for each type of wave.  Distance in this case is the radius of a circle centered on the seismograph location.  To triangulate the epicentre, at least two more seismographs in different locations are needed – in reality there can be hundreds of seismographs that allow the calculation.  If circles are drawn about each seismograph, each with a different radius, they will intersect at the epicentre.  Variations in the structure of the earth mean that seismic wave velocities can vary, so that the circles may not all intersect at a single point.  However, the large number of seismographs around the world means that location of the epicentre is usually accurate.

P and S wave travel times and triangulation to find an epicenter

The November 14, 2016 earthquake in Kaikoura, New Zealand was a M7.8 event.  We live about 500km north of the epicentre.  Almost on the stroke of midnight we felt definite shaking (P-waves).  A few seconds later the shaking increased significantly.  At this point trees were swaying and water was slopping over the edge of the pool; this was due to ground roll from the slower surface waves.  Other than an adrenaline rush, there was no damage for us; farther south it was a very different story.  One spectacular outcome was the abrupt, lateral 6m shift along part of the New Zealand coastline; the same coast was also uplifted 1-2m.

Contour map of land movement and displacement from the aftermath of 2016 Kaikoura earthquake, NZ

I’ve always thought that New Zealand is a great place to witness geology in action.  But sometimes it can go a bit too far.

An excellent technical paper on the Kaikoura event by Ian Hamling (GNS Science) and a host of co-authors, has been published in Science, March 26, 2017.

 

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin