You are on page 1of 18

Lecture 12, January 30, 2003

Reminder Quiz 1 in the tutorial session Tuesday – up to and including fins (ch1-
ch3).

Last Day,
• Determined the fin heat rate
• Defined fin efficiency and effectiveness
• Looked at temperature distributions in fins
• Introduced the Biot number

Today:
• Look at CFD simulations of fins (sneak intro to convection)
• Look at numerical solution (control volume) technique
• Finned arrays

CFD (Computional Fluid Dynamics) Results

The following results are obtained using a commercial CFD code, Fluent. This code
solves the Navier-Stokes equations, coupled with conservation of energy to obtained
detailed solutions of pressure, temperature and velocity fields. From these fields, we can
calculate whatever is of interest for our particular problem.

Note that these simulations were carried out very quickly without performing the detailed
checks and refinements which are required when using CFD for real design/analysis
work. Accordingly, the results will certainly be representative of the physics, but will not
be quantitatively perfect (especially when turbulence is modelled). The results are
however perfect to illustrate the points that are made in this section.

We will use this code to solve laminar and turbulent flow over a single fin. We will
explore:
• Fin effectiveness
• How convection correlations are determined
• How the temperature field is coupled to the velocity field in convection
problems.

Let’s begin with a uniform air stream impinging on a hot vertical wall. The air is at
293K, while the wall is at 353 K. Initially, let the uniform air velocity be 0.98 m/s,
which will correspond to a laminar flow. The following figure depicts the resulting
velocity vectors computed by Fluent. Note that the flow is from right to left, and that
the hot wall is the left wall. Since the air cannot pass through the wall, it is deflected
upwards, and exits the domain vertically at the top wall. Note the interesting velocity
pattern in the lower left corner. Here the velocities are very low, and the flow is
recirculating. This will have implications on the energy field.

The velocity vectors look nice, but a much better way to visualize a two dimensional flow
field is to compute the stream function, as discussed in Mech 341. The next figure
depicts contours of the stream function. Each contour line is a line upon which the
stream function is constant – this is a streamline. The significant feature of a streamlines,
which make them a fantastic tool for visualizing flow fields, is that they are everywhere
parallel to the velocity vector. This means that no flow can pass through a streamline.
Pick any two streamlines, and follow them from the inlet to the outlet. What you are
looking at the true path that the fluid takes to traverse our domain. Since the flow rate is
constant, and this is an incompressible flow, we can glean even more from this plot.
When the streamlines are close together, the air is moving faster than when the
streamlines are further apart. Therefore, we can clearly see in the figure that the air is
decelerated as it approaches the hot wall and turns the corner, and it accelerates again as
it exits the domain. Again, I will defer detailed discussion to both Mech 341, and the
convection section of the course, but it is precisely where the flow is decelerating that
flow separation and regions of recirculation are likely to form. These regions will have
an enormous impact on heat transfer, as we shall see.
Let’s look closely at the velocity vectors in the lower left corner. It is clear from this
figure that the air is moving in roughly a circular pattern in the corner. The details of
this “flow structure” will change considerably with the addition of a fin along the bottom
section of the hot wall, and especially when the flow becomes turbulent.
Now, let’s look at the resulting temperature field. The next figure depicts temperature
contours over the entire domain. Most of the domain contains air at the free stream
temperature, 293 K. Adjacent to the hot wall though the air is clearly heated by the hot
wall. Notice that the temperature contours are very different where the air is recirculating
compared to where the flow is “attached” to the hot wall. The recirculating air is heated
at the hot wall, and carries this energy away from the wall. Accordingly, the temperature
contours are wildly distorted along the paths taken by the air.

This fact is even more apparent in the next two figures which are simply increasingly
magnified views of the recirculation zone.
As we zoom in really close, note that that the temperature contour lines are very nearly
parallel to the hot wall very close the wall, where the velocities are very nearly zero (no-
slip condition at the wall). Remember from Fourier’s law, that heat energy will be
transferred from regions of higher temperature to regions of lower, and that this will
occur in a direction perpendicular to the temperature contour lines (i.e. in the opposite
direction of the temperature gradient). It is clear that energy is being transferred from the
wall to the air.

Notice too, that the contour lines are perpendicular to the bottom boundary. I have
chosen this to be a symmetry boundary, which physically means that the geometry and
flow is mirrored about this line. If that is the case, then there can be no heat transfer
across this line since there can be no temperature gradient there (since the temperature is
the same at the same distance from the line). Temperature contour lines intersecting a
boundary at 90O like this indicate that there is no heat transfer through this boundary.
Now Lets add a fin, made of a metallic material with a conductivity of 202 W/mK, and
use the same free stream velocity. Notice that the velocity vectors look similar, but there
is definitely more distortion near the fin,
Looking closely at the fin tip, we see that the velocity vectors closest to the fin are
actually directed towards the inlet – the fluid nearest the fin surface is moving against the
main flow! This means that the flow has “separated” from the fin surface very near the
fin tip, and this the recirulation zone extends along the entire length of the fin.

Looking now near the lower left corner, we see a fairly similar plot compared to when
there was no fin, though the flow certainly has been perturbed by the fin. Note that
structure is evident within the recirculation region. Flow comes down from the vertical
sidewall, and turns to flow along the fin in the reverse direction from the bulk flow.
Some of this flow is directed upwards fairly quickly, to form the circular recirculation,
while some of it continues along the fin. Again, this will have implications on the heat
transfer.

Again, the streamlines are a usefull tool for describing the general flow.
Let’s now look at the temperature contours. Again, the temperature contours follow the
general features of the frecirculating flow. The prime difference between this case, and
the case with no fin, is that there are many temperature contour lines (nearly) parallel to
the fin surface, indicating that there is significant heat transfer from the fin surface to the
air. Note too, that the high temperature region extends further out into the flow in the
middle of the fin compared to the base and the tip sections. This should not be surprising
since the air velocities in this region were the lowest.

This feature is increasingly evident as we magnify the contours. It is very clear that the
moving air is carrying heat energy with it – clearly demonstrating how convection
enhances heat transfer compared to conduction alone. Finally, not that the lone
temperature contour in the fin is practically a vertical line. This suggests that heat
transfer is practically 1-D along the direction of the fin – as we would expect from
calculating the Biot Number.
We will now calculate the heat transferred from the hot wall to the air, and break this out
into two sections. The first section comprising the fin only (remember qfin is calculated
by determing how much heat energy is conducted in at the base as this must be dissipated
at the surface of the fin) and the second comprising the rest of the wall. The fin is 0.5 cm
high, while the rest of the wall shown here is 20 cm high. This means that the fin cross
section is 1/40th the cross section of the remaining wall. Further, we will vary the air
velocity, to see the effect that this has on heat transfer. In order to describe the different
air velocities then, we will calculate a Reynolds number at the end of the fin (which is 15
cm long).
ρVL
Re L =
µ
where L= 0.15 m, ρ=1.225 kg/m is the air density, and µ = 1.8 x 10-5 kg/(ms) is the air
3

viscosity. Our 0.98 m/s velocity, presented above corresponds to ReL=10,000. A


summary of the heat rates for various cases are given in the following table. Notice that
when there is no fin present there is very little heat transfer through the area
corresponding to where the fin will go (0.855 W), adding the fin (and keeping the air
velocity constant) increases this to 43.5 W, for a fin effectiveness over 50. Note too
under these conditions that the fin is dissipating almost half as much energy as the
remaining wall, which has 40 times the cross sectional area! This clearly demonstrates
why fins are used. In fact, as the velocity increases, this gets even better

ReL Fin heat rate [W] Wall heat Fin effeciveness


rate [W] [-]
No Fin
(u=0.98m/s) 0.855 102.16
2,500 23.17 46.53 27.1
5,000 30.78 70.52 36.0
10,000 43.5 96.92 50.9
20,000 68.42 118.61 80.0
100,000 358.9 1422.79 419.8

The first 5 cases were assumed to be laminar, while the final case, at ReL=100,000, was
assumed to be turbulent. Notice that for the laminar cases, when we double the flow rate,
we can about a 50% increase in heat dissipated. However, when we jumped from
ReL=20,000 to ReL=100,000, a fivefold increase, we got a very similar increase in heat
dissipated. Turbulence changes the flow field enormously and provides for much more
heat transfer. This occurs in two ways. First, the turbulent flow creates far more mixing,
and hence is more effective at carrying hot fluid away from the surface of the fin to the
cooler surrounding fluid. Second, the increased energy in the flow is able to radically
alter the flow structure, resulting in much smaller recirculation regions. The following
figure shows the velocity vectors for the turbulent flow case (ReL=100,000). Notice that
the recirclations regin is clearly smaller.

If we again focus on the tip region of the fin, we see that now the flow is attached to the
fin – the velocity vectors close to the fin surface point in the same direction as the main
flow. The recirculation zone is limited to a much smaller area near the base of the fin.
The temperature field is also radically altered by the turbulent flow. Now, the high
temperature region is confined very close to the fin virtually everywhere. Since the fin is
now dissipating considerable heat energy, there is also a mich greater temperature drop
within the fin itself. Again, we can see that temperature contours inside the fin are
vertical lines, indicating that heat transfer is 1-D in this case. If we tried this again with a
fin having a much higher Biot number (lower conductivity) we would see a very different
story.
If we keep everything else constant, and reduce the fin conductivity to 0.1 W/mK (as in
the next two figures), the temperature field again changes radically, and the heat
dissipated through the fin drops to 4.6 W from 358.9 W.
Notice that the temperature of the fin has dropped to 293 K (the ambient) about halfway
along its length, meaning that the rest of the material is simply wasted. Also note that the
temperature contours in the fin are very two dimensional. Clearly, a 1-D analysis would
not work in this case. Note the value of the Biot number in indicating this before we
have done any analysis.
Finally, lets calculate a local value of a convection coefficient at each point along the fin.
Note that this is easy to do since we know the entire temperature field, and we can
calculate the heat rate at each and every point by calculating the gradient of the
temperature distribution. At the fin surface, the fluid velocity is zero (no-slip condition),
and heat transfer is purely by conduction at this point. Therefore, the heat rate is
Calculated using Fourier’s law. We then compare this with Newton’s law of cooling,
q = hA(Ts - T∞). We know the area of each control volume, and we know the local
surface temperature and our reference ambient temperature. It is therefore a simple
calculation to determine a local h value. The next figure plots the variation of the local
convection coefficient over the fin for the various cases. Note that the assumption of a
constant h, that we made when considering the fin only, is clearly a big approximation.
In general, the convection coefficients are high near the fin tip, and drop as the base is
approached. In the laminar flow cases, though it the effect of the recirculating air is
clearly apparent as h drops to a minimum and then increases again.
How do we actually solve the 1-D fin equation
numerically?
Stepping back to what we did last day with MATLAB, how do the provided programs
actually solve for the temperature distribution? Examine the programs available on the
web site (unifin.m for example). The script ‘finfun2004.m’ is a script that produces all
the outout given in the powerpoint slides last day. The following description should help
you understand exactly how the functions work.

Now, applying conservation of energy for a steady 1D system with no generation,

Express this in a convenient form, identifying the unknowns,

And, considering each control volume, we have exactly as many unknowns as equations
and we can solve for the temperature distribution and hence the fin heat rate.
Fin Arrays (section 3.6.5 in text)
How do we extend our analysis to include multiple fins?
We shall develop resistance networks to describe the geometry of interest.
First define a thermal resistance for a single fin based on what we have already done with
fins (numerically or analytically)

And reform this equation to define a thermal resistance for a fin.

Both the fins, and the area between them will dissipate heat, and so we have to consider
the spaces as well. If there are N fins, then the total area dissipating heat is

The total heat rate is then,

These equations, and the concept of thermal resistance networks allow us to analyze any
array of fins, and even the effects of contact resistance. Be very careful though about the
approximation of constant convection coefficient.

You might also like