Tag Archives: ductile flow

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