8.1 Turbulent and Non-Newtonian Flows#
Lecture 8.1
Stephen Neethling
Table of Contents#
Outline#
Reynolds number and the onset of turbulence
Turbulent flow in pipes
Modelling turbulence
Other Rheologies
The simple balances from the earlier lectures are not the end of the story even for flows in simple geometries – Why?
An explicit assumption in these derivations is that the flow is steady (no time dependency)
What happens if there is a perturbation in the flow?
If this perturbation grows, it means that the flow can not be steady and that this assumption is invalid
Perturbation will be damped if viscous forces are stronger than the inertial force associated with the perturbation
Reynolds Number#
Reynolds number represents the balance between inertial and viscous forces
If Reynolds number is small perturbations are suppressed – Laminar Flow
If Reynolds number is large perturbations grow – Turbulent Flow
Reynolds Number – Flow in Pipe#
Pipe diameter as characteristic length scale
Average velocity as characteristic velocity
Laminar: Re < 2100
Transition: 2100 < Re < 4000
Turbulent: Re > 4000
Flow in Pipe#
As Reynolds number increases beyond transition average flow profile goes from parabolic to closer to plug flow
Laminar boundary layer near the wall – Gets thinner as the Reynolds number increases
Thickness of boundary layer also influenced by surface roughness
Flow in Pipes - The Engineering Approach#
Fanning Friction factor:
Ratio of shear stress exerted on the fluid to its kinetic energy:
For fully developed flow the shear stress on the wall must balance the pressure drop:
Combine to give Fanning friction factor:
Friction factor is a function of Reynolds Number and pipe roughness:
For laminar flow \( f = \frac{16}{re} \)
For turbulent flow empirical relationships or charts are used
Note that that Moody friction factor is often used instead of the Fanning friction factor, but it is simply 4 times larger
Example#
A horizontal pipe needs to have water flow through it at 0.1m3 /s
𝜇 = 0.001 𝑃𝑎. 𝑠
𝜌 = 1000 𝑘𝑔/𝑚3
The pipe has a diameter of 30cm and can be assumed to be smooth
If the pipe is 100m long, what pressure drop is required?
Flow given a Pressure Drop#
More complex to calculate unless laminar
Calculation tactic:
Guess a fluid velocity
Calculate Reynolds number
Look up the friction factor
Use the specified pressure drop to calculate a new fluid velocity based on the friction factor
Repeat from step 2 until sufficiently converged
Example#
If we halve the pressure drop calculated in the previous example, what will the new flowrate be?
Modelling Other Turbulent Flows#
Pipe flow is a very special case for which lots of experimental data is available
Can use empirical relationships to accurately predict behaviour What about turbulent flow in more complex geometries?
Can readily rely on empirical data, especially if we want to use modelling for design
We need to be able to understand and model turbulence
What Happens in Turbulent FLow?#
At high Reynolds numbers large scale flow structures and instabilities will develop smaller scale perturbations
This cascade of larger perturbations generating smaller flow structures will continue down until the flow structures are small enough that viscosity can “win”
The Kolmogorov length scale
At this finest level the energy in the flow is dissipated as heat via the viscosity
Direct Numerical Simulation (DNS)#
Solve Navier-Stokes equation with enough resolution to resolve all the eddies down to their laminar cores
Very computationally expensive – prohibitively so for all but the lowest Reynolds numbers and smallest systems
We can get a rough estimate of the resolution required based on Kolmogorov length scales and dimensional arguments
Water at 1 m/s in a 1 m channel would require a resolution of about 10 µm (about 1015 elements in a cubic box)!
Turbulence Modelling#
As DNS is only useable on a small range of problems, we need to model the effect of turbulence
Turbulence dissipates more energy than the macroscopic strain rate would suggest
Need to calculate this extra dissipation
3 main approaches
Simply increase the viscosity!
Trivial to do – often used in large scale simulations such as ocean modelling
Assume that you can separate the turbulent and mean components of the flow
Reynold Average Navier Stokes (RANS) type models
We will look at the k-ε model as it is the most commonly used of this type
Resolve as much as you can and then use a model for the bits you can’t resolve
Large Eddy Simulations (LES)
Both RANS and LES approaches involve the calculation of an additional turbulent viscosity that needs to be added to the underlying fluid viscosity
Mean and Fluctuating Flow Components#
Turbulent structures are typically embedded within macroscopic flow structures
In analysing and modelling turbulence it is thus useful to distinguish between these flow components
Assume that the flow at any point can be decomposed into a mean component and a fluctuating component:
where 𝐮 is the instantaneous velocity, 𝐔 is the average velocity and 𝐮′ is the fluctuating component of the velocity
Reynolds Stress Tensor#
The averaged equation is similar to the original except for the inclusion of the divergence of a rank 2 tensor:
Known as the Reynolds Stress Tensor
Need to relate the Reynolds Stress Tensor to the mean flow
Know as the turbulence closure problem
No right way to do this, but some approximations are better than others
The Reynolds Stress Tensor represents the transport of momentum due to the fluctuations in the flow
In the standard Navier-Stokes equation the viscous stress term represents the “diffusion” of momentum
We can therefore think of Reynolds Stress Tensor in terms of a diffusion of momentum due to velocity fluctuations – a turbulent eddy viscosity
Turbulent Eddy Viscosity#
Write the Reynolds Stress Tensor in terms of the macroscopic strain rate:
This means that you can combine the liquid viscosity and the turbulent losses into a single effective viscosity:
In turbulence modelling it is often more convenient to use the kinematic viscosity
We now need to calculate this turbulent eddy viscosity
Zero Equation Models#
The simplest type of turbulence model
Assume no explicit transport of turbulence
Mainly based on dimensional arguments
\(l_0\) is a turbulent length scale, \(t_0\) is a turbulent time scale
Zero equation models are not very good in steady state RANS type models where the turbulent length scale is not well defined
Often adequate in LES models where the turbulent structures are resolved down to the grid resolution
Unresolved turbulent length scale similar to mesh resolution
Less transport of small scale turbulence before it is dissipated
If we assume that we know the turbulent length scale, we still need the turbulent time scale from some macroscopic flow property
Two obvious candidates
Magnitude of the strain rate or magnitude of the vorticity
Both have units of inverse time
Based on strain rate – Smagorinsky mode
\(C_S\) is a tuneable constant. Typical values are 0.1-0.2
Based on vorticity – Baldvin-Lomaz model
Less widely used than Smagorinsky – Use in modelling boundary layers
In boundary layers 𝑘 is a function of distance from the wall
Two Equation Model#
We will revisit zero equation models in the context of LES
For RANS modelling zero equation models are generally inadequate
Need to consider the generation, transport and dissipation of turbulence
Two quantities can be used to model this
Turbulent kinetic energy - 𝑘
Turbulent dissipation rate - 𝜀
So called 𝑘 − 𝜀 turbulence model
Other models exist, but this is one of the most widely used and illustrates the idea
Turbulent Kinetic Energy - \(k\)#
The fluctuating components of the turbulence stores energy
This is known as the turbulent kinetic energy, 𝑘, which, per mass of fluid, can be written as follows:
Note that the turbulent kinetic energy is therefore closely related to the Root Mean Square (RMS) of the velocity distribution
This turbulent kinetic energy is generated at the large scale and dissipated at the small scale
It can also be transported around the system by the mean flow
Turbulent Dissipation Rate - \(\varepsilon \)#
This is equal to the viscous losses associated with the fluctuating component of the velocity
I.e. the rate at which the turbulent kinetic energy is converted into thermal energy
While this is how the energy is dissipated, this equation is not directly useful as we do not know the fluctuation components (or their gradients)
Equations for \(k\) and \(\varepsilon\)#
Both the turbulent kinetic energy (𝑘) and the dissipation rate (𝜀) can be transported, generated and destroyed and are represented by a pair of coupled ODEs
Don’t worry too much about how these are derived – note, though, the additional complexity associated with modelling turbulence
The semi-empirical constants are typically assigned the following values:
Boundaries in Turbulence Modelling#
At inlets both 𝑘 and 𝜀 need to be assigned values
If appropriate values are not assigned, long inlet regions will be required
At walls 𝑘 is zero (no slip), but the transition typically occurs over a very narrow boundary layer
Typically need to model this boundary layer as it is usually computationally very expensive to explicitly resolve
Large Eddy Simulations - LES#
Transient simulations that resolve as much of the dynamics as possible
Zero equation models are usually appropriate
No time averaging or smoothing of equations above the grid resolution
If unresolved resolution is small then the assumption of no turbulent transport between generation and dissipation is appropriate
Can use Smagorinsky model with length scale as the grid resolution
RANS vs LES#
Other Rheologies#
Thus far we have only considered Newtonian rheology
Shear stress proportional to the strain rate
A good approximation for many fluids, importantly including water and virtually all gases
Other Rheologies are possible
Shear stress a more complex function of strain rate
Can also be a function of strain history
Thixotropic fluids
Shear thinning#
Apparent viscosity decreases with strain rate
Quite common
Dense suspensions, emulsions, foams etc…
Sometimes called pseudo-plastic fluids
Some physical origins:
Structures in the fluid can break down at high shears
E.g. some suspensions where particles can form flocs
Molecules align with or are extended in the shear direction
E.g. molten plastics
Particles in a fluid align with the flow and so move past one another more easily
Yield Stress#
An extreme example of shear thinning are materials with a yield stress
Below a yield stress a material does not flow and acts like a solid
Above the yield stress the material flows and acts like a liquid
Quite common
It is really a matter of the size of the yield stress
Logical extension of plastic deformation in a solid
Shear thickening#
Apparent viscosity increases with strain rate
Much less common
Classic example is corn starch in water
Some physical origins:
Structures in the fluid jam and can’t pass one another as easily at high strain rates
Lots of current interest in shear-thickening fluids
Flexible protection and armouring
Modelling non-Newtonian Rheologies#
Need to incorporate these rheologies into the Navier-Stokes equation
Require a relationship between shear stress and strain rate tensors
Often easier to incorporate rheology as an apparent viscosity that depends on strain rate
Note that in most solvers it is the velocity (and thus strain rate) that is assumed to be known
Power law fluid#
The simplest model for both shear thinning and shear thickening
The separation of the strain rate into two terms is so that stress tensor points in the correct direction
if \(n=1\) this is Newtonian
if \(n>1\) this is shear thickening
if \(n<1\) this is shear thinning
If the velocity is in the y direction and only changes in the x direction
Bingham Plastic#
The simplest type of yield behaviour is to assume that there is a yield stress followed by a linearly increasing stress with increasing strain rate
If the velocity is in the y direction and only changes in the x direction
Free Surface Flow on an Inclined Plane#
This is the problem that we solved earlier for a Newtonian fluid
Let us consider a Bingham plastic with a yield stress \(𝜏_0\)
Calculation :
The shear stress distribution in this problem is the same as that for the Newtonian fluid and can be obtained in the same manner:
Note that again 𝑥 is defined as the distance below the free surface
As the 𝑥𝑧 component is the only non-zero shear stress:
If we assume that the geometry is defined such that 𝜌𝑔cos 𝛽 is a positive number, then \(|𝜏_{𝑥𝑧}| = − 𝜏_{𝑥𝑧} \text{and} |\frac{du_z}{dx}| = -\frac{du_z}{dx} \) (note that if \(𝜏_{𝑥𝑧}\) was positive, which would have been the case if we had defined 𝑥 as being the distance above the bottom of the flow, then \(𝜏_{𝑥𝑧} = 𝜏_{𝑥𝑧} \text{and} |\frac{du_z}{dx}| = -\frac{du_z}{dx} \)
We need to do this in two regions
A deep region when \(-𝜏_{𝑥𝑧} > 𝜏_0\) and a shallow region where \(-𝜏_{𝑥𝑧} < 𝜏_0\)
Let \(𝑥_0\) be the depth below the liquid surface of this transition:
In the deep region (\( x > x_0\)):
Which can be integrated to give:
B can be onbtained from no slip at the bottom ( \(v_z = 0 ~~ \text{at} ~~ x = h\))
The velocity in the top portion is constant and can be obtained by setting \(x = x_0\):
Yielding following overall velocity profile:
Where: \(x_0 = \frac{\tau_0}{\rho g \cos(\beta)}\)