Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

The Journal of The Textile Institute

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/tjti20

A study of drag reduction with textile roughness


on a cyclist model

X. Y. Hsu, Jiun-Jih Miau, T. H. Ku, J. J. Chen, W. C. Yuan, Y. H. Lai, Y. R. Chen, Y. J.


Chen, C. H. Tseng, C. H. Chen, S. S. Jan, Y. S. Ciou, Y. Chen & C. W. Chiu

To cite this article: X. Y. Hsu, Jiun-Jih Miau, T. H. Ku, J. J. Chen, W. C. Yuan, Y. H. Lai, Y. R.
Chen, Y. J. Chen, C. H. Tseng, C. H. Chen, S. S. Jan, Y. S. Ciou, Y. Chen & C. W. Chiu (2022)
A study of drag reduction with textile roughness on a cyclist model, The Journal of The Textile
Institute, 113:6, 1206-1230, DOI: 10.1080/00405000.2021.1950978

To link to this article: https://doi.org/10.1080/00405000.2021.1950978

© 2021 The Author(s). Published by Informa


UK Limited, trading as Taylor & Francis
Group.

Published online: 26 Jul 2021.

Submit your article to this journal

Article views: 1265

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=tjti20
THE JOURNAL OF THE TEXTILE INSTITUTE
2022, VOL. 113, NO. 6, 1206–1230
https://doi.org/10.1080/00405000.2021.1950978

ARTICLE

A study of drag reduction with textile roughness on a cyclist model


X. Y. Hsua, Jiun-Jih Miaub , T. H. Kuc, J. J. Chenb, W. C. Yuanb, Y. H. Laib, Y. R. Chenb, Y. J. Chenb,
C. H. Tsengb, C. H. Chend, S. S. Janb, Y. S. Cioua, Y. Chena and C. W. Chiue
a
Taiwan Textile Research Institute, New Taipei City, Taiwan; bNational Cheng Kung University, Tainan, Taiwan; cFu Jen Catholic University,
New Taipei City, Taiwan; dNational Pingtung University of Science and Technology, Pingtung, Taiwan; eNational Taiwan University of
Science and Technology, Taipei, Taiwan

ABSTRACT ARTICLE HISTORY


Efforts were made to examine the effect of textile roughness for the reduction of aerodynamic drag Received 16 July 2020
experienced on a cyclist model. First of all, flow visualization experiments were conducted in a water Accepted 6 April 2021
channel with a 1/5 scale cyclist model to identify the flow separation lines on the contoured surface.
KEYWORDS
Subsequently, based on the information learned, the idea of delaying flow separation on the model
Textile roughness;
surface was implemented by means of sport suits. Seven sport suits featuring different textile materials aerodynamic drag; cyclist
and local roughness patterns were tested on a full-scale cyclist model in wind tunnel experiments, in model; critical
addition to a reference case without a sport suit. From the CD data deduced, we identified a critical transition phenomenon
condition in most of the cases studied, featuring a least drag coefficient occurred at a Reynolds num-
ber within the flow speed range studied. For the best case found, the drag coefficient is amounted
7.5% less than that of the reference case at the same Reynolds number. Furthermore, the Cp and Cprms
values reduced from the pressure measurements on the cyclist model with the sport suits unveil a
trend that the reduction of drag actually infers the retreat of the local flow separation lines toward
the rear side of the body accompanied with less severe the flow unsteadiness.

Introduction Given the posture of a cyclist model fixed, this study aims
to examine the effect of textile roughness on delaying flow sep-
In cycling aerodynamics, the drag produced by a riding cyclist
aration on the contoured surface. In the literature, a number of
gets pronounced over the other components of the total drag
studies have modeled the flow phenomenon using single circu-
as the speed increases. For example, Kyle and Burke (1984)
lar cylinder or multiple circular cylinders roughened with textile
pointed out that at a speed higher than 8.9 m/s, 90% of the
materials (Chowdhury et al., 2009, 2014; Oggiano et al., 2007,
total drag is due to the aerodynamic drag, of which 70% is pro-
duced by the cyclist. Extensive studies on the aerodynamic drag 2009). Recently, Hsu et al. (2019) performed an experimental
of cyclists can be found in the literature, encompassing a wide work with 38 fabric samples tested on a circular cylinder in a
range of topics. To name a few, research topics include the pos- wind tunnel experiment, and concluded that the roughness of
tures of cyclists (Blocken et al., 2018a; Defraeye et al., 2010, the textile materials could be expressed in terms of the
2011; Fintelman et al., 2014), dynamic motions of the cyclist Nikuradse’s sand-grain roughness (Nikuradse, 1937). By this
legs (Crouch et al., 2014, 2016a, 2016b) and multiple cyclists at finding, the effect of textile roughness on the drag crisis phe-
racing (Blocken et al., 2013, 2018b; Defraeye et al., 2014). nomenon could be described by the empirical relations origin-
From the perspectives of fluid dynamics, a cyclist can be ally proposed for other types of roughness elements (Basu,
regarded as a three-dimensional bluff body, from which 1985). As for the drag crisis phenomenon of flow over a circu-
flow gets separated and developed into a wake region down- lar cylinder, a significant number of works can be found in the
stream (Crouch et al., 2014, 2016a, 2016b, 2017; Miau et al., literature (Achenbach, 1968; Bearman, 1969; Farell & Blessman,
2020). According to a flow visualization experiment made in 1983; Roshko, 1993; Schewe, 1983; Wieselsberger, 1922;
a water channel with a 1/5 scale cyclist model, Miau et al. Zdravkovich, 1997). Basically, the phenomenon is known for a
(2020) pointed out that the limiting streamline pattern drastic reduction in the drag coefficient within a small range of
(Lighthill, 1963) on the surface provided valuable informa- Reynolds numbers, and is very sensitive to free stream turbu-
tion concerning the development of three-dimensional flow lence and surface roughness (Achenbach, 1971; Nakamura &
separation from the model. Miau et al. (2020) further com- Tomonari, 1982; Wieselsberger, 1922; Zdravkovich, 1997).
mented that delaying the flow separation would lead to In examining the effect of textile roughness on the drag
reduction of the aerodynamic drag and suppression of the coefficients of a cyclist and a circular cylinder, Brownlie
wake region behind the cyclist model. et al. (2009) found that the drag reduction caused by a sport

CONTACT Jiun-Jih Miau jjmiau@mail.ncku.edu.tw Department of Aeronautics and Astronautics, National Cheng Kung University, Tainan, Taiwan.
ß 2021 The Author(s). Published by Informa UK Limited, trading as Taylor & Francis Group.
This is an Open Access article distributed under the terms of the Creative Commons Attribution-NonCommercial-NoDerivatives License (http://creativecommons.org/licenses/by-nc-nd/4.
0/), which permits non-commercial re-use, distribution, and reproduction in any medium, provided the original work is properly cited, and is not altered, transformed, or built upon in
any way.
THE JOURNAL OF THE TEXTILE INSTITUTE 1207

suit on the cyclist could be up to 7.6%, whereas the rough- water-channel and wind-tunnel experiments, respectively.
ness effect on the reduction of the drag of a circular cylinder The cyclist models were made by a 3-D printer, with the con-
could be up to 50% provided that the flow was experienced tour data provided by TTRI (Taiwan Textile Research
through a critical transition process. Brownlie et al. (2009) Institute). For the water-channel experiments, a bike model
emphasized that the drag crisis could be triggered by the tex- was made by the same 3-D printer to fit the 1/5 scale cyclist
tile roughness on a cyclist, thus resulting in significant reduc- model, whereas for the wind-tunnel experiments a commer-
tion of drag. In addition, Debraux et al. (2011) examined the cial bike was selected for the full-scale model. It is worthwhile
data of the effective frontal area of an elite cyclist, which was to mention that the full-scale model was too large to be
defined as a product of the projected area of the cyclist and printed in one piece. Actually, it was segmented into 23
the drag coefficient. The data reported by Debraux et al. pieces for printing, then assembled afterwards.
(2011) were obtained in a wind-tunnel experiment with the Detailed dimensions of a 1/5 scale cyclist model are

cyclist at a static position for speeds over a range of 4.2 m/s shown in Figure 2. Notably, its pitch angle is 25 , which is
to 27.8 m/s. A least drag coefficient was identified at a speed defined as the angle between the torso of the model indi-
between 11.1 m/s and 13.9 m/s; the reduction was amounted cated and the horizontal axis aligned in the direction of the
to 10% of the drag coefficient at the lowest speed 4.2 m/s. As incoming flow. The torso length, L, is denoted as the char-
noted, the extents of drag reduction reported in Debraux acteristic length in this study; L ¼ 135 mm for the 1/5 scale
et al. (2011) and Brownlie et al. (2009) are comparable. model. Moreover, the coordinate system employed for this
The present study was carried out experimentally and study is indicated in the figure with the origin located at the
numerically. First, flow visualization experiments were per- surface of the pelvis region, where x, y and z denote the
formed in a water channel with a 1/5 scale cyclist model to streamwise, lateral and vertical directions, respectively.
examine the complex flow phenomenon near the model surface.
Particular attention was paid to identify the locations of the flow
separation lines on the surface. In parallel to the flow visualiza- Methods for the water-channel experiments
tion experiment, numerical work was conducted to obtain the
A closed-return type water channel was employed for the
time-mean flow distribution around the cyclist model, from
flow visualization experiments, with a test section of 0.6 m
which the wall shear stress distribution on the model surface
by 0.6 m in cross section and 2.5 m in length. The cyclist
was deduced to supplement the findings of the water-channel
model was situated 1 m downstream from the inlet of the
flow visualization which basically provided the qualitative infor-
test section, where the turbulence intensity measured was
mation of the flow phenomenon. Second, experiments were car-
about 1% of the incoming flow velocity at 0.09 m/s. As
ried out in a wind tunnel with a full-scale model to obtain the
shown in Figure 3(a), the cyclist model and the bike were
drag force using a balance and the real-time pressure measure-
situated on a supporting plate, with a blockage ratio of 6.8%
ments on the contoured body surface. The results of the model
in total. Care was taken prior to the experiments to verify
with seven sport suits together with a reference case without
that the disturbance introduced by the ground plate had lit-
sport suit were compared. A major finding is that the results of
the drag measurements confirm the effectiveness of drag reduc- tle effect on the flow downstream. This was confirmed by
tion due to textile roughness. The results of pressure measure- the dye-streak visualization image shown in Figure 3(b),
ments further provide insights into the roughness effect on unveiling that the disturbance produced from the leading
modifications of the boundary-layer flows on the local surfaces. edge had little impact on the flow downstream.
Flow visualization experiments were performed using the
dye-streak, oil-film and ink-dot methods. For the dye-streak
Methodology method, the dye was introduced upstream from the model
to reveal the flow motions in the region of interest. The
The cyclist models
outer diameter of the dye injection tube was 1 mm, that
Figure 1(a) and 1(b) shows the photos of a 1/5 scale model the Reynolds number based on the freestream velocity and
and a full-scale model at a dropped position for the the diameter of the tube was about 60. Note that the

Figure 1. The photos of (a) a 1/5 scale model with a bike for the water-channel experiment and (b) a full-scale model with a bike for the wind-tunnel experiment.
1208 X. Y. HSU ET AL.

Figure 2. Detailed dimensions of the 1/5 scale model with the coordinate system defined for this study.

Figure 3. (a) A photo taken for the test section of the water channel. (b) A dye streak introduced upstream in the test section proving that flow over the leading
edge of the ground plate did not produce additional disturbance.

disturbance generated by the injection tube resulted in a wake campus was employed for the present study. The wind tun-
behind the tube. Under the flow condition, the disturbance gen- nel is equipped with two test sections, including a primary
erated is categorized as a laminar wake (Roshko, 1954; Van test section which is 2.6 m by 4 m in cross section at the
Dyke, 1982), whose velocity deficit would be diffused by the vis- inlet and 36.5 m long, and the other which is 2.6 m by 6 m
cous effect mainly. Thus, the disturbance produced could be in cross section at the inlet and 21 m long. Driven by an
regarded as insignificant to flow motion. For the oil-film axial fan with a motor rated 450 kW, the maximum speed
method, a white water-color paint was applied on the surface of in the primary test section could reach 36 m/s. At 2.5 m
the model. After the model was placed in the flow for a length downstream from the inlet of the primary test section,
of time until most of the paint material on the windward side where the cyclist model was situated, the non-uniformity of
of the model was carried away, the flow separation lines would the flow was verified less than 0.4% of the free stream vel-
be revealed clearly. As for the ink-dot flow visualization tech- ocity (Kao, 2005; Miau et al., 2013). In the same cross-sec-
nique, the same water-color paint was applied on the model tional plane, the free stream turbulence intensity measured
surface in dots at the pre-determined grid points. Until the dot- by a single normal hot-wire probe was about 0.35% for the
ted paint dried for a certain length of time in the air, the model free stream velocities at 10 and 20 m/s, respectively. The
was placed in the test section. As a result, a limiting streamline thickness of the boundary layer developed on the floor at
pattern would be emerged from the appearance of the ink-dots this streamwise location was about 60 mm at the free stream
due to the viscous effect of flow near the surface. velocity 10 m/s.
Experiments were carried out for the full-scale cyclist
model shown in Figure 1(b), on which 32 pressure taps out
Methods for the wind-tunnel experiments
of 75 in total were selected for measurements. See Figure 4
The ABRI (Architecture and Building Research Institute) for the locations of the pressure taps selected. Each pressure
wind tunnel (Miau et al., 2013) at the NCKU Kuei-Ren tap was connected to a pressure sensor, AMS-5812-000-D-B,
THE JOURNAL OF THE TEXTILE INSTITUTE 1209

Figure 4. Indications of 32 pressure taps on the model surface with the frontal, top and side views.

Analog Microelectronics GmbH, featuring a maximum (2019) demonstrated that applying a local roughness pattern
frequency response of 500 Hz. Prior to the experiment, on the cyclist model without textile material could have sig-
calibration was made for each of the sensors to obtain a nificant impact on drag reduction. The idea was that if a
relation between the output voltages and the known pres- roughness pattern could be placed on the model surface at a
sures, which were provided by a Druck DPI-610 portable location appropriate to disturbing the boundary-layer flow
pressure calibrator. effectively, then the delay of flow separation would be
For the experiment, a self-made force balance system was remarkable. Chen et al. (2019) compared the results of ten
employed to obtain the aerodynamic drag of the full-scale different roughness patterns and found that a pattern called
model together with the bike, with an area blockage ratio the skeleton triangle, which is shown in Figure 6, was the
of 4.2% in total. In this study, no correction was made on most effective for drag reduction. In Chen et al. (2019), the
the drag force measured with regard to the area blockage roughness patterns were applied on the shoulders, the upper
effect. The balance was designed to measure a one-compo- arms, two sides of the waist and the outer sides of the upper
nent force in the streamwise direction. As shown in Figure legs at locations upstream, but close to, the flow separation
5(a), it was equipped with two S-Beam load cells, CLB- lines as learned from the flow visualization experiments in
200NA, Tokyo Measuring Instruments Lab., which were the water channel. The thickness of each of the roughness
installed in opposite directions to account for the forces in patterns was 1 mm. A maximum reduction amounted to
both directions. Calibration for each of the load cells was 6.7% was found in comparison with the drag coefficient of
made prior to the wind-tunnel experiments. Figure 5(b) the model without a roughness pattern at the same
and 5(c) shows the calibration curves of the two load Reynolds number. The Reynolds number is defined as Re¼
cells, respectively. For each of the load cells, the linearity VrefL/m; m denotes the kinematic viscosity.
of calibration could be examined by the correlation coeffi- The skeleton triangular pattern proposed by Chen et al.
cient enclosed. (2019) was implemented in the design of sport suits in this
In the wind tunnel, a Pitot tube was situated at the inlet study. Seven sport suits were made for tests. See Figure 7
of the test section to measure the free stream velocity, Vref, for the photos of the sport suits, named Cases 1R, 2R, 2RT,
which is denoted as the reference velocity for experiment. It 3R, 3RT, 4R and 4RT. The characteristic features of the
was reduced from the output of a differential pressure trans- sport suits1R, 2R, 2RT, 3R and 3RT are further described in
ducer, Validyne DP-103, calibrated with the Druck DPI-610 Table 1, which were made of four types of textile materials,
portable pressure calibrator mentioned above. A thermom- named Types A, B, C and D. On the other hand, the textile
eter was installed in the test section to monitor the tempera- materials of sport suits 4R, 4RT are identical and uniform
ture of the flow. For each set of the measurements, the in roughness, as described in Table 2.
output signals from the 32 pressure transducers, the force Roughness of the textile materials is deemed critical to
balance, the differential pressure transducer with the Pitot modify the boundary-layer flow on the model surface.
tube and the thermometer were sampled simultaneously by Roughness measurements of the textile materials were car-
the analog to digital converter modules, NI-9215, at a rate ried out with a VK-9710 3 D laser scanning confocal micro-
of 1 kHz per channel for 120 s. scope. Each image obtained was reduced for a quantity
named S10z (ISO 25178, 2012), according to the definition
Design of textile roughness below.
S10z ¼ S5p þ S5v (1)
The sport suits tested in the wind tunnel experiments were
made of different types of textile materials with addition of where S5p and S5v are the average heights of the five high-
local roughness patterns. In an earlier study, Chen et al. est peaks and the deepest valleys, respectively, within an
1210 X. Y. HSU ET AL.

Figure 5. (a) A schematic view of the self-made force balance, (b) the calibration results of load cell A, and (c) the calibration results of load cell B.

Figure 6. Illustrations of the roughness pattern called skeleton triangle (Chen et al., 2019). (a) The design drawing, (b) a view of the roughness pattern made of
emulsion, and (c) the patterns applied on the shoulders, arms, waist and upper legs of the cyclist model (Chen et al., 2019).
THE JOURNAL OF THE TEXTILE INSTITUTE 1211

Figure 7. The photos of the sport suits for the seven cases: Cases 1R, 2R, 2RT, 3R, 3RT, 4R and 4RT. Note that the characteristics of the sport suits are also described
in Tables 1 and 2.

area of 1.5 mm by 1 mm measured. The relative roughness for each of the types, the non-uniform roughness can be
given in Tables 1 and 2 is defined as the S10z value normal- highlighted by the roughness measurements in two selected
ized by the reference length, L ¼ 0.675 m, the torso length regions, named a and b, which are characterized by differ-
of the full-scale model. ent textures. Moreover, in each of the photos, a dashed-line
As indicated in Table 1, the textile materials of Types B, area marks a repeated pattern of the non-uniform rough-
C and D are non-uniform in roughness. Details about the ness. A ratio of this area to a square of 20 mm by 20 mm
roughness patterns are further shown in Figure 8. Notably, indicated by the solid line is provided for reference.
1212 X. Y. HSU ET AL.

Figure 7. Continued

For Cases 2RT, 3RT and 4RT, the skeleton triangle pat- the quantities of time-mean value F and root-mean square
terns (Chen et al., 2019) were implemented. The thicknesses value Frms, respectively. See the expressions below.
of the patterns varied in different cases, with the thicknesses
1X N
of 2.45 mm, 1.85 mm and 0.5 mm for Cases 2RT, 3RT and F¼ F ðiÞ (2)
4RT, respectively. In addition, a case named Case O denot- N i¼1
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
ing the model without a sport suit was studied. The results u
u 1 X N
of Case O will be referenced for comparison later. Frms ¼t ðFFÞ2 (3)
N  1 i¼1

Statistical analysis of the experimental data F(i): the real-time force value at the data point i sampled;
N: the number of the data points in a sampled
In this study, the data sampled within a length of time for
record, N ¼ 12000.
analysis, named a sampled record, is treated as a stationary
Subsequently, the drag coefficient, CD, and its root-mean
random process (Bendat & Piersol, 1991). The methods of
square value, CDrms, can be reduced according to the expres-
data analysis are presented below. The force-balance data
sions below.
obtained from a wind tunnel experiment can be reduced to
THE JOURNAL OF THE TEXTILE INSTITUTE 1213

Table 1. Characteristics of the sport suits 1R, 2R, 2RT, 3R and 3RT, each of which was made of a combination of different types of textile materials as indicated.
(a) The textile materials used in the sport suits
Fabric type Weight (gsm) Fiber content Fiber structural Relative roughness (S10z/L) Remarks
Type A 130 86% polyester Warp knitting 5.88  104 Uniform roughness
14% elastane
Type B 180 82% polyester Warp knitting a: 9.24  104 Non-uniform roughness
18% elastane b: 7.24  104
(See Figure 8a for the sampled areas a and b)
Type C 150 80% polyester Warp knitting a: 9.04  104 Non-uniform roughness
20% elastane b: 5.6  104
(See Figure 8b for the sampled areas a and b)
Type D 110 94% polyester Warp knitting a: 7.61  104 Non-uniform roughness
6% elastane b: 6.12  104
(See Figure 8c for the sampled areas a and b)
(b) Descriptions of the design of the sport suits (See also Figure 7)
Areas of the model surface with different fabric types
Inner side of the
Shoulders and upper arms and Local
Front of the outer side of Rear of the the two sides of roughness
Case upper body upper arms upper body the upper body pattern
1R A C C D None
2R A A C D None
2RT A A C D Yes (2.45 mm thick)
3R B B A D None
3RT B B A D Yes (1.85 mm thick)

Table 2. Characteristics of sport suits 4R and 4RT, whose base textile materials were identical and uniform in roughness.
Case Weight (gsm) Fiber content Yarn number Fiber structure Relative roughness (S10z/L) Local roughness pattern
4R 200 Nylon/Spandex ¼ 80/20 Nylon: 40 D; Spandex: 40 D Warp knitting 5.56  104 none
4RT 200 Nylon/Spandex ¼ 80/20 Nylon: 40 D; Spandex: 40 D Warp knitting 5.56  104 Yes (0.5 mm thick)

F Prms
CD ¼ 1 (4) CPrms ¼ 1 (9)
2 qair Vref
2
q 2
2 air ref At
V
Frms In the expressions above, Pstatic denotes the static pres-
CDrms ¼1 (5)
q 2
2 air Vref At sure of the freestream flow. In addition, a real-time pressure
In the expressions above, qair denotes the density of the coefficient, Cpr, is defined below.
air; At denotes the frontal area of the cyclist model together PðiÞPstatic
with the bike, 0.438 m2. CPr ðiÞ ¼ (10)
2 qair Vref
1 2
Similarly, the pressure data obtained at each of the pres-
sure taps on the model surface in a wind tunnel experiment
can be reduced to the time-mean value P and the root- CFD method for flow simulation
mean square value Prms, respectively. See the expressions
below. Numerical simulation of flow around the cyclist model
X
N without a bike was made to examine the phenomenon of
1
P¼ P ðiÞ (6) flow separation from the body surface. A commercial
N i¼1 software ANSYS Fluent was employed to solve the RANS
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u (Reynolds-Averaged Navier-Stokes) equations using the
u 1 X N
Prms ¼t ðPPÞ2 (7) SIMPLE algorithm. In solving the equations, a k-x SST
N  1 i¼1 (Shear-Stress Transport) turbulence model was adopted.
See Figure 9(a) for the computation domain. The dimen-
P(i) denotes the real-time pressure value at the data point
sions were defined according to the test section of the
i sampled.
water channel. A 1/5 scale cyclist model with no bike was
Subsequently, the pressure coefficient, Cp, and its root-
mean square value, CPrms, can be reduced according to the situated in the domain. An adaptive grid system generated
expressions below. by the ANSYS ICEM software was employed, which con-
sisted of seven million cells. The grid system is illustrated
PPstatic in Figure 9(b). Care was taken with the mesh size
CP ¼ 1 (8)
2 qair Vref
2
near the model to ensure that the distance from the first
grid point to the surface of the body, in terms of yþ, a
1214 X. Y. HSU ET AL.

Figure 8. The non-uniform textile materials of (a) Type B, (b) Type C and (c) Type D shown in Table 1. In each photo, a and b indicate the regions where the
roughness measurements were made. The laser scanning images of the measurements corresponding to the two regions are enclosed. The percentage value indi-
cated in each of the photos is the ratio of the area enclosed by the dashed line characterizing the repeated pattern of roughness versus the area enclosed by the
solid line, 20 mm by 20 mm.

non-dimensional distance based on the wall-friction vel- Results and discussion


ocity (White, 1974) and the kinematic viscosity, be less
Results of the water-channel experiments
than 1. For the computation made at Re¼ 1.4  104 , the
height of the first cell from the surface was 1.5  104 m. The results of flow visualization of the 1/5 scale model
More information regarding the present computation obtained in the water channel at Re ¼ 1.4  104 together
method can be found in Tseng (2019). with the CFD results obtained at the same Reynolds number
THE JOURNAL OF THE TEXTILE INSTITUTE 1215

Figure 9. The setup for the numerical simulation at Re ¼ 1.4  104. (a) The computation domain for simulating the flow over a 1/5 scale cyclist model in the test
section of the water channel. (b) An illustration of the adaptive grid system employed for the numerical simulation.

Figure 10. (a) Oil-film visualization on the back surface of the model; (b) an illustration of the counter-rotating vertical flows over the back of the model (Miau et
al., 2020); (c) the wall shear-stress distribution on the back surface obtained by the present numerical simulation; (d) the ink-dot flow visualization on the back sur-
face; (e) the skin friction distribution on the back surface; (f) and (g) the dye visualization of flow around the head of the model, respectively.

are shown in Figure 10. Note that the flow visualization In Figure 10(a), the oil-film flow visualization results
experiments were made for the model together with a sting unveil the formation of the flow separation lines on the rear
support and a ground plate on the floor of the test section. side of the shoulders, which were extended toward the cen-
On the other hand, the CFD simulation was carried out for ter of the back of the model. This appearance is associated
the model only, for the sake of simplicity. with two large-scale counter-rotating vortical flows induced
1216 X. Y. HSU ET AL.

Figure 11. (a) Oil-film visualization on the right side of the model; (b) wall shear-stress distribution by numerical simulation on the right side of the model; (c) oil-
film visualization on the left side of the model; (b) wall shear-stress distribution obtained by numerical simulation on the left side of the model.

by flow separation from the shoulders and the upper body, separation lines can be clearly discerned. Basically, the flow
subsequently re-attached at the center of the back surface. separation lines revealed from the shear stress distribution
For reference, a numerical plot of the streamline distribu- are conformed with those seen in the experimental flow
tion in Figure 10(b) (Miau et al., 2020) for a cyclist model visualization indicated by dashed lines in the numerical
at the hoods position illustrates the formation of the vortical image. This comparison gives a strong support to the pre-
flows over the back surface. This flow characteristic was also sent numerical method, which is capable of unveiling the
mentioned in Crouch et al. (2014) and Crouch et al. (2017) major flow characteristics around the cyclist model.
for models at different postures. Figure 10(c) shows the More details concerning the formation of the separation
numerical results of the wall shear stress distribution on lines can be seen with the ink-dot visualization image and
the back surface of the present model. In this plot, the the corresponding numerical results shown in Figure 10(d)
THE JOURNAL OF THE TEXTILE INSTITUTE 1217

Figure 12. The distributions of CD and CDrms versus Re for all the cases studied. The CD and CDrms distribution of Case 3R are highlighted with solid symbols.

and 10(e), respectively. Both images clearly indicate that the appearance of the diffused dye behind the head, one could
formation of a flow separation line is resulted from a con- further make an estimation about the lateral extent of the
vergence of the limiting streamlines nearby. Physically, by wake region.
the law of conservation of mass, a separation line is formed As a counterpart of Figure 10, Figure 11 presents the
due to the convergence of the surrounding fluid near the results of flow visualization obtained from the right and left
surface, leading to a lift of the fluid from the surface sides of the cyclist model together with the CFD simulation
(Lighthill, 1963). Furthermore, the complicated flow separ- results. The dashed lines marked in the photos indicate the
ation patterns revealed are attributed to the contoured sur- flow separation lines identified by the flow visualization,
face of the model. Apart from the separation lines identified which can be referenced for comparison. Two interesting
on the back of the model, two spiral separations formed on aspects of the flow characteristics learned from the experi-
the two sides of the pelvis are noted in the figure. Each of mental and numerical results are described below. First, the
the spiral separations is associated with the convergence flow separation lines along the two sides of the waist and
of the limiting streamlines to a focal point, followed by a lift the outer sides of the two legs basically follow the topology
of the fluid from the focal point (Lighthill, 1963). of the three-dimensional contoured surface as the cyclist
A flow separation region appeared on the rear side of the model is facing the incoming flow. Second, the appearances
head of the model caught our attention, which can be fur- of the flow separations lines on the upper right and left
ther elaborated with the photos in Figure 10(f) and 10(g). arms bear a similarity to a situation of flow over a normal
The photo in Figure 10(f) was taken from a top view of the circular cylinder. This observation strongly infers that flows
model with the dye streaks released upstream from the around the upper arms can be modeled by flow over a cir-
model. As seen, the flow around the head surface was cular cylinder, which will be elaborated later.
entrained into a wake region, which was extended over the Above findings enlighten an important aspect in sport
rear side of the head and the neck region. A side view suit design. Namely, if a sport suit with a roughness element
shown in Figure 10(g) further indicates that a reversed flow could have significant impact on delaying the flow separ-
was developed from the neck towards the head, subse- ation lines on the model, the reduction of drag would be
quently shedding downstream. In accordance with the remarkable. Accordingly, the sport suits of Cases 2RT and
1218 X. Y. HSU ET AL.

Figure 13. (a) Cp and (b) Cprms distributions versus Re at pressure taps 4-4, 4 and 4-6 on the right shoulder of the model for all cases studied. The Cp and Cprms
distributions of Cases 1R and 2RT are highlighted with solid symbols, which are connected by dashed lines respectively. In each of the plots, the distribution of
Case O is plotted with a dashed line for reference.
THE JOURNAL OF THE TEXTILE INSTITUTE 1219

Figure 13. Continued

3RT were designed with the skeleton triangle patterns measurements were made, the CD value shown actually is
(Chen et al., 2019) on the shoulders, waist and arms, the mean of the repeated measurements, whereas the upper
whereas the Case 4RT was designed with the skeleton tri- and lower error bars indicate the highest and lowest CD val-
angle patterns on the shoulders and arms. See also Figure 7 ues obtained, respectively. As a counterpart of the plot of
for the locations of the roughness patterns on the the CD distributions, a plot of the distributions of CDrms, i.e.,
sport suits. the standard deviation associated with the CD values, is
given in Figure 12(b). Similarly, each of the CDrms value
plotted correspond to the mean of multiple measurements,
Results of the wind-tunnel experiments accompanied with the error bars to indicate the range of the
CDrms values among the measurements.
The CD distributions versus Re for all the eight cases studied For Case O, the CD values are the highest compared to
in the wind-tunnel experiment are presented in Figure the seven cases with the sport suits for the Reynolds num-
12(a). Note that for each of the CD values plotted, the bers higher than 5.5  105. As noted, at Re ¼ 3.9  105 the
accompanied error bars indicate the range of uncertainty. uncertainty associated with the plotted value was pronoun-
Owing to that at each of the flow speeds measured multiple cedly higher than those at higher Reynolds numbers due to
1220 X. Y. HSU ET AL.

Figure 14. (a) Cp and (b) Cprms distributions versus Re at pressure taps 4-8, 4-10 and 4-14 on the left shoulder of the model for all cases studied. The Cp distribu-
tions of Cases 1R and 2RT are highlighted with solid symbols, which are connected by dashed lines respectively. The Cprms distributions of Case 1R in the three plots
of (b) are highlighted with solid symbols. In each of the plots, the distribution of Case O is plotted with a dashed line for reference.
THE JOURNAL OF THE TEXTILE INSTITUTE 1221

Figure 14. Continued

the contamination of background noise in the measured sig- O with small differences only. In fact, the distribution of
nals. For Case O, Re¼ 5.5  105 to 10  105, the CD values Case 2R is the only one among the seven cases with sport
appear to decrease monotonically with Re, from 0.71 to suits that shows no appreciable reduction in drag for Re
0.67. On the other hand, the CDrms distribution of Case O higher than 5.5  105. Based on this finding, one can say
shows a trend of decreasing with Re up to 9  105, but that the surface roughness at the shoulders plays a critical
reversed the trend at higher Re. This infers that the flow role to reduce drag at high Reynolds numbers, and in Case
unsteadiness gets more pronounced at Re beyond 9  105. 2R the textile material at the shoulders could be too smooth
Discussion on the CD data of the seven cases with sport to have a significant impact on delaying flow separation.
suits is made below. Comparing the CD distributions of A comparison of the CD distributions of Cases 2R and
Cases 1R and 2R, whose sport suits feature different materi- 2RT, whose textile materials are identical, unveils that the
als at the shoulders but identical in other areas, the differ- latter with the skeleton triangular patterns implemented has
ence is rather significant. Namely, the CD distribution of a significant impact on drag reduction. In the CD distribu-
Case 1R shows a significant reduction compared to that of tion of 2RT, a minimum CD value can be identified at Re
Case O for Re higher than 5.5  105. On the other hand, the between 7  105 and 8  105, beyond which the CD value
distribution of Case 2R appears to be similar to that of Case gets increased. This appearance infers the existence of a
1222 X. Y. HSU ET AL.

Figure 15. The real-time Cpr traces obtained at three pressure taps, 4-8, 4-10 and 4-14, of Case 1R at Re¼ 4.6  105. All three traces consistently indicate that at
the time instant 100 s, the laminar separation bubble is formed on the left shoulder.

critical condition under which the CD value reaches a min- value about 6  105. Evidently, this finding is contrary to
imum. The corresponding Reynolds number can be named that learned from Cases 2R, 2RT, 3R and 3RT above. In
as the critical Reynolds number (Hsu et al., 2019). A similar summary, the comparisons made above suggest that the
finding was pointed out by Basu (1985) in reviewing the effectiveness of the local roughness patterns deserves careful
experimental results of flows over roughened two-dimen- investigation for each of the individual cases.
sional circular cylinders. Physically speaking, the critical To better understand the physical mechanism of surface
flow phenomenon in flow over a two-dimensional circular roughness involved in drag reduction, it is worthwhile to
cylinder is accompanied with the formation of separation examine the pressure data obtained on the model surface.
bubbles on the circular cylinder (Zdravkovich, 1997; In the following, the results of the pressure measurements
Bearman, 1969). Likewise, one would expect that separation made on the shoulders, the right waist and the right arm
bubbles could be formed on the surface of the cyclist model, are presented for discussion. Similar to Figure 12, each of
although the model is much more complicated in a 3-D the Cp or Cprms values plotted in the following figures is the
contoured shape. More findings in this regard will be pre- mean of the values obtained from multiple measurements,
sented later. accompanied with the error bars to indicate the range of
Cases 3R and 3RT feature sport suits with the same tex- variations among the measurements.
tile materials, but the latter includes the skeleton triangular Figures 13 and 14 present the distributions of Cp and
patterns. A comparison of the CD distributions of the two Cprms on the right and left shoulders, respectively, versus
cases reveals a difference similar to that seen in Cases 2R Reynolds number for all the cases studied. The locations of
and 2RT that the artificial roughness pattern has a positive the pressure taps on the right shoulder, named 4-4, 4 and 4-
impact on drag reduction, even though the thickness of the 6, and on the left shoulders, named 4-8, 4-10 and 4-14, are
roughness pattern of Case 3RT is less than that of Case shown in the figures for reference.
2RT. Meanwhile, it is worthwhile to mention that the crit- In comparing the Cp distributions of all the eight cases,
ical CD value obtained in Case 3RT appears to be the lowest one finds a feature in common in the two figures that the
among all the cases presented in the figure. Note that in distributions of Case O show the least negative. In viewing
Figure 12(a), the CD distribution of Case 3RT is highlighted that flow separations were developed on both shoulders
with solid symbols. In this case, the critical condition (Figure 10), this appearance infers that for all the seven
occurred about Re¼ 6.4  105, which is lower than the crit- cases with sport suits the flow separations on the shoulders
ical Reynolds number in Case 2RT mentioned above. were actually delayed downstream relative to the Case O.
Referring to the CD value of Case O at the same Reynolds Further explanation in this regard can be made below.
number, the reduction in Case 3RT is amounted to 7.5%. Delay of flow separation on the upper surface of a shoulder
Comparing the CD distributions of Cases 4R and 4RT, it is equivalent to a situation that the boundary-layer flow per-
is found that the addition of the skeleton triangular patterns sists over the shoulder surface a longer distance, which
at the shoulders and arms in fact has a negative impact on basically is convex in shape. Thus, the attached flow keeps
drag reduction at Reynolds numbers beyond the critical accelerating along the surface, leading to lower static
THE JOURNAL OF THE TEXTILE INSTITUTE 1223

Figure 16. Cp and (b) Cprms distributions versus Re at pressure taps (a) 13-1 and (b) 13-6 on the upper right waist of the model for all the cases studied. The Cp
and Cprms distributions of Case 2RT at the pressure tap 13-1 in (a) are highlighted with solid symbols, which are connected by dashed lines. The Cp and Cprms distri-
butions of Case 2RT and 3RT at the pressure tap 13-6 in (b) are highlighted with solid symbols, which are connected by dashed lines. In each of the plots, the distri-
bution of Case O is plotted with a dashed line for reference.

pressure at the pressure taps measured. Based on this rea- where the flow unsteadiness is mainly associated with the
soning, one can further argue that among the seven cases phenomenon of vortex shedding (Bloor, 1964; Gerrard,
with sport suits Cases 1R and 2RT appear to be the most 1966; Miau et al., 2006). In Figures 13 and 14, it is noted
effective in delaying the flow separation over the shoulders, that for all the cases studied, the Cprms values are about 0.1
as the Cp values show the most negative. In the figures, the in general. This observation infers that the unsteadiness of
Cp distributions of the two cases are highlighted with flow over both shoulders was not affected by the sport suits
solid symbols. appreciably. Nevertheless, special attention is paid to Case
The value of Cprms basically indicates the degree of flow 1R at Re¼ 4.6  105 in Figure 14(b), where the Cprms values
unsteadiness at the pressure tap measured. Presumably, the at the three pressure taps 4-8, 4-10 and 4-14 on the left
flow unsteadiness is predominated by flow separation on the shoulder appear distinctly higher. The distributions of Case
shoulders, a situation like flow over a circular cylinder 1R in the three plots of Figure 14(b) are highlighted with
1224 X. Y. HSU ET AL.

Figure 16. Continued

solid symbols. To gain further insights into this finding, the tripped to promote turbulent momentum mixing, thus delay
real-time pressure coefficient traces in terms of Cpr are pre- flow separation effectively.
sented in Figure 15 for examination. In the figure, the three Discussion on the characteristics of the separated flow
real-time traces consistently indicate an abrupt change from around the right waist of the cyclist model can be made by
1 to 1.5 at t ¼ 100 s. Physically, this appearance infers examining the Cp and Cprms data obtained at pressure taps
that a separation bubble was formed on the left shoulder at 13-1 and 13-6. For all the eight cases studied, the distribu-
the time instant and remained afterwards. The formation of tions of Cp and Cprms versus Re are presented in Figure 16
a separation bubble on the shoulder would alter the pressure for comparison. Several interesting features of the flow char-
distribution remarkably, as evidenced in flow over a circular acteristics deserve attention here. First, on the pressure data
cylinder in the critical regime (Farell & Blessman, 1983; obtained at the pressure tap 13-1, the Cp values of Case O
Gaster, 1969; Lin et al., 2011; Miau et al., 2011; Tani, 1964). show a trend of increasing with Re up to 10  105, and the
More discussions on the Cp distributions of Cases 1R and corresponding Cprms values maintain at a relatively high
2RT in Figures 13(a) and 14(a) can be made below. In level compared to the rest cases, and reach a peak around
Figure 13(a), it is noticed that the Cp distributions of Case Re¼ 7  105. These observations suggest that for Case O,
1R at the pressure taps 4-4, 4 and 4-6 consistently show a flow separation around the waist is accompanied with
significant drop at Re¼ 4.6  105. Similar finding is noted strong flow unsteadiness, which gets less significant at Re
in Figure 14(a) for the Cp distributions of Cases 1R and beyond 7  105. Since the separated flow developed from
2RT at the pressure taps 4-8, 4-10 and 4-14. Remarkable the waist is merged into the wake flow behind the back of
uncertainties associated with the Cp values of Case 1R at the cyclist model, the appearance of high Cprms values could
Re¼ 4.6  105 in all the plots in Figure 14(a) as well as the be intimately associated with the flow unsteadiness of the
corresponding Cprms plots in Figure 14(b) infer the unsteadi- wake flow behind the back of the model. Based on this rea-
ness of the formation of separation bubble. The effect of soning, one could further argue that since the Cprms values
delaying flow separation is seen at this Reynolds number show a trend of decreasing with Re, the flow unsteadiness
and beyond. For Case 2RT, in viewing that the roughness associated with the wake gets less severe as Re increased
elements situated on the windward side of both shoulders beyond 7  105. For the seven cases with sport suits, the Cp
(Figure 7), the boundary layer flows on the shoulders were distributions show less variations with Re and the Cprms
THE JOURNAL OF THE TEXTILE INSTITUTE 1225

Figure 17. (a) Cp and (b) Cprms distributions versus Re at pressure taps 21, 22, 21-1 and 21-4 on the upper right arm of the model for all the cases studied. In the
figure, the Cp and Cprms distributions of Case 2R are highlighted with solid symbols, which are connected by dashed lines; the Cp and Cprms distributions of Case 4R
are highlighted with solid symbols. In each of the plots, the distribution of Case O is plotted with a dashed line for reference.

distributions stay at lower values compared to that of Case that pressure tap 13-6 be located in the wake region. The Cp
O. These findings infer that the wake flow behind the back distribution of Case O appears to be the lowest in value
of the cyclist model tends to be reduced in size with less among the eight cases studied, implying that the flow separ-
unsteadiness, compared to Case O. Examining the data in ation line in this case would be located farthest upstream
the figure, one may further obtain an idea that the most compared to the other cases. This can be reasoned with our
remarkable change is due to Case 2RT, whose Cp and Cprms knowledge about flow around a bluff body. Delaying flow
distributions at the pressure tap 13-1 are highlighted with separation on a contoured surface would result in an
the solid symbols in Figure 16(a). increase in base pressure, such as in the case of flow over a
Second, at pressure tap 13-6, for each of the cases the Cp circular cylinder (Achenbach, 1968; Igarashi, 1978; Roshko,
distribution versus Re appears to be almost constant. 1993). Moreover, a numerical study by Phung et al. (2017)
Meanwhile, the Cprms values stay low. These findings suggest on flow over a cyclist model at the hoods position using a
1226 X. Y. HSU ET AL.

Figure 17. Continued

Large-Eddy-Simulation (LES) method indicated that the flow 13-6 are highlighted in Figure 16(b) with solid symbols. A
separation lines on the contoured surface of the waist region summary based on above findings at two pressure taps sug-
were actually wandering in time, and the drag coefficient in gests that Case 2RT is the most effective in delaying flow sep-
real time varied accordingly. Following this reasoning, one aration in the waist region of the model.
can argue that as far as delaying the flow separation around The Cp and Cprms data obtained at pressure taps 21, 22,
the waist region is concerned, Case 2RT or 3RT appears to 21-1 and 21-4 on the right arm of the cyclist model are pre-
be the most effective, accompanied with the least flow sented in Figure 17 for discussion. Notably, the locations of
unsteadiness measured. Conceivably, the addition of the local pressure taps 21 and 22 are on the windward and rearward
roughness patterns at the waist in the two cases had a signifi- sides, respectively. Pressure taps 21-1 and 21-4 are located
cant impact on the flow development. The Cp and Cprms dis- on the outer side of the arm where the flow separation phe-
tributions of the two cases corresponding to the pressure tap nomenon would be anticipated.
THE JOURNAL OF THE TEXTILE INSTITUTE 1227

Figure 17. Continued

The Cp and Cprms data obtained at pressure tap 22 basic- hand, attention is drawn to Case 2R. The distribution of
ally reflect the characteristics of flow in the wake region Cprms versus Re unveils a trend that the flow unsteadiness
behind the arm. For Case O, the Cp values appear to be appears to be more severe than that of Case O at Re beyond
increased with Re and approach to a level about 1 for Re 7  105. Meanwhile, the Cp and Cprms distributions of Case
beyond 7  105, meanwhile the Cprms values appear to be 2R show no appreciable variations with Re, which infers no
decreased remarkably for Re beyond 7  105. These findings evidence about the occurrence of the critical transition phe-
strongly suggest that flow separation be delayed further nomenon. This can be also learned from the Cp data
rearward as Re gets increased up to 7  105. On the other obtained from pressure taps 21-1 and 21-4 that for Case 2R
1228 X. Y. HSU ET AL.

Figure 18. The real-time Cpr traces obtained at the three pressure taps, 21, 21-1 and 21-4, of Case 4R at Re¼ 6.2  105. The traces obtained at pressure taps 21-1
and 21-4 show fluctuations between the levels of -1 and -3, inferring switching off and on regarding the formation of the separation bubble on the arm surface.

the Cp values stay about 1.5 for Re beyond 7  105, which pressure tap 21-4 shows that it levels off at about 2.5 for
is attributed to no formation of separation bubble on the Re¼ 6  105 and beyond. These findings suggest that in this
outer surface of the right arm. Also noted from the Cp distri- case the separation bubble stay on the arm surface for Re
butions of the pressure tap 21 in the figure, the Case 2R beyond 6  105. The corresponding Cprms distributions
appears to have the least negative values and remain almost appear to stay at low values in this Re range, implying that
constant, slightly less negative than 1, over the Reynolds the separation bubble is stably situated on the arm. In the
number range studied. In the figure, the Cp and Cprms distri- figure, the Cp and Cprms distributions of Case 4R are high-
butions of Case 2R are highlighted with the solid symbols. lighted with solid symbols.
For Case 4R, the Cprms data obtained at Re¼ 6  105, at
pressure taps 21-1 and 21-4, show distinctly high values
Conclusions
compared to those of the other cases. Interestingly, this phe-
nomenon was confirmed by a repeated measurement. To Fruitful results were obtained in this study, in terms of the
gain better understanding on these data, the real-time Cpr flow visualization on a 1/5 scale model made in a water
traces of one set of the measurements at the two pressure channel supplemented by the results of numerical simula-
taps are presented in Figure 18, in which the real-time trace tion, and the drag force and pressure measurements made
at the pressure tap 21 is also included for reference. in a wind tunnel on a full-scale cyclist model with seven
Notably, high Cprms values are attributed to pressure fluctua- sport suits. These findings lead to the following conclusions.
tions with large amplitude characterized by two levels. The
two levels in fact correspond to on and off, respectively, 1. The CD data of six out of the seven cases with sport
concerning the formation of a separation bubble on the arm suits clearly indicate that a critical condition with
surface. This has been seen in flow over a circular cylinder, remarkable drag reduction can be reached within the
for which switching between states appears to be very pro- Reynolds number range studied.
nounced in the critical transition range (Bearman, 1969; Lin 2. Comparing the CD data obtained for the seven cases
et al., 2011; Miau et al., 2011; Zdravkovich, 1997). To be with sport suits indicates that both the base textile
more specific, in the real-time Cpr traces, the level of 3 materials and the local roughness pattern are crucial to
infers the presence of a separation bubble on the arm sur- the reduction of aerodynamic drag. Evaluation on the
face. On the other hand, the pressure coefficient at the level effectiveness should be made for each of the individual
of 1 infers no presence of a separation bubble. The large cases, whereas no general guidelines could be recom-
pressure fluctuations imply that the arm of the cyclist model mended. The best result found in this study is due to
would be experiencing a remarkably unsteady side force. the Case 3RT that the drag reduction is amounted to
For Case 4R, the Cp distribution at pressure tap 21-1 7.5% at Re¼ 6.4  105, compared to that of Case O.
appears to drop to the lowest value about 2.2 at Re¼ 3. By examining the Cp and Cprms data together with the
6  105, and then recovers to a less negative value about CD data obtained for all the cases studied, reduction of
1.6 at higher Re. Meanwhile, the Cp distribution at drag evidently is accompanied with the delay of flow
THE JOURNAL OF THE TEXTILE INSTITUTE 1229

separation on the model and lower the flow unsteadi- Brownlie, L., Kyle, C., Carbo, J., Demarest, N., Harber, E., MacDonald,
ness around the model. R., & Nordstrom, M. (2009). Streamlining the time trial apparel of
cyclists: he Nike Swift Spin project. Sports Technology, 2(1–2),
4. The present findings of separation bubbles on the local 53–60. https://doi.org/10.1080/19346182.2009.9648499
surfaces of the cyclist model provide a directly support Chen, J. J., Miau, J. J., & Tseng, C. H. (2019). Flow over a complex
to the statement made in the literature (Brownlie et al., bluff body: Investigation of flow visualization of the cyclist. 14th
2009) that the critical transition phenomenon could International Symposium on Advanced Science and Technology in
have a significant impact on drag reduction of the aero- Experimental Mechanics, 1–4 November, 2019, Tsukuba, Japan.
Chowdhury, H., Alam, F., Mainwaring, D., Subic, A., Tate, M., Forster,
dynamic flow around a cyclist body. D., & Beneyto-Ferre, J. (2009). Design and methodology for evaluat-
ing aerodynamic characteristics of sports textiles. Sports Technology,
2(3–4), 81–86. https://doi.org/10.1080/19346182.2009.9648505
Acknowledgements Crouch, T. N., Burton, D., Brown, N. A. T., Thompson, M. C., &
The authors would like to express their sincere appreciation to Taiwan Sheridan, J. (2014). Flow topology in the wake of a cyclist and its
effect on aerodynamic drag. Journal of Fluid Mechanics, 748, 5–35.
Textile Research Institute for providing the 3-D contour data of the
https://doi.org/10.1017/jfm.2013.678
cyclist model for this study and Architecture Building Research
Crouch, T. N., Burton, D., LaBry, Z. A., & Blair, K. B. (2017). Riding
Institute for the technical assistance in the wind tunnel experiments.
against the wind: A review of competition cycling aerodynamics.
Sincere appreciation also goes to Mr. Tzu-Liang Chen who designed
Sports Engineering, 20(2), 81–110. https://doi.org/10.1007/s12283-
the force balance for the wind tunnel experiments. Funding support by
017-0234-1
Ministry of Science and Technology, Taiwan, under the project num-
Crouch, T. N., Burton, D., Thompson, M. C., Brown, N. A., &
ber: 108-2627-H-006-003 for this work is gratefully acknowledged.
Sheridan, J. (2016a). Dynamic leg-motion and its effect on the aero-
dynamic performance of cyclists. Journal of Fluids and Structures,
65, 121–137. https://doi.org/10.1016/j.jfluidstructs.2016.05.007
Disclosure statement Crouch, T., Burton, D., Venning, J., Thompson, M., Brown, N., &
No potential conflict of interest was reported by the authors. Sheridan, J. (2016b). A comparison of the wake structures of scale
and full-scale pedalling cycling models. Procedia Engineering, 147,
13–19. https://doi.org/10.1016/j.proeng.2016.06.182
Debraux, P., Grappe, F., Manolova, A. V., & Bertucci, W. (2011).
ORCID
Aerodynamic drag in cycling: Methods of assessment. Sports
Jiun-Jih Miau http://orcid.org/0000-0003-3655-9917 Biomechanics, 10(3), 197–218. https://doi.org/10.1080/14763141.
2011.592209
Defraeye, T., Blocken, B., Koninckx, E., Hespel, P., & Carmeliet, J.
References (2010). Aerodynamic study of different cyclist positions: CFD ana-
lysis and full-scale wind-tunnel tests. Journal of Biomechanics, 43(7),
Achenbach, E. (1968). Distribution of local pressure and skin friction 1262–1268. https://doi.org/10.1016/j.jbiomech.2010.01.025
around a circular cylinder in cross-flow up to Re ¼ 5  106. Journal Defraeye, T., Blocken, B., Koninckx, E., Hespel, P., & Carmeliet, J.
of Fluid Mechanics, 34(4), 625–639. https://doi.org/10.1017/ (2011). Computational fluid dynamics analysis of drag and convect-
S0022112068002120 ive heat transfer of individual body segments for different cyclist
Achenbach, E. (1971). Influence of surface roughness on the cross-flow positions. Journal of Biomechanics, 44(9), 1695–1701. https://doi.org/
around a circular cylinder. Journal of Fluid Mechanics, 46(2), 10.1016/j.jbiomech.2011.03.035
321–335. https://doi.org/10.1017/S0022112071000569 Defraeye, T., Blocken, B., Koninckx, E., Hespel, P., Verboven, P., Nicolai,
Basu, R. I. (1985). Aerodynamic forces on structures of circular cross- B., & Carmeliet, J. (2014). Cyclist drag in team pursuit: Influence of
section. Part 1. Model-scale data obtained under two-dimensional cyclist sequence, stature, and arm spacing. Journal of Biomechanical
conditions in low-turbulence streams. Journal of Wind Engineering Engineering, 136(1), 011005. https://doi.org/10.1115/1.4025792
and Industrial Aerodynamics, 21(3), 273–294. https://doi.org/10. Farell, C., & Blessman, J. (1983). On critical flow around smooth circu-
1016/0167-6105(85)90040-6 lar cylinders. Journal of Fluid Mechanics, 136(1), 375–391. https://
Bearman, P. W. (1969). On vortex shedding from a circular cylinder in doi.org/10.1017/S0022112083002190
the critical regime. Journal of Fluid Mechanics, 37(3), 577–585. Fintelman, D. M., Sterling, M., Hemida, H., & Li, F. X. (2014).
https://doi.org/10.1017/S0022112069000735 Optimal cycling time trial position models: Aerodynamics versus
Bendat, J. S., & Piersol, A. G. (1991). Random data analysis and meas- power output and metabolic energy. Journal of Biomechanics, 47(8),
1894–1898. https://doi.org/10.1016/j.jbiomech.2014.02.029
urement procedures (2nd ed.). John Wiley & Sons.
Gaster, M. (1969). The structure and behavior of separation bubbles. No.
Blocken, B., Defraeye, T., Koninckx, E., Carmeliet, J., & Hespel, P.
3595.
(2013). CFD simulations of the aerodynamic drag of two drafting
Gerrard, J. H. (1966). The mechanics of the formation region of vorti-
cyclists. Computers & Fluids, 71, 435–445. https://doi.org/10.1016/j.
ces behind bluff bodies. Journal of Fluid Mechanics, 25(2), 401–413.
compfluid.2012.11.012
https://doi.org/10.1017/S0022112066001721
Blocken, B., van Druenen, T., Toparlar, Y., & Andrianne, T. (2018a).
Hsu, X. Y., Miau, J. J., Tsai, J. H., Tsai, Z. X., Lai, Y. H., Ciou, Y. S.,
Aerodynamic analysis of different cyclist hill descent positions. Shen, P. T., Chuang, P. C., & Wu, C. M. (2019). The aerodynamic
Journal of Wind Engineering and Industrial Aerodynamics, 181, roughness of textile materials. The Journal of the Textile Institute,
27–45. https://doi.org/10.1016/j.jweia.2018.08.010 110(5), 771–779. https://doi.org/10.1080/00405000.2018.1518636
Blocken, B., van Druenen, T., Toparlar, Y., Malizia, F., Mannion, P., Igarashi, T. (1978). Flow characteristics around a circular cylinder with
Andrianne, T., Marchal, T., Maas, G.-J., & Diepens, J. (2018b). a slit. Bulletin of JSME, 21(154), 656–664. https://doi.org/10.1299/
Aerodynamic drag in cycling pelotons: New insights by CFD simu- jsme1958.21.656
lation and wind tunnel testing. Journal of Wind Engineering and Kao, Y. M. (2005). Calibration of the ABRI environment wind tunnel
Industrial Aerodynamics, 179, 319–337. https://doi.org/10.1016/j. and experimental study of 2-D bluff body aerodynamic flows [Master
jweia.2018.06.011 thesis]. Department of Aeronautics and Astronautics, National
Bloor, M. S. (1964). The transition to turbulence in the wake of a cir- Cheng Kung University.
cular cylinder. Journal of Fluid Mechanics, 19(02), 290–304. https:// Kyle, C. R., & Burke, E. (1984). Improving the racing bicycle.
doi.org/10.1017/S0022112064000726 Mechanical Engineering, 106(9), 34–45.
1230 X. Y. HSU ET AL.

Lighthill, M. L. (1963). Introduction, boundary layer theory. In L. Oggiano, L., Troynikov, O., Konopov, I., Subic, A., & Alam, F. (2009).
Rosenhead (Ed.), Laminar boundary layers (Chap. II). Oxford Aerodynamic behavior of single sport jersey fabrics with different
University Press. roughness and cover factors. Sports Engineering, 12(1), 1–16. https://
Lin, Y. J., Miau, J. J., Tu, J. K., & Tsai, H. W. (2011). Non-stationary, doi.org/10.1007/s12283-009-0029-0
three-dimensional aspects of flow around a circular cylinder at crit- Phung, M. V., Miau, J. J., Lin, S. Y., & Li, S. R. (2017). A research of
ical Reynolds numbers. AIAA Journal, 49(9), 1857–1870. https://doi. wake development in Cycling Hood Position. 12th International
org/10.2514/1.J050674 Symposium on Advanced Science and Technology in Experimental
Miau, J. J., Chen, Z. L., & Hu, C. C. (2013). The characteristics of the Mechanics, 1–4 November 2017, Kanazawa, Japan.
ABRI wind tunnel. In S. B. Chaplin (Ed.), Wind tunnels: Design/construc- Roshko, A. (1954). On the development of turbulent wakes from vortex
tion, types and usage limitations (pp. 1–33). Nova Science Publishers. streets. NACA Report, 1191.
Miau, J. J., Li, S. R., Tsai, Z. X., Phung, M. V., & Lin, S. Y. (2020). On Roshko, A. (1993). Perspectives on bluff body aerodynamics. Journal of
the aerodynamic flow around a cyclist model at the hoods position. Wind Engineering and Industrial Aerodynamics, 49(1–3), 79–100.
Journal of Visualization, 22, 35–47. https://doi.org/10.1016/0167-6105(93)90007-B
Miau, J. J., Tsai, H. W., Lin, Y. J., Tu, J. K., Fang, C. H., & Chen, M. Schewe, G. (1983). On the force fluctuations acting on a circular cylin-
(2011). Experiment on smooth circular cylinders in cross-flow in der in cross flow from subcritical up to transcritical Reynolds num-
the critical Reynolds number regime. Experiments in Fluids, 51(4), bers. Journal of Fluid Mechanics, 133, 265–283. https://doi.org/10.
949–967. https://doi.org/10.1007/s00348-011-1122-2 1017/S0022112083001913
Miau, J. J., Tu, J. K., Chou, J. H., & Lee, G. B. (2006). Sensing flow Tani, I. (1964). Low-speed flows involving bubble separations. Progress
separation on a circular cylinder by MEMS thermal-film sensors. in Aerospace Sciences, 5, 70–103. https://doi.org/10.1016/0376-
AIAA Journal, 44(10), 2224–2230. https://doi.org/10.2514/1.17408 0421(64)90004-1
Nakamura, Y., & Tomonari, Y. (1982). The effects of surface roughness Tseng, C. H. (2019). Research of the flow structure and experiment data
on the flow past circular cylinders at high Reynolds numbers. on full and small scale cyclist model [Master thesis]. Department of
Journal of Fluid Mechanics, 123, 363–378. https://doi.org/10.1017/ Aeronautics and Astronautics, National Cheng Kung University.
S0022112082003103 Van Dyke, M. (1982). An album of fluid motion. The Parabolic Press.
Nikuradse, J. (1937). Laws of flow in rough pipes. NACA Technical White, F. M. (1974). Viscous fluid flow. McGraw-Hill, Inc.
Memorandum 1292. Wieselsberger, C. (1922). New data on the law of fluid resistance.
Oggiano, L., Saetran, L., Løset, S., & Winther, R. (2007). Reducing the NACA Technical Note, 84.
athlete’s aerodynamic resistance. Journal of Computational and Zdravkovich, M. M. (1997). Flow around circular cylinders:
Applied Mechanics, 8(2), 163–173. Fundamentals (Vol. 1, pp. 1–18). Oxford University Press.

You might also like