Vous êtes sur la page 1sur 22

ARTICLE IN PRESS

J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Contents lists available at ScienceDirect

Journal of Wind Engineering


and Industrial Aerodynamics
journal homepage: www.elsevier.com/locate/jweia

The ow around high speed trains


Chris Baker 
School of Civil Engineering, University of Birmingham, Edgbaston, Birmingham, B15 2TT, UK

a r t i c l e in fo

abstract

Article history:
Received 4 March 2009
Accepted 18 November 2009
Available online 23 December 2009

This paper considers aspects of the aerodynamic behaviour of high speed trains. It does not specically
address the many aerodynamic problems associated with such vehicles, but rather attempts to describe,
in fundamental terms, the nature of the ow eld. The rationale for such an approach is that the ow
elds that exist are the primary cause of the aerodynamic forces on the train and its components which
result in a whole range of aerodynamic issues. This paper thus draws on a wide range of model scale
and full scale experimental and computational work and attempts to build up a comprehensive picture
of the ow eld. Attention is restricted to trains in the open air (i.e. tunnel ows will not be considered)
for both still air conditions and crosswind conditions. For still air conditions the ow eld will be
described for a number of ow regions i.e.

Keywords:
High speed trains
Aerodynamics
Cross winds
Boundary layers
Wakes
Slipstreams

 around the nose of the train;


 along the side, roof and underbody of the train;
 the wake of the train;
Calculations of the nature of the wind relative to the train will be presented for a variety of train
speeds and wind speeds. For crosswind conditions, the nature of the ow eld around typical trains,
including surface pressure distributions, will be presented. In addition the aerodynamic admittances/
weighting functions for different types of train will be discussed. Finally some remarks will be made as
to the relevance of the data that has been presented to current issues in train aerodynamics.
& 2009 Elsevier Ltd. All rights reserved.

1. Introduction
This paper aims to set out a description of the ow eld around
high speed trains in the open air. It will approach this from a fairly
fundamental point of view, and will not specically address
practical issues and problems associated with the aerodynamic
behaviour of trains, although these will be briey discussed at the
end of the paper. Such an approach is adopted in the hope that
such a description will clarify the basic ow mechanisms that
exist around high speed trains, and will thus inform future
consideration of a range of more practical issues.
The basic tools in the study of train aerodynamics are full scale
testing, wind tunnel testing and CFD calculations, as indeed is the
case in other elds of aerodynamics. In the case of the study of
train aerodynamics, all of these approaches are fraught with
difculties. Full scale measurements have to be made in very
turbulent ows and very often a large number of runs have to be
carried out to enable the mean and unsteady ow patterns to be
elucidated. Baker et al. (2001) describes the technique of
ensemble averaging through which the results of a large

 Tel.: + 44 1214145067; fax: + 44 1214143675.

E-mail address: c.j.baker@bham.ac.uk


0167-6105/$ - see front matter & 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2009.11.002

number of runs are considered together. Results obtained using


this technique will be used extensively in what follows. Both wind
tunnel tests and CFD calculations are made difcult because of the
large length/height ratio of high speed trains that makes wind
tunnel models or computational grids very long and thin, and
which both require specialist techniques and expertise. This point
having been made however, experimental and computational
techniques will not be discussed at any length in what follows,
although where the nature of the technique has the potential to
seriously affect the results that are being presented then this will
be pointed out.
Section 2 discusses the aerodynamics of high speed trains in
conditions of zero cross wind. The discussion is framed in terms of
three ow regions, viz.

 the nose region around the front of the train;


 the boundary layer region along the length of the train (for the
train side, train roof and train underbody);

 the wake region behind the train.


This scheme is based on that developed by the author in Baker
et al. (2001), although in this paper the number of ow regions is
reduced from the ve in Baker et al. (2001) to the three listed

ARTICLE IN PRESS
278

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Fig. 1. Velocities in the nose region of the ICE service train (z0 =0.5 m).

Fig. 2. Pressure time history measured during the passage of two ETR 500 trains
(x axis is an arbitrary time) (Mancini and Malfatti 2001).

Fig. 3. Results of the potential ow calculations of Sanz-Andres and SantiagoProwald (2002) (Pressure coefcient traces are shown for 2D and 3D computations. The x axis parameter T is the time from the passing of the nose of the train
normalised by train speed and distance from the centre of the train).

abovethe upstream and nose region in Baker et al. (2001) being


considered together here, and the near wake and far wake regions
in Baker et al. (2001) being similarly combined. For each of the

Fig. 4. Synthesis of boundary layer measurements (Brockie and Baker, 1990;


Schetz, 2001; Sockel, 1996).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

ow regions the work of the author and his co-workers, and the
work of other investigators are considered to develop as complete
a picture as possible of the ow eld around the train.
Section 3 then goes on to consider the ow eld around trains
in a cross wind. This begins by a consideration of the nature of the
wind ow relative to the train in terms of the mean velocity
prole, turbulence prole and power spectrum. A qualitative
picture of the ow around trains is then developed from a
consideration of the work of a number of authors, and the nature
of the pressure distribution around high speed trains is also
discussed, in terms of both steady and unsteady surface pressures.
Finally the way in which these pressures sum to give cross wind
forces and moments is discussed in terms of the aerodynamic
admittances and aerodynamic weighting functions. Some concluding remarks are then made in Section 4 and the implications
of the results for current issues in high speed train aerodynamics
are set out.

2. The ow around trains with no cross wind


2.1. The nose region
In this section the ow upstream and just downstream of the
nose of high speed trains will be considered. In this region the
variations of air velocity and pressure are essentially inviscid. A
typical variation in air velocity is shown in Fig. 1 around the front of

279

a 14 car ICE service train. This data was obtained from trackside
anemometry in full scale experiments designed to measure the
slipstreams around such trains. The experiments are reported in
outline (RAPIDE Consortium, 2001) and discussed in considerable
length by the author and his co-workers in Sterling et al. (2008b).
Data from these experiments will be used extensively in what
follows to illustrate a number of effects. The data in Fig. 1 is an
ensemble average of the data from 17 train passes. This data was
aligned (at the point corresponding to the peak of the velocity trace
shown in the gure) and the data at all other points averaged over
all the runs. Thus x, the position along the train, is dened as
measured from this peak in velocity. The lateral distance y0 is
dened as the distance from the rail edge, and the vertical distance
z0 as the distance from the top of the rail. For the results shown the
velocities were measured at trackside with no platform present
(z0 =0.5 m). The air velocity data, u, is divided by the train speed, v,
to give the normalised value U. From Fig. 1 the velocity peak can be
seen to be sharply dened and, as would be expected, decreases
away from the train. The standard deviation of the ensemble is
small in all casesof the order of 0.020.03, which indicates that in
this ow region there is little run to run variation.
The velocity changes illustrated in Fig. 1 are accompanied by
pressure changes. Fig. 2 shows typical pressure changes caused by
an ETR 500 (Mancini and Malfatti, 2001). These measurements
were obtained from train passing tests carried out as part of the
major EU TRANSAERO project. It can be seen that there is a rapid
increase and then decrease in pressure around the train nose.

0.4
0.3

y' = 1.16m

0.2

y' = 1.5m
y'= 2.42m

0.1
0

-50

50 100 150 200 250 300 350 400 450 500


x (m)

Fig. 5. ICE service car velocity measurements (z0 = 0.5 m; measurements made at trackside with no platform) (Sterling et al., 2008b).

0.3
0.25
x=50m

0.2
U

x=100m
0.15

x=200m
x=300m

0.1

x=350m

0.05
0
0

y (m)
Fig. 6. ICE service car velocity measurements (z0 =0.5 m; measurements made at trackside) (Sterling et al., 2008b).

ARTICLE IN PRESS
280

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Again for any particular train, this effect is highly repeatable from
run to run.
As such ows are inviscid they can be well predicted by
reasonably simple calculation methodsas shown Fig. 3 below
from the potential ow calculations of Sanz-Andres and SantiagoProwald (2002). More complex panel methods can be used to
calculate the details of the pressure and velocity variations
around train nose shapes of different types (such as the results

for the Euler method shown in Fig. 2). The blunter the nose shape,
the higher are the velocity and pressure disturbances.
2.2. The boundary layer region
2.2.1. Train side
Over the last few decades a number of investigators have made
boundary layer measurements on trains, using conventional train

Fig. 7. Boundary layer parameters for ICE service train (Sterling et al., 2008b).

Turbulence intensity

0.3

0.2
Trackside y'=1.16m,
z'=0.5m
Platform y=1m,
z=1m

0.1

0
0

100

200

300

400

x (m)
Fig. 8. Turbulence intensity for boundary layers on the side of the ICE service train (Sterling et al., 2008b).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

0.8

Trackside y'=1.16m,
z'=0.5m
Platform y = 1m, z=1m

Autocorrelation

0.6

0.4

0.2

0
0

0.5

1.5

2.5

-0.2
Lag time (sec)
Fig. 9. Autocorrelations for boundary layers on the side of the ICE service train (Sterling et al., 2008b).

Fig. 10. Local skin friction coefcients for HST model and full scale tests of Brockie and Baker (1990).

Fig. 11. Boundary layer displacement thickness on 1/76th model scale HST roof from Brockie and Baker (1990).

281

ARTICLE IN PRESS
282

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

based pitot probes, hot wire probes etc. These tests have been
carried out at both full scale and model scale for a variety of train
types. From these experiments it is possible to derive standard
boundary layer parameters such as the displacement thickness
and the form parameter (although note that these are formally
derived for classical two dimensional boundary layers, rather than
the complex three dimensional ows found around trains). The
data from some of these experiments is summarised in Fig. 4
below, with data from the wind tunnel and full scale tests of
Brockie and Baker (1990) for the UK HST, and the data correlation
of model scale results given in Schetz (2001), reporting the earlier

Fig. 12. Boundary layer on train roofs from moving model tests for 1/25 scale ICE
model (Baker et al., 2001): (a) displacement thickness; (b) form parameter.

work of Sockel (1996) for a variety of other trains. All dimensions


given in these gures are the equivalent full scale values. Note
that the results of Schetz (2001) and Sockel (1996) are for the
actual, somewhat loosely dened, boundary layer thickness.
Schetz (2001) notes that the ratio between this thickness and
the displacement thickness is between 8 and 11, i.e. one order of
magnitude.
It can be seen that all the model scale results are broadly
consistent with each other, and show rstly a steady development
in the total boundary layer thickness and the displacement
thickness along the length of the train, and secondly values of the
form parameter that are signicantly below the value of 1.4 that
one would expect for an equilibrium boundary layer. The
measurements in Brockie and Baker (1990), together with a
consideration of the momentum integral equation, suggest that
the side wall (train vehicle side) boundary layer is very far from
two dimensional, with a divergence of the ow up the side of the
train and a convergence over the roof (see below). The full scale
HST results are however somewhat different, and show little
growth along the train, although the form parameters are
consistent with the model scale measurements.
A different method of obtaining information on the state of the
boundary layer on the train comes from measurements made
using stationary anemometers at the trackside or on platforms.
The measurements that were made on the German ICE have
already been described above (RAPIDE Consortium, 2001; Sterling
et al., 2008b). Fig. 5 shows the measurements that were made at
all positions along the train. The inviscid velocity peaks around
the nose described in the last section can be seen around x =0 m,
but the velocities in these peaks can be seen to be small in
comparison to the boundary layer velocities. At each distance
away from the train the velocity increases steadily along the train
up to the wake region around x = 350 m. (This region will be
discussed in detail below).
Fig. 6 shows the same data, but plotted in the conventional
boundary layer velocity prole format for different distances
along the length of the train. There can be seen to be a gradual
thickening of the boundary layer as x increases, as would be
expected.
Fig. 7 shows the boundary layer displacement thicknesses and
form parameters obtained from this data. In addition these

Fig. 13. Velocity traces beneath Korean high speed train (Kwon and Park, 2006).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

parameters are also given for a similar set of full scale


measurements made above a station platform, and for a set of
model scale experiments made at half train height without a
platform simulation (Sterling et al., 2008b). Note that for the
platform experiments, z is dened as the distance from the top of
the platform, and y the distance from the train side. It can be seen
that the displacement thicknesses in all three cases grow along
the length of the train, with the trackside full scale values being
larger than those for the other experiments. This is not surprising,
as the former measurements were taken in a region close to the
ground exposed to the aerodynamically rough bogies, whereas
the latter were obtained from regions closer to smoother areas of
the train. The form parameters are again signicantly less than
the equilibrium values as in Fig. 4. Perhaps the major point to
emerge from this data are the large values of displacement
thickness near the front of the vehicle in the full scale
measurements, suggesting a major ow disturbance around the
nose that is not replicated in the model scale measurements
shown in either Figs. 4 or 7. It could be conjectured that this is due

283

to the existence of an unrealistic laminar boundary layer near the


vehicle nose in the model tests that does not exist at the much
higher Reynolds numbers of the full scale tests.
Turbulent boundary layers such as those on the side of the
train are also characterised by their unsteady ow characteristicswith the magnitude of the turbulence being characterised
by the turbulence intensity, and the scale by parameters such as
the integral time or length scales. In terms of the ensemble
average data, velocity data from stationary probes that has been
obtained for the ICE service train, the turbulence intensity can be
approximated by (1  ensemble standard deviation)/ensemble
mean. Fig. 8 shows a plot of this parameter along the train for
both the measurements made at trackside and those made above
the platform at broadly equivalent positions. It can be seen that in
both cases the turbulence intensity is more or less constant along
the train, although, as would be expected, the value is
signicantly higher for the trackside measurements than for the
platform measurements. The values are of the order of 0.050.1,
which are typical values for at plate boundary layers. Fig. 9

Fig. 14. Vertical and horizontal velocity proles beneath Korean high speed train (Kwon and Park, 2006).

ARTICLE IN PRESS
284

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

shows the autocorrelations of velocity for these two cases. From


these plots the integral length and time scales (the scales that
contain the most turbulence energy) can be found to be 4.7 m and
0.067 s for the trackside measurements and 4.1 m and 0.059 s for
the platform measurements. The integral length scale is thus of
the order of 20% of the length of an individual carriage.
The nal boundary layer parameter that is of interest is the
skin friction coefcient (Cf), as the surface skin friction determines
to a large extent the overall drag of the train. Fig. 10 shows the
local skin friction coefcient for the HST model and full scale
results of Brockie and Baker (1990). It can be seen that, as is to be
expected, the skin friction is very dependent upon scale and
indicates the necessity for as large a scale as possible in either
wind tunnel tests or computations if the drag is to be accurately
predicted. It is of interest to note that most of the individual
values of local skin friction fall on or below the accepted smooth
wall correlations of skin friction and local Reynolds number for
ow over a two dimensional at plate, indicating again the nonequilibrium, three dimensional nature of the side wall boundary
layers. Similar coefcients could also be determined from the
model scale moving model measurements for the ICE reported in
Baker et al. (2001), through a t of the logarithmic law to the
velocity proles. The authors of this paper acknowledge that this
procedure was not an accurate one, but average values of 0.0029
were obtained, which is again somewhat below the smooth ow
at plate value at that Reynolds number. These values are broadly
consistent with those shown in Fig. 10. Similarly the full scale ICE
slipstream measurements can be used to obtain values of skin
friction, through the tting of a logarithmic law to the velocity
proles from which the shear velocity and thus the skin friction
coefcient can be determined. The log law is however a poor t to
many of the velocity proles and thus this procedure is an
approximate one. The average value for the measurements, above
the platform for the centre of the train from x = 100 to 300 m
where the log law t was reasonable is 0.0046, rather higher than
would be expected from the results presented above. For the
measurements at trackside, exposed to the rough bogies, the
equivalent skin friction coefcient is, predictably, much higher at
0.038.
2.2.2. Train roof
There is not a great deal of experimental data for the boundary
layer development over the train roof. Fig. 11 shows a compilation
of the conventional boundary layer measurements of Brockie and
Baker (1990). This data indicates a displacement thickness that is

considerably thicker than the corresponding side wall boundary


layers, which leads to the conclusion that there is a diverging ow
up the side of the wall and a converging ow over the roof of the
train.
Baker et al. (2001) describes some measurements made using
the alternative approach of stationary probes and a moving
model, for a 1/25th scale ICE vehicle. The results for displacement
thickness and form parameter are shown in Fig. 12. The x axis in
these gures is a normalised time, with the normalisation being
with vehicle speed and vehicle length. T= 0 is when the train nose
passes, and T= 4 is when the tail of the four coach train passes.
There can be seen to be little difference in the displacement
thickness results between the side wall data (already included in
Fig. 7) and the roof data, which is not consistent with the results
of Fig. 10. Again the form parameters are signicantly below the
equilibrium boundary layer value of 1.4.

2.2.3. Train underbody


The ow underneath high speed trains has not been extensively studied in the past, but has recently come to prominence as
a result of the ballast ying issue that can cause a variety of
problems to the train and to the track. The investigations into this
problem have resulted in a number of investigations taking place
which have measured the velocity and pressure eld beneath high

Fig. 16. Velocity proles measured below Shinkansen train (Ido et al., 2008). The
different curves indicate different sections along the train, with M1 being at the
front of the vehicle and M5 at the rear.

Fig. 15. Pressure coefcients beneath Eurostar train (Quinn and Hayward, 2008).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

speed trains. The results of a number of these investigations are


considered herethose of Kwon and Park (2006) for the Korean
high speed train, Quinn and Hayward (2008) for the Eurostar
trains on the Channel Tunnel Rail Link, and the experiments
reported in Ido et al. (2008) for the Japanese Shinkansen train.
Fig. 13 shows the velocities beneath the Korean train for a number
of heights, measured at full scale with pitot probes. The train
speed was 300 kph, about 83 m/s. The velocities at various
distances from the upper rail surface are shown. It can be seen
that these velocities reach high levels of around 40% of the train

285

speed at heights of 0.18 m. A small peak in the velocities can be


seen as the train nose passes, but for most of the traces the
velocities are highly turbulent and uctuating as would be
expected for the rough under train environment. Fig. 14 shows
vertical and horizontal velocity proles for the same tests. The
results for the vertical prole imply some sort of annular ow
prole (with an inexion point between the track and the train),
rather than a conventional boundary layer prole. The velocities
can also be seen to peak on the train centre line and fall off
towards the outside of the track.

Fig. 17. Helical vortices in the wake of trains (Baker et al., 2001; Sockel, 1996; FLUENT).

ARTICLE IN PRESS
286

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

The velocity and pressure elds were measured beneath the


Eurostar train and are reported in Quinn and Hayward (2008). The
velocity traces are similar to those described above and are not shown
here. The pressure trace is shown in Fig. 15, both for the entire train
and for the rst quarter of the train. One unit of normalised time
represents the passage of one carriage of the train. These are
ensemble averages of 20 nominally identical train passes. A
prominent nose maximum and minimum can be seen in both
traces that is similar to those shown in Section 2.1. However regular
minima are also seen along the vehicle, that seem to be associated
with the passage of the middle of each coach between the bogies. A
tail peak can also be seen. From the Eurostar velocity data it was also
possible to obtain autocorrelation functions, and thus the integral
length scales and time scales. The length scales were found to be
between 1.8 and 2.0 m, and the time scales between 0.02 and 0.03 s.
These values are rather smaller than those for the side boundary
layer, as may be expected because the ow is signicantly
constrained.
The ow eld beneath a full scale 16 car Shinkansen test train
and a three car wind tunnel model are reported in Ido et al. (2008).
The wind tunnel tests were carried out with a moving ground belt
beneath the middle car only, and this may have some inuence on
the results that are shown in Fig. 16 below. The results of
Fig. 16 look rather different from those of Fig. 14, but this is
because of the frame of reference that is used and in fact they
are identical in form. Those of Fig. 16 show clearly the inexion
point prole that could only be inferred from Fig. 14. There is
good agreement between the model scale and full scale data. Ido
et al. (2008) also goes on to look at a number of different
underbody geometriesbogie fairings etcand shows that the
underbody ow is, unsurprisingly, quite sensitive to geometric
changes.
Finally, similar information is also presented for tests on the ICE in
Kaltenbach et al. (2008). However the velocity prole information
has been normalised in an arbitrary way, and is not amenable
for comparison with the other results presented here. It is also of
interest to note that the ballast ying problem has also driven wind
tunnel and computational research on the ow in and around bogies
(Ido et al., 2008; Kaltenbach et al., 2008). This work is at an early
stage, and is very challenging both experimentally and computationally, with the results being very sensitive to both the experimental set up and the computational conguration (including the
turbulence model). Further progress can be expected in this area in
the future.

2.3. The train wake


In the last few decades there has been a great deal of work
carried out on the nature of the wakes of road vehiclesprimarily
with a view to being able to minimise drag which is of course
dominated by wake effects for road vehicles. It is clear from the
literature that the precise nature of the wake varies from vehicle
to vehicle, but there seem to be a relatively small number of ow
mechanisms that can exist in some combination or othershear
layer separations, longitudinal helical ows, vortex streets and a
separation cavity (Vino et al., 2005; Sims-Williams et al., 2001;
Nouzawa et al., 1992). All of these phenomena are subject to
instabilities with Strouhal numbers (based on vehicle velocity and
some representative frontal dimension) between 0.05 and 0.4.
Now whilst the tail shape of trains is rather different to those of
cars, one might expect equally complex ows. The major physical
difference will be that the thickness of the boundary layer around
the train will be relatively much greater than for cars, and thus
any separated shear layers will also be much thicker.
With these points in mind, let us now consider the experimental data and computational work that is available to describe
the wakes of high speed trains. Fig. 17 shows experimental
velocity vectors for a model TGV (Paradot et al., 1999) and a
model ICE (Baker et al., 2001), together with a visualisation of the
ow around an ICE using standard RANS CFD methods (FLUENT).
It can be seen in each case that there is strong evidence of helical
vortices behind the train, which extend a considerable distance

Fig. 19. ICE 1/25th model scale wake measurements (y is the distance from the
side of the train, h is the train height) (Baker et al., 2001).

Fig. 18. Computations of wake oscillations for the ICE (Schulte-Werning et al., 2003).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

into the vehicle wake. Schulte-Werning et al. (2003) takes the CFD
work somewhat further and investigates the unsteadiness of
these helical structures through the use of unsteady RANS
techniques. The results are illustrated in Fig. 18. A well dened
oscillation can be seen with a Strouhal number of 0.14.
Now the full scale slipstream data presented in Sterling et al.
(2008b) has been analysed intensively to attempt to determine
the wake characteristics. Fig. 5 shows the ensemble average
velocity traces, and for those nearest to the train there can be
seen to be a noticeable peak in the velocity in the near wake (i.e.
350400 m). These velocities were measured 1 m above the track
with no platform in place. The equivalent model scale experiments of Baker et al. (2001), measured half way up body height,
do not show this peak. The experiment and simulation of Fig. 17
indicate that the helical vortex occurs close to the ground, and it
thus seems likely that this is what is observed in the full scale
slipstream measurements. Further investigation of the full scale
data however showed that the technique of ensemble averaging
was, in this case, actually hiding a physical effect, and that this
peak did not appear on around half of the individual velocity
traces that were used. Careful re-analysis revealed that the
slipstream measurements were picking out some type of vortex
shedding oscillation with a Strouhal number of 0.11, which is
close to the computational value of 0.14 reported above. It thus
seems that the helical vortices in the train wake undergo

y/h = 0.033

0.1

0.05

0
0

50

100

287

some sort of regular oscillation with a Strouhal number of around


0.110.14.
The situation is further complicated however when the model
scale data of Baker et al. (2001) is considered in more detail. The
wake velocities just behind the train showed a very great deal of
run to run variation with high ensemble standard deviations. A
wavelet investigation of the turbulence scales suggested peaks at
two Strouhal numbers0.03 (hypothesised to be due to wake
cavity pumping) and 0.5 (taken to be due to shear layer
instability). As the model measurements were made higher up the
train than the full scale measurements, it may be that the
unsteady helical vortex motion is more important close to the
ground, and other types of unsteadiness higher up the vehicle.
Further work is required in this area.
Up to this point we have implicitly been considering the near
wake of trainswithin a few vehicle heights of the train tail. One
would expect that any large scale ow structures that exist would
decay fairly rapidly and that the far wake of the train would show
a gradual decrease in velocity. This effect can be seen in the
ensemble average velocity traces for the ICE in Fig. 5 for a value of
x4420 m (the tail of the train is at x= 365 m). The velocity
measurements made nearest the train decay steadily, whilst the
measurements made furthest away rst increase and then
decreaseshowing the lateral spread of the wake. This effect
can also be seen in the model scale wake measurement for the ICE
of Fig. 19 below Baker et al. (2001). The x axis variable is a time
normalised by the vehicle speed and the length of a single car of
the four car model train. In this case the origin is at the train tail. A
model for the longitudinal, lateral and vertical velocities in this
decaying wake was developed in Baker (2001a) based on the
similarity method of Eskridge and Hunt (1979). The expressions
for the velocities are simple algebraic functions of the distance
along and across the wake, and governed by two parameters.
Fig. 20 shows the best t curves to the model scale ICE data. The
agreement can be seen to be good, and suggest that the wake
velocities are self similar when expressed in a suitably
dimensionless format.

150

T
3. The ow around trains with a cross wind

y/h = 0.1333

0.1

3.1. The natural wind relative to a train

0.05

0
0

50

100

150

T
y/h = 0.40

0.1

0.05

0
0

50

100

150

T
Fig. 20. Best t curves to wake velocities using model of Baker (2001a)
((a) y/h= 0.033, (b) y/h= 0.2, (c) y/h=0.533).

In this section we consider how the natural wind appears to a


train moving though it, for the fundamental case of a train moving
over a at ground. We assume for the sake of simplicity that the
wind is normal to the train track and that the vertical velocity
prole is given by the usual logarithmic law with a surface
roughness length of z0 ESDU (1985). We further assume that the
turbulence intensity is given by the values given in that same
reference, and is again a function of the surface roughness length.
To calculate the velocity prole and turbulence intensity with
respect to a train the train velocity must also be considered. This
is straightforward and the method has been frequently rehearsed
in the past (Baker, 1991a for example) and one can easily obtain
the velocity prole, yaw angle, and turbulence intensity relative
to a moving train. Sample calculations are shown in Fig. 21 below
for a cross wind speed u, 3 m above the ground of 20 m/s and
various train speeds (n), for a value of z0 of 0.03 m, typical of a
rural upstream fetch. The turbulence intensity with respect to the
train is dened as the atmospheric turbulence level divided by the
wind velocity relative to the train. It can be clearly seen that at
low train speeds, the velocity prole takes on the appropriate
boundary layer form, with a signicant velocity variation across
the train height, the yaw angle is close to 901 and the turbulence
intensity is high. At high train speeds, the velocity prole is more

ARTICLE IN PRESS
288

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Height (m)

5
v = 0 m/s

v = 20 m/s

v = 40 m/s
v = 60 m/s

v = 80 m/s

1
0
0

0.5
1
1.5
Normalised velocity relative to train

Height (m)

5
v = 0 m/s

v = 20 m/s

v = 40 m/s

v = 60 m/s

v = 80 m/s

0
0

20

40

60

80

100

Yaw angle relative to train (degrees)


6

Height (m)

5
v = 0 m/s

v = 20 m/s

v = 40 m/s
v = 60 m/s

v = 80 m/s

1
0
0

0.1
0.2
0.3
0.4
Turbulence intensity relative to train

Fig. 21. The wind characteristics seen by a moving train over a at ground
(u= 20 m/s).

uniform, and the yaw angle is low and varies somewhat over the
height of the train. The turbulence intensity is lowof the order
of a few percent. Clearly these results have implications for the
type of wind tunnel or computational simulation that is usedif
there is a desire to model trains moving at high speeds, then these
tests can reasonably be carried out in low turbulence wind
tunnels, with no velocity shear (although the yaw angle twist
would not be simulated). If low vehicle speed conditions are
required, then atmospheric turbulence and shear needs to be
simulated, although it needs to be recognised that any simulation
will only model one specic set of wind speed/vehicle speed
conditions. The same comments can be made for the critical case
of a train on an embankment, where the wind prole speeds up
close to the ground, although in this case the wind shear and
turbulence intensities are rather less in all cases.
As well as the mean velocity proles and the turbulence
intensity, the turbulence spectrum experienced by a moving train
will be different to that experienced by a stationary train. This
was investigated in Cooper (1985) and typical results are shown
in Fig. 22. In this gure the spectra are shown for a range of train
speed/wind speed ratios from 0 to innity, for a pure cross wind.
The x axis is a frequency normalised with the atmospheric
turbulence length scale and the wind velocity relative to the train,
and the y axis is the spectral density normalised with the
frequency and the wind velocity variance. Plotted in this way the
spectra show a remarkable level of similarity, which thus implies
that, to a rst approximation, they scale on the wind velocity
relative to the train.
Whilst the above gures give some indication of the wind
statistics relative to the train, it is not unusual when looking at
train cross wind stability to specify, in varying degrees of detail,
an extreme wind gust, on the basis that such wind gusts will
cause train stability problems. The question thus arises as to the
nature of these gusts, which vary both in space and in time. The
author and his co-workers have in recent years investigated a
number of velocity and surface pressure datasets obtained on the
University of Birmingham Wind Engineering eld site at Silsoe in
Bedfordshire, and, of particular relevance to the current discussion, have adopted the technique of ensemble averaging of
extreme gusts i.e. identifying the gusts in the time series and

Fig. 22. Wind spectra relative to a moving train over a at ground (Cooper, 1985).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

289

Fig. 24. The Chinese hat gust prole (CEN, 2007).

Fig. 23. Extreme gust proles (Sterling et al., 2006) (ds indicates dataset number;
velocities normalised by the corresponding standard deviation): (a) streamwise
velocity component; (b) lateral velocity component; (c) vertical velocity
component.

then averaging the gusts to produce an average time series. Fig. 23


shows the average streamwise, lateral and vertical velocity
uctuations (normalised by the standard deviation) measured
1 m above ground level at the site (Sterling et al., 2006) (although
the results are similar to those measured at a range of heights up
to 10 m). It can be seen that the peak streamwise gusts, have a
magnitude of about three to four times the standard deviation of
the wind velocity uctuations and last for about one second either
side of the maximum peak. These gusts are superimposed upon a
longer term gust with a magnitude of around one standard
deviation above the mean. The corresponding vertical uctuations
show a negative peak at the maximum gust, indicating that these
gusts are associated with sweep events in the atmospheric
boundary layer. In terms of the spatial correlation of such gusts,
the results show that the short term peaks are only correlated
over heights of a few metres, and extreme pressure coefcients

measured on a 2 m high horizontal wall at the same site (Baker,


2001b), also indicate that the horizontal correlation of such peaks
is similar. However the longer term, underlying and less intense
peaks, seem to be correlated over lengths of tens of metres.
There are two general approaches in dealing with extreme
gusts in crosswind stability considerationsthe rst is to dene a
time period for these gusts of, say 1 or 3 s, on the basis that such a
gust period would be required for any train to blow over. It can be
seen that such an approach would signicantly lter the short
term gust peak shown in Fig. 23. Another approach is to dene a
specic type of peak, such as the Chinese hat used in CEN (2007)
and shown in Fig. 24. This is a spatially distributed peak only, and
lacks the temporal variation shown in Fig. 24, although its spatial
spread of around 100 m either side of the peak corresponds
roughly to the longer term base peak referred to above. This
gust prole has been obtained from a study of the time and spatial
averaged wind statistics and as such reects all the information
within these statistics rather than just the information describing
the extreme gusts. Note further that when it is applied a spatial
lter of one vehicle length is used. Either of the above approaches
is clearly a simplication of the nature of the gust, although both
have practical merits.
An alternative approach to using a simplied extreme gust has
been outlined in Ding et al. (2008) and Baker et al. (2008b) where
a correctly spatially and temporally correlated wind eld has been
simulated to match the required wind correlations. The wind
velocities seen by a train as it passes through this simulated wind
eld will then contain the necessary spatial and temporal
information concerning the extreme gusts in a more realistic way.

3.2. Wake ow structure in cross winds


In the last few decades a number of wind tunnel tests and
computational simulations have elucidated the ow structure
around high speed trains in cross winds. The wind tunnel work of
Mair and Stewart (1985), Copley (1987). Chiu (1991), Robinson
and Baker (1990) in the 1970s and 1980s used an idealised train
model, and it was found that the dominant ow pattern was a
series of inclined wake vortices such as might be found around a
missile at high angles of attack (Fig. 25a). At higher yaw angles a
more conventional vortex shedding pattern was observed, with
some switching of the ow at intermediate angles. Recent
computations, using both RANs and LES techniques on various
ICE congurations (Hemida, 2006; Wu, 2004; Diedrichs, 2005)
have also shown similar patterns (Figs. 25bd). There is broad
agreement both qualitatively and quantitatively between the

ARTICLE IN PRESS
290

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Fig. 25. Wind tunnel tests and CFD computations of wake vortex ows behind trains in a cross wind. (aidealised train (Copley, 1987); b2 car ICE (Hemida, 2006);
cICE (Diedrichs, 2005)).

Fig. 26. Computed and measured surface streamline patterns (Chiu, 1991;
Hemida, 2006).

wind tunnel tests and the CFD resultssee the surface streamline
patterns from Chiu (1991), Hemida (2006) in Fig. 26. There is
some evidence, particularly from the LES results of wake
unsteadiness (Hemida, 2006) that indicates two modes of wake
unsteadinessa horizontal wake oscillation with a Strouhal
number of 0.1, and a weak vortex shedding motion with a
Strouhal number of 0.150.2. It has to be said however that these
frequencies are not always well dened in the LES simulations,
and do tend to vary somewhat with the type of calculation used.
Measurements of a different kind are reported in Baker et al.
(2007) which presents data from moving model experiments on a
four car ICE train, with rather a crude cross wind generator placed
normal to the track. The x axis unit is time normalised by train
velocity and carriage lengthwith zero being the point at which

the train nose passes the measurement point, and 4 being the
point at which the train tail passes that point. The y axis is the
slipstream velocity normalised by the train velocity. The slipstream velocities were measured in the wake of the vehicle and
ensemble averaged in a manner that has been previously
described. These results are shown in Fig. 27 in two
formatsone for the velocities themselves (Fig. 27a) and one
for the velocities minus the velocities measured with no cross
wind (Fig. 27b). In the rst of these the crosswind magnitude can
be seen before the train nose has passed, and the inviscid nose
peak is clearly visible. This is followed by a dip in the velocities,
due to the sheltering effect of the train. A maximum in the
velocity can then be seen before a gradual decay. In the wake the
crosswind velocities are again seen. In the alternative method of
presentation any value of the relative velocity that exceeds the
upstream wind speed indicates an enhancement of that wind
speed by the train wake. This can be seen to occur at a position
that corresponds to the passing of the second car of a four car
vehicle, with lower values elsewhere. Baker et al. (2007) argues
that this is consistent with the presence of inclined vortices in the
wake of the train. Finally Fig. 28 shows the maximum wake
velocities (one second averages) that were measured for a wide
variety of high speed trains in full scale test in the UK, normalised
by the train speed. There is, inevitably, a great deal of scatter but it
can be seen that there is a general increase in the maximum
normalised velocities as the cross wind speed increases.

3.3. Pressure distributions in cross winds


The idealised train low turbulence wind tunnel measurements
of Mair and Stewart (1985), Copley (1987), Chiu (1991) made
extensive measurements of the pressure distributions around the

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

291

Normalised mean velocities -with cross wind


0.5
0.4
0.125m

0.3

0.75m
0.2

2.0m

0.1
0
-2

0.3

Normalised velocity

0.2
0.125m
0.1

0.75m
2.0m

0
-2

-0.1
Normalised time
Fig. 27. Wake velocity measurements from moving model experiments for ICE model (Baker et al., 2007).

Normalised velocity

0.400

0.300

0.200

0.100

0.000
0

4
Wind speed (m/s)

Fig. 28. Maximum slipstream velocities for high speed passenger trains (Baker et al., 2007).

model, and sample results are shown in Fig. 29a Chiu (1991)
together with the results of simple panel method calculations
Fig. 29b shows equivalent measurements for a 2 car ICE 2 model
(Wu, 2004), which also shows the LES calculations of Diedrichs
(2005). Pressure coefcients are shown on loops around the
vehicle, for a variety of different distance (x) from the train nose,
normalised with the model diameter D or length L. The results are
plotted with a negative pressure coefcient in the positive
direction. For all values of x that are not near the nose, it can be
seen there is a suction peak on the windward roof corner (451 in
Fig. 29a, 315 degrees in Fig. 29b), small suctions over the rest of
the roof, leeside and underside, and a positive pressure coefcient
on the windward wall. Near the nose however, for small values of
x there is a suction peak on the leeward side, that can be expected

to give a major contribution to the overall side force. In general


the numerical solutions show a good level of agreement with the
experiments.
A similar set of results, from Baker et al. (2008a) are shown for
a wind tunnel model of a UK Class 365 electrical multiple unit
(e.m.u.) in Fig. 30. These measurements were made in a
simulation of the atmospheric boundary layer i.e. in a highly
turbulent ow. They are plotted with the positive pressure
coefcient in the positive direction in a manner similar to, but
confusingly different from, Fig. 29. The pressure coefcients are
shown on loops around the vehicle (with loop B being near the
front of the leading vehicle and loop H near the back of that
vehicle. Results are shown for yaw angles of 451 and 901 (results
for other yaw angles are given in Baker et al., 2008a). The large

ARTICLE IN PRESS
292

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Fig. 29. Pressure distributions around trains (aIdealised train, lines are panel method results and points are experimental results (Chiu, 1991), front of windward face is
at 01; b2 car ICElines are LES results of Hemida (2006), points are experimental results of Wu (2004), front of windward face is at 2701).

suction peak on the windward corner can again be seen for all of
the loops at both yaw angles. Fig. 30 also shows the standard
deviation of the pressure coefcients, which gives an indication of

the steadiness or otherwise of the ow. It can be seen that this


parameter peaks at the windward roof corner, perhaps indicating
some unsteady separated ow in this region. Fig. 31 shows a

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

293

Fig. 29. (Continued)

proper orthogonal decomposition (POD) analysis of the pressure


detail. A full description of this type of analysis is given in Holmes
et al. (1997), Baker (2000). It shows that the rst and most
energetic POD mode seems to be concentrated at the windward
roof of the vehicle, and thus physically corresponds to the
pressure uctuations on the windward roof. The second mode
has the characteristics of the mean pressure distribution and thus
physically corresponds to quasi-steady pressure uctuations
around the model. A fuller discussion of this analysis is given in
Baker et al. (2008a).

3.4. Aerodynamic admittances and weighting functions


The overall forces on trains are composed of the sum of the
mean and uctuating pressures over the train surface. One would
expect that small scale uctuations that only affect small parts of
the surface of the vehicle would be damped out in this summation
(integration) process, and thus the force uctuations would be
smaller than the upstream velocity or pressure uctuations. This
effect is important in calculating cross wind stability in a number
of related ways. If calculations of stability are carried out in the
frequency domain to use the terminology of Baker (1991b, c) (as
in Xu and Ding, 2006, for example) then the force coefcient
spectra can be obtained from the wind spectra through the use of
aerodynamic admittance functions, which are effectively normalised ratios of force coefcient spectra to wind spectra. Fig. 32

shows the curves for aerodynamic admittance from the work of


Cooper (1985) that corresponds to the velocity spectra relative to
the vehicle shown in Fig. 22. Aerodynamic admittance values are
plotted against normalized frequency, for a ratio of train speed to
normal wind speed of 4.0, and for two different vehicle heights. It
can be seen that all the curves (for different ratios of the vehicle
length to the turbulence integral length scale) all approach unity
at low frequencies (showing that large turbulent eddies are
correlated over the entire vehicle) but fall rapidly at high
frequencies for the reason set out above, and thus will act as a
lter on high frequency wind uctuations, and thus the energy
available to excite suspension frequencies that lie in this range.
Similar experimental curves, from the Class 365 e.m.u. results
mentioned above are shown in Fig. 33 Sterling et al. (2008a). The
reduced frequency is the actual frequency normalized by the wind
tunnel velocity and the carriage length. It can be seen that some of
the experimental curves tend to values other than unity at low
frequencies. Sterling et al. (2008a) suggest that this is because of a
number of reasonsprimarily because the concept of admittance
implicitly assumes that the ow is not signicantly affected by
the body on which the forces are measured, and because the
streamline passing through the reference point is not the
signicant streamline dened as the streamline on which the
turbulent characteristics mostly inuence the unsteady forces on
the body. Sterling et al. (2008a) shows that this streamline is
usually signicantly below the reference position of 3 m, which is
not unreasonable. That being said, as long as the same reference

ARTICLE IN PRESS
294

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Fig. 30. Mean and standard deviations of pressure coefcient distributions around a model class 365 e.m.u: (a) 901 yaw; (b) 451 yaw.

Fig. 31. POD analysis of uctuating pressures around a model class 365 e.m.u. (Baker et al., 2008a).

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

295

6
5
4

Side force

Lift force

2
1
0

0.2

0.4
0.6
Time (sec)

0.8

6
Fig. 32. Aerodynamic admittances from Cooper (1985) (L = vehicle length,
H= vehicle height, xLu =turbulent length scale, VT = train speed, x axis is frequency
normalised with turbulence length scale and wind velocity relative to the train).

5
4

Side force

Lift force

2
1
0

0.2

0.4
0.6
Time (sec)

0.8

Fig. 34. Aerodynamic weighting functions for the class 365 e.m.u (left v= 55.9 m/s,
u= 15 m/s, right v= 17.9 m/s, u= 15 m/s).

Fig. 33. Experimental measurements of aerodynamic admittances for Class


365 e.m.u side force coefcients (x axis is frequency normalised with length of
vehicle and wind velocity relative to the train).

position is used in any model tests in which the admittances were


measured, and in any calculation using these admittances, then
the calculations will be valid. Now assuming suitable values of the
physical parameters that determine the admittance functions (i.e.
wind and vehicle velocities, vehicle length and turbulence length
scales, indicates that the admittance falls to values of 0.5 at
around 12 Hz, suggesting that wind uctuations above these
values will be ltered out by the lack of correlation of surface
pressures.
Now whilst this frequency domain analysis is useful, in any
calculation of train behaviour that considers suspension dynamic
effects including discrete track irregularities, the time domain
equivalent is required Baker et al. (2008b). This is known as the
aerodynamic weighting function and can be shown to be the
Fourier transform of the aerodynamic admittance, and vice versa.
The derivation of such functions from experimental data is far
from straightforward, and is discussed in detail in Baker et al.
(2008b). Typical values are given in Fig. 34 for side and lift force
coefcient weighting functions for the class 365 e.m.u referred to
above. Effectively these functions weight the relative wind
velocities and train force coefcients over a period leading up to
the time of application. The main points to note about these
functions are rstly that the values of weighting function for the
high speed case are smaller than for the low speed case, which
reects the fact that the wind velocity uctuations relative to the
train are smaller the faster the train velocity, and secondly that

they are only non-zero for values less than around 0.5 s i.e. the
train side and lift force coefcients are fully determined by the
values of the relative wind velocity over the previous 0.5 s. These
effects are illustrated in the results of Fig. 35, which show wind
time series generated by the method of Ding et al. (2008) and the
side forces and the lift forces for the class 365 corresponding to
the weighting functions shown in Fig. 34. The ltering effect of the
higher frequency uctuations in velocity is very clear, particularly
at the higher vehicle speeds.
Finally if calculations are carried out in the amplitude
domain, then the extreme values of the force coefcients,
based on extreme values of the forces and extreme values of the
upstream velocities, might also be expected to be less than the
mean values of these coefcients, as the high frequency uctuations will have been ltered out. This effect is illustrated in Fig. 36,
for the static wind tunnel model tests carried out on a UK Class
390 Pendolino train (Baker et al., 2004) for the extreme
coefcients based on a gust averaging time of 1 s. At rst sight
the fact that the extreme/mean ratio is signicantly below unity
seems inconsistent with the admittance and weighting function
results above which suggest that such effects due to lack of
correlation of surface loads, should be conned to averaging times
less than 0.5 to 1.0 s. However more recent calculations carried
out by the author suggest that the results of Fig. 36 are affected, to
a signicant degree, by the specication of wind velocities at the
reference streamline rather than on the signicant streamline
dened above. Nonetheless, as with the admittances, if the same
reference height is used in any stability calculations as was used
in the experiments that measured extreme values, then the
results will still be valid.

4. Concluding remarksimplications for practical issues


In this section some brief remarks are made on the implications of the ow eld descriptions of previous sections for a
number of practical issues of current concern.

ARTICLE IN PRESS
296

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Fig. 35. Wind velocity and force time histories for the class 365 e.m.u (left v= 55.9 m/s, u= 15 m/s, right v= 17.9 m/s, u= 15 m/s).

Fig. 36. Ratio of extreme to mean side force coefcients for the Class 390
Pendolino full scale experiments (Baker et al., 2004)legend indicates different
train/track congurations.

(a) It is clear that the boundary layer measurements made on


model vehicles are not fully consistent with those made at full
scale, partly due to the expected effects of scale (i.e. Reynolds
number) and partly because at full scale there would appear
to be a very rapid growth of the boundary layer near the nose.
These effects will affect the skin friction drag, and thus the

overall drag on the train, and the nature of any model scale
slipstream measurements in the boundary layer region. More
work is required in this area to determine the appropriate
way of representing the train boundary layer at model scale in
both wind tunnel and moving model experiments.
(b) The integral time scale at the train side is less than 0.1 s, and
thus ow unsteadiness in this region is unlikely to have any
effect on the stability of passengers or trackside workers, who
have a minimum response time of around 0.3 s (Jordan et al.,
2008). Instability in this region will rather be caused by mean
ow effects. (Note however that this conclusion is only true
for smooth high speed passenger trainSterling et al. (2008b)
shows that for freight trains the integral time scales on the
train side are rather larger and turbulent ows along such
trains could affect passenger and trackside worker stability.)
(c) The integral time scales in the underbody gap are very short at
0.02 to 0.03 s, with associated integral length scales of the order
of 2 m. Unpublished calculations of ballast ight paths beneath
the train by the author suggest time scales of the order of 0.3 s
and path lengths of the order of 3 m. This mismatch of times and
scales suggest that the ight path of train ballast will be
primarily determined by the mean ow eld beneath the train.
(d) The large scale unsteady ow structures in the near wake of the
train have time and length scales that are potentially hazardous
to waiting passengers, and in addition are a potentially large

ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

cause of train drag. It is possible that careful optimisation of train


nose/tail design could reduce the intensity of these vortices and
help in the alleviation of both issues.
(e) Trains in cross winds experience quite different wind conditions
(in terms of shear and turbulence) when they are operating at full
speed than when they are operating at low speed. There is a
divergence of view as to which effects are most important i.e. is
the critical condition when trains are moving at line speed, as
trains ought to operate in all weather conditions, or can it be
expected that in such conditions effects external to the railway
environment (such as ying debris) will cause trains to operate at
low speeds, and potentially to come to a stand? The answer to
this question will determine which type of wind tunnel test is
most appropriate to determine cross wind forces and moments
for the purposes of risk calculations.
(f) The longitudinal vortices in the wake of trains in cross winds can
cause an enhancement of slipstream velocities. It is possible that
the cross wind condition represents the most serious safety risk
when considering the effects of train slipstreams.
(g) When modelling the effects of cross winds on trains, if short
period gusts are modelled with time scales of less than 0.5 s,
or if suspension effects with frequencies greater than 1 or 2Hz
are modelled, then it is necessary to include the ltering
effects represented by the aerodynamic admittances or
aerodynamic weighting functions. If effects with rather larger
time scales are being considered (say the gross overturning of
a vehicle with a time scale of 2 to 3 s), then the force
uctuations on the train can be assumed to be quasi-steady
and to follow the wind velocity uctuations.

Acknowledgements
Much of the material presented in this paper has been taken
from the results of the authors collaborators over the last two
decades (research students and fellows, former and present
colleagues), who are too numerous to name individually. Nonetheless their implicit contribution to this work is gratefully
acknowledged.

References
Baker, C.J., 1991a. Ground vehicles in high cross windsPart I Steady aerodynamic
forces. Journal of Fluids and Structures 5, 6990.
Baker, C.J., 1991b. Ground vehicles in high cross windsPart 2 Unsteady
aerodynamic forces. Journal of Fluids and Structures 5, 91111.
Baker, C.J., 1991c. Ground vehicles in high cross windsPart 3 The interaction of
aerodynamic forces and the vehicle system. Journal of Fluids and Structures 5,
221241.
Baker, C.J., 2000. Aspects of the use of the technique of orthogonal decomposition
of surface pressure elds. Wind and Structures 3, 2.
Baker, C.J., 2001a. Flow and dispersion in vehicle wakes. Journal of Fluids and
Structures 15 (7), 10311060.
Baker, C.J., 2001b. Unsteady wind loading on a wall. Wind and Structures 4 (5),
413440.
Baker, C.J., Dalley, S.J., Johnson, T., Quinn, A., Wright, N.G., 2001. The slipstream and
wake of a high speed train. Proceedings of the Institution of Mechanical
Engineers F Journal of Rail and Rapid Transit 215, 8399.
Baker, C.J., Lopez-Calleja, F., Jones, J., Munday, J., 2004. Measurements of the cross
wind forces on trains. Journal of Wind Engineering and Industrial Aerodynamics 92, 547563.
Baker, C.J., Sterling, M., Johnson, T., Figura-Hardy, G., Pope, C., 2007. The effect of
crosswinds on train slipstreams. In: International Conference on Wind
Engineering, Cairns, Australia.
Baker, C.J., Sterling, M., Bouferrouk, A., ONeil, H., Wood, S., Crosbie, E., 2008a.
Aerodynamic forces on multiple unit trains in cross winds. In: Proceedings of
the BBAA VI, Milano, Italy, July 2024.
Baker, C.J., Bouferrouk, A., Perez, J., Iwnicki, S.D., 2008b. The integration of cross
wind forces into train dynamic calculations. World Congress on Rail Research,
Seoul, S Korea.
Brockie, N.J.W., Baker, C.J., 1990. The aerodynamic drag of high speed trains.
Journal of Wind Engineering and Industrial Aerodynamics 34, 273290.

297

CEN, 2007. Railway applicationsAerodynamicsPart 6: requirements and test


procedures for cross wind assessment. DRAFT prEN 14067-6.
Chiu, T.W., 1991. A two dimensional second order vortex panel method for the
ow in a cross wind over a train and other 2D bluff bodies. Journal of Wind
Engineering and Industrial Aerodynamics 37, 4364.
Cooper, R.K., 1985. Atmospheric turbulence with respect to moving ground vehicles.
Journal of Wind Engineering and Industrial Aerodynamics 17, 215238.
Copley, J.M., 1987. The 3-D ow around railway trains. Journal of Wind
Engineering and Industrial Aerodynamics 26, 21.
Diedrichs, B., 2005. Computational methods for crosswind stability of railway
trains. KTH Engineering Sciences report TRITA AVE 2005:27.
Ding, Y., Sterling, M., Baker, C.J., 2008. Train stability in cross winds, a new
approach? Proceedings of the Institute of Mechanical Engineers Part F: Journal
of Rail and Rapid Transport 222 (1), 8597.
ESDU, 1985. Characteristics of atmospheric turbulence near the ground: part 2
Single point data for strong windneutral atmosphere, Data Item 85020,
Engineering Sciences Data Unit, London.
Eskridge, R.E., Hunt, J.C.R., 1979. Highway modelling part 1: Prediction of velocity
and turbulence elds in the wakes of vehicles. Journal of Applied Meteorology
18, 387400.
FLUENT. Unsteady ow behind a high speed train by C Heine and G Matschke,
FLUENT web site; /http://www.uent.com/about/news/newsletters/02v11i1/
a8.htmS.
Hemida, H.N., 2006. Large eddy simulation of the ow around simplied high
speed trains under side wind conditions, Licentiate of Engineering thesis,
Chalmers University of Technology, Goteborg, Sweden.
Holmes, J.D., Sankaran, R., Kwok, K.C.S., Syme, M.J., 1997. Eigenvector modes of
uctuating pressure on low rise building models. Journal of Wind Engineering
and Industrial Aerodynamics 6971, 697707.
Ido, A., Saitou, S., Nakade, K., Iikura, S., 2008. Study on under-oor ow to reduce
ballast ying phenomena. World Congress on Rail Research, Seoul, Paper
number S2.3.4.2.
Jordan, S.C., Johnson, T., Sterling, M., Baker, C.J., 2008. Evaluating and modelling the
response of an individual to a sudden change in wind speed. Building and
Environment 43, 15211534.
Kaltenbach, H.-J., Gautier, P.-E., Agirre, G., Orellano, A., Schroeder-Bodenstein, K.,
Testa, M., Tielkes, Th., 2008. Assessment of the aerodynamic loads on the
trackbed causing ballast projection: results from the DEUFRAKO project
Aerodynamics in Open Air (AOA), World Congress on Rail Research, Seoul,
paper number S2.3.4.1.
Kwon, H.B., Park, C.S., 2006. An experimental study on the relationship between
ballast ying phenomenon and strong wind under high speed train. World
Congress on Rail Research Montreal, Paper T3.3.2.3.
Mair, W.A., Stewart, A.J., 1985. The ow past yawed slender bodies, with and
without ground effects. Journal of Wind Engineering and Industrial Aerodynamics 18, 301.
Mancini, G., Malfatti, A., 2001. Full scale measurements on high speed train ETR
500 passing in open air and in tunnels of Italian high speed line. TRANSAERO A
European Initiative on Transient Aerodynamics for Railway System Optimization (Notes on Numerical Fluid Mechanics and Multidisciplinary Design
(NNFM)) by Burkhard Schulte-Werning, Remi Gregoire, Antonio Malfatti, and
Gerd Matschke, pp. 101122.
Nouzawa, T., Hiasa, K., Nakamura, T., Kawamoto, A., Sato, H., 1992. Unsteady wake
analysis of the aerodynamic drag of a notchback model with critical afterbody
geometry. SAE Congress, Detroit, Paper 929202.
Paradot, N., Talcotte, C., Willaime, A., Guccia, L., Bouhadana, J.-L., 1999.
Methodology for computing the ow around a high speed train for drag
estimation and validation using wind tunnel experiments. World Congress on
Rail Research, Tokyo.
Quinn, A., Hayward, M., 2008. Full scale aerodynamic measurements underneath a
high speed train. In: Proceedings of the BBAA VI, Milano, Italy, July 2024.
RAPIDE Consortium, 2001. Synthesis report of RAPIDE project. Aerodynamics
workshop, Koln.
Robinson, C.G., Baker, C.J., 1990. The effect of atmospheric turbulence on trains.
Journal of Wind Engineering and Industrial Aerodynamics 34, 251272.
Sanz-Andres, A., Santiago-Prowald, J., 2002. Train-induced pressure on
pedestrians. Journal of Wind Engineering and Industrial Aerodynamics 90,
10071015.
Schetz, J.A., 2001. Aerodynamics of high speed trains. Annual Review of Fluid
Mechanics 33, 371414.
Schulte-Werning, B., Heine, C., Matschke, G., 2003. Unsteady wake characteristics
of high speed trains. PAMM Proceedings Applied Maths and Mechanics 2,
332333.
Sims-Williams, D., Dominy, R., Howell, J., 2001. An investigation into large scale
unsteady structures in the wake of real and idealised hatchback car models.
SAE Congress, Detroit, Paper 2001-01-1041.
Sockel, H., 1996. The aerodynamics of trains. In: Schetz, J.A., Fuhs, A.E. (Eds.),
Handbook of Fluid Dynamics and Fluid Machinery. Wiley, New York, pp.
17211741.
Sterling, M., Baker, C., Quinn, A., Hoxey, R., Richards, P., 2006. An investigation of
the wind statistics and extreme gust events at a rural site and implications for
wind tunnel testing. Wind and Structures 9 (3), 193216.
Sterling, M., Baker, C.J., Bouferrouk, A., ONeil, H., Wood, S., Crosbie, E., 2008a.
An investigation of the aerodynamic admittances and aerodynamic
weighting functions of trains. In: Proceedings of the BBAA VI, Milano, Italy,
July 2024.

ARTICLE IN PRESS
298

C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277298

Sterling, M., Baker, C.J., Jordan, S.C., Johnson, T., 2008b. A study of the slipstreams
of high speed passenger trains and freight trains. Proceedings of the Institute
of Mechanical Engineers Part F: Journal of Rail and Rapid Transport 222 (2),
177193.
Vino, G., Watkins, S., Mousley, P., Watmuff, J., Prasad, S., 2005. Flow structures in the
near wake of the Ahmed model. Journal of Fluids and Structures 20, 673695.

Wu, D., 2004. Predictive prospects of unsteady detached-eddy simulations


in industrial external aerodynamic ow simulations. Diploma thesis,
Lehrstuhl fur Strmungslehre und Aerodynamishes Institute, Aachen,
Germany.
Xu, Y.L., Ding, Q.S., 2006. Interaction of railway vehicles with track in cross-winds.
Journal of Fluids and Structures 22, 295314.

Vous aimerez peut-être aussi