You are on page 1of 8

Modeling and Analysis of the Technical Performance of DC-Motor

Electric Bicycle Drives Based on Bicycle Road Test Data


Annette Muetze, Ying Chin Tan
Dept. of Electrical and Computer Engineering
University of WisconsinMadison, Madison, WI 53706 USA
muetze@engr.wisc.edu, yctan@wisc.edu
AbstractElectric motor powered bicycles have been making
their way into the U.S. market for about two decades. Custom-
designed electric bicycles allow common issues such as high cost
and weight to be overcome. To this aim, customized modeling
tools are required. The paper discusses the modeling of a direct-
drive dc-motor electric bicycle drive system for technical per-
formance evaluation, where the operating cycle proles used are
based on actual road tests. It is explained how the measurement
data can be processed and coupled with the usual model
of a direct-drive dc-drive system, thereby extending common
modeling approaches. Then, the different riding proles are
analyzed with the developed tool. The results both illustrate the
ability of such low-cost drives to serve for commuting purposes
with moderate driving styles and their limits to support rather
sporty rides.
Index TermsPower-assisted bicycle, direct-drive, efciency,
modeling, performance evaluation.
I. INTRODUCTION
Electric motor powered bicycles have been making their
way into the U.S. market for about two decades. Such electric
bicycles can be used for a large variety of purposes, including
serving as a vehicle for police or law enforcers, a guide bike
during races, and for leisurely rides and commuting (e.g. [1]
[8]). When designing electric bicycles (including their electric
drive systems) and trying to overcome common issues such
as high bicycle cost and weight, it is important that the drive
be most efcient over a given operating cycle; this leads to
custom-designed electric bicycles such as city bicycles, hill
bicycles, distance bicycles, and speedy bicycles. In this
paper, a model of an electric bicycle drive is presented that
can be used to evaluate the technical performance of a given
drive system both instantaneously and over a whole operating
cycle. A very unique characteristic of the model is that the
operating cycle proles used as input are based on actual road
tests. Special emphasis is based on the coupling techniques of
these experimental data to the different modules of the drive
model. For example, both the command and the load torque are
required in the model, but only one measured value per time
step is available. Furthermore, the speed has been measured,
which, in the conventional model, is a function of the net
torque available at the shaft. As far as the authors know, this
is the rst time that such an approach on analysis and design
of electric bicycle drives is reported on in the literature. The
model allows investigation of:
1) Instantaneous drive parameters (e.g. currents, voltages,
torques, battery loading, efciencies).
2) Overall technical performance of the drive over a given
driving cycle (e.g. efciency, energy consumption).
3) Inuence of the parameters of the different drive com-
ponents on the technical performance and other output
parameters.
First, in the style of a concise review, the different elements
of the conventional model are presented (Section II). Then,
the drive simulation technique is discussed. This includes the
test vehicle and data recording, as well as the data processing
and model extension to analyze the technical performance
of a drive using such measured riding proles (Section III).
Here, special emphasis is put on the different ways of using
the measurement data as input into the model. In the next
section, following a short overview of the four measured
riding proles, the performances of different dc-motor drives
is analyzed using the presented model (Section IV). Both
the advantages and limits of the drives as well as of the
modeling approach are shown. The ndings are summarized
and prospects of future work are given at the end (Section V).
II. MODEL STRUCTURE AND IMPLEMENTATION
A. Objectives
The model is designed for the investigation of both the
instantaneous and the overall performance of direct-drive dc-
motor electric bicycle drive systems under different riding
conditions. To this aim, it allows:
1) Investigation of various instantaneous drive parameters,
such as motor current, voltage, torque, remaining battery
energy, and system efciency.
2) Overall performance evaluation of the drive over a given
driving cycle, such as system efciency and total power
consumption.
3) Investigation of the inuence of controller, battery, and
motor parameters on the different drive parameters. For
example, battery internal resistance, motor inductance
and resistance. Thereby, modications to better meet
the demands of custom-designed electric bicycles can
be identied and veried.
Here, only direct-drive systems with brushed dc-motors of
fully-powered electric bicycles are considered. However, the
model can be extended at a later stage to include brushless
dc motors and/or the additional control of the output power
due to the required human-to-motor power ratio of pedelec-
type electric bicycles. With pedelec-type electric bicycles,
only a certain pre-determined, speed-dependent fraction of the
bicycle propulsion power is delivered by the drive.
1574 1-4244-0743-5/07/$20.00 2007 IEEE
B. Structure
The overall model is organized in submodels that are
implemented as individual submodules. It consists of the four
today well-known main components of such drive systems
(Sections II-C through II-E):
1) Electric motor.
2) Controller, itself consisting of three submodules:
a) Transformation of the user demand into the corre-
sponding duty cycle.
b) Control of the switching based on the duty cycle.
c) A capacitor to provide a stiff voltage at the switch
throw terminal.
3) Battery.
4) Electromechanical system that relates net torque and
acceleration, using a simplied single-wheel represen-
tation.
and of a series of technical performance modules (Section II-
F). A sketch of the electric components, 1) to 3), and their
main interactions is shown in Fig. 1, the equivalent circuit in
Fig. 2.
Battery Motor Controller
Battery
voltage
Battery
current
Demand
voltage
Feedback
current
D
e
m
a
n
d
Throttle
Fig. 1. Overview of the three electric modules and their main interactions.
emf
= k
v
w
R
b
V
batt
C
V
batt
V
c I
c
R
a
L
a
B J
T
net
= k
t
I
a
- T
load
w
I
a
I
batt
I
batt
-I
c
Battery
DC Motor
Controller
Fig. 2. Equivalent circuit of the system.
The model is implemented using the commercially available
software package MATLAB
R
Simulink
R
. All parameters are
dened globally.
C. Motor model including mechanical equation
The motor including the mechanical equation are modeled
in the today well established way consisting of:
1) The electrical equivalent circuit ((1) - (3))
V
a
= R
a
I
a
+L
a
dI
a
dt
+k
v
(1)
EMF = k
v
(2)
T
e
= k
t
I
a
(3)
with the armature voltage, current, resistance, and induc-
tance V
a
, I
a
, R
a
, and L
a
, the back-EMF voltage given by
the product of back-EMF constant k
v
and rotor speed ,
the electrical torque T
e
and torque constant k
e
.
and
2) The mechanical equivalent circuit (4):
T
net
= T
e
T
load
= J
d
dt
+B (4)
where T
net
, J, and B are the net torque, inertia, and
damping coefcient respectively.
The corresponding block is shown in Fig. 3.
Electrical
1
R
a
+sL
a
1
B+sJ
k
t
Electro-
mechanical
Torque
constant
I
a
T
e
Speed
T
load
k
v
Back-emf constant
V
cmd
V-limiter
V
a
Fig. 3. Motor model block including mechanical equation (constant motor
ux).
D. Controller model
The controller controls the power ow from the battery
source to the motor. The controller model module consists
of three sub-modules (Fig. 4):
1) Transformation of the user demand into the switch
control (Fig. 4(a)):
The controller controls the switching of the switch that
connects the motor terminals either to the battery voltage
V
batt
or to zero voltage. The torque command given
by the rider T
cmd
is transformed into the corresponding
current and compared to the armature current. The error
is controlled to become zero with a Proportional Integral
(PI) PWM controller. The output D of the controller
takes on the values 0 or 1.
2) Control of the switch (Fig. 4(b)):
The instantaneous throw voltage V
cmd
and current I
cmd
that result from the torque command T
cmd
and the actual
point of operation of the motor are obtained from the
output of the controller D, the armature voltage V
a
, and
the armature current I
a
. The command voltage V
cmd
is
used to control the motor (Fig. 3), whereas the current
I
cmd
is required to calculate the voltage across the
voltage stiffening capacitor.
3) A capacitor to provide a stiff voltage at the switch throw
terminal (Fig. 4(c)):
The capacitor voltage V
c
is obtained by integration of
the capacitor current which is the difference between
the battery and throw currents I
batt
and I
cmd
.
E. Battery model
The battery model block comprises the battery voltage V
batt
and the battery internal resistance R
b
. Temperature and load
dependence are not considered. The battery voltage interacts
with the capacitor voltage, and the battery current is obtained
from the voltage drop across the battery internal resistance
(Fig. 5).
1575
T
cmd
1
k
t
R
a
I
a
R
a
PI-Controller
D Comparator
0/1
(a) Processing of user demand.
D
V
c
0
V
cmd
D
I
a
0
I
cmd
(b) Switch.
I
batt
I
cmd
1
C
V
c
V
batt
1
s
x
0
(c) Voltage stiffening capacitor.
Fig. 4. Controller model block including processing of the user demand.
V
batt
V
c
1
R
b
I
batt
Fig. 5. Battery model block.
F. Technical performance modules
The following parameters are investigated with the technical
performance modules:
1) Instantaneous electrical input power P
in
.
2) Instantaneous mechanical output power P
mech
.
3) Input energy (battery energy output) W
in
(accumulated
since the beginning of the driving cycle).
4) Battery capacity W
batt0
.
5) Battery remaining energy W
batt
(t).
6) Instantaneous drive efciency .
7) Average drive efciency over driving cycle
ave
.
The equations implementing these seven aspects are simple
and well-established. Therefore, not all technical performance
modules are shown in Fig. 6.
V
batt
P
in
I
batt
T
e
P
mech
w
P
in
W
in
1
3600
1
s
P
mech
P
in
P
mech
P
in
~
P
in
h 100
if execution
1) 2)
3)
6)
Fig. 6. Selected technical performance modules.
For the realization of the technical performance blocks,
several if-execution blocks are used. This is required to prevent
the occurrence of undened numbers caused by divisions by
zero. For example, at the beginning of a simulation (t = 0),
unless initialized otherwise, both the electrical input power
P
in
and the mechanical output power P
mech
are zero, and
calculation of the efciency would result in an undened
number. Using the if-execution block, this is avoided as
follows (nominator N and denominator D):
The input D is processed as follows to become

D:
1) N = 0, D = 0 :

D is set to be a non-zero constant so
that N/

D = 0.
2) N = 0, D = 0 :

D takes the value

D(t) = D(t t)
so that N/

D = inf.
3) Other than above:

D = D remain unchanged.
For the example of the instantaneous drive efciency (case
6) it is N = P
mech
, D = P
in
, and

D =

P
in
.
III. DRIVE SCENARIO SIMULATION TECHNIQUE
A. Test vehicle description
For the experimental investigation, an electric bicycle with
a brushed dc-motor installed in the front hub, a controller, a
thumb throttle, and a battery pack is used (Fig. 7). This bicycle
is a commercially available bicycle that has been available in
the laboratory. All experiments were carried out using this test
vehicle. The electric hub-motor in the front wheel is not used
during the measurements, yet, using this bicycle, the actual set
up of an electric bicycle is represented.
Fig. 7. Electric bicycle test set-up used for the experimental investigation.
The torque and speed are directly measured in the hub of
the rear wheel of the test bicycle, using a Power Tap
R
hub
(Fig. 8) [9]. The measurement information is transmitted to
the CPU through the receiver. For all measurements, the tire
pressure was kept at 50,...,60 psi, which is typical for bicycles
that are used for leisure and commuting and that are commonly
not re-inated before each ride. The anemometer that can be
seen in Fig. 7 is not used for the measurements discussed in
this paper.
B. Data recording
The riding proles were recorded (measured) in terms of
power, P
PT
, torque, T
PT
, and ground speed,
PT
. The sampling
interval is set at its minimum time, t
s
= 1.26s. Table I shows
a sample data set.
1576
Fig. 8. Power Tap
R
hub [9] used for the experimental investigation.
TABLE I
SAMPLE DATA SET FROM POWER TAP
R
Time Torque Speed Power Distance
Minutes Nm km/h W km
0 0 0 0 0
0.021 6.1 4.1 21 0.006
0.042 4.5 6.5 24 0.008
0.063 3.7 6.5 20 0.01
C. Data processing and model extension
The measurements of the riding interval proles as de-
scribed above (Section III-B) are inputs to the model using the
Simulink
R
signal builder block. In addition to the torque T
PT
and the speed
PT
, the speed difference
PT
and acceleration
(d/dt)
PT
=
PT
/t
s
between two sampling intervals are
used for this purpose.
1) Command torque and load torque: The model contains
the load and command torques T
load
and T
cmd
as inputs (Figs. 3
and 4). However, only the torque produced by the rider is
available in the form of T
PT
, along with the speed
PT
. As
this torque shall be produced by the electric motor, it is used
as input for the torque command, T
cmd
= T
PT
= T
cmd,PT
.
The load torque T
load
needs to be derived from the measure-
ment data at each sampling interval. Two different approaches
are taken:
1) Delay of T
load
when compared with T
command
:
The load torque at time t, T
load
(t), is approximated
by the command torque of the previous time interval,
T
load
(t +t
s
) = T
cmd
(t), or, T
load
(t) = T
cmd
(t t
s
).
2) Approximation of T
load
using the mechanical equation:
The load torque T
load
is derived using the mechanical
equation (4) that becomes
T
load,PT
= T
cmd,PT
B
PT
+J

PT
t
s
. (5)
Both approaches have been implemented. For this purpose,
the motor and controller models (Figs. 3 and 4) have been
modied to use the measured data T
cmd,PT
and, in the case
of approach no. 2,
PT
and (d/dt)
PT
as inputs. From these,
the new parameters T
load,PT
, D
PT
, V
cmd,PT
, and I
cmd,PT
are
derived. Here, the

signies that a values is obtained


through processing of the measured data T
PT
and
PT
. Fig. 9
shows selected elements of the modied controller model
block for method 2.
1
k
t
R
a
I
a
R
a
T
cmd,PT
D
PT*
V
cmd,PT*
PT
d
dt
w
J
I
cmd,PT*
T
load,PT*
Fig. 9. Modied controller model block (selected) with the command and
load torques derived from the experimentally obtained riding proles; the

signies that a value is derived from recorded (measured) data.


Approach 1 completely ignores the outside drive parameters
that contribute to the riding prole, whereas approach 2
includes some of these through the simplied single-wheel
model. Even though the slopes of the route are not considered
explicitly, some information is included in the acceleration
that is taken from the recorded (measured) data (d/dt)
PT
.
Therefore, we are using approach 2 in the following.
2) Speed: The recorded (measured) speed
PT
can be used
as input into the model to compute:
1) The back-EMF.
2) The rolling friction feedback in the mechanical equation.
Both approaches are implemented so that the parameters
that are derived from the motor speed are computed from the
proper speed, even at operating points when the motor cannot
produce the commanded torque due to limitations imposed by
the stator current or back-EMF. The modied model of the
dc-motor and the electromechanical system with
PT
used as
input for both computations is shown in Fig. 10. With this
approach, the command signals that the drive receives are
completely decoupled from the ability of the drive to produce
the command torque at a given speed as well as from errors
introduced by the single-wheel model behind the mechanical
equation.
k
v
V
cmd,PT*
w
PT
T
e0
1
s
x
0
1
L
a
1
R
a
1
k
t
k
t
I
a
T
e
T
net
w
0
1
s
x
0
1
J
B
T
load,PT*
w
sim
Fig. 10. Modied model of dc-motor and electromechanical system (constant
motor ux) with the command voltage, load torque, and the motor speed for
the computation of the back-EMF and the rolling friction feedback in the
mechanical equation derived from the experimentally obtained riding proles;
the

signies that a value is derived from measured data. Compare with


Fig. 3.
3) Efciency calculations: With the measured speed
PT
available and being used as input to one or both feedback
loops, it can be of interest to only concentrate on behavior and
efciency of the motor before the simplied single-wheel
model of the mechanic system (mechanical equation). This
can be notably of interest with congurations and operating
1577
points where the drive cannot deliver the required torque at
a given speed. Therefore, a second module to calculate the
mechanical power

P
mech
is implemented that uses the electrical
motor torque T
e
and the recorded (measured) speed
PT
as
input (Fig. 11).
T
e
w
sim
P
mech
P
mech
T
e
w
PT
~
Fig. 11. Modication of P
mech
module to be used when simulated speed,

sim
, and recorded (measured) speed,
PT
, are not in agreement.
IV. RESULTS
A. Riding interval proles
The test-bicycle described above (Section III-A) was used
to obtain riding interval proles as follows: The measure-
ments were designed to obtain road data of real-life appli-
cations. To this aim, riding interval proles, proles no. 1
through 4, of four different riders and with intervals of 15 to 25
minutes were recorded, where the bicycle was used for a short
leisurely ride, grocery shopping, or commuting (Table II).
TABLE II
CHARACTERISTICS OF RECORDED RIDING PROFILES
Prole Rider Riding P
mech,max
P
mech,ave
T
max
T
ave
no. weight interval
kg min W W Nm Nm
1 50 18 204.0 35.6 27.9 4.7
2 75 16 389.1 133.9 40.8 8.2
3 85 22 368.6 66.3 26.4 5.9
4 95 25 857.0 179.0 50.2 9.9
Exemplarily, the measured torque versus time and power
versus time characteristics of riding proles 1 and 4 are
shown in Figs. 1215. They illustrate the spread of the prole
characteristics, and thus requirements on the drive, as riding
proles 1 and 4 are those with the lowest and highest torque
and power demands respectively (It should be noted that the
maximum speed of the riding prole 4 exceeds the speed limit
for low-speed electric bicycles according to U.S. law, which
is 20mph.). The characteristics of riding proles 2 and 3 are
in-between these two extremes.
B. Simulation results
1) Parameters: The four rides were analyzed using the
developed model and example-case values of a dc-drive sys-
tem. The base example-case values were adapted from the
commercially available electric bicycle that was also used as
test vehicle for the riding prole measurements (Section III-A,
Fig. 7). The bicycle has a 24V brushed dc-hub motor and a
battery system of two 12V, 12Ahr lead acid batteries.
In this work, we seek to identify the limits of the drives
and the simulation techniques and not to optimize one single
drive conguration for one given riding prole. Therefore, we
select by intention a comparatively large value of the arma-
ture resistance to account for low-cost motors and additional
resistance of the connections. The parameters of the motor
0
10
20
30
40
50
0 5 10 15 20 25
Time [min]
Torque [Nm]
Fig. 12. Measured torque versus time of riding prole no. 1 (same scales
as Fig. 14 by intention), T
max
= 27.9Nm, T
ave
= 4.7Nm.
0
100
200
300
400
500
600
700
800
900
0 5 10 15 20 25
Time [min]
Power [W]
Fig. 13. Measured power versus time of riding prole no. 1 (same scales as
Fig. 15 by intention), P
max
= 204.0W, P
ave
= 35.6W.
0
10
20
30
40
50
0 5 10 15 20 25
Time [min]
Torque [Nm]
Fig. 14. Measured torque versus time of riding prole no. 4, T
max
=
50.2Nm, T
ave
= 9.9Nm.
are set to R
a
= 1, L
a
= 1mH, k
v
= k
t
= 1Nm/A, and
R
batt
= 2 12m. The parameters of the PI-controller were
designed to achieve 0.06s rise time, 5% overshoot,0.2s settling
time, and 30rad/sec bandwidth (K
p
= 100, K
i
= 0.3). We set
J = 10kg/m
2
according to [10].
1578
0
100
200
300
400
500
600
700
800
900
0 5 10 15 20 25
Time [min]
Power [W]
Fig. 15. Measured power versus time of riding prole no. 4, P
max
= 857.0W,
P
ave
= 179.0W.
Using these base example-case values of the different
parameters as starting point, we vary selected parameters
throughout the analysis. When doing so, the abilities of the
modeling technique are analyzed using the performance of
the low-cost example-case drive for illustration. For reasons
of comparison, only the rst 840s (14min) of the rides are
simulated. The results illustrate well how the performances of
a given drive system depends on the characteristic of a given
riding prole.
2) Speed: A comparison of the simulated and the recorded
(measured) speed shows the limitations given by the simplied
single-wheel model of the mechanical system: Exemplarily,
Fig. 16 shows an extract of riding prole no. 1. The simplied
model does not reproduce the speed (dynamic behavior) of
the drive correctly, notably for fast speed slopes, but the drive
itself is able to produce the required torque as can be seen
from Fig. 17.
It is important to recall thatas a result of the way the
recorded data are used as inputs to the modelthe command
signals the motor receives are completely decoupled from
inaccuracies of the speed simulation (Section III-C.2). Corre-
sponding simulations were also carried out for the other riding
proles, with different values of the lumped inertia as small
as J = 0.1kgm/s
2
, and of the controller, leading to similar
results. In order to correct these inaccuracies, more complex
models of the mechanical system, such as those suggested in
[10] or [11], would need to be implemented and adapted to
work with the measurement data as inputs.
As a result of the limits of the speed simulation, the output
efciency needs to be calculated via the modied performance
submodule, using the recorded (measured) speed
PT
and the
electric torque T
e
(Fig. 11).
3) Torque: Operating points can occur where the drive
cannot produce the command torque due limitations imposed
by the current or the back-EMF. Whenever the command
torque exceeds the torque limit of the drive at its current
speed, deep speed dips can be seen in the simulated speed

sim
when compared with the recorded (measured) one,
PT
.
For illustration, the simulation of riding prole no. 1 is
50 100 150 200
0
5
10
15
Time [s]
S
p
e
e
d

[
r
a
d
/
s
]


w
PT
w
sim
Fig. 16. Recorded and simulated speed w
PT
and w
sim
of riding prole no. 1
(40s-230s): The simplied model does not reproduce the speed (dynamic
behavior) of the drive correctly, notably for fast speed slopes, but the drive
itself is able to produce the required torque as can be seen from Fig. 17.
50 100 150 200
10
5
0
5
10
15
20
25
30
Time [s]
T
o
r
q
u
e

[
N
m
]


T
cmd,PT
T
e
Fig. 17. Command and electrical torques T
cmd,PT
and T
e
of riding pro-
le no. 1 (40s-230s): The drive is able to produce the required torque.
shown in Fig. 18, where several of such dips can be seen (at
approximately 30s, 250s, 400s, 440s, 480s, 615s, and 720s).
The inability of the drive to produce the command torque can
further be seen in Fig. 19, where the electric torque does not
reach the command torque whenever a speed dip can be seen
in Fig. 18.
These limits of the drive are even more evident with the
simulations of the riding prole no. 4. (Figs. 20 and 21).
If not the recorded (measured) speed
PT
but the simulated
one
sim
were used as inputs into the feedback loops, and in a
real case, such decreases in speed would cause a decrease of
back-EMF and thus increase of current and torque. However,
as the motor does not receive the simulated, but the higher
recorded (measured) speed as inputs in the simulations, the
simulated torque remains low, further decreasing the simulated
speed. Therefore, drawing incorrect conclusions from these
rapid drops in the simulated speed drops must be avoided.
1579
0 200 400 600 800
0
5
10
15
Time [s]
S
p
e
e
d

[
r
a
d
/
s
]


w
PT
w
sim
Fig. 18. Recorded and simulated speed w
PT
and w
sim
of riding prole no. 1
(rst 840s); at approximately 30s, 250s, 400s, 440s, 480s, 615s, and 720s, the
command torque exceeds the torque limit of the drive at the current speed.
0 200 400 600 800
10
5
0
5
10
15
20
25
30
Time [s]
T
o
r
q
u
e

[
N
m
]


T
cmd,PT
T
e
Fig. 19. Command and electrical torques T
cmd,PT
and T
e
of riding pro-
le no. 1 (rst 840s); at approximately 30s, 250s, 400s, 440s, 480s, 615s, and
720s, the command torque exceeds the torque limit of the drive at the current
speed.
Of course, such speed simulation results with speed dips
where the motor cannot produce the command torque can even
less be used for energy consumption calculations than those
discussed above (Section IV-B.2). Therefore, the modied
performance submodule (Fig. 11) needs to be used.
The orders of magnitudes are illustrated with the following
simple approximation: With torque and back-EMF constant of
k
t
= k
v
= 1Nm/A, the back-EMF is 10V at 10rad/s speed.
Neglecting a possible current limit of the machine, 14V are left
for the voltage drop at the resistance(s), giving approximately
14Nm torque.
4) Efciencies: First, we consider the following two com-
puted values for each of the four riding proles: (i) The battery
energy output, which is the energy input to the drive, W
in
,
and is calculated from the battery voltage and current V
batt
and I
batt
(Fig. 6.1). (ii) The drive output energy W
out,
which
is calculated from the recorded (measured) speed
PT
and the
electric torque T
e
(Fig. 11) (Table III).
0 50 100 150 200
30
20
10
0
10
20
Time [s]
S
p
e
e
d

[
r
a
d
/
s
]


w
PT
w
sim
Fig. 20. Recorded and simulated speed w
PT
and w
sim
of riding prole no. 4
(rst 200s): The drive is not able to produce the required torque. The
discrepancy is that large that the simulated speed becomes negative.
0 200 400 600 800
10
0
10
20
30
40
50
Time [s]
T
o
r
q
u
e

[
N
m
]


T
cmd,PT
T
e
Fig. 21. Command and electrical torques T
cmd,PT
and T
e
of riding pro-
le no. 4 (rst 840s): The drive is not able to produce the required torque.
TABLE III
RESULTS OF DRIVE SCENARIO SIMULATIONS
Prole W
in
W
out,
W
out,
/W
in
W
prole
W
out,
/W
prole
no. [Wh] [Wh] [Wh]
1 29.36 9.36 0.31 12.03 0.78
2 34.48 15.45 0.45 16.36 0.95
3 37.03 22.37 0.60 24.93 0.90
4 30.00 17.29 0.58 54.76 0.32
From these two values, the ratio W
out,
/W
in
can be calcu-
lated, which has the form of an equivalent efciency. This
quantity does not consider if the drive is able to produce the
required torque, but it only considers the torque the motor can
deliver. As both the input and the output energy are derived
from the recorded speed, these computations are not affected
by the inaccuracies of the simulated speed. Here, relative large
amounts of time during which the drive is operated at high
speed translates into higher values of this ratio for the different
riding proles.
1580
Next, the energy requirement of the riding proles, W
prole
,
each for the rst 840s, are considered (Table III), from which
a second ratio, W
out,
/W
prole
, can be derived. From this
quantity conclusions on how much of the riding prole can
be produced by the drive can be drawn.
In the model, the armature and battery currents I
a
and I
batt
can be both positive and negative. As a result, the simulations
include the possibility of generating mode. This becomes
notably obvious with riding prole no. 4. At operating points
where the back-EMF that is computed from the recorded (mea-
sured) speed
PT
exceeds 24rad/s, the back-EMF is larger than
the maximum command voltage, the armature current becomes
negative (Fig. 10), eventually translating into a negative battery
current I
batt
(Figs. 4 and 5) that recharges the battery.
By comparing W
out
and W
prole
, conclusions on how much
of the riding prole can be produced by the drive can be drawn.
These results do not change with a change of J, as the
different energy computations and the command signals are
decoupled from the simulated speed. Furthermore, the simu-
lations with different values of the controller, K
p
= 2 and
K
i
= 0.5, were carried out, leading to the same results.
Regarding the energy requirements of the different riding
proles in a more general way, most proles could be sup-
plied from one or 1.5 laptop-size batteries, when assuming a
conservative estimate of 30% overall efciency.
V. CONCLUSIONS
Different technical performance criteria of electric bicycle
drive systems for given operating cycles can be evaluated
via a model implementation that uses riding proles based
on actual road tests. Such analysis can contribute to de-
signing better, custom-designed electric bicycles and thereby
overcoming common issues such as high cost and weight.
The implementation shows how the characteristic of a given
drive depends on the characteristic of the riding prole. The
presented tool can be used to develop more general answers to
operating areas of drives with respect to the different torque-
speed combinations and their derivations with respect to time
as they can occur with electric bicycles.
APPENDIX
Appendix A: List of abbreviations
Acronym Denition
DC Direct Current
EMF Electro-Magnetic Force
PI Proportional-Integral
PWM Pulse-Width Modulation
Appendix B: List of symbols
Name Description
B rolling resistance
I
a
armature current
I
batt
battery current
I
cmd
current resulting from command duty
I
cmd,PT
current resulting from command duty,
derived from measured values
J inertia
List of symbols continued
Name Description
k
v
back-EMF constant
k
t
torque constant
L
a
armature inductance
P
in
input power
P
mech
mechanical power
t time
R
a
armature resistance
R
b
battery resistance
T
cmd
command torque
T
cmd,PT
command torque, derived from recorded values
T
PT
recorded (measured) torque
T
e
electric torque
T
load
load torque
T
load,
load torque, derived from recorded values
T
net
net torque
T
PT
measured torque
V
batt
battery voltage
V
c
dc-link voltage
V
cmd
command voltage
V
cmd,PT
command voltage, derived from recorded values
W
batt
battery remaining energy
W
batt0
battery capacity
W
in
energy input
W
mech
mechanical energy
speed

PT
recorded (measured) speed
REFERENCES
[1] B. Kumar and H. Oman, Power control for battery-electric bicycles,
Proceedings of NAECON 93 - National Aerospace and Electronics
Conference, vol. 1, pp. 428-434, Dayton, OH, May 24-28, 1993.
[2] W.C. Morchin, Battery-powered electric bicycles, Proceedings of
Northcon 94, pp. 269-274, Seattle, WA, October 11-13, 1994.
[3] E.A. Lomonova, A.J.A. Vandenput, J. Rubacek, B. dHerripon, and
G. Roovers, Development of an improved electrically assisted bicycle,
Proceedings of 37th IEEE IAS Annual Meeting, pp. 384-389, Pittsburgh,
PA, October 13-18, 2002.
[4] F.E. Jamerson, Electric bikes worldwide 2002: with electric scooters &
neighborhood EVs, Naples, Fla: Electric Battery Bicycle Co, 2002.
[5] A. Muetze, A.G. Jack, and B.C. Mecrow, Brushless-dc motor using soft
magnetic composites as a direct drive in an electric bicycle, Proceedings
of the 9th European Conference on Power Electronics and Applications
(EPE), paper no. 350, Graz, 2001.
[6] D.G. Wilson, J. Papadopoulos, and F.R. Whitt, Bicycling science,
Cambridge, Mass: MIT Press, 2004.
[7] P. Fairley, Chinas cyclists take charge: electric bicycles are selling by
the millions despite efforts to ban them, IEEE Sprectrum, vol. 42, no. 6,
pp. 54-69, June 2005.
[8] A. Muetze and Y.C. Tan, Performance evaluation of electric bicylces,
Proceedings of 40th IEEE IAS Annual Meeting, pp. 2865-2872, Hong
Kong, October 2-6, 2005.
[9] Power Tap
R
is by Graber Products, Inc., 5253 Verona Road, Madison,
WI USA, http://www.cycle-ops.com.
[10] S. Lee and W. Ham, Self stabilizing strategy in tracking control of
unmanned electric bicycle with mass balance, Proceedings of IEEE
Intellingent Robots and Systems (RSJ), vol. 3, pp. 2200-2205, September
30 - October 4, Lausanne, 2002.
[11] L. Li, F.Y. Wang, and Q. Zhou, Integrated longitudinal and lateral
tire/road friction modeling and monitoring for vehicle motion control,
IEEE Transactions on Intelligent Transportation Systems, vol. 7, no. 1,
pp. 1-19, March 2006.
1581

You might also like