§6.7 Turbulent boundary layers 315Figure 6.17 Fluctuation of u and other quantities in a turbu-lent pipe flow.. §6.7 Turbulent boundary layers 317Figure 6.18 The shear stress, τyx, in a
Trang 1one or two enormous vortices of continental proportions These hugevortices, in turn, feed smaller “weather-making” vortices on the order ofhundreds of miles in diameter These further dissipate into vortices ofcyclone and tornado proportions—sometimes with that level of violencebut more often not These dissipate into still smaller whirls as they inter-act with the ground and its various protrusions The next time the wind
blows, stand behind any tree and feel the vortices In the great plains,
where there are not many ground vortex generators (such as trees), youwill see small cyclonic eddies called “dust devils.” The process continuesright on down to millimeter or even micrometer scales There, momen-tum exchange is no longer identifiable as turbulence but appears simply
as viscous stretching of the fluid
The same kind of process exists within, say, a turbulent pipe flow athigh Reynolds number Such a flow is shown in Fig.6.17 Turbulence
in such a case consists of coexisting vortices which vary in size from asubstantial fraction of the pipe radius down to micrometer dimensions.The spectrum of sizes varies with location in the pipe The size andintensity of vortices at the wall must clearly approach zero, since thefluid velocity goes to zero at the wall
Figure6.17shows the fluctuation of a typical flow variable—namely,velocity—both with location in the pipe and with time This fluctuationarises because of the turbulent motions that are superposed on the aver-
age local flow Other flow variables, such as T or ρ, can vary in the same
manner For any variable we can write a local time-average value as
where T is a time that is much longer than the period of typical
fluctua-tions.9 Equation (6.81) is most useful for so-called stationary processes—ones for whichu is nearly time-independent.
If we substitute u = u + u in eqn (6.81), where u is the actual local
velocity and u is the instantaneous magnitude of the fluctuation, weobtain
9Take care not to interpret this T as the thermal time constant that we introduced
in Chapter 1 ; we denote time constants are asT
Trang 2§6.7 Turbulent boundary layers 315
Figure 6.17 Fluctuation of u and other quantities in a
turbu-lent pipe flow
This is consistent with the fact that
u or any other average fluctuation= 0 (6.83)since the fluctuations are defined as deviations from the average
We now want to create a measure of the size, or lengthscale, of
turbu-lent vortices This might be done experimentally by placing two
velocity-measuring devices very close to one another in a turbulent flow field
When the probes are close, their measurements will be very highly
corre-lated with one one another Then, suppose that the two velocity probes
are moved apart until the measurements first become unrelated to one
another That spacing gives an indication of the average size of the
tur-bulent motions
Prandtl invented a slightly different (although related) measure of the
lengthscale of turbulence, called the mixing length, He saw as an
average distance that a parcel of fluid moves between interactions It
has a physical significance similar to that of the molecular mean free
path It is harder to devise a clean experimental measure of than of the
Trang 3correlation lengthscale of turbulence But we can still use the concept of
to examine the notion of a turbulent shear stress.
The shear stresses of turbulence arise from the same kind of tum exchange process that gives rise to the molecular viscosity Recallthat, in the latter case, a kinetic calculation gave eqn (6.45) for the lami-nar shear stress
where was the molecular mean free path and u was the velocity
differ-ence for a molecule that had travelled a distance in the mean velocity
gradient In the turbulent flow case, pictured in Fig.6.18, we can think ofPrandtl’s parcels of fluid (rather than individual molecules) as carrying
the x-momentum Let us rewrite eqn (6.45) in the following way:
• The shear stress τ yx becomes a fluctuation in shear stress, τ yx ,resulting from the turbulent movement of a parcel of fluid
• changes from the mean free path to the mixing length
• C is replaced by v = v + v , the instantaneous vertical speed of the
fluid parcel
• The velocity fluctuation, u , is for a fluid parcel that moves a
dis-tance through the mean velocity gradient, ∂u/∂y It is given by
Trang 4§6.7 Turbulent boundary layers 317
Figure 6.18 The shear stress, τyx, in a laminar or turbulent flow
Notice that, while u = v = 0, averages of cross products of fluctuations
(such asu v or u 2) do not generally vanish Thus, the time average of
the fluctuating component of shear stress is
In addition to the fluctuating shear stress, the flow will have a mean shear
stress associated with the mean velocity gradient, ∂u/∂y That stress is
µ(∂u/∂y), just as in Newton’s law of viscous shear.
It is not obvious how to calculatev u (although it can be measured),
so we shall not make direct use of eqn (6.86) Instead, we can try to model
v u From the preceding discussion, we see that v u should go to zero
when the velocity gradient, (∂u/∂y), is zero, and that it should increase
when the velocity gradient increases We might therefore assume it to be
proportional to (∂u/∂y) Then the total time-average shear stress, τ yx,
can be expressed as a sum of the mean flow and turbulent contributions
that are each proportional to the mean velocity gradient Specifically,
Trang 5where ε m is called the eddy diffusivity for momentum We shall use this
characterization in examining the flow field and the heat transfer.The eddy diffusivity itself may be expressed in terms of the mixinglength Suppose that u increases in the y-direction (i.e., ∂u/∂y > 0).
Then, when a fluid parcel moves downward into slower moving fluid,
it has u (∂u/∂y) If that parcel moves upward into faster fluid, the sign changes The vertical velocity fluctation, v , is positive for anupward moving parcel and negative for a downward motion On average,
u and v for the eddies should be about the same size Hence, we expectthat
Turbulence near walls
The most important convective heat transfer issue is how flowing fluidscool solid surfaces Thus, we are principally interested in turbulence nearwalls In a turbulent boundary layer, the gradients are very steep nearthe wall and weaker farther from the wall where the eddies are largerand turbulent mixing is more efficient This is in contrast to the gradualvariation of velocity and temperature in a laminar boundary layer, whereheat and momentum are transferred by molecular diffusion rather thanthe vertical motion of vortices In fact,the most important processes inturbulent convection occur very close to walls, perhaps within only afraction of a millimeter The outer part of the b.l is less significant.Let us consider the turbulent flow close to a wall When the boundarylayer momentum equation is time-averaged for turbulent flow, the result
Trang 6§6.7 Turbulent boundary layers 319
In the innermost region of a turbulent boundary layer — y/δ 0.2,
where δ is the b.l thickness — the mean velocities are small enough
that the convective terms in eqn (6.90a) can be neglected As a result,
∂τ yx /∂y 0 The total shear stress is thus essentially constant in y and
must equal the wall shear stress:
τ w τ yx = ρ (ν + ε m ) ∂u
Equation (6.91) shows that the near-wall velocity profile does not
de-pend directly upon x In functional form
u = fnτ w , ρ, ν, y (6.92)
(Note that ε mdoes not appear because it is defined by the velocity field.)
The effect of the streamwise position is carried in τ w, which varies slowly
with x As a result, the flow field near the wall is not very sensitive
to upstream conditions, except through their effect on τ w When the
velocity profile is scaled in terms of the local value τ w, essentially the
same velocity profile is obtained in every turbulent boundary layer.
Equation (6.92) involves five variables in three dimensions (kg, m, s),
so just two dimensionless groups are needed to describe the velocity
where the velocity scale u ∗ ≡3τ w /ρ is called the friction velocity The
friction velocity is a speed characteristic of the turbulent fluctuations in
the boundary layer
Trang 7Equation (6.91) can be integrated to find the near wall velocity profile:
To complete the integration, an equation for ε m (y) is needed
Measure-ments show that the mixing length varies linearly with the distance from
the wall for small y
where κ = 0.41 is called the von Kármán constant Physically, this says that the turbulent eddies at a location y must be no bigger that the dis-
tance to wall That makes sense, since eddies cannot cross into the wall
The viscous sublayer Very near the wall, the eddies must become tiny;
and thus ε m will tend to zero, so that ν ε m In other words, inthis region turbulent shear stress is negligible compared to viscous shearstress If we integrate eqn (6.94) in that range, we find
= (u ∗ )2y ν
a large increase in the wall shear stress)
The log layer Farther away from the wall, is larger and turbulent
shear stress is dominant: ε m ν Then, from eqns (6.91) and (6.89)
Trang 8§6.7 Turbulent boundary layers 321
Assuming the velocity gradient to be positive, we may take the square
root of eqn (6.97), rearrange, and integrate it:
du =
2
τ w ρ
for B 5.5 Equation (6.99) is sometimes called the log law
Experimen-tally, it is found to apply for (u ∗ y/ν) 30 and y/δ 0.2.
Other regions of the turbulent b.l For the range 7 < (u ∗ y/ν) < 30,
the so-called buffer layer, more complicated equations for , ε m , or u are
used to connect the viscous sublayer to the log layer [6.7, 6.8] Here,
actually decreases a little faster than shown by eqn (6.95), as y3/2 [6.9]
In contrast, for the outer part of the turbulent boundary layer (y/δ
0.2), the mixing length is approximately constant: 0.09δ Gradients
in this part of the boundary layer are weak and do not directly affect
transport at the wall This part of the b.l is nevertheless essential to
the streamwise momentum balance that determines how τ w and δ vary
along the wall Analysis of that momentum balance [6.2] leads to the
following expressions for the boundary thickness and the skin friction
To write these expressions, we assume that the turbulent b.l begins at
x = 0, neglecting the initial laminar region They are reasonably accurate
for Reynolds numbers ranging from about 106 to 109 A more accurate
Trang 9formula for C f, valid for all turbulent Rex, was given by White [6.10]:
C f (x) = ! 0.455
ln(0.06 Re x )"2 (6.102)
Like the turbulent momentum boundary layer, the turbulent thermalboundary layer is characterized by inner and outer regions In the in-ner part of the thermal boundary layer, turbulent mixing is increasinglyweak; there, heat transport is controlled by heat conduction in the sub-layer Farther from the wall, a logarithmic temperature profile is found,and in the outermost parts of the boundary layer, turbulent mixing is thedominant mode of transport
The boundary layer ends where turbulence dies out and uniform stream conditions prevail, with the result that the thermal and momen-tum boundary layer thicknesses are the same At first, this might seem
free-to suggest that an absence of any Prandtl number effect on turbulentheat transfer, but that is not the case The effect of Prandtl number isnow found in the sublayers near the wall, where molecular viscosity andthermal conductivity still control the transport of heat and momentum
The Reynolds-Colburn analogy for turbulent flow
The eddy diffusivity for momentum was introduced by Boussinesq [6.11]
in 1877 It was subsequently proposed that Fourier’s law might likewise
be modified for turbulent flow as follows:
q = −k ∂T
∂y −
another constant, whichreflects turbulent mixing
where ε h is called the eddy diffusivity of heat This immediately suggests
yet another definition:
turbulent Prandtl number, Prt ≡ ε m
Trang 10§6.8 Heat transfer in turbulent boundary layers 323
Equation (6.103) can be written in terms of ν and εm by introducing Pr
and Prt into it Thus,
Before trying to build a form of the Reynolds analogy for turbulent
flow, we must note the behavior of Pr and Prt:
• Pr is a physical property of the fluid It is both theoretically and
actually near unity for ideal gases, but for liquids it may differ from
unity by orders of magnitude
• Pr t is a property of the flow field more than of the fluid The
nu-merical value of Prtis normally well within a factor of 2 of unity It
varies with location in the b.l., but, for nonmetallic fluids, it is often
near 0.85
The time-average boundary-layer energy equation is similar to the
time-average momentum equation [eqn (6.90a)]
and in the near wall region the convective terms are again negligible This
means that ∂q/∂y 0 near the wall, so that the heat flux is constant in
y and equal to the wall heat flux:
The constant A depends upon the Prandtl number It reflects the thermal
resistance of the sublayer near the wall As was done for the constant
B in the velocity profile, experimental data or numerical simulation may
be used to determine A(Pr) [6.12,6.13] For Pr≥ 0.5,
A(Pr) = 12.8 Pr 0.68 − 7.3 (6.109)
Trang 11To obtain the Reynolds analogy, we can subtract the dimensionlesslog-law, eqn (6.99), from its thermal counterpart, eqn (6.108):
T w − T (y)
q w /(ρc p u ∗ ) − u(y) u ∗ = A(Pr) − B (6.110a)
In the outer part of the boundary layer,T (y) T ∞ and u(y) u ∞, so
C f = A(Pr) − B (6.110d)Rearrangment of the last equation gives
q w
(ρc p u ∞ )(T w − T ∞ ) = C f 2
1+ [A(Pr) − B]4C f 2 (6.110e)The lefthand side is simply the Stanton number, St= h (ρc p u ∞ ) Upon substituting B = 5.5 and eqn (6.109) for A(Pr), we obtain the Reynolds-Colburn analogy for turbulent flow:
1+ 12.8Pr0.68 − 1 4C f 2 Pr≥ 0.5 (6.111)This result can be used with eqn (6.102) for Cf , or with data for C f,
to calculate the local heat transfer coefficient in a turbulent boundary
layer The equation works for either uniform T w or uniform q w This isbecause the thin, near-wall part of the boundary layer controls most ofthe thermal resistance and that thin layer is not strongly dependent onupstream history of the flow
Equation (6.111) is valid for smooth walls with a mild or a zero
pres-sure gradient The factor 12.8 (Pr 0.68 − 1) in the denominator accounts
for the thermal resistance of the sublayer If the walls are rough, thesublayer will be disrupted and that term must be replaced by one thattakes account of the roughness (see Sect.7.3)
Trang 12§6.8 Heat transfer in turbulent boundary layers 325
Other equations for heat transfer in the turbulent b.l.
Although eqn (6.111) gives an excellent prediction of the local value of h
in a turbulent boundary layer, a number of simplified approximations to
it have been suggested in the literature For example, for Prandtl numbers
not too far from unity and Reynolds numbers not too far above transition,
the laminar flow Reynolds-Colburn analogy can be used
The best exponent for the Prandtl number in such an equation actually
depends upon the Reynolds and Prandtl numbers For gases, an exponent
of−0.4 gives somewhat better results.
A more wide-ranging approximation can be obtained after
introduc-ing a simplifed expression for C f For example, Schlichting [6.3, Chap XXI]
shows that, for turbulent flow over a smooth flat plate in the low-Re range,
Somewhat better agreement with data, for 2× 105 Re x 5 × 106, is
obtained by adjusting the constant [6.15]:
k L
dx
... class="text_page_counter">Trang 12< /span>
§6.8 Heat transfer in turbulent boundary layers 325
Other equations for heat transfer. ..
Trang 10§6.8 Heat transfer in turbulent boundary layers 323
Equation (6.103) can... Pr 0.68 − 7.3 (6.109)
Trang 11To obtain the Reynolds analogy, we can subtract the