Download as pdf or txt
Download as pdf or txt
You are on page 1of 22

ARTICLE IN PRESS

J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

Contents lists available at ScienceDirect

Journal of Wind Engineering


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

The flow 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: This paper considers aspects of the aerodynamic behaviour of high speed trains. It does not specifically
Received 4 March 2009 address the many aerodynamic problems associated with such vehicles, but rather attempts to describe,
Accepted 18 November 2009 in fundamental terms, the nature of the flow field. The rationale for such an approach is that the flow
Available online 23 December 2009
fields that exist are the primary cause of the aerodynamic forces on the train and its components which
Keywords: result in a whole range of aerodynamic issues. This paper thus draws on a wide range of model scale
High speed trains and full scale experimental and computational work and attempts to build up a comprehensive picture
Aerodynamics of the flow field. Attention is restricted to trains in the open air (i.e. tunnel flows will not be considered)
Cross winds for both still air conditions and crosswind conditions. For still air conditions the flow field will be
Boundary layers
described for a number of flow regions i.e.
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 flow field 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 number of runs are considered together. Results obtained using


this technique will be used extensively in what follows. Both wind
This paper aims to set out a description of the flow field around tunnel tests and CFD calculations are made difficult because of the
high speed trains in the open air. It will approach this from a fairly large length/height ratio of high speed trains that makes wind
fundamental point of view, and will not specifically address tunnel models or computational grids very long and thin, and
practical issues and problems associated with the aerodynamic which both require specialist techniques and expertise. This point
behaviour of trains, although these will be briefly discussed at the having been made however, experimental and computational
end of the paper. Such an approach is adopted in the hope that techniques will not be discussed at any length in what follows,
such a description will clarify the basic flow mechanisms that although where the nature of the technique has the potential to
exist around high speed trains, and will thus inform future seriously affect the results that are being presented then this will
consideration of a range of more practical issues. be pointed out.
The basic tools in the study of train aerodynamics are full scale Section 2 discusses the aerodynamics of high speed trains in
testing, wind tunnel testing and CFD calculations, as indeed is the conditions of zero cross wind. The discussion is framed in terms of
case in other fields of aerodynamics. In the case of the study of three flow regions, viz.
train aerodynamics, all of these approaches are fraught with
difficulties. Full scale measurements have to be made in very
 the nose region around the front of the train;
turbulent flows and very often a large number of runs have to be
 the boundary layer region along the length of the train (for the
carried out to enable the mean and unsteady flow patterns to be
train side, train roof and train underbody);
elucidated. Baker et al. (2001) describes the technique of
 the wake region behind the train.
‘‘ensemble averaging’’ through which the results of a large

This scheme is based on that developed by the author in Baker


 Tel.: + 44 1214145067; fax: + 44 1214143675. et al. (2001), although in this paper the number of flow regions is
E-mail address: c.j.baker@bham.ac.uk reduced from the five in Baker et al. (2001) to the three listed

0167-6105/$ - see front matter & 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.jweia.2009.11.002
ARTICLE IN PRESS
278 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

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 flow calculations of Sanz-Andres and Santiago-


Prowald (2002) (Pressure coefficient traces are shown for 2D and 3D computa-
tions. 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).

above—the upstream and nose region in Baker et al. (2001) being


considered together here, and the near wake and far wake regions Fig. 4. Synthesis of boundary layer measurements (Brockie and Baker, 1990;
in Baker et al. (2001) being similarly combined. For each of the Schetz, 2001; Sockel, 1996).
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 279

flow regions the work of the author and his co-workers, and the a 14 car ICE service train. This data was obtained from trackside
work of other investigators are considered to develop as complete anemometry in full scale experiments designed to measure the
a picture as possible of the flow field around the train. slipstreams around such trains. The experiments are reported in
Section 3 then goes on to consider the flow field around trains outline (RAPIDE Consortium, 2001) and discussed in considerable
in a cross wind. This begins by a consideration of the nature of the length by the author and his co-workers in Sterling et al. (2008b).
wind flow relative to the train in terms of the mean velocity Data from these experiments will be used extensively in what
profile, turbulence profile and power spectrum. A qualitative follows to illustrate a number of effects. The data in Fig. 1 is an
picture of the flow around trains is then developed from a ‘‘ensemble average’’ of the data from 17 train passes. This data was
consideration of the work of a number of authors, and the nature aligned (at the point corresponding to the peak of the velocity trace
of the pressure distribution around high speed trains is also shown in the figure) and the data at all other points averaged over
discussed, in terms of both steady and unsteady surface pressures. all the runs. Thus x, the position along the train, is defined as
Finally the way in which these pressures sum to give cross wind measured from this peak in velocity. The lateral distance y0 is
forces and moments is discussed in terms of the aerodynamic defined as the distance from the rail edge, and the vertical distance
admittances and aerodynamic weighting functions. Some con- z0 as the distance from the top of the rail. For the results shown the
cluding remarks are then made in Section 4 and the implications velocities were measured at trackside with no platform present
of the results for current issues in high speed train aerodynamics (z0 =0.5 m). The air velocity data, u, is divided by the train speed, v,
are set out. to give the normalised value U. From Fig. 1 the velocity peak can be
seen to be sharply defined and, as would be expected, decreases
away from the train. The standard deviation of the ensemble is
2. The flow around trains with no cross wind small in all cases—of the order of 0.02–0.03, which indicates that in
this flow region there is little run to run variation.
2.1. The nose region The velocity changes illustrated in Fig. 1 are accompanied by
pressure changes. Fig. 2 shows typical pressure changes caused by
In this section the flow upstream and just downstream of the an ETR 500 (Mancini and Malfatti, 2001). These measurements
nose of high speed trains will be considered. In this region the were obtained from train passing tests carried out as part of the
variations of air velocity and pressure are essentially inviscid. A major EU TRANSAERO project. It can be seen that there is a rapid
typical variation in air velocity is shown in Fig. 1 around the front of increase and then decrease in pressure around the train nose.

0.4

0.3
y' = 1.16m
0.2 y' = 1.5m
U

y'= 2.42m

0.1

0
-50 0 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

0.2 x=50m
x=100m
0.15
U

x=200m
x=300m
0.1
x=350m

0.05

0
0 1 2 3
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) 277–298

Again for any particular train, this effect is highly repeatable from for the Euler method shown in Fig. 2). The blunter the nose shape,
run to run. the higher are the velocity and pressure disturbances.
As such flows are inviscid they can be well predicted by
reasonably simple calculation methods—as shown Fig. 3 below 2.2. The boundary layer region
from the potential flow calculations of Sanz-Andres and Santiago-
Prowald (2002). More complex panel methods can be used to 2.2.1. Train side
calculate the details of the pressure and velocity variations Over the last few decades a number of investigators have made
around train nose shapes of different types (such as the results boundary layer measurements on trains, using conventional train

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

0.3
Turbulence intensity

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) 277–298 281

0.8

0.6 Trackside y'=1.16m,


z'=0.5m

Autocorrelation
Platform y = 1m, z=1m

0.4

0.2

0
0 0.5 1 1.5 2 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 coefficients 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).
ARTICLE IN PRESS
282 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

based pitot probes, hot wire probes etc. These tests have been work of Sockel (1996) for a variety of other trains. All dimensions
carried out at both full scale and model scale for a variety of train given in these figures are the equivalent full scale values. Note
types. From these experiments it is possible to derive standard that the results of Schetz (2001) and Sockel (1996) are for the
boundary layer parameters such as the displacement thickness actual, somewhat loosely defined, boundary layer thickness.
and the form parameter (although note that these are formally Schetz (2001) notes that the ratio between this thickness and
derived for classical two dimensional boundary layers, rather than the displacement thickness is between 8 and 11, i.e. one order of
the complex three dimensional flows found around trains). The magnitude.
data from some of these experiments is summarised in Fig. 4 It can be seen that all the model scale results are broadly
below, with data from the wind tunnel and full scale tests of consistent with each other, and show firstly a steady development
Brockie and Baker (1990) for the UK HST, and the data correlation in the total boundary layer thickness and the displacement
of model scale results given in Schetz (2001), reporting the earlier thickness along the length of the train, and secondly values of the
form parameter that are significantly 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 flow 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 profile 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. 12. Boundary layer on train roofs from moving model tests for 1/25 scale ICE Fig. 7 shows the boundary layer displacement thicknesses and
model (Baker et al., 2001): (a) displacement thickness; (b) form parameter. 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) 277–298 283

parameters are also given for a similar set of full scale to the existence of an unrealistic laminar boundary layer near the
measurements made above a station platform, and for a set of vehicle nose in the model tests that does not exist at the much
model scale experiments made at half train height without a higher Reynolds numbers of the full scale tests.
platform simulation (Sterling et al., 2008b). Note that for the Turbulent boundary layers such as those on the side of the
platform experiments, z is defined as the distance from the top of train are also characterised by their unsteady flow character-
the platform, and y the distance from the train side. It can be seen istics—with the magnitude of the turbulence being characterised
that the displacement thicknesses in all three cases grow along by the turbulence intensity, and the scale by parameters such as
the length of the train, with the trackside full scale values being the integral time or length scales. In terms of the ensemble
larger than those for the other experiments. This is not surprising, average data, velocity data from stationary probes that has been
as the former measurements were taken in a region close to the obtained for the ICE service train, the turbulence intensity can be
ground exposed to the aerodynamically rough bogies, whereas approximated by (1  ensemble standard deviation)/ensemble
the latter were obtained from regions closer to smoother areas of mean. Fig. 8 shows a plot of this parameter along the train for
the train. The form parameters are again significantly less than both the measurements made at trackside and those made above
the equilibrium values as in Fig. 4. Perhaps the major point to the platform at broadly equivalent positions. It can be seen that in
emerge from this data are the large values of displacement both cases the turbulence intensity is more or less constant along
thickness near the front of the vehicle in the full scale the train, although, as would be expected, the value is
measurements, suggesting a major flow disturbance around the significantly higher for the trackside measurements than for the
nose that is not replicated in the model scale measurements platform measurements. The values are of the order of 0.05–0.1,
shown in either Figs. 4 or 7. It could be conjectured that this is due which are typical values for flat plate boundary layers. Fig. 9

Fig. 14. Vertical and horizontal velocity profiles beneath Korean high speed train (Kwon and Park, 2006).
ARTICLE IN PRESS
284 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

shows the autocorrelations of velocity for these two cases. From considerably thicker than the corresponding side wall boundary
these plots the integral length and time scales (the scales that layers, which leads to the conclusion that there is a diverging flow
contain the most turbulence energy) can be found to be 4.7 m and up the side of the wall and a converging flow over the roof of the
0.067 s for the trackside measurements and 4.1 m and 0.059 s for train.
the platform measurements. The integral length scale is thus of Baker et al. (2001) describes some measurements made using
the order of 20% of the length of an individual carriage. the alternative approach of stationary probes and a moving
The final boundary layer parameter that is of interest is the model, for a 1/25th scale ICE vehicle. The results for displacement
skin friction coefficient (Cf), as the surface skin friction determines thickness and form parameter are shown in Fig. 12. The x axis in
to a large extent the overall drag of the train. Fig. 10 shows the these figures is a normalised time, with the normalisation being
local skin friction coefficient for the HST model and full scale with vehicle speed and vehicle length. T= 0 is when the train nose
results of Brockie and Baker (1990). It can be seen that, as is to be passes, and T= 4 is when the tail of the four coach train passes.
expected, the skin friction is very dependent upon scale and There can be seen to be little difference in the displacement
indicates the necessity for as large a scale as possible in either thickness results between the side wall data (already included in
wind tunnel tests or computations if the drag is to be accurately Fig. 7) and the roof data, which is not consistent with the results
predicted. It is of interest to note that most of the individual of Fig. 10. Again the form parameters are significantly below the
values of local skin friction fall on or below the accepted smooth equilibrium boundary layer value of 1.4.
wall correlations of skin friction and local Reynolds number for
flow over a two dimensional flat plate, indicating again the non-
equilibrium, three dimensional nature of the side wall boundary 2.2.3. Train underbody
layers. Similar coefficients could also be determined from the The flow underneath high speed trains has not been exten-
model scale moving model measurements for the ICE reported in sively studied in the past, but has recently come to prominence as
Baker et al. (2001), through a fit of the logarithmic law to the a result of the ‘‘ballast flying’’ issue that can cause a variety of
velocity profiles. The authors of this paper acknowledge that this problems to the train and to the track. The investigations into this
procedure was not an accurate one, but average values of 0.0029 problem have resulted in a number of investigations taking place
were obtained, which is again somewhat below the smooth flow which have measured the velocity and pressure field beneath high
flat 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 fitting of a logarithmic law to the velocity
profiles from which the shear velocity and thus the skin friction
coefficient can be determined. The log law is however a poor fit to
many of the velocity profiles 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 fit 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 coefficient 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 Fig. 16. Velocity profiles measured below Shinkansen train (Ido et al., 2008). The
of the conventional boundary layer measurements of Brockie and different curves indicate different sections along the train, with M1 being at the
Baker (1990). This data indicates a displacement thickness that is front of the vehicle and M5 at the rear.

Fig. 15. Pressure coefficients beneath Eurostar train (Quinn and Hayward, 2008).
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 285

speed trains. The results of a number of these investigations are speed at heights of 0.18 m. A small peak in the velocities can be
considered here—those of Kwon and Park (2006) for the Korean seen as the train nose passes, but for most of the traces the
high speed train, Quinn and Hayward (2008) for the Eurostar velocities are highly turbulent and fluctuating as would be
trains on the Channel Tunnel Rail Link, and the experiments expected for the rough under train environment. Fig. 14 shows
reported in Ido et al. (2008) for the Japanese Shinkansen train. vertical and horizontal velocity profiles for the same tests. The
Fig. 13 shows the velocities beneath the Korean train for a number results for the vertical profile imply some sort of annular flow
of heights, measured at full scale with pitot probes. The train profile (with an inflexion point between the track and the train),
speed was 300 kph, about 83 m/s. The velocities at various rather than a conventional boundary layer profile. The velocities
distances from the upper rail surface are shown. It can be seen can also be seen to peak on the train centre line and fall off
that these velocities reach high levels of around 40% of the train 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) 277–298

The velocity and pressure fields were measured beneath the 2.3. The train wake
Eurostar train and are reported in Quinn and Hayward (2008). The
velocity traces are similar to those described above and are not shown In the last few decades there has been a great deal of work
here. The pressure trace is shown in Fig. 15, both for the entire train carried out on the nature of the wakes of road vehicles—primarily
and for the first quarter of the train. One unit of normalised time with a view to being able to minimise drag which is of course
represents the passage of one carriage of the train. These are dominated by wake effects for road vehicles. It is clear from the
ensemble averages of 20 nominally identical train passes. A literature that the precise nature of the wake varies from vehicle
prominent nose maximum and minimum can be seen in both to vehicle, but there seem to be a relatively small number of flow
traces that is similar to those shown in Section 2.1. However regular mechanisms that can exist in some combination or other—shear
minima are also seen along the vehicle, that seem to be associated layer separations, longitudinal helical flows, vortex streets and a
with the passage of the middle of each coach between the bogies. A separation cavity (Vino et al., 2005; Sims-Williams et al., 2001;
tail peak can also be seen. From the Eurostar velocity data it was also Nouzawa et al., 1992). All of these phenomena are subject to
possible to obtain autocorrelation functions, and thus the integral instabilities with Strouhal numbers (based on vehicle velocity and
length scales and time scales. The length scales were found to be some representative frontal dimension) between 0.05 and 0.4.
between 1.8 and 2.0 m, and the time scales between 0.02 and 0.03 s. Now whilst the tail shape of trains is rather different to those of
These values are rather smaller than those for the side boundary cars, one might expect equally complex flows. The major physical
layer, as may be expected because the flow is significantly difference will be that the thickness of the boundary layer around
constrained. the train will be relatively much greater than for cars, and thus
The flow field beneath a full scale 16 car Shinkansen test train any separated shear layers will also be much thicker.
and a three car wind tunnel model are reported in Ido et al. (2008). With these points in mind, let us now consider the experi-
The wind tunnel tests were carried out with a moving ground belt mental data and computational work that is available to describe
beneath the middle car only, and this may have some influence on the wakes of high speed trains. Fig. 17 shows experimental
the results that are shown in Fig. 16 below. The results of velocity vectors for a model TGV (Paradot et al., 1999) and a
Fig. 16 look rather different from those of Fig. 14, but this is model ICE (Baker et al., 2001), together with a visualisation of the
because of the frame of reference that is used and in fact they flow around an ICE using standard RANS CFD methods (FLUENT).
are identical in form. Those of Fig. 16 show clearly the inflexion It can be seen in each case that there is strong evidence of helical
point profile that could only be inferred from Fig. 14. There is vortices behind the train, which extend a considerable distance
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 geometries—bogie fairings etc—and shows that the
underbody flow 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 profile 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 flying problem has also driven wind
tunnel and computational research on the flow 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 computa-
tionally, with the results being very sensitive to both the experi-
mental set up and the computational configuration (including the
turbulence model). Further progress can be expected in this area in Fig. 19. ICE 1/25th model scale wake measurements (y is the distance from the
the future. 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) 277–298 287

into the vehicle wake. Schulte-Werning et al. (2003) takes the CFD some sort of regular oscillation with a Strouhal number of around
work somewhat further and investigates the unsteadiness of 0.11–0.14.
these helical structures through the use of unsteady RANS The situation is further complicated however when the model
techniques. The results are illustrated in Fig. 18. A well defined scale data of Baker et al. (2001) is considered in more detail. The
oscillation can be seen with a Strouhal number of 0.14. wake velocities just behind the train showed a very great deal of
Now the full scale slipstream data presented in Sterling et al. run to run variation with high ensemble standard deviations. A
(2008b) has been analysed intensively to attempt to determine wavelet investigation of the turbulence scales suggested peaks at
the wake characteristics. Fig. 5 shows the ensemble average two Strouhal numbers—0.03 (hypothesised to be due to wake
velocity traces, and for those nearest to the train there can be cavity ‘‘pumping’’) and 0.5 (taken to be due to shear layer
seen to be a noticeable peak in the velocity in the near wake (i.e. instability). As the model measurements were made higher up the
350–400 m). These velocities were measured 1 m above the track train than the full scale measurements, it may be that the
with no platform in place. The equivalent model scale experi- unsteady helical vortex motion is more important close to the
ments of Baker et al. (2001), measured half way up body height, ground, and other types of unsteadiness higher up the vehicle.
do not show this peak. The experiment and simulation of Fig. 17 Further work is required in this area.
indicate that the helical vortex occurs close to the ground, and it Up to this point we have implicitly been considering the near
thus seems likely that this is what is observed in the full scale wake of trains—within a few vehicle heights of the train tail. One
slipstream measurements. Further investigation of the full scale would expect that any large scale flow structures that exist would
data however showed that the technique of ensemble averaging decay fairly rapidly and that the far wake of the train would show
was, in this case, actually hiding a physical effect, and that this a gradual decrease in velocity. This effect can be seen in the
peak did not appear on around half of the individual velocity ensemble average velocity traces for the ICE in Fig. 5 for a value of
traces that were used. Careful re-analysis revealed that the x4420 m (the tail of the train is at x= 365 m). The velocity
slipstream measurements were picking out some type of vortex measurements made nearest the train decay steadily, whilst the
shedding oscillation with a Strouhal number of 0.11, which is measurements made furthest away first increase and then
close to the computational value of 0.14 reported above. It thus decrease—showing the lateral spread of the wake. This effect
seems that the helical vortices in the train wake undergo 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
y/h = 0.033 the four car model train. In this case the origin is at the train tail. A
0.1 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
0.05 along and across the wake, and governed by two parameters.
U

Fig. 20 shows the best fit 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
0 dimensionless format.
0 50 100 150
T
3. The flow around trains with a cross wind
y/h = 0.1333
0.1 3.1. The natural wind relative to a train

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


train moving though it, for the fundamental case of a train moving
U

0.05
over a flat ground. We assume for the sake of simplicity that the
wind is normal to the train track and that the vertical velocity
profile is given by the usual logarithmic law with a surface
0 roughness length of z0 ESDU (1985). We further assume that the
0 50 100 150 turbulence intensity is given by the values given in that same
T reference, and is again a function of the surface roughness length.
To calculate the velocity profile and turbulence intensity with
y/h = 0.40 respect to a train the train velocity must also be considered. This
0.1 is straightforward and the method has been frequently rehearsed
in the past (Baker, 1991a for example) and one can easily obtain
the velocity profile, yaw angle, and turbulence intensity relative
to a moving train. Sample calculations are shown in Fig. 21 below
0.05
U

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 defined as the atmospheric turbulence level divided by the
0
wind velocity relative to the train. It can be clearly seen that at
0 50 100 150
low train speeds, the velocity profile takes on the appropriate
T
boundary layer form, with a significant velocity variation across
Fig. 20. Best fit curves to wake velocities using model of Baker (2001a) the train height, the yaw angle is close to 901 and the turbulence
((a) y/h= 0.033, (b) y/h= 0.2, (c) y/h=0.533). intensity is high. At high train speeds, the velocity profile is more
ARTICLE IN PRESS
288 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

uniform, and the yaw angle is low and varies somewhat over the
6
height of the train. The turbulence intensity is low—of the order
5 of a few percent. Clearly these results have implications for the
v = 0 m/s type of wind tunnel or computational simulation that is used—if
Height (m)

4 v = 20 m/s there is a desire to model trains moving at high speeds, then these
3 v = 40 m/s tests can reasonably be carried out in low turbulence wind
v = 60 m/s
2 tunnels, with no velocity shear (although the yaw angle twist
v = 80 m/s
would not be simulated). If low vehicle speed conditions are
1
required, then atmospheric turbulence and shear needs to be
0 simulated, although it needs to be recognised that any simulation
0 0.5 1 1.5 will only model one specific set of wind speed/vehicle speed
Normalised velocity relative to train conditions. The same comments can be made for the critical case
of a train on an embankment, where the wind profile speeds up
close to the ground, although in this case the wind shear and
6
turbulence intensities are rather less in all cases.
5 As well as the mean velocity profiles and the turbulence
v = 0 m/s intensity, the turbulence spectrum experienced by a moving train
Height (m)

4
v = 20 m/s will be different to that experienced by a stationary train. This
3 was investigated in Cooper (1985) and typical results are shown
v = 40 m/s
2 v = 60 m/s
in Fig. 22. In this figure the spectra are shown for a range of train
speed/wind speed ratios from 0 to infinity, for a pure cross wind.
1 v = 80 m/s
The x axis is a frequency normalised with the atmospheric
0 turbulence length scale and the wind velocity relative to the train,
0 20 40 60 80 100 and the y axis is the spectral density normalised with the
Yaw angle relative to train (degrees) frequency and the wind velocity variance. Plotted in this way the
spectra show a remarkable level of similarity, which thus implies
that, to a first approximation, they scale on the wind velocity
6
relative to the train.
5 Whilst the above figures give some indication of the wind
v = 0 m/s
statistics relative to the train, it is not unusual when looking at
Height (m)

4
v = 20 m/s
train cross wind stability to specify, in varying degrees of detail,
3 v = 40 m/s
an extreme wind gust, on the basis that such wind gusts will
v = 60 m/s
2 cause train stability problems. The question thus arises as to the
v = 80 m/s
1 nature of these gusts, which vary both in space and in time. The
author and his co-workers have in recent years investigated a
0 number of velocity and surface pressure datasets obtained on the
0 0.1 0.2 0.3 0.4
University of Birmingham Wind Engineering field site at Silsoe in
Turbulence intensity relative to train Bedfordshire, and, of particular relevance to the current discus-
Fig. 21. The wind characteristics seen by a moving train over a flat ground sion, have adopted the technique of ensemble averaging of
(u= 20 m/s). extreme gusts i.e. identifying the gusts in the time series and

Fig. 22. Wind spectra relative to a moving train over a flat ground (Cooper, 1985).
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 289

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

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 considerations—the first is to define 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 significantly filter the short
term gust peak shown in Fig. 23. Another approach is to define a
specific 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 profile has been obtained from a study of the time and spatial
averaged wind statistics and as such reflects all the information
within these statistics rather than just the information describing
the extreme gusts. Note further that when it is applied a spatial
filter of one vehicle length is used. Either of the above approaches
is clearly a simplification of the nature of the gust, although both
have practical merits.
An alternative approach to using a simplified extreme gust has
been outlined in Ding et al. (2008) and Baker et al. (2008b) where
a correctly spatially and temporally correlated wind field has been
simulated to match the required wind correlations. The wind
Fig. 23. Extreme gust profiles (Sterling et al., 2006) (ds indicates dataset number; velocities seen by a train as it passes through this simulated wind
velocities normalised by the corresponding standard deviation): (a) streamwise
velocity component; (b) lateral velocity component; (c) vertical velocity
field will then contain the necessary spatial and temporal
component. information concerning the extreme gusts in a more realistic way.

then averaging the gusts to produce an average time series. Fig. 23 3.2. Wake flow structure in cross winds
shows the average streamwise, lateral and vertical velocity
fluctuations (normalised by the standard deviation) measured In the last few decades a number of wind tunnel tests and
1 m above ground level at the site (Sterling et al., 2006) (although computational simulations have elucidated the flow structure
the results are similar to those measured at a range of heights up around high speed trains in cross winds. The wind tunnel work of
to 10 m). It can be seen that the peak streamwise gusts, have a Mair and Stewart (1985), Copley (1987). Chiu (1991), Robinson
magnitude of about three to four times the standard deviation of and Baker (1990) in the 1970s and 1980s used an idealised train
the wind velocity fluctuations and last for about one second either model, and it was found that the dominant flow pattern was a
side of the maximum peak. These gusts are superimposed upon a series of inclined wake vortices such as might be found around a
longer term gust with a magnitude of around one standard missile at high angles of attack (Fig. 25a). At higher yaw angles a
deviation above the mean. The corresponding vertical fluctuations more conventional vortex shedding pattern was observed, with
show a negative peak at the maximum gust, indicating that these some switching of the flow at intermediate angles. Recent
gusts are associated with ‘‘sweep’’ events in the atmospheric computations, using both RANs and LES techniques on various
boundary layer. In terms of the spatial correlation of such gusts, ICE configurations (Hemida, 2006; Wu, 2004; Diedrichs, 2005)
the results show that the short term peaks are only correlated have also shown similar patterns (Figs. 25b–d). There is broad
over heights of a few metres, and extreme pressure coefficients agreement both qualitatively and quantitatively between the
ARTICLE IN PRESS
290 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

Fig. 25. Wind tunnel tests and CFD computations of wake vortex flows behind trains in a cross wind. (a—idealised train (Copley, 1987); b—2 car ICE (Hemida, 2006);
c—ICE (Diedrichs, 2005)).

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 slip-
stream 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
formats—one for the velocities themselves (Fig. 27a) and one
for the velocities minus the velocities measured with no cross
wind (Fig. 27b). In the first 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
Fig. 26. Computed and measured surface streamline patterns (Chiu, 1991; upstream wind speed indicates an enhancement of that wind
Hemida, 2006). 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
wind tunnel tests and the CFD results—see the surface streamline that this is consistent with the presence of inclined vortices in the
patterns from Chiu (1991), Hemida (2006) in Fig. 26. There is wake of the train. Finally Fig. 28 shows the maximum wake
some evidence, particularly from the LES results of wake velocities (one second averages) that were measured for a wide
unsteadiness (Hemida, 2006) that indicates two modes of wake variety of high speed trains in full scale test in the UK, normalised
unsteadiness—a horizontal wake oscillation with a Strouhal by the train speed. There is, inevitably, a great deal of scatter but it
number of 0.1, and a weak vortex shedding motion with a can be seen that there is a general increase in the maximum
Strouhal number of 0.15–0.2. It has to be said however that these normalised velocities as the cross wind speed increases.
frequencies are not always well defined 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. 3.3. Pressure distributions in cross winds
(2007) which presents data from moving model experiments on a
four car ICE train, with rather a crude cross wind generator placed The idealised train low turbulence wind tunnel measurements
normal to the track. The x axis unit is time normalised by train of Mair and Stewart (1985), Copley (1987), Chiu (1991) made
velocity and carriage length—with zero being the point at which extensive measurements of the pressure distributions around the
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 291

Normalised mean velocities -with cross wind


0.5

0.4

0.3 0.125m
0.75m
0.2
2.0m
0.1

0
-2 0 2 4 6

0.3

0.2
Normalised velocity

0.125m
0.1 0.75m
2.0m

0
-2 0 2 4 6

-0.1
Normalised time

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

0.400

0.300
Normalised velocity

0.200

0.100

0.000
0 2 4 6 8
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) to give a major contribution to the overall side force. In general
together with the results of simple panel method calculations the numerical solutions show a good level of agreement with the
Fig. 29b shows equivalent measurements for a 2 car ICE 2 model experiments.
(Wu, 2004), which also shows the LES calculations of Diedrichs A similar set of results, from Baker et al. (2008a) are shown for
(2005). Pressure coefficients are shown on loops around the a wind tunnel model of a UK Class 365 electrical multiple unit
vehicle, for a variety of different distance (x) from the train nose, (e.m.u.) in Fig. 30. These measurements were made in a
normalised with the model diameter D or length L. The results are simulation of the atmospheric boundary layer i.e. in a highly
plotted with a negative pressure coefficient in the positive turbulent flow. They are plotted with the positive pressure
direction. For all values of x that are not near the nose, it can be coefficient in the positive direction in a manner similar to, but
seen there is a suction peak on the windward roof corner (451 in confusingly different from, Fig. 29. The pressure coefficients are
Fig. 29a, 315 degrees in Fig. 29b), small suctions over the rest of shown on loops around the vehicle (with loop B being near the
the roof, leeside and underside, and a positive pressure coefficient front of the leading vehicle and loop H near the back of that
on the windward wall. Near the nose however, for small values of vehicle. Results are shown for yaw angles of 451 and 901 (results
x there is a suction peak on the leeward side, that can be expected 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) 277–298

Fig. 29. Pressure distributions around trains (a—Idealised train, lines are panel method results and points are experimental results (Chiu, 1991), front of windward face is
at 01; b–2 car ICE—lines 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 steadiness or otherwise of the flow. It can be seen that this
the loops at both yaw angles. Fig. 30 also shows the standard parameter peaks at the windward roof corner, perhaps indicating
deviation of the pressure coefficients, which gives an indication of some unsteady separated flow in this region. Fig. 31 shows a
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 293

Fig. 29. (Continued)

proper orthogonal decomposition (POD) analysis of the pressure shows the curves for aerodynamic admittance from the work of
detail. A full description of this type of analysis is given in Holmes Cooper (1985) that corresponds to the velocity spectra relative to
et al. (1997), Baker (2000). It shows that the first and most the vehicle shown in Fig. 22. Aerodynamic admittance values are
energetic POD mode seems to be concentrated at the windward plotted against normalized frequency, for a ratio of train speed to
roof of the vehicle, and thus physically corresponds to the normal wind speed of 4.0, and for two different vehicle heights. It
pressure fluctuations on the windward roof. The second mode can be seen that all the curves (for different ratios of the vehicle
has the characteristics of the mean pressure distribution and thus length to the turbulence integral length scale) all approach unity
physically corresponds to quasi-steady pressure fluctuations at low frequencies (showing that large turbulent eddies are
around the model. A fuller discussion of this analysis is given in correlated over the entire vehicle) but fall rapidly at high
Baker et al. (2008a). frequencies for the reason set out above, and thus will act as a
filter on high frequency wind fluctuations, and thus the energy
available to excite suspension frequencies that lie in this range.
3.4. Aerodynamic admittances and weighting functions Similar experimental curves, from the Class 365 e.m.u. results
mentioned above are shown in Fig. 33 Sterling et al. (2008a). The
The overall forces on trains are composed of the sum of the reduced frequency is the actual frequency normalized by the wind
mean and fluctuating pressures over the train surface. One would tunnel velocity and the carriage length. It can be seen that some of
expect that small scale fluctuations that only affect small parts of the experimental curves tend to values other than unity at low
the surface of the vehicle would be damped out in this summation frequencies. Sterling et al. (2008a) suggest that this is because of a
(integration) process, and thus the force fluctuations would be number of reasons—primarily because the concept of admittance
smaller than the upstream velocity or pressure fluctuations. This implicitly assumes that the flow is not significantly affected by
effect is important in calculating cross wind stability in a number the body on which the forces are measured, and because the
of related ways. If calculations of stability are carried out in the streamline passing through the reference point is not the
‘‘frequency domain’’ to use the terminology of Baker (1991b, c) (as ‘‘significant streamline’’ defined as the streamline on which the
in Xu and Ding, 2006, for example) then the force coefficient turbulent characteristics mostly influence the unsteady forces on
spectra can be obtained from the wind spectra through the use of the body. Sterling et al. (2008a) shows that this streamline is
aerodynamic admittance functions, which are effectively normal- usually significantly below the reference position of 3 m, which is
ised ratios of force coefficient spectra to wind spectra. Fig. 32 not unreasonable. That being said, as long as the same reference
ARTICLE IN PRESS
294 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

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

Fig. 31. POD analysis of fluctuating 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) 277–298 295

6
5
4
Side force
3
Lift force
2
1
0
0 0.2 0.4 0.6 0.8 1
Time (sec)

6
5
Fig. 32. Aerodynamic admittances from Cooper (1985) (L = vehicle length,
H= vehicle height, xLu =turbulent length scale, VT = train speed, x axis is frequency 4
normalised with turbulence length scale and wind velocity relative to the train). Side force
3
Lift force
2
1
0
0 0.2 0.4 0.6 0.8 1
Time (sec)

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).

they are only non-zero for values less than around 0.5 s i.e. the
train side and lift force coefficients 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
Fig. 33. Experimental measurements of aerodynamic admittances for Class the weighting functions shown in Fig. 34. The filtering effect of the
365 e.m.u side force coefficients (x axis is frequency normalised with length of higher frequency fluctuations in velocity is very clear, particularly
vehicle and wind velocity relative to the train). at the higher vehicle speeds.
Finally if calculations are carried out in the ‘‘amplitude
domain’’, then the ‘‘extreme values’’ of the force coefficients,
position is used in any model tests in which the admittances were based on extreme values of the forces and extreme values of the
measured, and in any calculation using these admittances, then upstream velocities, might also be expected to be less than the
the calculations will be valid. Now assuming suitable values of the mean values of these coefficients, as the high frequency fluctua-
physical parameters that determine the admittance functions (i.e. tions will have been filtered out. This effect is illustrated in Fig. 36,
wind and vehicle velocities, vehicle length and turbulence length for the static wind tunnel model tests carried out on a UK Class
scales, indicates that the admittance falls to values of 0.5 at 390 Pendolino train (Baker et al., 2004) for the extreme
around 1–2 Hz, suggesting that wind fluctuations above these coefficients based on a gust averaging time of 1 s. At first sight
values will be filtered out by the lack of correlation of surface the fact that the extreme/mean ratio is significantly below unity
pressures. seems inconsistent with the admittance and weighting function
Now whilst this frequency domain analysis is useful, in any results above which suggest that such effects due to lack of
calculation of train behaviour that considers suspension dynamic correlation of surface loads, should be confined to averaging times
effects including discrete track irregularities, the ‘‘time domain’’ less than 0.5 to 1.0 s. However more recent calculations carried
equivalent is required Baker et al. (2008b). This is known as the out by the author suggest that the results of Fig. 36 are affected, to
aerodynamic weighting function and can be shown to be the a significant degree, by the specification of wind velocities at the
Fourier transform of the aerodynamic admittance, and vice versa. reference streamline rather than on the significant streamline
The derivation of such functions from experimental data is far defined above. Nonetheless, as with the admittances, if the same
from straightforward, and is discussed in detail in Baker et al. reference height is used in any stability calculations as was used
(2008b). Typical values are given in Fig. 34 for side and lift force in the experiments that measured extreme values, then the
coefficient weighting functions for the class 365 e.m.u referred to results will still be valid.
above. Effectively these functions weight the relative wind
velocities and train force coefficients over a period leading up to
the time of application. The main points to note about these 4. Concluding remarks—implications for practical issues
functions are firstly that the values of weighting function for the
high speed case are smaller than for the low speed case, which In this section some brief remarks are made on the implica-
reflects the fact that the wind velocity fluctuations relative to the tions of the flow field descriptions of previous sections for a
train are smaller the faster the train velocity, and secondly that number of practical issues of current concern.
ARTICLE IN PRESS
296 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

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).

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 flow 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
flow effects. (Note however that this conclusion is only true
for smooth high speed passenger train—Sterling et al. (2008b)
shows that for freight trains the integral time scales on the
train side are rather larger and turbulent flows along such
Fig. 36. Ratio of extreme to mean side force coefficients for the Class 390
trains could affect passenger and trackside worker stability.)
Pendolino full scale experiments (Baker et al., 2004)—legend indicates different
train/track configurations. (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 flight paths beneath
the train by the author suggest time scales of the order of 0.3 s
(a) It is clear that the boundary layer measurements made on and path lengths of the order of 3 m. This mismatch of times and
model vehicles are not fully consistent with those made at full scales suggest that the flight path of train ballast will be
scale, partly due to the expected effects of scale (i.e. Reynolds primarily determined by the mean flow field beneath the train.
number) and partly because at full scale there would appear (d) The large scale unsteady flow structures in the near wake of the
to be a very rapid growth of the boundary layer near the nose. train have time and length scales that are potentially hazardous
These effects will affect the skin friction drag, and thus the to waiting passengers, and in addition are a potentially large
ARTICLE IN PRESS
C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298 297

cause of train drag. It is possible that careful optimisation of train CEN, 2007. Railway applications—Aerodynamics—Part 6: requirements and test
nose/tail design could reduce the intensity of these vortices and procedures for cross wind assessment. DRAFT prEN 14067-6.
Chiu, T.W., 1991. A two dimensional second order vortex panel method for the
help in the alleviation of both issues. flow in a cross wind over a train and other 2D bluff bodies. Journal of Wind
(e) Trains in cross winds experience quite different wind conditions Engineering and Industrial Aerodynamics 37, 43–64.
(in terms of shear and turbulence) when they are operating at full Cooper, R.K., 1985. Atmospheric turbulence with respect to moving ground vehicles.
Journal of Wind Engineering and Industrial Aerodynamics 17, 215–238.
speed than when they are operating at low speed. There is a Copley, J.M., 1987. The 3-D flow around railway trains. Journal of Wind
divergence of view as to which effects are most important i.e. is Engineering and Industrial Aerodynamics 26, 21.
the critical condition when trains are moving at line speed, as Diedrichs, B., 2005. Computational methods for crosswind stability of railway
trains. KTH Engineering Sciences report TRITA AVE 2005:27.
trains ought to operate in all weather conditions, or can it be
Ding, Y., Sterling, M., Baker, C.J., 2008. Train stability in cross winds, a new
expected that in such conditions effects external to the railway approach? Proceedings of the Institute of Mechanical Engineers Part F: Journal
environment (such as flying debris) will cause trains to operate at of Rail and Rapid Transport 222 (1), 85–97.
ESDU, 1985. Characteristics of atmospheric turbulence near the ground: part 2
low speeds, and potentially to come to a stand? The answer to
Single point data for strong wind—neutral atmosphere, Data Item 85020,
this question will determine which type of wind tunnel test is Engineering Sciences Data Unit, London.
most appropriate to determine cross wind forces and moments Eskridge, R.E., Hunt, J.C.R., 1979. Highway modelling part 1: Prediction of velocity
for the purposes of risk calculations. and turbulence fields in the wakes of vehicles. Journal of Applied Meteorology
18, 387–400.
(f) The longitudinal vortices in the wake of trains in cross winds can FLUENT. Unsteady flow behind a high speed train by C Heine and G Matschke,
cause an enhancement of slipstream velocities. It is possible that FLUENT web site; /http://www.fluent.com/about/news/newsletters/02v11i1/
the cross wind condition represents the most serious safety risk a8.htmS.
Hemida, H.N., 2006. Large eddy simulation of the flow around simplified high
when considering the effects of train slipstreams. speed trains under side wind conditions, Licentiate of Engineering thesis,
(g) When modelling the effects of cross winds on trains, if short Chalmers University of Technology, Goteborg, Sweden.
period gusts are modelled with time scales of less than 0.5 s, Holmes, J.D., Sankaran, R., Kwok, K.C.S., Syme, M.J., 1997. Eigenvector modes of
fluctuating pressure on low rise building models. Journal of Wind Engineering
or if suspension effects with frequencies greater than 1 or 2Hz and Industrial Aerodynamics 69–71, 697–707.
are modelled, then it is necessary to include the filtering Ido, A., Saitou, S., Nakade, K., Iikura, S., 2008. Study on under-floor flow to reduce
effects represented by the aerodynamic admittances or ballast flying phenomena. World Congress on Rail Research, Seoul, Paper
number S2.3.4.2.
aerodynamic weighting functions. If effects with rather larger
Jordan, S.C., Johnson, T., Sterling, M., Baker, C.J., 2008. Evaluating and modelling the
time scales are being considered (say the gross overturning of response of an individual to a sudden change in wind speed. Building and
a vehicle with a time scale of 2 to 3 s), then the force Environment 43, 1521–1534.
Kaltenbach, H.-J., Gautier, P.-E., Agirre, G., Orellano, A., Schroeder-Bodenstein, K.,
fluctuations on the train can be assumed to be quasi-steady
Testa, M., Tielkes, Th., 2008. Assessment of the aerodynamic loads on the
and to follow the wind velocity fluctuations. 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.
Acknowledgements Kwon, H.B., Park, C.S., 2006. An experimental study on the relationship between
ballast flying phenomenon and strong wind under high speed train. World
Congress on Rail Research Montreal, Paper T3.3.2.3.
Much of the material presented in this paper has been taken Mair, W.A., Stewart, A.J., 1985. The flow past yawed slender bodies, with and
from the results of the author’s collaborators over the last two without ground effects. Journal of Wind Engineering and Industrial Aero-
dynamics 18, 301.
decades (research students and fellows, former and present Mancini, G., Malfatti, A., 2001. Full scale measurements on high speed train ETR
colleagues), who are too numerous to name individually. None- 500 passing in open air and in tunnels of Italian high speed line. TRANSAERO A
theless their implicit contribution to this work is gratefully European Initiative on Transient Aerodynamics for Railway System Optimiza-
tion (Notes on Numerical Fluid Mechanics and Multidisciplinary Design
acknowledged. (NNFM)) by Burkhard Schulte-Werning, Remi Gregoire, Antonio Malfatti, and
Gerd Matschke, pp. 101–122.
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
References geometry. SAE Congress, Detroit, Paper 929202.
Paradot, N., Talcotte, C., Willaime, A., Guccia, L., Bouhadana, J.-L., 1999.
Baker, C.J., 1991a. Ground vehicles in high cross winds—Part I Steady aerodynamic Methodology for computing the flow around a high speed train for drag
forces. Journal of Fluids and Structures 5, 69–90. estimation and validation using wind tunnel experiments. World Congress on
Baker, C.J., 1991b. Ground vehicles in high cross winds—Part 2 Unsteady Rail Research, Tokyo.
aerodynamic forces. Journal of Fluids and Structures 5, 91–111. Quinn, A., Hayward, M., 2008. Full scale aerodynamic measurements underneath a
Baker, C.J., 1991c. Ground vehicles in high cross winds—Part 3 The interaction of high speed train. In: Proceedings of the BBAA VI, Milano, Italy, July 20–24.
aerodynamic forces and the vehicle system. Journal of Fluids and Structures 5, RAPIDE Consortium, 2001. Synthesis report of RAPIDE project. Aerodynamics
221–241. workshop, Koln.
Baker, C.J., 2000. Aspects of the use of the technique of orthogonal decomposition Robinson, C.G., Baker, C.J., 1990. The effect of atmospheric turbulence on trains.
of surface pressure fields. Wind and Structures 3, 2. Journal of Wind Engineering and Industrial Aerodynamics 34, 251–272.
Baker, C.J., 2001a. Flow and dispersion in vehicle wakes. Journal of Fluids and Sanz-Andres, A., Santiago-Prowald, J., 2002. Train-induced pressure on
Structures 15 (7), 1031–1060. pedestrians. Journal of Wind Engineering and Industrial Aerodynamics 90,
Baker, C.J., 2001b. Unsteady wind loading on a wall. Wind and Structures 4 (5), 1007–1015.
413–440. Schetz, J.A., 2001. Aerodynamics of high speed trains. Annual Review of Fluid
Baker, C.J., Dalley, S.J., Johnson, T., Quinn, A., Wright, N.G., 2001. The slipstream and Mechanics 33, 371–414.
wake of a high speed train. Proceedings of the Institution of Mechanical Schulte-Werning, B., Heine, C., Matschke, G., 2003. Unsteady wake characteristics
Engineers F Journal of Rail and Rapid Transit 215, 83–99. of high speed trains. PAMM Proceedings Applied Maths and Mechanics 2,
Baker, C.J., Lopez-Calleja, F., Jones, J., Munday, J., 2004. Measurements of the cross 332–333.
wind forces on trains. Journal of Wind Engineering and Industrial Aero- Sims-Williams, D., Dominy, R., Howell, J., 2001. An investigation into large scale
dynamics 92, 547–563. unsteady structures in the wake of real and idealised hatchback car models.
Baker, C.J., Sterling, M., Johnson, T., Figura-Hardy, G., Pope, C., 2007. The effect of SAE Congress, Detroit, Paper 2001-01-1041.
crosswinds on train slipstreams. In: International Conference on Wind Sockel, H., 1996. The aerodynamics of trains. In: Schetz, J.A., Fuhs, A.E. (Eds.),
Engineering, Cairns, Australia. Handbook of Fluid Dynamics and Fluid Machinery. Wiley, New York, pp.
Baker, C.J., Sterling, M., Bouferrouk, A., O’Neil, H., Wood, S., Crosbie, E., 2008a. 1721–1741.
Aerodynamic forces on multiple unit trains in cross winds. In: Proceedings of Sterling, M., Baker, C., Quinn, A., Hoxey, R., Richards, P., 2006. An investigation of
the BBAA VI, Milano, Italy, July 20–24. the wind statistics and extreme gust events at a rural site and implications for
Baker, C.J., Bouferrouk, A., Perez, J., Iwnicki, S.D., 2008b. The integration of cross wind tunnel testing. Wind and Structures 9 (3), 193–216.
wind forces into train dynamic calculations. World Congress on Rail Research, Sterling, M., Baker, C.J., Bouferrouk, A., O’Neil, H., Wood, S., Crosbie, E., 2008a.
Seoul, S Korea. An investigation of the aerodynamic admittances and aerodynamic
Brockie, N.J.W., Baker, C.J., 1990. The aerodynamic drag of high speed trains. weighting functions of trains. In: Proceedings of the BBAA VI, Milano, Italy,
Journal of Wind Engineering and Industrial Aerodynamics 34, 273–290. July 20–24.
ARTICLE IN PRESS
298 C. Baker / J. Wind Eng. Ind. Aerodyn. 98 (2010) 277–298

Sterling, M., Baker, C.J., Jordan, S.C., Johnson, T., 2008b. A study of the slipstreams Wu, D., 2004. Predictive prospects of unsteady detached-eddy simulations
of high speed passenger trains and freight trains. Proceedings of the Institute in industrial external aerodynamic flow simulations. Diploma thesis,
of Mechanical Engineers Part F: Journal of Rail and Rapid Transport 222 (2), Lehrstuhl fur Strmungslehre und Aerodynamishes Institute, Aachen,
177–193. Germany.
Vino, G., Watkins, S., Mousley, P., Watmuff, J., Prasad, S., 2005. Flow structures in the Xu, Y.L., Ding, Q.S., 2006. Interaction of railway vehicles with track in cross-winds.
near wake of the Ahmed model. Journal of Fluids and Structures 20, 673–695. Journal of Fluids and Structures 22, 295–314.

You might also like