You are on page 1of 26

Fundamentals of Magnetohydrodynamics

Takehiko Yokomine (Liquid metal MHD)

P. A. Davidson, “An Introduction to Magnetohydrodynamics,”


Cambridge texts in applied mathematics, Cambridge University Press, 2001

I‘m Prof. Yokomine in charge of the lecture.


This lecture is based on the contents of this book. If you want to know more,
please refer to this book.

1
Magnetohydrodynamics is the academic discipline which studies the dynamics of electrically
conducting fluids. The word magnetohydrodynamics is derived from magneto- meaning magnetic
field, and hydro- meaning liquid, and -dynamics meaning movement.

The idea of magnetohydrodynamics is that magnetic fields can induce currents in a moving conductive
fluid, which create forces on the fluid, and also change the magnetic field itself. The set of equations
which describe magnetohydrodynamics are a combination of the Navier-Stokes equations of fluid
dynamics and Maxwell's equations of electromagnetism. These differential equations have to be
solved simultaneously, either analytically or numerically. Because magnetohydrodynamics is a fluid
theory, it cannot treat kinetic phenomena, i.e., those in which the existence of discrete particles, or of a
non-thermal distribution of their velocities, is important.

The fluid in question must be electrically conducting and non-magnetic, which limits us to liquid metals,
hot ionized gases (plasmas) and strong electrolytes (salt water, KOH solution).

The mutual interaction of a magnetic field, B, and a velocity field, u, arises partially as a result of the
laws of Faraday and Ampère, and partially because of the Lorentz force experienced by a current-
carrying body.

Magnetohydrodynamics is the academic discipline which studies the dynamics of


electrically conducting fluids. The word magnetohydrodynamics is derived from
magneto- meaning magnetic field, and hydro- meaning liquid, and -dynamics
meaning movement

The idea of magnetohydrodynamics is that magnetic fields can induce currents in


a moving conductive fluid, which create forces on the fluid, and also change the
magnetic field itself. The set of equations which describe MHD are a combination
of the Navier-Stokes equations of fluid dynamics and Maxwell's equations of
electromagnetism. These differential equations have to be solved simultaneously,
either analytically or numerically. Because magnetohydrodynamics is a fluid
theory, it cannot treat kinetic phenomena, that is, those in which the existence of
discrete particles, or of a non-thermal distribution of their velocities, is important.

The fluid in question must be electrically conducting and non-magnetic, which


limits us to liquid metals, hot ionized gases, say plasmas, and strong electrolytes,
such as salt water and KOH solution.

The mutual interaction of a magnetic field, B vector, and a velocity field, u vector,
arises partially as a result of the laws of Faraday and Ampère, and partially
because of the Lorentz force experienced by a current-carrying body.

2
Classification of ”電磁流体力学”

Compressible Plasma

Electromagnetic
electric
fluid MHD
conductive

Incompressible
FHD

Non‐conductive

EHD

MHD: Magneto‐hydro‐dynamics
FHD: Ferro‐hydro‐dynamics (磁性流体力学)
EHD: Electro‐hydro‐dynamics (電気粘性流体力学)

In general, the dynamics of electromagnetic fluid, that is “電磁流体” in Japanese,


can be broadly classified into four categories.
Plasma is a dynamics of compressible electromagnetic fluid that was learned in
Murakami-sensei's lecture.
MHD, which is abbreviation for magnetohydrodynamics is primarily a dynamics
dealing with incompressible conductive fluids.
In addition, there are Ferrohydrodynamics, for short “FHD” that treats magnetic
fluids, “磁性流体” in Japanese and Electrohydrodynamics, for short “EHD” that
handles electrorheologicalfluids, “ER流体” in Japanese, both of which are being
developed as multifunctional fluids.
Electromagnetic fluid dynamics in Japanese “電磁流体力学” in the broad sense
includes all these.

3
Hallmark of MHD effects
(1) The relative movement of conducting fluid and a magnetic field causes an emf
(of order |uXB|) to develop in accordance with Faraday’s law of induction.
(emf = electromotive force)
Wire loop
B In general, electrical
currents will ensue, the
current density being of
order of J=(uXB),  is the
J pulling electrical conductivity

JxB
b
(2) These induced currents, J must, according to Ampère’s law, give rise to the
second, induced magnetic field, b. This adds to the original magnetic field and
the change is usually such that the fluid appears to ‘drag’ the magnetic field
lines along with it.

(3) The combined magnetic field (imposed B + induced b) interacts with the
induced current density to give rise to a Lorentz’s force (per unit volume), JXB.
This acts on the conductor and is generally directed so as to inhibit the relative
movement of magnetic field and the fluid.

Let‘s briefly review the MHD effect here.

The relative movement of conducting fluid and a magnetic field causes an


electromotive force, emf, of which order |uXB|, to develop in accordance with
Faraday’s law of induction.
In general, electrical currents will ensue, the current density being of order of J=σ
(uXB), where σ is the electrical conductivity

These induced currents, J must, according to Ampère’s law, give rise to the
second, induced magnetic field, b. This adds to the original magnetic field and the
change is usually such that the fluid appears to ‘drag’ the magnetic field lines
along with it.

The combined magnetic field, that is, imposed B plus induced b, interacts with
the induced current density to give rise to a Lorentz’s force per unit volume, JXB.
This acts on the conductor and is generally directed so as to inhibit the relative
movement of magnetic field and the fluid.

4
Importance of characteristic length
Induced current density
Imposed magnetic field depends on the product of
(1) Typical velocity of the motion, u j   uB
A modest current density spread over a large
(2) Conductivity of the fluid,  and area can produce a high magnetic field, whereas
(3) Characteristic length, l the same current density spread over a small
area induces only a weak magnetic field.

ul determines the ratio of the induced field to the applied field:

ul → ∞ (Ideal conductors), Imposed magnetic field ~ Induced magnetic field


“Freezing together” of magnetic field and medium

Astrophysical MHD: High conductivity of plasmas, vast characteristic length

ul → 0, Imposed magnetic field remains relatively unperturbed.

Liquid metal MHD: u leaving B unperturbed

Liquid metals:
 : ~ 106 -1・m-1, u: ~ 1 m・s-1, J: ~ a few A・cm-2, l : ~ 0.1 m [Laboratory scale]
→ Induced magnetic field, b, is usually found to be negligible
by comparison with the imposed magnetic field, B.

Imposed magnetic field B depends on the product of typical velocity of motion, u,


conductivity of the fluid σ, and characteristic length l.
A modest current density spread over a large area can produce a high magnetic
field, whereas the same current density spread over a small area induces only a
weak magnetic field. It is therefore the product σul which determines the ratio of
the induced field to the applied magnetic field.

In the limit as σul tends to infinity, that is in the case of ideal conductors, induced
magnetic field is comparable with imposed magnetic field. This is so-called,
“Freezing together” of magnetic field and medium.
In the case that σul tends to zero, the imposed magnetic field remains relatively
unperturbed. In general, this condition applies to liquid metals, where u leaves B
unperturbed.

The values of each parameters for typical liquid metal are like these. The
products σul is quite small, so that induced magnetic field, b, is usually found to
be negligible by comparison with the imposed magnetic field, B. Nevertheless, it
should be emphasized that an imposed magnetic field can substantially alter the
velocity field of liquid metals.

5
From Electromagnetics to MHD
The only difference between MHD and conventional electrodynamics (ED) lies in the fluidity
of the conductor.

Some important parameters in ED and MHD:

Magnetic Reynolds number: Re m   ul Rem is a dimensionless parameter of


conductivity.
Alfvèn velocity [m・s-1]: va  B 
1 : permeability of free space
Magnetic damping time [s]:    B 2   : electrical conductivity
: density of conducting medium

When Rem is large, the magnetic field lines act rather like elastic bands frozen into the conducting
medium.
Small disturbances of medium tend to result in near-elastic oscillations, with the
magnetic field providing the restoring force for vibration [dissipation process]. In a fluid,
this results in Alfvèn waves, which turns out to have a frequency of ~ va/l.

When Rem is small, u has little influence on B, the induced magnetic is negligible.
The kinetic energy converts into heat via Joule dissipation. The relevant time scale is the damping
time, , rather than l/va.

The only difference between MHD and conventional electrodynamics lies in the
fluidity of the conductor.
I show some important parameters in ED and MHD, that is, Magnetic Reynolds
number, Alfvèn velocity and Magnetic damping time. The μ is the permeability of
free space, σ is the electrical conductivity and ρ is the density of conducting
medium. Magnetic Reynolds number is a dimensionless parameter of
conductivity.

When Magnetic Reynolds number is large, the magnetic field lines act rather like
elastic bands frozen into the conducting medium. Namely, small disturbances of
medium tend to result in near-elastic oscillations, with the magnetic field providing
the restoring force for vibration. So that this is the dissipation process. In a fluid,
this results in Alfvèn waves, which turns out to have a frequency of ω ~ va/l.

On the other hand, when Rem is small, u has little influence on B and the
induced magnetic is negligible. The kinetic energy converts into heat via Joule
dissipation. In this case, the relevant time scale is the magnetic damping time, τ,
rather than l/va.

6
• Reynolds number • Magnetic Reynolds number
UL UL UL
Re  Rem  
 m 1 
inertia term v v
  advection term v B
viscous term  2 v  
diffusion term  m 2 B
in Navier‐Stokes Eq. in Induction Eq.

• Mach number • Magnetic Mach number


U U
M Mm 
a UA
local flow velocity flow velocity
 
speed of sound Alfvèn velocity

Since the dimensionless number Rem has appeared, for reference, I show a
comparison with the dimensionless number between fluid mechanics and MHD.

7
Astrophysics v.s. Liquid‐metal
In liquid metal MHD, the primary role of B is to
dissipate mechanical energy on time scale of .

As for Rem «1 case, Ohm’s law plays critical role.


This means that the high resistivity of the circuit,
i.e., the current and hence induced field are small.
We consider B ~ constant (imposed field: B0).
The electric field must be irrotational

B 0
E   0
t
Alfvén waves:
A magnetic field behaves like a Ohm’s law and the Lorentz force per unit volume
plucked string, transmitting a simplify to
J      u  B 0 
transverse inertial wave with a phase
velocity of va.
As for high Rem case, Faraday’s law
plays critical role. This is a common F  J  B0
feature of astrophysical MHD.
 is the electrostatic potential.)

As for high Rem case, Faraday’s law plays critical role, where a magnetic field
behaves like a plucked string, transmitting a transverse inertial wave with a phase
velocity of va.
This is Alfvén waves, which is a common feature of astrophysical MHD.

In liquid metal MHD, the primary role of B is to dissipate mechanical energy on


time scale of τ as shown in previous page.
As for case Rem is much lea than 1, Ohm’s law plays critical role. This means
that the high resistivity of the circuit, that is, the current and hence induced field
are small. Now, we consider imposed filed B0, that equals to B, is nearly constant.
Because the electric field must be irrotational, the following equation is derived.
Then, Ohm’s law and the Lorentz force per unit volume simplify to these
equations.

8
Johann Carl Friedrich Gauss Michael Faraday James Clerk Maxwell
(1777~1855) (1791~1867) (1831~1879)
Charles-Augustin de Coulomb
(1736~1806)

Georg Simon Ohm Jean-Baptiste Biot Hendrik Antoon Lorentz


André-Marie Ampère
(1789~1854) (1774~ 1862) (1853~1928)
(1775~1836)
No Photo: Félix Savart (1791~1841)

Governing Equations of Electrodynamics

From here, let’s go over the governing equations of electrodynamics and fluid
dynamics.

9
Electric Field and Lorentz Force
A particle moving with velocity u and carrying a charge q is, in general, subject to
three electromagnetic forces:

f  qE s  qEi  qu  B
Es : Electrostatic field
Ei : Induced Electric field (Induced by the change of magnetic field)
 The first term is the electrostatic force.
 The second term is the force which the charge experiences in the presence of
a time-varying magnetic field.
 The third term is the Lorentz force.
And
e
  Es  (Gauss’s law)   Ei  0 (zero divergence)
0
B
  E s  0 (Coulomb’s law)   Ei   (Faraday’s law)
t
0 is permittivity of free space
It is convenient to define the total electric field as: E  E s  Ei

f  q (E  u  B )

A particle moving with velocity u and carrying a charge q is, in general, subject to
three electromagnetic forces as follows, where Es is the electrostatic field and Ei
is the induced electric field which induced by the change of magnetic field.
The first term of this equation reveals the electrostatic force. The second term
means the force which the charge experiences in the presence of a time-varying
magnetic field. And the third term is the Lorentz force.

Also, these four equations are important, that is, Gauss’s law, zero divergence,
Coulomb’s law and Faraday’s law, respectively. ε0 is permittivity of free space.
It is convenient to define the total electric field as E=Es+Ei. In that case, the
electromagnetic force can be rewritten by using E.

10
Ohm’s Law
Ohm’s Law
J   E  u  B
In MHD, we are less concerned with the forces on individual charges than the bulk force
acting on the medium.
Charge Density e   q
Current Density J  qu :summed over a unit volume
Force per unit volume: F   e E  J  B
The first term in this equation is negligible: F  J  B
Why it can be neglected?
e
J   (Conservation of Charge)
t
By taking the divergence of both side of Ohm’s law, and using Gauss’s law   Ei  0 , we find
e e 0
     u  B   0  e  :Charge Relaxation Time (around 10-18[s] for a typical conductor)
t  e 
Here, we are interested in events which take  e
Neglect e by comparison with  e   0    u  B 
place on a time-scale much longer than e. t e
Pseudo-static eq.
u e
We have e ~  0 uB l , while Ohm’s law requires E ~ J  . e E ~  0 uB l  J   ~
JB
l
Evidently, since u e l ~1018 , the Lorentz force completely dominates above equation. (and   J  0 )

This is well-known Ohm’s law.


In liquid metal MHD, we are less concerned with the forces on individual charges
than the bulk force acting on the medium. So that, The force per unit volume can
be written as F=ρeE+J×B. The first term in this equation is negligible, so F can be
rewritten as F=J×B.

Check the bottom half of the page for details on why the first term can be ignored.

11
Ampère’s Law
The Ampère-Maxwell Equation
 E  rhs
 B   J  0 
 t 
・     0 and Gauss' Law,
Displacement Current ( introduced by Maxwell ) this yields
 
In MHD, this correction is not needed. Because: ・J   0  ・E   e  0
t t
e E
 0 and  0 is also small in conductor.
t t
E  0 J J Differential form of
0
t
~
 t
~ e
t
 J (by Ohm’s law)  B  J Ampère’s law

Biot-Savart Law
 J  x   r
B x   d 3 x r  x  x
4 r3

This is equivalent to ∇×B = μJ , reveals the true character


of Ampère’s law.

The Ampère-Maxwell Equation is as follows. Here, the red underlined part


represents the displacement current which has been introduced by Maxwell.
However, in the case of liquid metal MHD, the time derivative of the charge
density is zero and displacement current is also quite smaller than the induced
current density J. So that, the Ampère-Maxwell equation with the correction with
the displacement current is simplified to the differential form of Ampère’s law

12
Reduced Form of Maxwell’s Equation for MHD
Maxwell’s Equation:
  E  e  0 (Gauss’ law)
B  0 (Solenoidal nature of B)   J  e t (charge conservation)
B
E   (Faraday’s law in differential form) F  q  E  u  B  (Lorentz force)
t
 E 
  B    J  0 (Ampère-Maxwell equation)
 t 
In MHD:  ρe plays no significant part.
 The contribution of ∂e/∂t to the charge conservation equation is negligible.
 The displacement currents are negligible.

Ampère’s law plus charge conservation   B  J J  0

B
Faraday’s law plus the solenoidal constraint on B E   B  0
t

Ohm’s law plus the Lorentz Force J   E  u  B F  JB

Although electrostatic forces are of no importance in MHD, they can lead to some unexpected
effects in those cases where they are significant.

This is the reduced form of Maxwell’s equations for liquid metal MHD that
summarizes the above discussion.
Although electrostatic forces are of no importance in liquid metal MHD, they can
lead to some unexpected effects in those cases where they are significant.

13
Claude Louis Marie Henri Navier
Leonhard Euler (1707~1783) (1785~1836) Sir George Gabriel Stokes
(1819~1903)

Osborne Reynolds (1842~1912) Ludwig Prandtl (1875~1953) Theodore von Kármán (1881~1963)

Governing Equations of Fluid Mechanics

14
Different Categories of Fluid Flow
The beginner in fluid mechanics is often bewildered by the many diverse categories of
fluid flow.
Potential flow
Boundary layers These subjects rarely exist in isolation,
Turbulence but rather interact in some complex way.
Vortex dynamics and so on.

Purpose of this part is to give some indication as to what expressions such as boundary
layers, turbulence and vorticity mean, how these subjects interact, and when they are
likely to be important in practice.

There are three very broad sub-divisions in the subject.

 The first relates to the issue of when a fluid may be treated as inviscid, and when the finite
viscosity possessed by all fluids (viscous fluids) must be taken into account.
 There is the sub-division between laminar flow and turbulent flow.
 The sub-division which occurs in fluid mechanics is between irrotational flow (∇×u=0 )
and rotational flow.

The beginner in fluid mechanics is often bewildered by the many diverse


categories of fluid flow, say potential flow, boundary layers, turbulence, vortex
dynamics and so on. These subjects rarely exist in isolation, but rather interact in
some complex way.
Purpose of this part is to give some indication as to what expressions such as
boundary layers, turbulence and vorticity mean, how these subjects interact, and
when they are likely to be important in practice.
There are three very broad sub-divisions in the subject.
The first relates to the issue of when a fluid may be treated as inviscid, and when
the finite viscosity possessed by all fluids (viscous fluids) must be taken into
account.
There is the sub-division between laminar flow and turbulent flow.
The sub-division which occurs in fluid mechanics is between irrotational flow
(∇×u=0 ) and rotational flow.

15
Quantify “Inertia”
Velocity Field u(x,t)
Steady flow: u is a function of x but not of t.
The speed of the fluid at any one point in space is
steady, the flow pattern does not change with time,
and the streamline represents particle trajectories R
for individual fluid “lumps”.

Let s be a curvilinear coordinate measured along C (a particular streamline), and


V(s) be the speed |u|.

dV V2
(acceleration of lump)  V eˆ t  eˆ n
ds R

R :curvature of the streamline


eˆ t , eˆ n :unit vector tangential and normal to the streamline

In general, the acceleration of a typical fluid element is of order |u2|/l, where l is a


characteristic length scale of the flow pattern.

First, let’s quantify “Inertia”.


We consider the velocity field u(x,t). Off course, the steady flow means that u is a
function of x but not of t. In this case, the speed of the fluid at any one point in
space is steady, the flow pattern does not change with time, and the streamline
represents particle trajectories for individual fluid “lumps”.
Let s be a curvilinear coordinate measured along C, which is a particular
streamline, and V(s) be the speed |u|. Then, the acceleration of lump equals to
the such RHS, where R is the curvature of the streamline and these two vectors
are unit vector tangential and normal to the streamline, respectively.
In general, the acceleration of a typical fluid element is of order |u^2|/l, where l is
a characteristic length scale of the flow pattern.

16
Quantify “Shear Stress”
In most fluids, this is quantified using an empirical
law known as Newton’s law of viscosity.
Newton’s law of viscosity says that a shear
stress,  , is directly proportional to angular
distortion rate, d  dt .
It is clear from the diagram that d  dt  du x dy.
u x
   (    )
y
In two-dimensional flow:
A glance at a right figure will confirm that the
additional contribution to d  dt is du y dx .
 u x u y 
 xy     
 y x 

Net horizontal shear force  xy  2ux


per unit volume fx   
y y 2

Next let’s quantify the “shear stress”.


In most fluids, the shear stress is quantified using an empirical law known as
Newton’s law of viscosity.
Newton’s law of viscosity says that a shear stress, τ , is directly proportional to
angular distortion rate, dγ/dt.
See the figure above on the right. It is clear from the diagram that dγ/dt=dux/dt.
So that, the shear stress can be expressed as follows.

In two-dimensional flow as shown in figure below on the right. A glance at the


figure will confirm that the additional contribution to dγ/dt is duy/dt, then the shear
stress in two-dimensional can be represented as follows. The net horizontal
shear force per unit volume, fx, can be expressed in the form of the second
derivative of velocity component in x-direction, ux.

17
Relative Sizes of Inertial and Viscous Forces
Viscous forces per unit volume are of the form of gradients in shear stress (  xy y ).
f ~  u l
2
l : Characteristic length scale normal to the streamlines

Inertial forces per unit volume are of the order of


fin ~   (acceleration) ~  u 2 l l : Typical geometric length scale

Ratio of the two is of order (=Inertial force/viscous force)


ul
Re  Re : Reynolds number

When Re is calculated using some characteristic geometric length scale, it is almost always very large.
( The viscosity of nearly all common fluids, include liquid metal, of the order of 10-6 m2/s. )
Because of the large size of Re, the viscosity effect seems to be not important.
However, this is extremely dangerous.

For example, inviscid theory predicts that a sphere sitting in a uniform cross-flow experiences no
drag (d’Alembert’s paradox) and this is clearly not the case, even for ‘thin’ fluids like air.

And more…

Let’s consider the relative sizes of inertial force and viscous force.

Viscous forces per unit volume are of the form of gradients in shear stress.
Therefore, the viscous force is the order of RHS according to the previous page,
where l⊥ is the characteristic length scale normal to the streamlines.
On the other hand, the inertial forces per unit volume are of the order of RHS,
where l is the typical geometric length scale.
Ratio of the two forces is represented as Reynolds number, Re.
When Re is calculated using some characteristic geometric length scale, it is
almost always very large. For example, the viscosity of nearly all common fluids,
include liquid metal, of the order of 10^-6 m^2/s.
Because of the large size of Re, the viscosity effect seems to be not important.
However, this simple conclusion is extremely dangerous.
For example, inviscid theory predicts that a sphere sitting in a uniform cross-flow
experiences no drag, which is so-called d’Alembert’s paradox, and this is clearly
not the case, even for ‘thin’ fluids like air.

18
Boundary Layers
No matter how small  might be, there are always regions near surfaces where the shear
stresses are significant, i.e. of the order of the inertial forces. (Boundary Layer)

Consider the flow over an aerofoil.


(Re for such a flow, based on the width of the
aerofoil, will be very large: Re ~ 108.)

• Away from the surface of the aerofoil, the fluid behaves as if it is inviscid.
• Close to the aerofoil, however, something else happens, and this is a direct result of a
boundary condition called the no-slip condition. (The fluid ‘sticks’ to the surface.)
There must be some transition region; from its free-stream value to zero on the surface.

Boundary Layer Thickness


Within the boundary layer, there must be some force. (Acting on the fluid which pulled the velocity down)
Viscous Force
Within the boundary layer the inertial ( u 2 l ) and viscous forces (  u  2 ) must be of similar order.
 l ~  ul  
1 2
 1  : Boundary layer thickness
Often (but not always) a high-Re flow may be divided into external, invisid flow plus one or more boundary layers.
(introduced by Prandtl in 1904)

Here , I will explain the boundary layer, briefly.


No matter how small dynamic viscosity  might be, there are always regions near surfaces where the shear
stresses are significant, that is , of the order of the inertial forces. This region is “Boundary Layer”.
Let’s consider the flow over an aerofoil like the right figure. The Reynolds number for such a flow, based
on the width of the aerofoil, will be very large, say Re ~ 10^8.
Away from the surface of the aerofoil, the fluid behaves as if it is inviscid. Close to the aerofoil, however,
something else happens, and this is a direct result of a boundary condition called the no-slip condition, in
which the fluid ‘sticks’ to the surface.
There must be some transition region from its free-stream value to zero on the surface.

Within the boundary layer, there must be some force acting on the fluid which pulls the velocity down
from the free-stream value to zero at the surface. The force which does this is the viscous force, and so
within the boundary layer the inertial and viscous forces must be of similar order.
Thus we see that, no matter how small we make ν , there is always some thin boundary layer in which
shear stresses are important. This is why an aerofoil experiences drag even when ν is very small.
This idea was introduced by Prandtl in 1904 and works well for external flow over bodies, particularly
streamlined bodies, but can lead to problems in confined flows. It is true that boundary layers from at the
boundaries in confined flows, and that shear stresses are usually large within the boundary layers and weak
outside the boundary layers. However, the small but finite shear in the bulk of a confined fluid can, over
long periods of time, have a profound influence on the overall flow pattern.

19
Laminar Flow and Turbulent Flow
It is an empirical
observation that at low
values of Re flows are Laminar flow
laminar, while at high
values of Re they are
turbulent.
This was first demonstrated in 1883
by O. Reynolds, who studied flow in
a pipe.


Transient flows


The transition from laminar to
turbulent flow occurs because, at a

certain value of Re, instabilities
develop in the laminar flow,
usually driven by the inertial force.
Turbulent flow

It is an empirical observation that at low values of Reynolds number flows are


laminar, while at high values of Reynolds number they are turbulent.
This was first demonstrated in 1883 by O. Reynolds, who studied flow in a pipe.
The transition from laminar to turbulent flow occurs because, at a certain value of
Re, instabilities develop in the laminar flow, usually driven by the inertial force.

20
If the front of the plate is stream-lined, and the
turbulence level in the external flow is low, the Flow visualization
cylinder
boundary layer usually starts off as laminar.

Often periodic (non-turbulent) fluctuations in the laminar flow


precede the onset of turbulence.

 At low values of Re (<1) : called “Creeping flow.”


 Re ~100 : These vortices start to peel off from the rear of the
cylinder in a regular, periodic manner. (called
“Karman’s vortex”)
 At a value of Re ~105, the flow at the rear of the cylinder
loses its periodic structure and becomes a turbulent wake.

https://www.youtube.com/watch?v=e1TbkLIDWys

If the front of the plate is stream-lined, and the turbulence level in the external
flow is low, the boundary layer usually starts off as laminar as shown in above
figure.
Often periodic, non-turbulent, fluctuations in the laminar flow precede the onset of
turbulence.
At low values of Re, that is, less than <1, the flow is called “Creeping flow.”
For Re is nearly 100, these vortices start to peel off from the rear of the cylinder
in a regular, periodic manner, which is called “Karman’s vortex”.
At a value of Re ~10^5, the flow at the rear of the cylinder loses its periodic
structure and becomes a turbulent wake.
A reference video has been uploaded to youtube.

21
Potential Flow and Vortical Flow
Potential Flow
There is a boundary layer, which is filled
with vorticity, and an external flow.
The flow upstream of the aerofoil is assumed to be
irrotational (free of vorticity), and since the vorticity
generated on the surface of the foil is confined to the
boundary layer, the entire external flow is irrotational.
The problem of computing the external motion is Easy to compute but
now reduced to solving two kinematic equations:   u  0   u  0 infrequent in nature
Vortical flows
Almost all real flows are laden with vorticity.
(The blood in our veins, the air in our lungs,
the wind blowing down the street, natural
convection in a room, and the flows in the
oceans and atmosphere…)
Very common but much more difficult to understand movie
In MHD
The Lorentz force generates vorticity in the interior of the fluid.
This makes MHD more difficult to understand, but on the other hand it makes it more attractive.
In MHD we have the opportunity to grab hold of the interior of a fluid and manipulate the flow.

The potential flow describes the velocity field as the gradient of a scalar function,
that is, the velocity potential. As a result, a potential flow is characterized by an
irrotational velocity field. The irrotationality of a potential flow is due to the curl of
the gradient of a scalar always being equal to zero.
See the right figure. There is a boundary layer, which is filled with vorticity, and an
external flow. The flow upstream of the aerofoil is assumed to be irrotational, and
since the vorticity generated on the surface of the foil is confined to the boundary
layer, the entire external flow is irrotational.
The problem of computing the external motion is now reduced to solving these
two kinematic equations. These equations are easy to compute but infrequent in
nature
Al most all real flows are laden with vorticity. For example, the blood in our veins,
the air in our lungs, the wind blowing down the street, natural convection in a
room, and the flows in the oceans and atmosphere, and so on. Thus, the vortical
flows are very common but much more difficult to understand.
In MHD, the Lorentz force generates vorticity in the interior of the fluid. This
makes MHD more difficult to understand, but on the other hand it makes it more
attractive.
In MHD, we have the opportunity to grab hold of the interior of a fluid and
manipulate the flow.

22
The Navier-Stokes Equation
Newton’s second law applied to a small blob of fluid of
volume δV :
Du
  V     p   V   ij x j   V
Dt
The mass of the element, V, times its acceleration, small blob
Du/Dt, equals the net pressure force acting on the of fluid
surface of the fluid blob.
The last term can be established by considering the forces
acting on a small rectangular element dxdydz.
We take the fluid to be Newtonian, so that the viscous stresses are given by the constitutive law:
 u x u y  Du
 xy      The conventional form of the N.S equation    p     2 u
 y x  Dt
We shall take the fluid to be incompressible so that the conservation of mass, expressed as     u     t
reduces to the so-called continuity equation:   u  0
The expression D/Dt represents the convective derivative.
It is the rate of change of a quantity associated with a given element of fluid.
(  t is the rate of change of a quantity at a fixed point in space.)
D  D 
 f    f    u    f  (for scalar function)  a    a    u   a  (for vector function)
Dt t Dt t

Here, I will introduce the Navier-Stokes equation.


See the figure for reference.
Newton’s second law applied to a small blob of fluid of volume δV.
The mass of the element, ρdV, times its acceleration, Du/Dt, equals the net
pressure force acting on the surface of the fluid blob.
The last term of the above equation can be established by considering the forces
acting on a small rectangular element dxdydz.
Now, we take the fluid to be Newtonian, so that the viscous stresses are given by
the constitutive law.
Thus, the conventional form of the Navier-Stokes equation can be described as
follows.
We shall take the fluid to be incompressible so that the conservation of mass
reduces to the so-called continuity equation, ∇・u=0.

23
Vorticity, Angular Momentum
Vorticity Field: ω    u  z   u y x  u x y 

 The rules governing the evolution of ω are somewhat


simpler than those governing u.
 Many flows are characterized by localized regions of
intense rotation (i.e. Vorticity).
 has nothing at all to do with the global rotation of a fluid.

Introduce some dynamics:


Consider the angular momentum, H, of a small material element that is instantaneously spherical.
1
H Iω I : moment of inertia of the blob
2 T : viscous torque acting
This angular momentum will change at a rate determined by DH on the sphere
 T
the tangential surface stresses (surface shear stress) alone. Dt
Dω DI
Convective derivative satisfies the usual rules of differentiation. I  ω  2 T
D Dt Dt
in cases where viscous stresses are negligible  Iω   0
Dt
  evolves independently of p.
 If  is initially zero, and the flow is inviscid, then  should
remain zero. (Potential flow theory)
 If I decreases in a fluid element (and =0) , the vorticity of
that element should increase.

Vorticity field can be expressed as this equation.


The rules governing the evolution of ω are somewhat simpler than those
governing u. Many flows are characterized by localized regions of intense
rotation.
Vorticity ω has nothing at all to do with the global rotation of a fluid.

Next, consider the angular momentum, H, of a small material element that is


instantaneously spherical.
This angular momentum will change at a rate determined by the tangential
surface alone. Thus, DH/Dt=nT, where nT is the viscous torque acting on the
sphere.
Convective derivative satisfies the usual rules of differentiation.
In cases where viscous stresses are negligible, the above equation can be
rewritten.
It should be noted that, the ω evolves independently of pressure. In addition, if ω
is initially zero, and the flow is inviscid, then ω should remain zero. This is
“Potential flow theory”.
If the moment of inertia decreases in a fluid element and n equals to zero, the
vorticity of that element should be increased.

24
Vorticity Equation
Navier-Stokes Equation  
 u 2 2   u    u    u   u    u  u  ω
u u
t
  u    u    p     2 u
t
 
 u  ω   u    u   p   u 2 2   2 u

ω (Bernoulli’s Function)
take the curl     u  ω    2 ω
t
( and B obey precisely the same evolution equation.)
u and  are both solenoidal, we have the vector relationship:   u  ω    ω   u   u    ω

  ω    u   2 ω
Dt

The terms on the right of this equation represent:


 Change in the moment of inertia of a fluid element due to stretching of that element
 Viscous torque on the element
In other words, the rate of rotation of a fluid blob may increase or decrease due to changes in its
moment of inertia, or change because it is spun up or slowed down by viscous stresses.

Here, the Vorticity equation is derived from the Navier-Stokes equation, so follow
the derivation process.
The terms on the right of this vorticity equation represents the change in the
moment of inertia of a fluid element due to stretching of that element and the
viscous torque on the element, respectively.
In other words, the rate of rotation of a fluid blob may increase or decrease due
to changes in its moment of inertia, or change because it is spun up or slowed
down by viscous stresses.

25
Advection and diffusion equation
Advection-diffusion equation for vorticity:
Advection-diffusion equation is so fundamental to MHD.
Consider two-dimensional flow: D z
The first term on the right of previous the equation now vanishes to yield   2 z
Dt
DT
Compare this with the equation governing the temperature, T, in a liquid.   2T
Dt
( : thermal diffusivity)
advection-diffusion equation
T T (for heat)
To illustrate the combined effect of these processes. u=(ux,0,0)  ux    2T
t x
Suppose that heat is injected into the fluid from a hot wire.
Hot wire Hot wire
 When the velocity is low and the conductivity high,
the isotherms around the wire will be almost
circular.
 When the conductivity is low, however, each fluid
element will tend to conserve its heat as it moves.
l : characteristic
The relative size of the advection to the diffusion term is given by the Peclet number: Pe  ul 
length scale
If the Peclet number is small, then the transfer of heat is diffusion-dominated.
When it is large, advection dominates.
The vorticity in a two-dimensional flow is advected and diffused just like heat.
Vorticity is advected by u and diffused by the viscous stresses.

Advection-diffusion equation for vorticity is so fundamental to MHD.


Now, consider two-dimensional flow.
The first term on the right of the vorticity equation appeared in previous page
vanishes to yield the following equation.
Compare this with the equation governing the temperature, T, in a liquid, that is
advection-diffusion equation for heat.
To illustrate the combined effect of these processes. u=(ux,0,0)
Suppose that heat is injected into the fluid from a hot wire.
When the velocity is low and the conductivity high, the isotherms around the wire
will be almost circular like left figure.
When the conductivity is low, however, each fluid element will tend to conserve
its heat as it moves like right figure.
The relative size of the advection to the diffusion term is given by the Peclet
number.
If the Peclet number is small, then the transfer of heat is diffusion-dominated.
When Peclet number is large, advection dominates.
The vorticity in a two-dimensional flow is advected and diffused just like heat.
That is, vorticity is advected by u and diffused by the viscous stresses.

26

You might also like