Equation of Motion and Vorticity Transport
Equation of Motion and Vorticity Transport
(6.1.2)
where the unit normal vector n points into the parcel exterior. Our next
task is to relate the rate of change of the parcel momentum to the fluid
density and velocity.
(6.1.3)
where u is the fluid velocity. The rate of change of linear momentum is
given by
(6.1.4)
Now, because the integration is over the volume of the parcel which is
not stationary but changes in time, interchanging the time differentiation
and the volume integration on the right-hand side of (6.1.4) is permissible
only if the time derivative is replaced by the material derivative D/Dt
under the integral sign. Thus,
(6.1.5)
Conservation of mass requires that the material derivative of the elemen-
tary mass drrip vanish, yielding the simplified expression
(6.1.6)
(6.1.7)
involving the point particle acceleration, the stress tensor, and the body
force. Explicitly, the x, y, and z components of (6.1.7) are given by
(6.1.8)
Equations (6.1.8) are valid irrespective of whether the fluid is compress-
ible or incompressible.
(6.1.9)
Figure 6.1.1 Illustration of a fluid parcel with an instantaneous rect-
angular shape, drawn with the solid line, in two-dimensional flow.
The fact that the parcel generally deforms to obtain a warped
shape, drawn with the dashed line, does not prevent us from ap-
plying Newton's law of motion in its integral form over the instan-
taneous parcel shape.
(6.1.10)
We note again that these equations are valid irrespective of whether the
fluid is compressible or incompressible.
(6.1.11)
and
(6.1.12)
The first integral on the right-hand side of (6.1.11) involves normal
stresses exerted on the vertical sides, and the second integral involves
shear stresses exerted on the horizontal sides; the converse is true for
(6.1.12).
In the case of steady unidirectional flow along the x axis, the point
particles move along the x axis with constant velocity and vanishing
acceleration, Du/Dt = O. The left-hand side of the equation of motion
(6.1.9) vanishes, yielding a balance between the hydrodynamic and body
force,
(6.1.13)
Restricting our attention to Newtonian fluids and using the consti-
tutive equation (4.5.4), we find the following:
• In the absence of axial and transverse stretching, dux/dx = O and
duy/dy = O, the normal stresses axx and ayy are equal to the
negative of the pressure p, axx = <jyy = —p.
• The shear stresses axy — cryx are independent of streamwise posi-
tion x, but may depend on the lateral position y.
Subject to these simplifications, the balance equations (6.1.11) and
(6.1.12) reduce to
(6.1.14)
where Ax = x% - x\ and Ay = j/2 — J/i- The second of equations (6.1.14)
is clearly satisfied when
(6.1.15)
reflecting the hydrostatic pressure variation. The first of equations (6.1.14)
is satisfied when
(6.1.16)
where G is a free parameter called the modified pressure gradient. Phys-
ically, G is determined by the physical mechanism driving the flow:
1. When G = O, the first of equations (6.1.16) shows that the pres-
sure variation in the x direction is hydrostatic, and the second
of equations (6.1.16) shows that the shear stress ayx is constant,
independent of y. This is the case of shear-driven flow.
(6.2.1)
where V-u = dux/dx+duy/dy+duz/dz is the divergence of the velocity.
If the fluid is incompressible, the second term on the right-hand side of
(6.2.1) is absent.
The x component of the vectorial expression (6.2.1) may be manip-
ulated to give
(6.2.2)
where the time derivative d/dt is taken keeping the spatial position fixed.
Combining the last four terms on the right-hand side of (6.2.2), we find
(6.2.3)
Working in a similar fashion with the y and z components of (6.2.1),
we derive the corresponding expressions
(6.2.4)
and
(6.2.5)
(6.2.7)
where summation of the repeated index j is implied. For example, the
x component of the divergence of M is given by
(6.2.8)
Subject to these definitions, equations (6.2.3) - (6.2.5) are expressed
by the collective form
(6.2.9)
where i = £,y, or z. The corresponding vector form is
(6.2.10)
(6.2.11)
(6.2.12)
where the unit normal vector n points outward from the parcel. In index
notation,
(6.2.13)
where summation is implied over the repeated index j.
(6.2.14)
The four integrals on the left- and right-hand side of (6.2.14) have the
following interpretation:
1. The first integral is the rate of change of the iih component of mo-
mentum of the fluid residing inside the control volume. At steady
state, this term vanishes.
2. The scalar UjUj of the second integrand on the left-hand side of
(6.2.14) is the component of the fluid velocity normal to the bound-
ary of the control volume; accordingly, the corresponding integral
expresses the rate of convective transport of the ith component of
the fluid momentum across the boundary of the control volume.
3. The first integral on the right-hand side of (6.2.14) is the ith com-
ponent of the surface force exerted on the boundary of the control
volume.
4. The second integral on the right-hand side of (6.2.14) is the ith
component of the body force exerted on the control volume.
We proceed now to discuss a particular application illustrating the
usefulness of the integral momentum balance in engineering analysis.
1
6.2.4 Flow through a sudden enlargement
Consider steady flow in a duct through the sudden enlargement, as
illustrated in figure 6.2.1, and identify the control volume with the sec-
tion of the duct confined between the vertical planes labelled 1 and 2.
1
ThIs example, and the one discussed in problem 6.2.1 was borrowed from the
pioneering text by Bird, Stewart, and Lightfoot cited in the bibliography, where a
collection of engineering problems are solved by use of integral mass, momentum, and
energy balances.
PIane l
Plane 2
Assuming that the density of the fluid is uniform and the velocity profile
is flat at the inlet and outlet, neglecting the shear stress at the walls,
approximating the normal stress at the inlet and outlet with the negative
of the pressure, assuming that the pressure at the washer-shaped area is
equal to the inlet pressure, and neglecting the effects of gravity, we find
that the integral momentum balance (6.2.14) at steady state simplifies
to
(6.2.15)
where Si and $2 are the cross-sectional areas of the inlet and outlet.
The three terms on the right-hand side of (6.2.15) are approximations to
the first integral on the right-hand side of (6.2.14) for the outlet, washer-
shaped area, and inlet. Mass conservation requires U\S\ = U^S^ solving
this equation for C/i, and substituting the result into (6.2.15), we find an
expression for the pressure drop
(6.2.16)
which predicts a rise in pressure in agreement with laboratory observa-
tion.
Plane 1 Plane 2
Problem
(6.3.1)
In index notation,
(6.3.2)
S
- V ' a' (6.3.3)
with Cartesian components
(6.3.4)
(6.3.5)
where the parentheses enclose the coordinates of the evaluation point.
Adding the two contributions, and factoring out the common product
AyAz, we obtain
(6.3.6)
Next, we observe that, in the limit as Ax tends to zero, the ratio of
the differences
(6.3.7)
tends to the partial derivative daxi/dx evaluated at the origin. Corre-
spondingly, the difference (6.3.6) reduces to
(6.3.8)
where the derivatives are evaluated at the origin.
Working in a similar fashion with pairs of sides that are perpendicular
to the y or z axis, and summing the contributions, we find that the left-
hand side of (6.3.2) takes the approximate form
(6.3.9)
where the quantity enclosed by the parentheses is evaluated at the origin.
Expression (6.3.9) is an approximation to the volume integral on the
right-hand side of (6.3.2), confirming the identity.
(6.3.10)
(6.3.11)
(6.3.12)
and
(6.3.13)
(6.3.14)
The terms enclosed by the parentheses on the left-hand sides are the
Cartesian components of the point particle acceleration; the right-hand
sides include the Cartesian components of the volume force due to the
hydrodynamic stresses and the components of the body force.
We emphasize again that equations (6.3.10) - (6.3.13) are valid for
both compressible and incompressible fluids.
(6.3.16)
The cylindrical polar components of the equation of motion are
pax = Y>x +pgx, pap = Za +pga, p Q1^ = S^ + p #<p, ( 5 3 ^ 7 )
where ax,aa, and a^ are the cylindrical polar components of the point
particle acceleration given in equations (2.8.10).
Using the alternative expressions (2.8.11), we find
(6.3.18)
Centrifugal force
Coriolis force
The negative of the first term on the right-hand side of the third of
equations (6.3.18), pua u^/'a, is the Coriolis force. This term expresses
an effective force in the (p direction when flow occurs in both the a and
(p directions. The Coriolis force arises, for example, in the flow due to a
spinning circular disk immersed in a tank of fluid.
E = Er er + £0 e# + S^ e<p. ,g g ig.
Using the rules of coordinate transformation and the chain rule of dif-
ferentiation, we find
(6.3.20)
(6.3.21)
where a r ,a#, and a^ are the spherical polar components of the point
particle acceleration given in equations (2.8.13).
Using the coordinate transformation rules and the chain rule of differen-
tiation, we find
(6.3.23)
(6.3.24)
where ar and CIQ are the plane polar components of the point particle
acceleration given in equations (2.8.15).
Alternative expressions are
(6.3.25)
involving, respectively, the centrifugal force and the negative of the Cori-
olis force on the right-hand side.
is the square of the magnitude of the velocity. The third term on the
left-hand side of (6.3.26) is called the vortex force. This force appears
when the vorticity vector is non-parallel to the velocity vector, otherwise
their cross product will vanish. A flow wherein the vorticity vector is
parallel to the velocity vector at every point, is called a Beltrami flow.
Problem
(6.4.1)
The equation of motion (6.3.10) then reduces to the Euler equation
(6.4.2)
The associated Eulerian form is
(6.4.3)
Explicitly, the three Cartesian components of (6.4.3) are given by
(6.4.4)
The cylindrical, spherical, and plane polar components of Euler's
equation follow readily from equations (6.3.17), (6.3.21), and (6.3.24),
by using the constitutive equations (4.7.7), (4.7.14), and (4.7.20) with
vanishing fluid viscosity.
Using the expression for the point particle acceleration shown on the
left-hand side of equation (6.3.26), we find an alternative form of Euler's
equation in terms of the vortex force,
(6.4.5)
where u2 = u^ + Uy + u^ is the square of the magnitude of the velocity.
Substituting this form into (6.4.6), and collecting all terms under the
gradient, we find
(6.4.8)
Since the spatial derivatives of the scalar quanitity enclosed by the
parentheses on the left-hand side vanish, this quantity must be spatially
uniform but possibly time-dependent. On the basis of this argument,
Euler's equation for irrotational flow reduces to Bernoulli's equation for
the irrotational flow of a uniform-density fluid,
(6.4.9)
where c(t) is an unspecified function of time.
(6.4.10)
Combining equations (6.4.9) and (6.4.10), we find
(6.4.11)
which provides us with the rate of change of the potential following a
point particle.
where 7 is the surface tension, and Km is the mean curvature of the free
surface. Applying equation (6.4.11) at a point on the free surface, and
using (6.4.12), we find that the rate of change of the potential following
a point particle is given by
(6.4.13)
Integrating this equation in time, we obtain a boundary condition for
the potential over the free surface.
and the potential energy due to the body force, all three per unit mass
of fluid. Bernoulli's equation requires that the sum of the three energies
be uniform thoughout the domain of flow.
Toricelli's law
Plane 1
U = ^Flgh, (6.4.15)
also describing the velocity of a rigid body in free gravitational fall.
(6.4.16)
Combining these equations to eliminate the pressure, and rearranging,
we find
(6.4.17)
Figure 6.4.4 Irrotational free-surface flow of a horizontal stream over
a hump.
Now, because the perturbation has been assumed small, the actual
velocities u\ and u^ may be replaced with the unperturbed velocities
Ui and C/2 in the numerator and denominator of the fraction on the
right-hand side of (6.4.17), yielding
(6.4.18)
(6.4.19)
which shows that the relative amplitude of the perturbation decays like
the square of the contraction ratio 3%/Si, confirming that the contraction
aids in the establishment of a uniform velocity profile.
(6.4.20)
where UQ is the upstream velocity, and do = d(x = -oo) is the upstream
depth. Combining now the mass conservation equation UQC[Q = u(x)d(x)
with equation (6.4.20) to eliminate u(x), and rearranging the resulting
expression, we derive a cubic equation for the reduced layer depth d(x) =
d(z)/d 0 ,
(6-4.21)
where h(x) = /i(x)/do is the reduced height of the hump. We have
introduced the dimensionless ratio
(6.4.22)
expressing the relative magnitude of inertial and graviational forces,
called the Proude number.
In practice, the Venturi flume is used to deduce the flow rate from
measurements of the deflection of the free surface from the horizontal
position. As the Froude number tends to zero, gravitational forces dom-
inate, and equation (6.4.21) has the obvious solution d(x) = 1 — h(x),
which shows that the free surface tends to become flat. On the other
hand, as the Proude number tends to infinity, inertial forces dominate,
and equation (6.4.21) has the obvious solution d(x) = 1, which shows
that the depth of the stream remains constant, and the free surface fol-
lows the topography of the hump. Intermediate values of the Froude
number yield free-surface profiles with a downward deflection (problem
c.6.4.1).
(6.4.23)
(6.4.24)
(6.4.25)
where the left-hand side is evaluated at the origin. Equation (6.4.25)
states that the rate of change of the quantity enclosed by the parentheses
on the left-hand side with respect to distance along the streamline should
vanish; thus, the quantity enclosed by the parentheses must remain con-
stant along the streamline. An equivalent mathematical statement is
(6.4.26)
Problems
Computer problems
(6.4.29)
for -1 < x < 1, where x = x/a is the reduced distance from the mid-
point. Compute and plot the reduced layer depth d against x for /IQ/do —
0.01, 0.05, 0.10 and Fr = 0.01, 0.1, 10, 100, and discuss the free-surface
shapes.
S = V - a = V - ( - p I + / i 2 E ) = -Vp + /^ 2 V • E, (6.5.1)
where I is the identity matrix.
Working in index notation, we find that the ith component of twice
the divergence of the rate of deformation tensor E on the right-hand side
of (6.5.1) is given by
(6.5.2)
(6.5.3)
S EE V • a = -Vp + n V 2 U.
(6.5.4)
(6.5.5)
(6.5.6)
whose three Cartesian components are:
(6.5.7)
The first term on the right-hand side of (6.5.6), equal to the negative
of the pressure gradient, represents the pressure force. The second term,
equal to the Laplacian of the velocity multiplied by the viscosity, is the
viscous force.
Working in index notation under the assumption that the fluid is
incompressible and thus the velocity field is solenoidal, V • u = O, we
find that the Laplacian of the velocity is equal to the negative of the curl
of the vorticity,
V 2 U = -V x a; (6-5.8)
(problem 6.5.1). An important consequence of this identity is that, if
the flow is irrotational, or if the vorticity is constant, or if the vorticity
field is irrotational, then the viscous force vanishes even though the fluid
may not be inviscid. In this case, the Navier-Stokes equation reduces to
Euler's equation which may be integrated to yield Bernoulli's equation
(6.4.9) for irrotational flow, or equation (6.4.26) for steady rotational
flow.
(6.5.9)
where m(t) is the strength of the point source. The no-penetration con-
dition at the surface of the bubble requires da/dt = ur(r — a), which
may be rearranged to give an expression for the strength of the point
source in terms of the bubble radius,
(6.5.10)
Substituting (6.5.10) into equations (6.5.9), we find
(6.5.11)
Referring now to Bernoulli's equation (6.4.9) for unsteady irrotational
flow, we compute the first and second terms on the left-hand side,
(6.5.12)
Substituting these expressions into Bernoulli's equation (6.4.9), and solv-
ing for the pressure, we find
(6.5.13)
Far from the bubble, the first and second terms on the right-hand side
of (6.5.13) vanish, and the pressure is given by the linear and possibly
time-dependent distribution Poo(x5 *) — P [c(t) + g • x],
The normal stress arr undergoes a jump across the surface of the
bubble, as determined by the bubble radius and surface tension 7. Using
the simplified version of the interfacial condition (4.3.13) for constant
surface tension, we find
(6.5.14)
where ps(t) is the pressure in the interior of the bubble, and Acm = I/a is
the mean curvature of the interface. Substituting the second of (6.5.11)
into the Newtonian constitutive equation orr — -p + 2 p, dur/dr, and
the result into (6.5.14), we find
(6.5.15)
Finally, we apply (6.5.13) at the surface of the bubble, evaluate the pres-
sure from (6.5.15), neglect hydrostatic variations over the diameter of the
bubble, and rearrange the resulting expression to obtain the generalized
Rayleigh equation
(6.5.16)
where x# is the location of the bubble center.
Equation (6.5.16) is a second-order nonlinear ordinary differential
equation governing the evolution of the bubble radius. To compute the
solution, we require the initial bubble radius, the initial rate of expansion
da/dt, and also the bubble pressure and pressure at infinity. The bubble
pressure may be further related to the bubble volume by means of an
appropriate equation of state provided by thermodynamics.
(6.5.17)
The cylindrical polar components of the Navier-Stokes equation arise
by substituting these expressions into the right-hand sides of (6.3.17) or
(6.3.18).
(6.5.19)
The plane polar components of the Navier-Stokes equation arise by sub-
stituting these expressions into the right-hand sides of (6.3.25).
Problems
(6.5.20)
(6.6.1)
Expanding out the derivatives on the left-hand side, and using the conti-
nuity equation dux/dx + duy/dy = O, we obtain the remarkably simpler
form
(6.6.2)
The left-hand side of (6.6.2) expresses the material derivative of the
vorticity which, according to equation (2.3.7), is equal to twice the rate
of change of the angular velocity of a small fluid parcel.
Next, we expand out the derivatives on the right-hand side of (6.6.2),
and expess the hydrodynamic volume force in terms of the stresses using
the two-dimensional counterparts of equations (6.3.4), and thus obtain
the general form of the vorticity transport equation for an incompressible
fluid,
(6.6.3)
Inviscid fluids
If viscous forces are insignificant, the shear stresses vanish, the normal
stresses are equal to the negative of the pressure, and the hydrodynamic
volume force is equal to the negative of the pressure gradient. Conse-
quently, the term enclosed by the square brackets on the right-hand side
of (6.6.3) makes no contribution, and the term expressing baroclinic pro-
duction obtains a simple form, yielding the vorticity transport equation
(6.6.4)
(6.6.5)
(6.6.6)
where v = IJL/p is a physical constant called the kinematic viscosity.
The following table displays the kinematic viscosities of water and air
at three temperatures. Note that the kinematic viscosity of air is higher
than that of water by two or three orders of magnitude. In contrast,
the viscosity of water is higher than that of air by one or two orders of
magnitude.
The right-hand side of (6.6.6) expresses diffusion of vorticity in the xy
plane. Like temperature or concentration of a species, vorticity spreads
out from regions of highly rotational flow, that is, regions where small
spherical parcels exhibit intense rotation, to regions of weakly rotational
or irrotational flow. The actual mechanism by which this occurs will
be exemplified in subsequent chapters with reference to unsteady and
boundary-layer flows.
Kinematic viscosity, v
20 1.004 15.05
40 0.658 18.86
80 0.365 20.88
(6.6.7)
Vortex stretching
(6.6.8)
This equation requires that the strength of the vorticity of a point par-
ticle, Up, be proportional to the distance of the point particle from the
axis of symmetry, a, so that the ratio between them remains constant
in time and equal to the initial value, as illustrated schematically in fig-
ure 6.6.2. This fundamental evolution law expresses a physical process
known as vortex stretching. The significance of vortex stretching will be
discussed in Chapter 11 in the framework of vortex flows.
6.6.3 Three-dimensional flow
(6.6.9)
where u2 = u*. + vfc + ILJ, is the square of the magnitude of the velocity.
To derive an evolution equation for the vorticity, we take the curl of
both sides of equation (6.6.9). A vector identity states that the curl of
the gradient of a smooth scalar function of position / vanishes,
V x V/ = O. (6.6.10)
The proof follows readily working in index notation: the ith component
of the left-hand side of (6.6.10) is given by
(6.6.11)
The symmetry of the second derivative on the right-hand side of (6.6.11),
combined with the inherent antisymmetry of the alternating tensor, re-
quires that the right-hand side vanish.
Using identity (6.6.10), we find that the curl of the second term on
the left-hand side of (6.6.9) and the curl of the first term on the right-
hand side of (6.6.9) both vanish. Invoking the definition u = V x u, we
obtain the vorticity transport equation for three-dimensional flow,
(6.6.12)
where v ~ IJL/p is the kinematic viscosity.
The second term on the left-hand side of (6.6.12) can be manipu-
lated to acquire a physical interpretation. In index notation, the ith
component of the vector expressed by this term is given by
(6.6.13)
Using property (2.3.9), we recast the right-hand side of (6.6.13) into the
form
(6.6.14)
An identity states that the divergence of the curl of a vector field vanish-
es; a consequence of this identity is that the vorticity field is solenoidal
(problem 2.3.4). Because the fluid has been assumed incompressible, the
velocity field is also solenoidal. Consequently, the second and third terms
on the right-hand side of (6.6.14) vanish. Substituting the result back
into equation (6.6.12), we find the desired vorticity transport equation
(6.6.15)
or in vector notation,
(6.6.16)
The left-hand side of (6.6.15) or (6.6.16) is the material derivative of the
vorticity, that is, the rate of change of the vorticity following the motion
of point particles.
Problems
f(x,y,z) = 0, (W.I)
(6.7.2)
where
(6.7.3)
is a scaling factor.
For example, if the first body is a sphere with diameter L\ centered
at the the point (xx^yc^zc)^ then
(6.7.4)
and
(6.7.5)
Setting the right-hand side of (6.7.5) equal to zero, we obtain the equa-
tion of a sphere with diameter L^ centered at the point (a xc, a yc, a ZQ).
Both fluids are assumed to be incompressible and Newtonian. Let pi
and //i be the density and viscosity of the first fluid, and p% and /^2 be
the density and viscosity of the second fluid.
Figure 6.7.1 Flows in two similar domains. If the Reynolds numbers
are equal, as shown in equation (6.7.6), the velocity and pressure
field of the second flow may be deduced from those of the first flow
by rescaling, and vice versa.
We will show that, when the values of the four control and physical
parameters characterizing the first flow, Li,C/i,pi, and p\, and the cor-
responding values of the control and physical parameters characterizing
the second flow, Z/2, ^2>P2> &nd /^5 are related by the equations
(6.7.6)
then the structure of the second flow may be inferred from the structure
of the first flow, and vice versa, by carrying out a simple computation
described as rescaling, as will be explained in the following subsection.
The ratio on the left-hand side of (6.7.6) is defined as the Reynolds
number of the first flow, and the the ratio on the right-hand side of
(6.7.6) is defined as the Reynolds number of the second flow.
6.7.2 Rescaling
To deduce the structure of the second flow from the structure of the
first flow, and vice versa, we introduce the dynamic pressure established
due to the flow, defined as the pressure deviation from the hydrostatic
distribution,
X2 = Q f X i . (6.7.8)
Equations (6.7.1) and (6.7.2) ensure that, if XI lies at the surface of the
body in the first flow, then X2 will lie at the surface of the body in the
second flow. In Section 6.7.3, we will show that, when relation (6.7.6)
is fulfilled, the velocity and dynamic pressure at the second point in the
second flow are related to those at the first point in the first flow by the
equations
First flow
(6.7.12)
and the dimensionless dependent variables
(6.7.13)
Solving for the dimensional variables in terms of their dimensionless
counterparts denoted by the hats, and substituting the result into the
Navier-Stokes equation and into the continuity equation, we find
(6.7.14)
and
Vi-ui-0, (6.7.15)
where Re\ = pi Ui LI/HI is the Reynolds number of the first flow shown
on the left-hand side of (6.7.6); we have introduced the dimensionless
gradient Vi = (d/dxi,d/dyi,d/dzi) and associated Laplacian operator
The far-field condition requires that, far from the body, the dimen-
sionless velocity uxi tends to unity, whereas uyi and uzi tend to van-
ish. The no-slip and no-penetration boundary conditions require that
the velocity vanishes at points (o;,j/,z) that satisfy equation (6.7.1) or,
equivalently, points (#1,3/1,21) that satisfy
/(Li £i, LI yi, LI 21) = O. (6.7.16)
Second flow
(6.7.17)
and the dimensionless dependent variables
(6.7.18)
Solving for the dimensional variables in terms of their dimensionless
counterparts, and substituting the result into the Navier-Stokes equa-
tion and into the continuity equation, we find
(6.7.19)
and
V2 • U2 = O, (6.7.20)
where Re2 = p2 U2 L2/'p>2 is the Reynolds number shown on the right-
hand side of (6.7.6); we have introdued the dimensionless gradient V2 =
(d/dx2,d/dy2,d/dz2) and associated Laplacian operator
V22 = (d2/dxld*/dyid2/dzl).
The far-field condition requires that, far from the body, the dimen-
sionless velocity uX2 tends, to unity, whereas uy2 and uZ2 tend to van-
ish. The no-slip and no-penetration boundary conditions require that
the velocity vanishes at points ( x , y , z ) that satisfy equation (6.7.2) or,
equivalently, points (£2, $2,^2) that satisfy
To this end, we compare one by one the equations and boundary con-
ditions governing the two flows in the dimensionless variables indicated
by a hat, and deduce the following:
These results suggest that, when the Reynolds numbers are equal,
the values of the dimensionless dependent variables in the two flows at
corresponding dimensionless times and corresponding dimensionless po-
sitions are equal. Setting, for example, P^(XI) — p® (£2), and using the
definitions (6.7.13) and (6.7.18), we find the second of relations (6.7.9)
subject to (6.7.10).
= pLU - LE.
M v ' (6.7.22)
-_x
X =
- -. y
y= Z=
-_z -_ tu
t=
L> L> L> ^T> (6.7.23)
and the dimensionless dependent variables
(6.7.24)
(6.7.25)
and
V-u-0. (6.7.26)
Our earlier analysis suggests that the structure of the flow depends on
L, C/, p and IJL collectively through the dimensionless Reynolds number,
in the sense of dynamic similitude expressed by equations (6.7.9) and
(6.7.10).
Characteristic scales
Stokes flow
R_ L*
^=TV (6.7.28)
where v is the kinematic viscosity. In the case of periodic flow with
angular frequency u due, for example, to an oscillating pressure gradient,
T may be identified with the period T = 2?r/a;.
Froude number
Fr = -?-,
V^' (6.7.29)
where g is the acceleration of gravity. In the case of flow over a hump
discussed in Section 6.4, the Froude number takes the specific form shown
in equation (6.4.22).
Bond number
Weber number
Problems