Tag Archives: Reynolds number

Fluid flow: Shields and Hjulström diagrams

Facebooktwitterlinkedininstagram

A schematic portrayal of some important functions used to determine the character of fluid flow and sedimentation.

Getting sedimentary grains moving in a flowing medium

Formation of granular sedimentary layers and many of the bedforms within requires movement of grains from one location to some other location, as bedload or suspended load. Initiation and maintenance of grain movement in a fluid (air or water) requires that the shear stress derived from the flowing fluid must overcome the opposing gravitational, viscous, and friction forces acting on each grain. The critical or threshold shear stress is that where grain movement begins.

An understanding of the conditions required to initiate grain movement is important for engineering problems involving fluid flow, such as the stability of bridge abutments in rivers, and for sedimentological interpretations of ancient environments. Questions like “What flow conditions were required to move the range of clast sizes in a pebbly sandstone” are central to sedimentary facies analysis.

A diagrammatic representation of the various forces acting on non-cohesive grains in a flowing fluid (air, water). The small black arrows around the grain boundaries represent viscous shear. A typical velocity profile (left) shows how bulk flow velocity increases with distance from the bed (blue arrows); the resulting pressure gradient is responsible for the lift forces acting on each grain. Lift forces may also develop from turbulence. The velocity profile shows the important differences in fluid flow at the sediment-water interface compared with that above the bed. Relatively continuous grain motion along a granular bed will take place when the drag and lift components exceed the opposing gravitational, viscous, and grain contact forces.

A diagrammatic representation of the various forces acting on non-cohesive grains in a flowing fluid (air, water). The small black arrows around the grain boundaries represent viscous shear. A typical velocity profile (left) shows how bulk flow velocity increases with distance from the bed (blue arrows); the resulting pressure gradient is responsible for the lift forces acting on each grain. Lift forces may also develop from turbulence. The velocity profile shows the important differences in fluid flow at the sediment-water interface compared with that above the bed. Relatively continuous grain motion along a granular bed will take place when the drag and lift components exceed the opposing gravitational, viscous, and grain contact forces.

The definition of initial grain movement is not as straight forward as it first appears. Does the threshold shear stress (or critical flow velocity) for grains on a sediment bed correspond to an initial nudge, or does the grain need to travel one or more revolutions of its circumference? In most natural sediment beds there is a range of grain sizes so which grain threshold stress applies – that responsible for movement of the coarsest, the mean, or median grain size? John Southard has given us an excellent summary of this problem.

Two graphical representations of these threshold conditions stand out; the iconic Shields and Hjulström diagrams. Both are empirical constructions derived from experimental data; both were developed in the 1930s – the original and modified forms of these diagrams are still used widely.

 

Shields diagram

Albert Frank Shields (1908 – 1974) was an American engineer whose experiments on the transport of granular sediment led to the formulation of the eponymous Shields Parameter (Θ) that expresses the shear stress (τ) required to initiate grain movement. The flume experiments were actually conducted in 1930s Germany and published in 1936 (Application of similarity principles and turbulence research to bed-load movement). The Shields diagram plots (Θ) against the Reynolds Number that describes the hydraulic conditions across the grain boundary. The original diagram specifies two fundamental domains: One of grain movement, the other where the threshold shear stress is not high enough to initiate movement. His experiments used grains of different densities – amber, lignite, granite, barite, and sand.

A modified version of Shields original 1936 graph (his Figure 6), showing the data envelope (grey) about his threshold curve (solid black line); the envelope encompasses grains of different densities. The threshold shear stress and boundary (grain) Reynolds Number are dimensionless. Shields added annotation to his original diagram - the bedforms he observed, the beginning of grain saltation, and bed erosion (he called it abrasion). The extended threshold curve at low Reynolds Numbers (dashed line), plus the turbulence boundaries were added by later workers (see Southard, 2021).

A modified version of Shields original 1936 graph (his Figure 6), showing the data envelope (grey) about his threshold curve (solid black line); the envelope encompasses grains of different densities. The threshold shear stress and boundary (grain) Reynolds Number are dimensionless. Shields added annotation to his original diagram – the bedforms he observed, the beginning of grain saltation, and bed erosion (he called it abrasion). The extended threshold curve at low Reynolds Numbers (dashed line), plus the turbulence boundaries were added by later workers (see Southard, 2021).

The Shields Parameter Θ can be written as:

Θ = τc.D2/(ρs – ρw)gD3

where τc = critical stress at the grain boundary; D = mean grain diameter, and ρ the density of the solid grains and water respectively. The value s – ρw)g is the submerged specific weight of a grain. The numerator τc.D2 is proportional to the fluid force acting on a grain; the denominators – ρw)gD3 is proportional to the weight of the grains. Θ is dimensionless because the shear stress (i.e., the pressure exerted on a grain) can be written as τ = ρw.gz (also the general form of the equation used to calculate hydrostatic and lithostatic pressures), where z is a characteristic depth or thickness and has the same units as grain diameter.

The Shields diagram plots (Θ) against the grain Reynolds Number. Use of the Reynolds Number (Re) is important because it relates inertial forces and dynamic fluid viscosity to the two fundamental types of flow – laminar and turbulent.

Re = ρw VD/μ

where V the mean velocity reflects shear rate and inertia forces, and μ is fluid viscosity that measures the resistance to shear. Re is dimensionless. In Reynolds’ original experiments D was the flow tube diameter; in the Shields diagram it corresponds to mean grain diameter. Thus, at low Re values viscous forces suppress turbulence and the flow is laminar. At high Re values inertial forces exceed viscous forces and flow is turbulent.

There have been several replottings, modifications, and recastings of Shields’ 1936 diagram. R.A. Bagnold (PDF available) published a US Geological Survey Report in 1966 where he considered the initiation of grain movement and the maintenance of a suspension load for finer-grained particles that incorporates Shields’ criteria. His entrainment diagram (his Figure 8) plots the non-dimensional threshold shear stress against (dimensional) grain diameter.

Bagnold's plot of non-dimensional threshold shear stress with actual grain diameter (modified here from his Figure 8), showing the domains of no grain motion, grain movement as part of the bedload, and the theoretical limits for particle suspension. The grey band that defines grain movement captures the general spread of experimental data. The lower bounding line is Shields threshold curve; the upper bounding line is Bagnold's calculated threshold curve.

Bagnold’s plot of non-dimensional threshold shear stress with actual grain diameter (modified here from his Figure 8), showing the domains of no grain motion, grain movement as part of the bedload, and the theoretical limits for particle suspension. The grey band that defines grain movement captures the general spread of experimental data. The lower bounding line is Shields threshold curve; the upper bounding line is Bagnold’s calculated threshold curve.

Recasting and replotting of the Shields criteria by M.C. Miller et al., 1977 is another frequently cited version. These authors extended the grain size Reynolds numbers by more than two orders of magnitude beyond those used by Shields. The non-dimensional variables are the same as in Shields’ original diagram. Their graph shows the spread of data (from various sources).

Miller et al., (1977) replotting of Shields’ non-dimensional threshold shear stress and the non-dimensional Reynolds Number over Re values about 2.5 orders of magnitude higher than those used by Shields. The authors used experimental data from multiple sources. The spread of data is indicated by the threshold envelope.

Miller et al., (1977) replotting of Shields’ non-dimensional threshold shear stress and the non-dimensional Reynolds Number over Re values about 2.5 orders of magnitude higher than those used by Shields. The authors used experimental data from multiple sources. The spread of data is indicated by the threshold envelope.

 

Hjulström’s diagram

Filip Hjulström’s (1902-1982) diagram is particularly useful for sedimentologists who prefer to think in terms of measurable flow velocities rather than shear stresses or shear velocities. It was published in 1939. It is still a popular reference. His diagram specifies three domains: erosion (net loss of sediment), deposition (net gain in sediment), and transport that may involve components of erosion or deposition.

[Hjulström, F. (1939). Transportation of detritus by moving water: Part 1. Transportation. In P. D. Trask (Ed.), Recent marine sediments. A Symposium: Tulsa, Oklahoma (pp. 531). Tulsa, OK: AAPG.]

Hjulström was a Swedish geographer who conducted experiments on flow and sediment transport for his PhD on the River Fyris. His experiments involved measuring the mean flow velocity at which grains of a specified diameter began to move; the experiments were conducted in a canal with water depth of one metre. Hjulström understood the problems of what constitutes a representative flow velocity – it can be taken as the surface flow velocity, the mean velocity over a specified depth, or more appropriately for this particular problem, the velocity at the grain boundary but this was difficult to measure (at least back in the 1939s). He chose mean velocity because it is relatively easy to measure. His experiments included a much greater range of grain sizes than those conducted by Shields – clay through cobble sizes.

A typical Hjulström plot delineating the principal domains of deposition, transport, and erosion. Modified from Nichols, 2009, Figure 4.5. I have added the region where particle cohesion influences erosion and transport. The domain boundaries, although drawn as solid lines, are in fact more fuzzy in their placement because of the spread of experimental data, and subtle but important differences among experimental conditions such as flow depths and, velocity profiles.

A typical Hjulström plot delineating the principal domains of deposition, transport, and erosion. Modified from Nichols, 2009, Figure 4.5. I have added the region where particle cohesion influences erosion and transport. The domain boundaries, although drawn as solid lines, are in fact more fuzzy in their placement because of the spread of experimental data, and subtle but important differences among experimental conditions such as flow depths and, velocity profiles.

The boundaries that describe each domain are based on flow depths of one metre, and average grain density of 2.5 – 2.6 g/cm3. The boundaries will shift for different water depths and grain densities. For example, domain boundaries will tend to move upwards for carbonate-dominated sediment (grain density 2.7 – 2.8 g/cm3) because greater shear stress is required to initiate grain movement. The boundary between bedload and suspension load transport will be diffuse depending on the textural and mineralogical character of the sediment, as well as the flow characteristics (e.g., compare the hydraulic behaviour of mica flakes to that of spherical quartz grains).

The lower curve separates depositional flows from transport flows. The upper V-shaped curve reflects the increasing cohesion of finer grained particles, particularly the clay and fine silt fractions of muds. Cohesion in this case is primarily a result of the large surface area relative to particle size, plus the residual electrical charges on clay mineral surfaces. Cohesive forces are also responsible for the elevated flow velocities required to erode consolidated muds. In fact, erosion of consolidated muds commonly produces chunks, or mud rip-ups rather than individual clay-silt particles.

Reading this graph is relatively straight forward as long as you are aware of the caveats such as grain shape and density, and whether flow is laminar or turbulent. For example, for any grain diameter, follow the vertical axis to determine the flow dynamics. It is also a useful diagram to compare the flow requirements for maximum clast sizes in a deposit with the requirements for mean grain size of that deposit.

Using Hjulström’s diagram to determine ball-park flow velocities for crossbedded, pebbly, glaciofluvial sandstone, where grain sizes range from medium sand (0.25 – 0.50 mm) to the upper size limit for pebbles (64 mm). The range of flow velocities for each size range is relatively broad, but the overall magnitude of velocities needed to transport these sediments ranges over about 2 orders of magnitude, from about 0.02 m/s to 3 m/s. Our paleoenvironmental interpretation of this deposit needs to account for this range of flow velocities, and for the rapid changes in flow energy from one crossbed to the next. For scale, the red dot is 26mm.

Using Hjulström’s diagram to determine ball-park flow velocities for crossbedded, pebbly, glaciofluvial sandstone, where grain sizes range from medium sand (0.25 – 0.50 mm) to the upper size limit for pebbles (64 mm). The range of flow velocities for each size range is relatively broad, but the overall magnitude of velocities needed to transport these sediments ranges over about 2 orders of magnitude, from about 0.02 m/s to 3 m/s. Our paleoenvironmental interpretation of this deposit needs to account for this range of flow velocities, and for the rapid changes in flow energy from one crossbed to the next. For scale, the red dot is 26mm.

Hjulström also contributed a paper Transportation of Detritus by Moving Water to an SEPM 1955 Special Publication No. 4 ‘Recent Marine Sediments’

 

This post is a companion to:

The hydraulics of sedimentation: Flow Regime

Sediment transport: Bedload and suspension load

Fluid flow: Froude and Reynolds numbers

Fluid flow: Stokes Law and particle settling

 

Plus related posts on Stratigraphy-sedimentology

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin

Fluid flow: Froude and Reynolds numbers

Facebooktwitterlinkedininstagram
Supercritical and subcritical flow domains manifested as standing waves and ripples

Supercritical and subcritical flow domains manifested as standing waves and ripples

19th century experiments that helped quantify the nature of fluid flow, surface waves, and bedforms.

It all depends on inertia, like the reluctance to get out of bed on a cold winter’s morning. But rather than feeling guilty, acknowledge that by sleeping in you are adhering to the mechanical Laws that prevent the universe from collapsing – the inertial forces that keep planets in orbit around their suns, and suns in motion through their galaxies.

Inertia is loosely defined as a force that resists the change in motion of a body; here motion refers to a vector that describes velocity and direction, and ‘body’ refers to pretty well anything composed of matter, including a body of fluid. The term was coined by astronomer Johannes Kepler (17th century); his erstwhile colleague Galileo demonstrated its qualities by experimenting with balls rolling along sloping surfaces.

However, it was Isaac Newton who codified the properties of inertia in his three Laws of Motion – apparently Newton credits Galileo with the discovery. The 1st Law, also called the Law of Inertia, states that the motion of a body will not change unless an external force acts on it (i.e., to accelerate, decelerate, or change its direction). The 2nd Law quantifies the relationship between an external force F, mass (m) and acceleration (a) as F = ma. And the 3rd Law states that when an external force is applied, there will be an equal and opposite force that resists the change in motion, i.e., an inertial force – also called the Action-Reaction Law.

Inertial forces depend on the mass of a body – the larger the mass, the greater the force. In fact, the concept of mass itself is based on inertia.

 

Inertia and fluid flow

Inertial forces are central to the quantification of fluid mechanics. We have William Froude (1810-1879) and Osborne Reynolds (1842-1912) to thank for their eponymous numbers (Froude number and Reynolds number) that describe the characteristic states of flow. And because these numbers are dimensionless, they allow experiments with models (e.g., wind tunnels, sediment flumes) that can be scaled to real-world fluid flow phenomena. Scaling can be applied to almost anything related to fluid flow – from the motion of a boat through water, to quantifying the formation of sediment bedforms or sediment gravity flows from small-scale sediment flume experiments. The importance of Froude and Reynolds numbers cannot be overstated.

 

Froude number

Froude’s influential paper of 1861 was published by the Institute of Naval Architects (PDF available). Froude had surmised that, to predict the behaviour of a ship moving through water, he would need to experiment with much smaller versions of ships, or models, that could be scaled to the behaviour of much larger vessels. Thus, Froude’s number was derived from experiments with model boats, a few metres long.

The number expresses the characteristics of flow, including surface waves and bedforms, as the ratio between inertial forces and gravitational forces:

                                                          Fr = V/√(g.D)

Where V is bulk flow velocity (having dimensional units L.T-1) that reflects the dominant effect of inertia on surface flows, and the component √(g.D) where g is the gravitational constant (units of L.T-2), and D is water depth (units of L). The denominator represents the speed of a surface gravity wave relative to the bulk flow velocity (√(g.D) simplifies to units of velocity). Whether the surface wave is faster, slower or the same speed as the bulk flow will depend on its resistance to move, or its inertia. Fr is dimensionless.

The numerical value of Fr is used to define three conditions of flow. If Fr = 1 (numerator = denominator), then any surface wave will remain stationary – it will not move upstream or downstream. This condition occurs when both the velocities and water depth are at critical values. Not surprisingly, this condition is called critical flow. A common manifestation of critical flow is the formation of stationary waves (or standing waves) above and usually in phase with antidune bedforms (i.e., upper flow regime).

 

A plot of the experimentally determined stability fields for bedforms, as a function of grain size and flow velocity. The transitions from one field to another are abrupt or gradual as indicated. Modified from Ashley, 1990, Figure 1 with minor additions.

A plot of the experimentally determined stability fields for bedforms, as a function of grain size and flow velocity. The transitions from one field to another are abrupt or gradual as indicated. Modified from Ashley, 1990, Figure 1 with minor additions.

When Fr < 1, inertial forces dominate, and the result is a subcritical condition – tranquil flow. This corresponds to lower flow regime bedforms such as ripples and larger dune structures.

When Fr > 1, gravitational forces dominate resulting in supercritical flow conditions. The corresponding stream flow surface conditions manifest as an acceleration of flow such that stationary waves break upstream (chutes – upper flow regime), commonly followed by a rapid decrease in flow and formation of a hydraulic jump where Fr < 1 (chute and pool conditions). A hydraulic jump is the region of turbulence that represents the transition from supercritical (laminar) flow to tranquil flow – as shown in the kitchen sink example below. Supercritical flow is also common in pyroclastic density currents.

A kitchen sink demonstration of the transition from laminar, supercritical flow to turbulent subcritical flow via a hydraulic jump.

A kitchen sink demonstration of the transition from laminar, supercritical flow to turbulent subcritical flow via a hydraulic jump.

The complexity of flow transitions in a small natural system is shown in this video clip of supercritical and subcritical (tranquil) domains in a small, shallow stream. The standing waves (left) represent critical conditions where the speed of the waves matches the stream flow velocity. Supercritical conditions downstream produce chutes. Downstream migrating ripples in the foreground indicate subcritical flow.

 

Reynolds number

Schematic representation of laminar and turbulent flow using hypothetical flow lines. The blue arrow (right) indicates mean flow velocity for turbulent flow.

Schematic representation of laminar and turbulent flow using hypothetical flow lines. The blue arrow (right) indicates mean flow velocity for turbulent flow.

Unlike Froude who was more concerned with the surface configurations of a flowing medium, Reynolds experiments in glass pipes were concerned with the bulk structure of flow, in particular the transition from laminar to turbulent flow (Reynolds, 1883, PDF available). To picture this, think of a flowing fluid as a set of flow lines. In laminar flow, the flow lines are parallel, or approximately so, and relatively straight. The flow velocity will be the same across each flow line. By contrast, turbulence is described by flow lines that constantly change direction and velocity. In a flowing stream this is manifested as eddies, boils, and breaking waves. In sedimentary systems, turbulence is an erosive process, and an important mechanism for maintenance of sediment suspension through water columns and in sediment gravity flows.

 

The video below shows the abrupt transition from laminar flow in the slightly sinuous trail of smoke, to turbulent flow above.

 

To understand the nature of the laminar-turbulent flow transition, Reynolds considered four variables:

  • Fluid density ρ (units of M.L-3).
  • Fluid viscosity (μ) that measures the resistance to shear and is strongly temperature-dependent. μ has units of M.(L.T)-1
  • Mean velocity of flow V, that reflects shear rate and inertia forces (units of L.T-1), and
  • Tube diameter D that influences the degree of turbulence (units of L).

Reynold’s number is written as:

                                                                  Re = ρVD/μ

that expresses the ratio of inertial (resistance) forces to viscous (resistance) forces. Re is dimensionless.

In his glass tube experiments, Reynolds systematically varied μ, V, and D (μ was varied by heating the water). For each combination he discovered that the transition from laminar to turbulent flow in water was abrupt, and consistently had Re values of about 12000. Reversing the experiment gave values of about 2000 for the transition from turbulent to laminar flow.

Reynolds’ original glassware used in his fluid flow experiments. Tube diameters ranged from 2.54 cm to 0.62 cm. Coloured dye was introduced through a funnel. In all experimental runs, the transition from laminar to turbulent flow was abrupt. These figures are from Reynolds’ 1863 paper.

Reynolds’ original glassware used in his fluid flow experiments. Tube diameters ranged from 2.54 cm to 0.62 cm. Coloured dye was introduced through a funnel. In all experimental runs, the transition from laminar to turbulent flow was abrupt. These figures are from Reynolds’ 1863 paper.

Re can be used to determine the kind of flow in large and small fluid systems. As a general rule:

  • Re values <2000 indicate laminar flow,
  • Re >4000 turbulent flow, and
  • the region in between these two extremes reflects transitional flow.

Flow in most open-surface geological and geomorphic systems tends to be turbulent, with familiar examples including channelized flow (river, tidal and submarine channels) and more open flow across broad expanses such as continental shelves. It also includes volcaniclastic systems like pyroclastic flows and surges. Experimental flow in flumes produces a variety of bedforms at Re values that range from about 4000 to >100,000.

Laminar flow at low velocities is probably responsible for deposition of lower flow-regime plane beds; Allen (1992) has suggested that laminar flow at higher velocities may be restricted to thin sheet floods. Fluids having high viscosity, such as glacial ice and lava, commonly exhibit laminar flow. The Re value in microscopic rock-fluid systems, such as intercrystal boundaries in diagenetic environments, will also be low because fluid viscosity will dominate in such confined spaces.

 

Comparing Froude and Reynolds numbers

Froude numbers express a relationship between the free-surface of a flow and the various waves and ruffles that form there, and bedforms at the sediment-water interface. Reynolds numbers deal to the bulk characteristics of flow – whether it has laminar or turbulent structure.

The numbers Fr and Re are like chalk and cheese – they are not comparable. Both are dimensionless ratios, but that’s where the similarity ends. Both functions depend on inertial forces (the resistance to do anything), but for Fr the inertial component is in the denominator, and for Re in the numerator. Thus, if inertial forces become dominant, the numerical value of Fr decreases and that for Re increases.

Both numbers have application well beyond the relatively narrow field of sedimentology. Both are used extensively in scaled models – Fr for elucidating the efficacy of movement through a fluid – boats through water, airplanes through air. Re is used extensively to describe fluid flow in biological systems.

Allen (1992) has given sedimentologists a diagram that generalizes the relationship between Fr and Re in terms of mean flow velocity and flow depth. The boundaries of the 4 domains correspond to critical flow transitions; subcritical (tranquil) to supercritical for Fr, and laminar to turbulent for Re. I have added the most common bedforms to these domains.

J.R.L. Allen’s (modified slightly from 1992, Fig. 1.21) plot showing four domains of fluid flow, the boundaries of which are defined by the laminar-turbulent flow transition (Reynolds), and the subcritical-supercritical flow transition (Froude).

J.R.L. Allen’s (modified slightly from 1992, Fig. 1.21) plot showing four domains of fluid flow, the boundaries of which are defined by the laminar-turbulent flow transition (Reynolds), and the subcritical-supercritical flow transition (Froude).

Literature

There are many publications on this topic, but I highly recommend two publications that provide greater detail of theory and practice on this and other topics in fluid flow and sedimentation:

John Southard’s excellent (open access), online Introduction to Fluid Motion and Sediment Transport.

J.R.L. Allen 1992 (and later editions) Principles of Physical Sedimentology (that no sedimentologist should be without).

 

Other posts in this series

Identifying paleocurrent indicators

Measuring and representing paleocurrents

Crossbedding – some common terminology

Sediment transport: Bedload and suspension load

The hydraulics of sedimentation: Flow regime

Fluid flow: Shields and Hjulström diagrams

Fluid flow: Stokes Law and particle settling

Facebooktwitterlinkedininstagram
Facebooktwitterlinkedin