Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

CFD estimation of heat transfer enhancement on a

cooling pipe in underground railway tunnels


Yew S. Ting1,2, Mark J. Gilbey1, John F. Missenden2
1. Parsons Brinckerhoff Ltd, Transportation
2. London South Bank University, Department of Engineering Systems

ABSTRACT

Cooling pipes within tunnels have been used to cool the Channel Tunnel. London
Underground (LU) wished to investigate a similar concept for cooling its tunnels,
although the space constraints dictate that smaller pipes would be required and so some
form of heat transfer augmentation from extended surfaces may be beneficial. The costs
of implementing cooling pipes within existing tunnels are likely to be high and hence
optimisation and confidence in cooling performance was required.

In the present study, a numerical 3D simulation was performed using a Computational


Fluid Dynamics (CFD) method to investigate the heat transfer performance of a cooling
pipe with and without extended surfaces, in a LU tunnel environment. A comparison was
made between the cooling pipe designs in order to identify the enhanced cooling benefits
that can be achieved using extended surfaces. The extended surface was found to
increase the cooling pipe performance two times.

1 INTRODUCTION

In an underground railway system, the heat generated by the train, its auxiliaries and the
body heat from passengers, is so great that excessive temperatures would develop when
limited cooling is available. The installation of in-car air-conditioning units could add
another major heat source in the tunnel, and therefore is not an ideal solution. The heat
from the underground networks must be rejected from the tunnel system in order to
maintain thermal comfort for passengers. Maidment and Missenden (2002) evaluated
groundwater as a sustainable cooling system for underground railways [1]. Ampofo et
al. (2004) conducted a series of studies in terms of passenger thermal comfort,
Nomenclature
Superscripts:
A area m2 ¯ average
h heat transfer coefficient W.m-2.°C-1
q conduction heat flux W.m-2 Subscripts:
Q total heat transfer rate W a ambient
T temperature °C p pipe
ΔT temperature difference °C f fin
u tunnel air velocity m.s-1 FLUENT heat transfer from CFD
o outer

investigation of heat load, and methods for delivering cooling to underground railway
systems. In these studies, different cooling methods have been proposed to reduce the
heat loads in the tunnel, and they showed that groundwater cooling delivered the highest
cooling power when compared to other methods [2][3][4].

The Channel Tunnel was the first refrigerated tunnel in the world. The installation of a
cooling system was needed because, at 50 km, the tunnels are too long for cooling to
take place naturally. The exceptional length of the tunnel coupled with its depth below
the sea bed limits the opportunities for ventilation and thermal exchange with the natural
environment. Figure 1 shows the Channel Tunnel installed with cooling water pipes with
nominal diameters of 400 and 300 mm for under-sea (19 km) and under-land (10 km –
UK, 3 km – French) sections, respectively [6].

Cooling
Cooling water
water pipes
pipes

Drains

Figure 1 - Cooling water pipes for cooling the Channel Tunnel (edited) [5][6]

Water cooling pipes have also been used in underground cable tunnels, for example the
Nankou-Karyoku Line of The Kansai Electric Power Co., Inc, Japan. The line was
composed of 3 circuits (2 cables/phase), 3.4 km in length, and has a total transmission
capacity of 1800 MW. It was verified that the air temperature in the tunnel was able to be
maintained at the maximum allowable value (37 °C) by applying the water pipe cooling
system, so that workers can perform maintenance tasks in the tunnel [7][8].
London Underground is currently undertaking a major programme to address thermal
conditions for passengers. This programme is known as the Cooling the Tube
Programme (CTP). As line upgrades raise the energy used and thus temperature of the
underground routes more heat will need to be removed from the tunnels to allow the
passengers to travel on the network without suffering from ill effects of heat. The
majority of the train heat is currently removed in two ways. The first is by ventilation
systems, both mechanical and also the piston effect induced by the trains. This accounts
for typically between 20 and 25% of the total energy removal. The second method of
heat removal is the heat sink effect of the ground. This accounts for approximately 75 to
80% of the total heat removed from the network. However, the ground surrounding an
underground railway is known to rise in temperature. The soil in London that surrounds
the deep tube lines is predominantly London clay. This acts as an effective insulator and
over the past 100 years has been absorbing (and continue to absorb) large quantities of
heat from the underground. Thermal equilibrium is ultimately reached, and the heat has
now left the ground thermally saturated and this is a contributory factor to the difficulty
in removing heat from London Underground [9]. This is because the ground now
releases heat in the winter and can be regarded as a large thermal flywheel [10].

In London Underground tunnels, the space constraints dictate that smaller pipes would
be required and some form of heat transfer augmentation from extended surfaces may be
beneficial. Therefore, the costs of implementing cooling pipes within existing tunnels are
likely to be high and hence optimisation and confidence in cooling performance were
required. The CTP Team conducted an initial literature review and could not find any
published heat transfer coefficients for cooling pipes in this application. A
Computational Fluid Dynamics model has the advantage of identifying the heat transfer
coefficient of pipes in tunnels which cannot be calculated accurately using the simplified
analogy due to several reasons and uncertainties. By using a CFD model, heat transfer
enhancement using extended surfaces may be identified leading to improved heat
transfer between the cooling pipe and the warm air within the tunnel. Therefore,
understanding the thermal performance of cooling pipes in tunnels becomes important to
the CTP, and the work reported in this paper seeks to contribute to this by modelling
cooling pipes in tunnels.

This paper investigates composite heat transfer coefficients on pipes and identifies an
optimum configuration.

2 MODELLING APPROACH

A three dimensional (3D) CFD model was generated to represent cooling pipes within
tunnels. The air within the model was divided into approximately eight hundred thousand
cells for which the CFD programme determined the airflow and temperature conditions
both in each cell and between objects within the domain. The FLUENT CFD package
(version 6.3) was used for simulation, visualisation, and analysis of fluid flow and heat
transfer.

2.1 Cooling Pipe Options


Two pipe configurations were considered, namely bare-pipe and finned-pipe, as shown
in Figure 2. Option A shows the simplest alternative of bare-pipes in the tunnels. The
bare-pipe represents a reference case from which the performance of the extended
surfaces could be judged. This pipe has an external diameter of the order of 150 mm,
meaning that the pipe wall thickness is likely to be about 5 mm. The pipe size is
expected to be close to the maximum size that could be installed and allows reasonable
circulation of air around the pipe. Clusters of smaller pipes are expected to inhibit heat
transfer by loss of air circulation around the pipe. Option B shows a variant whereby the
heat transfer surface area was increased by welding fins (300 mm height and 12.5 mm
thick) onto the pipe wall. The fins would be continuous along the length of the pipe. The
fins are inclined by 15° in sympathy with the tunnel walls and distant from the wall
surface.

Option A Option B
(Bare Pipe) (Bare Pipe with Fins)

Figure 2: Options for pipes within tunnels

In underground tunnels, fixtures will be situated at regular distances close enough to give
sufficient structural support for each pipe section. However, detail of the fixtures was not
expected to affect the boundary layer around the pipe, and therefore fixtures were not
included in the modelling. In the present models, the distance between the pipe outer
surface and the tunnel flange surface was 36 mm.

2.2 Tunnel Air Conditions


Figure 3 shows a typical air velocity profile over time for a tunnel segment in
underground railway, which simulated using the Subway Environment Simulation (SES)
software. As can be seen, the velocity fluctuates considerably as each train passes.

12.5

10
Air Velocity (m.s-1)

7.5

2.5

0
0 60 120 180 240 300 360 420 480
-2.5
Time (s)

Figure 3: Typical tunnel segment velocity profile

Each model was used for three simulations differing only in the air flow rate generated
by different instances in the passage of trains through the tunnel. Referring to Figure 3, a
low air flow rate at 0.36 m.s-1 was specified to represent the background or residual air
movement generated by trains generally. A high air flow rate at 11 m.s-1 was specified to
represent the maximum air movement generated just ahead of, or just behind a train
currently travelling at its top speed. In recognition that the train generated air flow rate in
a tunnel was cyclic, and follows approximately that of a sine-wave, an intermediate air
flow rate at 6 m.s-1 was specified so that determination of the cyclic effect on the
composite heat transfer coefficient can be accounted for by a curve-fit.

Although the air flow patterns about the cooling pipe(s) will differ considerably at the
point where a train was adjacent, this circumstance exists for up to approximately 6
seconds and was small relative to the typical headway of 120 seconds. An alternative
explanation can be developed by comparing train length (130 m for a six car train) with a
typical tunnel length (2 km). Both the adjacent passing time to headway, and the train to
tunnel length ratio, were approximately 5% to 6.5%. This means that modelling the train
generated air flow through the tunnel was a reasonable approximation, and there was no
need to include the physical presence of the train.

2.3 Tunnel Wall Geometry


The effect of the structural ribs upon the airflow and turbulence near the tunnel walls in
cast iron lined tunnels was believed to have a significant effect on the air velocity near to
the tunnel wall, leading to a non-uniform forced air convection near to the pipe surface
facing the tunnel wall. Therefore, this effect was thought to be significant upon the heat
transfer coefficient. Figure 4 shows an example of an underground railway tunnel with a
cooling pipe clamped to the cast iron tunnel lining.

Cast iron tunnel


lining with flanges

Pipe clamped to
wall

Tunnel floor

Figure 4: Example of cast iron tunnel linings for Piccadilly line

2.4 Heat Transfer at Surfaces


In the thermo-fluid analysis, the non-uniform tunnel air velocity rate surrounding the
pipe surface and the use of extended surfaces (fins) created a non-uniform pipe/fin air-
side surface temperature. Hence, the inclusion of thermal conductance of the pipe wall
and extended surfaces in the CFD model was needed. The boundary condition on the
inside surface of the pipe was a uniform water temperature since temperature gradients
inside the water were expected to be small.

In the present study, the tunnel walls, tunnel floor, and pipe inner wall were treated as
isothermal surfaces. The temperature of the tunnel wall and floor surfaces was estimated
at 28.5 °C (1.5 °C below the initial tunnel air temperature of the simulation). The
temperature of the pipe inner surface was estimated at 12 °C, which represents the
assumed average temperature for the water cooling medium.

The boundary physics for tunnel wall, floor and pipe were no-slip and zero wall
thickness. See Table 1 for the material thermal properties that were used in the present
simulations.

Table 1: Materials thermal properties

Specific Heat Thermal Dynamic


Density, ρ
Materials Capacity, cp Conductivity, k Viscosity, μ
(kg.m-3)
(J.kg-1.°C-1) (W.m-1.°C-1) (kg.m-1.s-1)
Tunnel Air 1.225 1006.43 0.0242 1.7894E-5
Mild Steel Pipe 7830 500 45.3 –
Cast Iron 7272 420 52 –
Source: FLUENT 6.3, ASHRAE (2005), Kreith and Bohn (2001)

3 CFD MODEL SETTING UP PROCESS

3.1 Simplification of the Model


A primary requirement of the model is that it was able to capture both the fully
developed boundary layer for the tunnel walls, and the fully developed boundary layer
for the outside of the cooling pipe(s) and extended surface(s). The wall effects of the
tunnel were considered to be important in this study, and hence the computer model
would be required to cover a sufficient length of the tunnel to provide a fully established
flow regime in advance of the area under consideration. As boundary layers on plain
surfaces can take beyond 20 hydraulic diameters to develop fully for turbulent flow; this
means a typical deep tube tunnel of 3.8 m diameter for example would require a tunnel
length of 76 m. Also, as it was necessary to use the FLUENT ‘enhanced wall treatment’,
instead of the ‘standard wall function’ to model the boundary layer on the pipe
accurately, this imposes a much more onerous constraint on near wall meshing, and a big
computational overhead is implied by the resulting very fine mesh requirement.

In order to prevent this from making the model impractical to use, a more efficient
means of achieving fully developed boundary layers was proposed. The calculation
domain can be shortened considerably by linking the extreme ends of the tunnel via a
periodic flow boundary condition. This means that the air flow out of the tunnel will be
replicated exactly, at the inlet of the tunnel, thus maintaining the velocity profile. Using
this approach, it was suggested that the domain can be reduced to about a few meters in
length, avoiding an impractically high cell count that would have been a consequence if
the very fine mesh around the cooling pipe(s) had occupied a long tunnel. The purpose of
the shortened domain with periodic flow boundaries was to develop and record the
velocity, pressure, and turbulence profiles from a quicker isothermal analysis for use
later as boundary values in a non-isothermal analysis using the same model. An
important assumption made in support of this strategy was that in reality, weak-
buoyancy from the small temperature differences, combined with a highly unidirectional
momentum field was dominated by the momentum field, and hence the resulting
boundary layer did not differ between an isothermal analysis and a non-isothermal
analysis.
Once the flow profiles had been determined from the isothermal analysis, solution of the
energy equation was activated in the model for non-isothermal analysis, and the inlet
boundary was changed from periodic to a velocity inlet using profiles of velocity and
turbulence quantities imported from the isothermal results. The tunnel exit boundary
condition was changed from periodic to an outlet pressure, using the pressure profile
imported from the isothermal results. This non-isothermal model used the field results
from the isothermal model as its initial field.

In the present CFD analysis, a very fine mesh (down to sub-mm) was required around
the outside of the cooling pipe(s). A fine mesh was required adjacent to the tunnel walls,
in order to resolve the tunnel boundary layer, and act as an intermediate step change
between the very fine mesh, and the general tunnel mesh that was relatively coarse for
air not in close proximity to any surface. For accuracy and economy, the mesh was
constituted of hexahedral cells only, tetrahedral cells should be avoided if at all possible.
Longitudinal aspect ratios should not exceed 10:1.

3.2 CFD Domain


The present models were created and meshed structurally using GAMBIT, as shown in
Figures 5 and 6.

Front View

Isometric View

Figure 5: Computational grids for bare-pipe in tunnel model (Model A)

The 3D models were composed of 546,648 and 290,720 cells for bare-pipe and finned-
pipe in tunnels, respectively. Bare-pipe in tunnel was meshed entirely with hexahedral
cells, and Finned-pipe in tunnel model was mixed with hexahedral (90.7% of total) and
wedge (9.3% of total) cells. Hexahedral cells were the preferred choice but because of
the complex geometry in some regions, other types of cells were used as appropriate.
The quality of the computational mesh; Models A and B comprised 96.08% and 93.69%,
respectively, of equisize skew in the ranges from 0 to 0.4. The CFD package states that
high quality meshes contain elements that possess average skew values of 0.4, where 0
describes a perfectly orthogonal element.

Front View

Isometric View

Figure 6: Computational grids for finned-pipe in tunnel model (Model B)

For the present models, the aspect ratio was less than 5 for almost the entire mesh, and
20-35 on the pipe boundary layer mesh. The high aspect ratio around the pipe surface
resulted in cells that were in-line with the general flow, and therefore not likely to affect
the simulation adversely. Overall, the FLUENT grid check was normal.

4 CALCULATING TRANSFER COEFFICIENTS BY CFD

The temperature and concentration gradients next to the surface must be accurately
known because they are used to calculate the mass and heat fluxes between the pipe
surface and the tunnel air. Therefore, FLUENT’s enhanced wall treatment was used in
order to model the fluid all the way to the wall. This near-wall modelling method
combines the two-layer model, where the viscosity-affected near-wall region is
completely resolved all the way to the viscous sub-layer, using enhanced wall functions.
In this approach, the whole domain is subdivided into a viscosity-affected region and a
fully turbulent region. The demarcation of the two regions is determined by a wall-
distance based turbulent Reynolds number [11].

The turbulent viscosity in the viscosity-affected near-wall region is determined with the
one-equation model of Wolfshtein (1969) [12]. The turbulent viscosities in the two
regions are smoothly blended using a blending function. To extend its applicability
throughout the near-wall region, FLUENT formulates the law of the wall (enhanced wall
function) as a single expression for the entire wall region by blending the linear (laminar)
and logarithmic (turbulent) laws of the wall using a function suggested by Kader (1981)
[13]. The latter is a near-wall modelling method that combines a two-layer model (the
viscosity-affected near-wall region is completely resolved all the way to the viscous sub-
layer) with enhanced wall functions. The near-wall mesh must be fine enough to resolve
the transport equations down to the laminar sub-layer, which is important if heat and
mass transfer coefficients are to be accurately determined from the gradients at the
surface.

4.1 Computational Study


According to the periodic flow decided in Section 3.1, the computational model was
introduced here. And such suppositions to the model were made:

(1) Physical properties of fluid as density, viscosity, specific heat and so on in


tunnel was constant.
(2) Fluid was incompressible, isotropic and continuous.
(3) Fluid was Newtonian.
(4) Fluid flowed in the longitudinal direction along the tunnel.
(5) The fluid flow in the periodic tunnel was periodically fully-developed.
(6) No net mass addition and other source terms exist in the tunnel during the fluid
flows through the tunnel.
(7) Gravitation was not taken into consideration because of weak-buoyancy from
the small temperature difference between surfaces in the model.

The numerical simulations were performed with three-dimensional, steady-state, and


turbulent flow system. Pressure-based solver and standard k-ε model were employed.
SIMPLE algorithm was used for pressure-velocity coupling, energy equation was
included once the flow profiles have been determined (fully-developed) from the
isothermal analysis, and second order upwind scheme was chosen to discretise
momentum and energy equations.

In summary, the following conditions were set up for the present simulations: the
working fluid was air; upstream bulk temperature was set at 30 °C, the inner pipe wall
was set at temperature of 12 °C; the tunnel walls and floor were set at temperature of
28.5 °C; all walls in tunnel were set as standard no slip conditions; and the mass rate of
flow in tunnel was inputted for low (0.36 m.s-1), medium (6 m.s-1) and high (11 m.s-1)
flow rates, respectively. The specified temperatures for all models were maintained as
constants during the steady state simulations. The steady-state solution was then assumed
achievable since the solutions converged at high level of convergence criteria.

5 RESULTS AND DISCUSSION

In the present study, fully developed turbulent flow was achieved in the tunnel model
with the highest velocity located at the centre of the flow and lowest velocity (near zero)
near to the walls (tunnel linings and floor). This was because the hydrodynamic
boundary layer of the fluid has developed due to wall friction effect and growth towards
the centre. The boundary layer development along the tunnel walls was quicker than
smooth wall channel because of the tunnel flanges which disturbed the boundary layer
and enhanced the turbulence in the flow.
With respect to the heat exchange between the tunnel warm air and cooling pipe surface,
it was interesting to notice that when the distance away from the cooling pipe was large
the tunnel air temperature remained uniform (bulk air temperature). The tunnel air
temperature in the region away from the cooling pipe did not significantly affected by
the convective heat transfer process. However, near the cooling pipe interface the tunnel
air showed high temperature gradients in that region, which means high heat exchange
on the cooling pipe surface.

In addition, the effect of non-uniform local air flow rate near to the pipe and fin surfaces
was observed in the present analysis (see Figure 7), causing non-uniform heat transfer
across the outer surfaces of the pipe and fins (see Figure 8). Here, the local air flow rate
was high near to the pipe and fin surfaces which facing the tunnel centre and low air
flow rate near to the pipe and fin surfaces which facing the tunnel wall.

Contours of Velocity Magnitude (m.s-1): Tunnel Wall

Tunnel Wall

High Low
High Low

Medium
Medium

Figure 7: Velocity contours surrounding the cooling pipes, in x-y plane

Contours of Static Temperature (°C): Tunnel Wall

Tunnel Wall

High Low

High Low

Medium Medium

Figure 8: Non-uniform heat transfer surrounding the cooling pipes, in x-y plane

The CFD model provides details for the pipe and fin surfaces heat transfer and average
and local heat transfer coefficients. Here, the pipe and fin’ average heat transfer
coefficients were computed from the CFD simulation as follows
QFLUENT QFLUENT
Pipe, h p ,o = ; Fin, h f ,o = (1)
Ap ,o (Ta − T p ,o ) A f ,o (Ta − T f ,o )

where QFLUENT is the total heat transfer rate from cooling pipe or fin surfaces (from CFD
model), Ap,o is the pipe outer surface area, Af,o is the fin outer surface area, Ta is the bulk
air temperature, and T p ,o and T f ,o represent the average temperature of the pipe and fin
surfaces, respectively.

Meanwhile, the pipe and fin’ local heat transfer coefficient were defined using

q FLUENT q
Pipe, h p ,o = ; Fin, h f ,o = FLUENT (2)
(Ta − T p ,o ) (Ta − T f ,o )

where qFLUENT is the local conduction heat flux from cooling pipe or fin surface, and Tp,o
and Tf,o represent the local temperature of the pipe and fin surfaces, respectively.

Figures 9 and 10 show the average heat transfer coefficients (W.m-2.°C-1) for the bare-
pipe and finned-pipe in tunnel models, respectively, with different tunnel air flow rates.
The average heat transfer coefficients for both pipe and fins were found to increase with
tunnel air velocity, but not directly proportional. In addition, the average heat flux (W.m-
2
) and heat transfer rate per meter length (W.m-1) of both the pipe and fins also increased
with tunnel air velocity, provided that there is a temperature difference between the
cooling pipe surfaces and it nearby surroundings.

35 800

700
Average Heat Transfer Coefficient (W.m .°C )

30
-1

2
y = -0.0496x + 3.4066x + 0.7056
-2

Heat Transfer Rate (W.m of pipe)

600
25
Average Heat Flux (W.m )
-2

-1

500
20
or

400

15
300

10
200

5
100

0 0
0 1 2 3 4 5 6 7 8 9 10 11 12
-1
Tunnel Air Velocity (m.s )

Heat transfer coefficient Heat flux


Heat transfer rate Poly. (Heat transfer coefficient)

Figure 9: Bare-pipe surface average heat transfer


In Figure 11, the average heat flux (W.m-2) of pipe was higher than fins because the
average heat transfer coefficient at the pipe surface was higher, whereas the heat transfer
rate per meter length (W.m-1) of fins was higher than pipe because of the larger cooling
surface areas of fins than pipe.

25
Average Heat Transfer Coefficient (W.m .°C )
-1

20
-2

2
y = -0.0349x + 2.3209x + 0.5988

15
Pipe
Fins
Poly. (Pipe)
10 2
y = -0.0447x + 2.1291x + 0.3498 Poly. (Fins)

0
0 1 2 3 4 5 6 7 8 9 10 11 12
-1
Tunnel Air Velocity (m.s )

Figure 10: Finned-pipe surface average heat transfer coefficients

450 300

400
250
Heat Transfer Rate (W.m of pipe/fins)

350
Average Heat Flux (W.m )
-2

300 200
Pipe
-1

250 Fins
150
Pipe
200
Fins
150 100

100
50
50

0 0
0 1 2 3 4 5 6 7 8 9 10 11 12
-1
Tunnel Air Velocity (m.s )

Figure 11: Finned-pipe surface average heat transfer


Compared to bare-pipe case, the heat transfer rate (W.m-1) of finned-pipe was higher
than bare-pipe by +103%, +37% and +20% for tunnel air flow rate of 0.36, 6 and 11 m.s-
1
, respectively. However, the benefit of finned-pipe was seemed to diminish with tunnel
air velocity, as a result of temperature rising inside the fins, as shown in Figure 8.

Figures 12 and 13(a) show the local heat transfer coefficients which vary significantly
with angular position around cylindrical surface (pipe outer surface) with the highest
occurring at the pipe surfaces facing the tunnel centre (polar coordinate of 0°) and then it
decreased toward the tunnel wall-side (polar coordinate of 180°). Figure 13(b) shows that
the local heat transfer coefficients around the inclined flat fin surfaces also varied
significantly with the high heat transfer coefficients at the surfaces facing the tunnel
centre and low heat transfer coefficients at the surfaces facing the tunnel walls. The local
heat transfer coefficient was at its peak near to the fin tip because of high air flow rate
near to the fin tip.

45

40
Local Heat Transfer Coefficient (W.m .°C )
-1

35
-2

u = 11
30
u = 11

25 u=6

u=6
20
90° u = 0.36

u = 0.36
15

0° 180°
10 (Air-side) (Tunnel wall-side)

Note: Solid lines denote pipe's upper-section;


5 90° Dash dotted lines denote pipe's lower-section

0
0 20 40 60 80 100 120 140 160 180
Polar Coordinate (°)

Figure 12: Bare-pipe local surface heat transfer coefficients

This result is to be expected in view of the positive dependence of heat transfer


coefficient on Reynolds number (dominated by high inertia force to viscous force), and
the much higher local flow rates at the surfaces that facing the tunnel centre, as seen in
Figure 7. Hence, the reduction in local heat transfer coefficients at the bare-pipe surfaces
made a difference by 25%, 43% and 44% (in average) at tunnel air velocity of 0.36, 6
and 11 m.s-1, respectively. For the finned-pipe case, the reduction in local heat transfer
coefficients at the pipe surfaces was 76, 77 to 75% (in average) at tunnel air velocity of
0.36, 6 and 11 m.s-1, respectively, and the reduction in local heat transfer coefficients at
the fin surfaces was 81, 77 to 78% (in average) for tunnel air flow rate of 0.36, 6 and 11
m.s-1, respectively.
40

(1st-q) 90° (2nd-q)


35
Local Heat Transfer Coefficient (W.m .°C )
-1

0° 180°
-2

30 (Air-side) (Tunnel wall-side)

(3rd-q) (4th-q)
25 90°

Note: Solid lines denote pipe's upper-half-section;


(a) 20
Dash dotted lines denote pipe's lower-half-section.

15

10

0
0 20 40 60 80 100 120 140 160 180

Polar Coordinate (°)

u = 11 (1st-q) u = 11 (2nd-q) u = 11 (3rd-q) u = 11 (4th-q)


u = 6 (1st-q) u = 6 (2nd-q) u = 6 (3rd-q) u = 6 (4th-q)
u = 0.36 (1st-q) u = 0.36 (2nd-q) u = 0.36 (3rd-q) u = 0.36 (4th-q)

45
21
(Tunnel
40 wall-side)
Local Heat Transfer Coefficient (W.m .°C )

30
-1

10
35
-2

(Air-side)

30 41 u = 11
1
u = 11
25 u=6
(b) u=6
Note: Solid lines denote
20 top fin; Dash dotted u = 0.36
lines denote bottom fin.
u = 0.36
15

10

0
1 3 5 7 9 11 13 15 17 19 21 23 25 27 29 31 33 35 37 39 41
Fin Coordinate

Figure 13: Finned-pipe local surface heat transfer coefficients


6 CONCLUSIONS

The CFD investigation in this report has studied the heat transfer performance of cooling
pipes for underground railway tunnel applications. A CFD simulation programme was
completed using 3D models with structured meshes, carefully constructed in order to
identify the composite heat transfer coefficient values for both bare-pipe and pipe with
extended surfaces. The comparisons of heat transfer performance between bare-pipe and
finned-pipe were also discussed in this paper.

The CFD analysis has shown the heat transfer benefits of using both the bare-pipes and
finned-pipes in tunnel environments. It was found that the heat transfer performance of
cooling pipe in tunnel increased with tunnel air velocity, but not directly proportional.

The non-uniform heat transfer was found across the outer surfaces of the pipe and fins
due to the effect of non-uniform local air flow rate near to the pipe and fins surfaces. The
local air flow rate was high near to the pipe and fin surfaces which face the tunnel centre
and low near to the surfaces which face the tunnel wall. Therefore, the local heat transfer
coefficient was high at the surfaces facing the tunnel centre and then decreased toward
the surfaces facing the tunnel walls. This was because the heat transfer coefficient is a
positive dependence on Reynolds number (dominated by high inertia force to viscous
force).

In its overall thermal performance, the finned-pipe showed a higher heat transfer rate
than a bare-pipe for all tunnel air flow rates. However, the benefit of using fins (extended
surfaces) on pipe seemed to diminish with tunnel air velocity, as a result of temperature
rising inside the fins and low heat flux around the fin tip area.

Recommendations to improve the usage of cooling pipes in LU’s tunnels would be; a
combination of bare-pipes and finned-pipes clamped along the tunnel walls is worth
consideration, for example, bare-pipes would be used in the locations where space is
limited for finned-pipes, so that LU’s tunnel networks could benefit from an energy
efficient and environmentally friendly water cooling system.

ACKNOWLEDGEMENTS

The authors wish to acknowledge the support of the UK Government Knowledge


Transfer Partnership (KTP) in funding this study. London Underground is also thanked
by the authors for permission to publish this work.

REFERENCES

(1) Maidment G.G. and Missenden J.F. (2002), “Evaluation of an underground


railway carriage operating with a sustainable groundwater cooling system,”
Internation Journal of Refrigeration, vol. 25, pp. 569-574.

(2) Ampofo F., Maidment G. and Missenden J. (2004), “Underground railway


environment in the UK – Part 1: Review of thermal comfort,” Applied Thermal
Engineering, vol. 24, pp. 611-631.
(3) Ampofo F., Maidment G. and Missenden J. (2004), “Underground railway
environment in the UK – Part 2: Investigation of heat load,” Applied Thermal
Engineering, vol. 24, pp. 633-645.

(4) Ampofo F., Maidment G. and Missenden J. (2004), “Underground railway


environment in the UK – Part 3: Methods of delivering cooling,” Applied
Thermal Engineering, vol. 24, pp. 647-659.

(5) Ellison B. (1995), “Cooling the Channel Tunnel,” Journal of the Mine
Ventilation Society of South Africa, vol. 48, no. 7, 9 pages.

(6) Fairbairn A.G. (1995), “Tunnel Cooling, Drainage and Fire-Fighting,”


Proceedings of the Institution of Civil Engineers, vol. 108, pp. 24-31.

(7) Hayashi M., Uchida K., Kumai W., Sanjo K., Mitani M., Ichiyanagi N. and Goto
T. (1989), “Development of water pipe cooling system for power cables in
tunnels,” IEEE Transactions on Power Delivery, vol. 4, no. 2, pp. 863-872.

(8) Kumai W., Hashimoto I. Ohsawa S., Mitani M. and Matsuda Y. (1994),
“Completion of high-efficiency water pipe cooling system for underground
transmission line,” IEEE Transactions on Power Delivery, vol. 9, pp. 585-590.

(9) Thompson J.A. (2006), “Sustainable cooling of underground railways through


enhancement of the heat sink effect,” PhD thesis, London South Bank
University.

(10) Thompson J.A., Gilbey M.J. and Maidment G.G. (2008-09), “Geothermal
cooling of underground railways – the opportunity,” Proceedings of the
Institution of Refrigeration, 12 pages.

(11) Pham Q.T., Trujillo F.J. and McPhail N. (2009), “Finite element model for beef
chilling using CFD-generated heat transfer coefficients,” International Journal of
Refrigeration, vol. 32, pp. 102-113.

(12) Wolfshtein M. (1969), “The velocity and temperature distribution in one-


dimensional flow with turbulence augmentation and pressure gradient,”
International Journal of Heat and Mass Transfer, vol. 12, pp. 301-318.

(13) Kader B.A. (1981), “Temperature and concentration profiles in fully turbulent
boundary layers,” International Journal of Heat and Mass Transfer, vol. 24, pp.
1541-1544.

(14) FLUENT 6.3 User’s Guide, Fluent Inc.

(15) 2005 ASHRAE Handbook – Fundamentals, SI Edition, ISBN: 1931862710.

(16) Kreith F. and Bohn M.S. (2001), “Principles of Heat Transfer,” Brooks/Cole,
New York, ISBN: 0-534-37596-0.

You might also like