ArticlePDF Available

Velocity Predictors for Predictive Energy Management in Hybrid Electric Vehicles

Authors:

Abstract and Figures

The performance and practicality of predictive energy management in hybrid electric vehicles (HEVs) are highly dependent on the forecast of future vehicular velocities, both in terms of accuracy and computational efficiency. In this brief, we provide a comprehensive comparative analysis of three velocity prediction strategies, applied within a model predictive control framework. The prediction process is performed over each receding horizon, and the predicted velocities are utilized for fuel economy optimization of a power-split HEV. We assume that no telemetry or on-board sensor information is available for the controller, and the actual future driving profile is completely unknown. Basic principles of exponentially varying, stochastic Markov chain, and neural network-based velocity prediction approaches are described. Their sensitivity to tuning parameters is analyzed, and the prediction precision, computational cost, and resultant vehicular fuel economy are compared.
Content may be subject to copyright.
IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 23, NO. 3, MAY 2015 1197
Velocity Predictors for Predictive Energy Management
in Hybrid Electric Vehicles
Chao Sun, Xiaosong Hu, Member, IEEE, Scott J. Moura, Member, IEEE, and Fengchun Sun
Abstract The performance and practicality of predictive
energy management in hybrid electric vehicles (HEVs) are highly
dependent on the forecast of future vehicular velocities, both in
terms of accuracy and computational efficiency. In this brief, we
provide a comprehensive comparative analysis of three velocity
prediction strategies, applied within a model predictive control
framework. The prediction process is performed over each
receding horizon, and the predicted velocities are utilized for
fuel economy optimization of a power-split HEV. We assume
that no telemetry or on-board sensor information is available for
the controller, and the actual future driving profile is completely
unknown. Basic principles of exponentially varying, stochastic
Markov chain, and neural network-based velocity prediction
approaches are described. Their sensitivity to tuning parameters
is analyzed, and the prediction precision, computational cost, and
resultant vehicular fuel economy are compared.
Index Terms—Artificial neural network (NN), comparison,
energy management, hybrid electric vehicle (HEV), model
predictive control (MPC), velocity prediction.
I. INTRODUCTION
SOPHISTICATED energy management strategies have
been developed to provide better fuel economy perfor-
mance in hybrid electric vehicles (HEVs) [1], [2]. This brief
intends to facilitate the performance of predictive energy
management through evaluating different horizon velocity
predicting approaches.
In the literature, dynamic programming (DP) and equiva-
lent consumption minimization strategy (ECMS) are crucial
in resolving the energy management problem for HEVs.
Globally, DP can ensure an optimal result when complete
knowledge of driving conditions is prescribed [3]. However,
the exact future power demand is usually unknown and the
computational burden is prohibitive. The DP solutions are
often realized offline and deployed as benchmarks [2]. The
ECMS is an instantaneous optimization for HEV energy
management [4], [5]. By defining an equivalent fuel cost
for battery energy, ECMS solves the optimal power split at
each time instant rather than over a time horizon. It has
Manuscript received May 9, 2014; accepted September 5, 2014. Date
of publication October 7, 2014; date of current version April 14, 2015.
Manuscript received in final form September 12, 2014. Recommended by
Associate Editor F. Vasca.
C. Sun was with the National Engineering Laboratory for Electric Vehicles,
Beijing Institute of Technology, Beijing 100081, China. He is now with the
Department of Mechanical Engineering, University of California at Berkeley,
Berkeley, CA 94720 USA (e-mail: chaosun.email@gmail.com).
X. Hu and S. J. Moura are with the Department of Civil and Environmental
Engineering, University of California at Berkeley, Berkeley, CA 94720 USA
(e-mail: xiaosong@chalmers.se; smoura@berkeley.edu).
F. Sun is with the National Engineering Laboratory for Electric
Vehicles, Beijing Institute of Technology, Beijing 100081, China (e-mail:
sunfch@bit.edu.cn).
Color versions of one or more of the figures in this paper are available
online at http://ieeexplore.ieee.org.
Digital Object Identifier 10.1109/TCST.2014.2359176
demonstrated that given an appropriate equivalence factor,
ECMS is comparable with DP [6]. Nevertheless, tuning the
equivalence factor is nontrivial. The ECMS variants, such as
telemetric-ECMS and adaptive-ECMS, are proposed to adjust
the equivalence factor based on information from telemetry
equipment or on-board sensors [7], [8]. Due to the uncertainty
of future driving profiles, endowing the controller with an
appropriate prediction ability can achieve better performance.
The forecast ability in the adaptive-ECMS is different from the
model predictive control (MPC) approach used in this brief.
The MPC has attracted increasing attention in the HEV
energy management research community. Given a finite
moving horizon, an MPC controller can maintain computa-
tional load within a practical range. An MPC controller solves
a short-term energy management problem at each time step,
via nonlinear programming [9], quadratic programming [10],
Pontryagin’s minimum principle [11], or DP [12]. The
performance of MPC strongly depends on the power reference
provided in each prediction horizon. The precision of future
power prediction is instrumental for the overall vehicle fuel
economy. Considering that road grade information is static, we
focus on the horizon velocity prediction. No telemetry devices
or environment detecting sensors are assumed, and the future
driving information is completely unknown.
We investigated three velocity predicting approaches. In [9],
we assumed power demand decreases exponentially over the
prediction horizon. The aim of this simple method was to
provide an intuitive understanding of how velocity predic-
tion affects fuel economy. To systematically investigate this
method, a generalized exponentially varying velocity predictor
is considered in this brief.
Markov-chain models are widely used for vehicular velocity
modeling [13], [14]. An MPC controller with a stochastic
Markov-chain velocity predictor is usually called stochastic
MPC (SMPC) [12], [15]. The 1-stage Markov-chain process
has proven to be effective in generating fixed-route driving
patterns. However, when comprehensive driving tasks are
considered, the accuracy of 1-stage Markov chain may decline.
On the other hand, the predicted power demand relies not
only on the present vehicle states, but also on the historical
values [16]. Typically, the more historical data used, the more
accurate the prediction. For SMPC, multistage Markov-chain
processes can hence be formulated to enhance the velocity
prediction accuracy.
Artificial neural networks (NNs) are a successful method for
time series forecasting [17]. Applications of NNs to predicting
city power load [18], driving handling behaviors [19], and traf-
fic flows [20] have verified its strong capability in predicting
nonlinear dynamic behaviors. In [21], we also employed NN
to predict the road type and traffic congestion to improve the
1063-6536 © 2014 IEEE. Personal use is permitted, but republication/redistribution requires IEEE permission.
See http://www.ieee.org/publications_standards/publications/rights/index.html for more information.
1198 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 23, NO. 3, MAY 2015
HEV fuel efficiency. Although tuning the NN parameters may
produce different prediction performances, it is easy to design
an NN-based velocity predictor with reasonable precision.
To the best of our knowledge, NN-based velocity predictor
for predictive energy management of HEVs has not been
investigated.
In this brief, we present a comparative study of three classes
of velocity predictors for MPC-based energy management of
a power-split HEV. This brief adds two original contributions
to the related literature. First, an NN-based velocity predictor
is implemented, for the first time, in MPC-based HEV energy
management. Three different NN structures are examined in
terms of prediction accuracy and consequent fuel economy.
Second, the generalized exponentially varying and Markov-
chain velocity predictors are investigated to fully explore
their potential. The above three classes of velocity predictors
are systematically compared in terms of prediction precision,
computation time, and consequent fuel consumption. Although
the foregoing contributions are made specifically for MPC-
based energy management of a power-split HEV, the proposed
approach extends to other HEV configurations.
The remainder of this brief is organized as follows.
In Section II, the problem formulation is introduced, including
the control-oriented vehicle model and nonlinear MPC opti-
mization strategy. Section III details the three different types of
horizon velocity predictors. Comparison results are illustrated
in Section IV. Finally, the conclusion is drawn in Section V.
II. PROBLEM FORMULATION
A. Control-Oriented Power-Split HEV Model
Power-split HEVs are dominant in the current HEV
market [14]. In a planetary gear set functioned power-split
HEV, the engine and motor/generator 1 (MG1) are connected
to the planet carrier and the sun gear, respectively. A torque
coupler is used to combine the ring gear with MG2 to power
the final drive. The kinematic constraint on the ring, carrier,
and sun gear angular velocities is given mathematically by
ωsS+ωrR=ωc(S+R)(1)
where Sand Rare the radii of the sun gear and the ring gear,
respectively. Angular speeds of the ring, sun, and carrier gears
are denoted as ωr,ωs,andωc, respectively. By neglecting
the inertia of the pinion gears and assuming that all the
powertrain shafts are rigid, inertial dynamics of the powertrain
is derived as
JMG1 dωMG1
dt =TMG1 +FS (2)
Jeng dωeng
dt =Teng F(S+R)(3)
JMG2 dωMG2
dt =TMG2 Taxle
gf+FR (4)
where JMG1,Jeng,andJMG2 are inertias of MG1, engine
and MG2, respectively; Teng =Tcis the engine torque,
TMG1 =Tsand TMG2 =Trare MG1 and MG2 torques
(positive in motoring mode), respectively; Frepresents the
internal force on pinion gears; gfis the gear ratio of the
final drive; and Taxle is the torque produced from powertrain
on the drive axle. To reduce the control-oriented model’s
complexity, we disregard the inertial dynamics, and use the
steady-state values of (2)–(4).MG2 torque and vehicle velocity
are given by
ωMG2 =gf
Rwheel V(5)
mdV
dt =Taxle +Tbrake
Rwheel +mgsin) 1
2ρACdV2
Crmg cos) (6)
where Rwheel is the wheel radius; Vis the vehicle velocity;
mis the vehicle mass; Tbrake is the friction brake torque; gis
the gravitational acceleration; θdenotes the road grade and
is assumed to be zero; (1/2 ACdis the aerodynamic drag
resistance; and Crrepresents the rolling resistance coefficient.
At each time instant, the controller computes an optimal
split between the engine, MG1 and MG2 to minimize fuel
consumption. Fuel flow rate of the engine ( ˙mfuel) and power
transfer efficiencies for MG1 and MG2 (ηMG1 and ηMG2)are
extracted from empirical maps
˙mfuel =ψ1eng,Teng )(7)
ηMG1 =ψ3MG1,TMG1)(8)
ηMG2 =ψ2MG2,TMG2)(9)
where ψ1,ψ2,andψ3are corresponding empirical maps.
Upper and lower boundaries of the battery state of
charge (SoC) need to be specified to maintain the battery
within a safe operating region and extend its life [22]. For
HEV energy management, the SoC is modeled as a single
state. Although more sophisticated battery models have been
developed to describe battery dynamics [23], [24], the internal
resistance model is still the most prevailing model for HEV
supervisory control due to its simplicity, described as
˙
SoC =−
Ibatt
Qmax (10)
Pbatt =Voc Ibatt I2
batt Rbatt (11)
where Ibatt and Qmax are battery current and maximum
capacity, respectively; Pbatt and Rbatt are battery power and
internal resistance; and Voc represents the open circuit voltage.
Positive Pbatt denotes discharge. The battery in a power-split
HEV is connected to a bidirectional converter to supply power
or recuperate energy from the electrical machines. Terminal
battery power is described by
Pbatt =PMG1/(ηMG1ηinv)kMG1 +PMG2/(ηMG2ηinv)kMG2 (12)
where PMG1 and PMG2 are MG1 and MG2 shaft powers,
respectively (positive in motoring mode); ηinv is the inverter
efficiency
ki=1,if Pi>0
1,if Pi0for i={MG1,MG2}. (13)
A complete description of the battery SoC dynamics can be
obtained from (9) to (13). Equations (1)–(13) summarize the
control model used for MPC. Throughout this brief, MPC is
applied to a detailed plant model furnished by the QSS-toolbox
developed at ETH Zürich [25].
SUN et al.: VELOCITY PREDICTORS FOR PREDICTIVE ENERGY MANAGEMENT IN HEVs 1199
B. Nonlinear MPC
A hierarchical MPC [26] is employed for fuel consumption
minimization, formulated as a constrained nonlinear optimiza-
tion problem and solved by DP at each time step. The engine
torque and speed are selected as control variables. Denoting x
as the state variable, uas the control variable, das the system
disturbance, and yas the output, the proposed control-oriented
powertrain model can be represented as
˙x=f(x,u,d), y=g(x,u,d)
with x=SoC, u=[Teng
eng]T,d=Vpredict ,y=
mfuel,Pbatt,TMG1
MG1,TMG2]T,whereVpredict is the future
velocity sequence provided by a velocity predictor in each
control horizon.
Considering a 1 s time step (t=1), at time step k,the
cost function Jkis formulated as
Jk=(k+Hp)t
kt
(mfuel(u(t))]2+λO(t)) dt (14)
where Hpis the prediction horizon length, which is herein
equal to the control horizon length [9]; O(t)is the engine
ON/OFF switching time; and λis the penalty for engine
ON/OFF switching. In addition, the following physical con-
straints must be enforced:
SoCmin SoC SoCmax
Tmin
eng Teng Tmax
eng
ωmin
eng ωeng ωmax
eng
Tmin
MG1 TMG1 Tmax
MG1
ωmin
MG1 ωMG1 ωmax
MG1
Tmin
MG2 TMG2 Tmax
MG2
ωmin
MG2 ωMG2 ωmax
MG2
Imin
batt Ibatt Imax
batt
Pmin
batt Pbatt Pmax
batt .(15)
For HEV structures, the final battery SoC is requested to be
the same as the initial SoC
SoC(T)=SoC(0). (16)
In this brief, the driving information of an assigned trip task
is completely unknown, which complicates the task of satisfy-
ing (16) over the global time horizon via MPC. This problem
can be addressed by approximating a cost-to-go function, or by
regulating the terminal SoC reference in each control horizon.
The second approach is used in this brief. Terminal SoC
references in all control horizons are specified to be SoC(0)
to guarantee the battery SoC converges to the initial value at
the end of the trip. This approach is conservative in utilizing
battery energy, but effective in enforcing constraint (16)
SoC(k+Hp)=SoC(0)(17)
where SoC(k+Hp)is the terminal SoC reference in control
horizon k.
The MPC controller is applied in the supervisory level of
the control architecture. The optimization task is solved using
DP when predicted vehicle velocities are provided in each
receding horizon. For compactness of notation, denote Vk=
V(kt). The control procedure is described as follows [27].
1) A horizon velocity predictor is used to estimate the
control horizon driving profile, based on current velocity
request Vkand historical velocities. Assume fpis the
velocity prediction function
Vpredict =fp(...,Vk3,Vk2,Vk1,Vk)
=Vk+1,Vk+2,...,Vk+Hp1.
2) Given Vpredict , calculate the optimal control policy
minimizing the objective function (14).
3) Apply the first element of the control policy. Feedback
states, and repeat the control procedure.
In this brief, we assume the target vehicle is equipped with-
out any radar, global positioning system or similar devices.
The road grade is zero and future driving profiles are com-
pletely unknown.
III. HORIZON VELOCITY PREDICTORS
Three horizon velocity predictors are investigated in
this brief: 1) generalized exponentially varying predictor;
2) Markov-chain-based predictor; and 3) artificial NN-based
predictor. The NN-based velocity predictor is implemented for
the first time in the HEV energy management problem, and is
compared with the other two approaches.
A. Exponentially Varying Velocity Predictors
The relationship between predicted future velocities and
total fuel consumption is complex. To obtain an intuitive
understanding of this relationship, exponentially varying future
velocities were considered in [28].
In each receding horizon, the exponentially varying horizon
velocities are formulated as
Vk+n=Vk×(1+ε)n,for n=1,2,...,Hp(18)
where Vkis the initial velocity at time step kand εis the
exponential coefficient. Different εvalues are considered to
examine the sensitivity of fuel economy to the predicted future
velocities.
B. Markov-Chain Velocity Predictors
Markov chain is an important methodology used in model-
ing driving velocities or power demands in HEVs [12]–[15].
In this brief, the Markov-chain states and emissions are defined
on discrete-valued domains given by vehicle velocity V(0
to 36 m/s) and vehicle acceleration αin (1.6to1.6m/s
2),
respectively.
Note that the driving behavior studied in this brief is com-
prehensive in an average sense. Thus, the Markov emission
probability distribution is computed from a comprehensive
dataset. Eight different driving cycles are included in this
dataset, considering both highway and urban driving scenarios.
Six of the sample cycles are standard driving cycles and the
other two are real collected driving data. Suppose the vehicle
velocity and acceleration are discretized into pand qintervals
and indexed by iand j, respectively. Velocity at time step kis
1200 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 23, NO. 3, MAY 2015
Fig. 1. One-stage Markov emission probability matrix (60×60).
Vk+n1, and next step acceleration is αk+n,wherenindexes
time in the receding horizon. The Markov-chain process is
defined by an emission probability matrix TRp×qwith
[T]ij =Pr[αk+n=αj|Vk+n1=Vi](19)
for i∈{1,..., p},j∈{1,...,q}and n∈{1,...,Hp}[29].
Suppose p=60 and q=60, the probability matrix extracted
from the sample dataset is shown in Fig. 1.
A random Markov-chain model can update its emission
probability matrix online by utilizing historical driving
data [29]. To reduce complexity and ensure fairness in the
comparison, we do not consider such adaptive Markov-chain
models in this brief. The complexity of the Markov-chain
model will also increase if more conditions, such as tempera-
ture, position, and road grade, are considered.
The precision of Markov-chain velocity predictors relies
heavily on the emission probability matrix. To increase the
prediction accuracy, 1-stage Markov-chain process can be
extended to multistage [30], using the following probabilities:
Pr[αk+n=αj|Vk+n1=Vi1
Vk+n2=Vi2,...,Vk+ns=Vis](20)
where sis the Markov-chain stage number. When the stages
increase, more historical speeds are needed for computing the
probability matrix, and the number of Markov states increases
exponentially. Thus, the resolution of the emission probability
matrix can be improved, leading to higher velocity prediction
precision. In the meantime, the sample velocity dataset for
generating multistage Markov-chain processes needs to cover
all possible input states. This requirement is inevitable and
may cause both the size of the emission probability matrix
and the online computation to grow substantially.
C. NN Velocity Predictors
The NNs can be trained to learn a highly nonlinear
input/output relationship by adjusting weights to minimize the
error between the actual and predicted output patterns of a
training set [31]. Three different types of network structures
are examined in this brief: 1) back propagation (BP-NN);
2) layer recurrent (LR-NN); and 3) radial basis function
NN (RBF-NN).
A three-layer BP-NN has a hierarchical feed forward
network structure, which is also the basis of other network
architectures. The input layer is used to receive and distribute
the input pattern, followed by a hidden layer that depicts the
nonlinearities of the input/output relationship. Output layer
yields the desired output patterns. For a BP-NN, the activation
function is hyperbolic tangent sigmoid function. The basic
formula of BP-algorithm is
a1=tan sig(n)=enen
en+enn=Wa0+b(21)
where a1and a0are the neural outputs of the current layer and
prior layer, respectively, nis the accumulator output, Wis the
weight, and bis the bias. The literature shows that a three-layer
BP-NN with a sigmoid function as the activation function can
approximate any nonlinear systems with arbitrary precision.
The main disadvantage of BP-NN is that it has a slow learning
convergence and is at risk of getting trapped into local minima.
With a self-connected hidden layer, the LR-NN has an
internal state, which allows the network to exhibit temporal
dynamic behavior [32]. However, this may increase the train-
ing convergence time. The basic formula of the recurrent layer
in a LR-NN is
a1(t)=tan sig(n(t)) =enen
en+en
n(t)=Wa0(t)+Wa1(tt)+b(22)
where Wis the feedback weight; a1(tt)is the delayed
output at time (tt).
An RBF-NN is a widely used feed forward network for time
series forecasting. In the standard approach to RBF-NN imple-
mentation, an RBF needs to be predefined at first. Then, the
number of hidden layer neurons is determined. An RBF-NN
usually has better convergence speed and performance
compared with BP-NN. In our simulation, the Gaussian
function is used as the RBF in the hidden layer to activate
the neurons, formulated as
a1=exp nc2
2b2n=Wa0+b(23)
where cis the neural net center and bis the spread width.
Both cand bcan be fit using a gradient descent method.
For velocity prediction purpose, the inputs of NNs are
historical velocity sequences, and the outputs are predicted
horizon velocity sequences. Each input–output pattern is com-
posed of a moving window of fixed length, which can be
expressed as
[Vk+1,Vk+2,...,Vk+Hp]= fNN(VkHh+1,...,Vk)(24)
where Hhis the dimension of the input velocity sequence and
fNN represents the nonlinear map function of an NN-based
predictor. In an MPC framework, the prediction horizon
length is Hp, so that the NN velocity predicting process is
Hp-step ahead.
The size of the NN depends on the number of input nodes
and the number of hidden nodes. Here, we only focus on eval-
uating the speed prediction performance of NNs with different
numbers of hidden layer nodes. The same driving dataset,
which is used for Markov emission probability computation,
is used for networking training. About 85% of the data is
employed as training sample to establish the network and the
rest 15% is used for performance validation.
SUN et al.: VELOCITY PREDICTORS FOR PREDICTIVE ENERGY MANAGEMENT IN HEVs 1201
Fig. 2. Sample (training) and testing (validation) dataset. Real HW means data real life highway driving trip. Real UB means real life urban driving trip.
IV. SIMULATION RESULTS AND COMPARISON
In this section, we provide a comprehensive comparative
analysis of the three velocity forecasting methods. Our dis-
course begins with an explanation of the training and val-
idation data. Then, a definition of performance metrics are
defined.
The eight driving cycle samples used for training include
four highway types and four urban types. The same data
is used both for probability computation in Markov-chain
approach and network training in NN approach. A different
set of eight driving cycles are used for testing and comparing
performance of the velocity predictors, which are standard
driving cycles HWFET, WVUINTER, UDDS, WVUSUB and
real collected driving data Real HW2, Real HW3, Real UB2,
and Real UB3. The real collected cycles in sample and testing
data are from [33] and [34]. All sample and test cycles are
concatenated for ease of presentation in Fig. 2.
The vehicle parameters and efficiency maps are obtained
from the Toyota Prius HEV model in Autonomie [35]. Sim-
ulation was performed on a personal computer with an Intel
Corel i7-3630QM CPU at 2.4GHz. The prediction and con-
trol horizon length Hpis specified to be 10, and the step
computation is completed within 0.75 s. The initial SoC and
final SoC in all the simulations are set as 0.60. Upper and
lower SoC bounds are 0.8 and 0.3, respectively. Simulations
for eight testing trips are conducted. Detailed results for the
UDDS testing cycle are provided in this section, along with
the performance distributions across all eight validation cycles.
The following four metrics are used to assess the tested
velocity predictors.
1) Average root mean squared error (RMSE, in
meter/second) of the predicted velocities in all of
the control horizons.
2) Online computation time Tcal (in microseconds) of the
velocity prediction process at each time step.
3) Violating frequency eof the velocity and acceleration
constraints from the predicted velocity sequences.
4) Compensated fuel consumption (in grams), which is the
most important metric to evaluate vehicle fuel economy.
Due to the property of prediction, MPC cannot guarantee an
exact final SoC constraint at the end of a global time horizon.
TAB LE I
SIMULATION RESULTS OF EXPONENTIALLY VARYING
VELOCITY PREDICTOR FOR UDDS
For fair comparison in the following analysis, the produced
fuel consumption results are compensated by converting
the final battery SoC deviation into equivalent engine fuel.
Simulation results for the UDDS cycle are illustrated in
Sections IV-A–IV-D. The same simulation procedure is
conducted for the other seven testing cycles. A comprehensive
fuel consumption performance comparison is demonstrated
in Section IV-E based on all simulation results.
A. Results of Exponentially Varying Predictor for UDDS
As shown in Table I, improved fuel efficiency can be
achieved when the predicted velocity decreases slightly, as
opposed to an increase or a large decrease. Terminal SoC of
MPC simulation becomes larger as εgrows. This is because
when the predicted velocity changes aggressively, the engine
tends to provide more power. Thus, more engine power will
be absorbed by the battery through the generator. The average
RMSE reaches its minimum when ε=−0.01, yielding the
best fuel economy.
B. Results of Markov-Chain Predictor for UDDS
Different types of driving cycles are included in the Markov
probability generation dataset: 1) four highway types and
2) four urban types. In this case, the driving behavior we
studied is comprehensive and blended, which is the main
reason that the 1-stage Markov-chain model has the worst
performance in characterizing future horizon velocities, as
1202 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 23, NO. 3, MAY 2015
TAB LE I I
SIMULATION RESULTS OF MARKOV-CHAIN PREDICTOR FOR UDDS
TABLE III
SIMULATION RESULTS OF NN-BASED PREDICTOR FOR UDDS
shown in Table II. The 1-stage model is, however, the most
computationally efficient.
For multistage Markov-chain models, one requires an often
prohibitively large set of driving data to identify the transition
matrix. This phenomenon exemplifies the poor applicability of
multistage Markov-chain velocity predictors. To address this
problem in realistic implementation, a sufficiently rich sample
database is required. In this brief, the emission probabilities
for unavailable high-stage Markov states are replaced by the
ones derived from lower-stage states to overcome the lack of
a large dataset. The maximum Markov stage in this simulation
is five, since when the stage number is larger than five, the
extracted probability matrix changes negligibly.
Simulation results of Markov-chain velocity predictor for
UDDS testing are shown in Table II, from which we can see
that more stages results in smaller RMSE. Hence, increased
fuel efficiency can be achieved via multistage Markov-chain
velocity predictors. However, in real-world implementation,
it is usually difficult to guarantee that all possible Markov
states be contained in the probability generation dataset. In
addition, the magnitude of the probability matrix for a mul-
tistage Markov-chain model scales exponentially. Therefore,
Markov-chain models with greater than three stages are
rarely used. Note that adaptive or self-learning improving
approaches are not considered here for fair comparison of the
three velocity predictors.
C. Results of NN-Based Predictor for UDDS
For NN-based predictors, the dimension of the input, Hh,is
specified to be 10. Although more neural nodes result in
higher training precision, excessive complexity can lead to
over-fitting. Table III shows the comparison results between
different types of NNs with different numbers of nodes.
Note that the error caused by velocity/acceleration constraint
violations (e) of the predicted horizon speed is included in the
average RMSE computation.
TAB LE I V
VELOCITY PREDICTOR COMPARISON FOR UDDS
The RBF-NN structure achieves better fuel economy than
the LR-NN or BP-NN velocity predictors, despite a similar
RMSE, since there are substantially fewer constraint
violations (e). Besides, RBF network structure needs much
less training time and tuning effort.
D. Velocity Predictor Assessment
The DP, deterministic MPC (DMPC) and ECMS are bench-
marks. In DMPC, the controller has full knowledge of the
horizon velocity profile at each time step. In ECMS, the
equivalence factor is tuned from the sampling cycles and
implemented in the testing trips as a constant (357.5). To
fully exploit the potential of ECMS, we also tuned the optimal
equivalence factor for each testing trip through trial-and-error.
This well-tuned ECMS is noted as optimal ECMS (OECMS)
and demonstrated for comparison.
The 1-stage and 5-stage Markov-chain velocity predictors
are selected to compare with the best velocity predictors from
the exponentially varying and NN types, denoted as 1-stage
MC, 5-stage MC, 0.01 EV, and RBF-100, respectively. These
four simulations together with benchmark results are shown
in Table IV.
1) Prediction Precision: As can be seen from Table IV,
the average RMSE of the 1-stage MC velocity predictor is
larger than that of the 0.01 EV velocity predictor. This
means that the 1-stage MC process behaves poorlyin modeling
comprehensive and blended driving cycles. The prediction
precision from 5-stage MC predictor is improved by 16%
compared with 1-stage MC, proving that employing more
states can increase the prediction precision. Consequently, the
average RMSE of RBF-100 is the minimum one as opposed
to the other approaches, although the constraints are violated
slightly by 0.18%. This indicates that the RBF-100 predictor
is preferable in modeling comprehensive driving behaviors.
The nature of each prediction method can be seen visually
in Fig. 3. This figure elucidates how a 1-stage MC is too
short-sited to capture velocity profiles longer than a few time
steps. One can also observe how the 5-stage MC roughly pre-
dicts constant acceleration. In contrast, the RBF-100 captures
microtrip-like behaviors better than the other three.
2) Fuel Consumption: From Table IV, we can see that the
RBF-100 predictor is the most energy-saving predictor, which
is reasonable as it achieves the minimum prediction precision.
Fig. 4 shows the SoC trajectories for DP, DMPC, and the
selected MPC controllers given in Table IV. The difference
SUN et al.: VELOCITY PREDICTORS FOR PREDICTIVE ENERGY MANAGEMENT IN HEVs 1203
Fig. 3. Prediction result of the four predictors in Table IV for UDDS. Predicted velocity [meter/second] is shown across the prediction horizon at each time
step.
Fig. 4. SoC trajectories of MPC approaches compared with DP result for
UDDS testing.
between the DP and MPC-based trajectories is due to DP being
a global optimization; whereas the MPC controller yields a
locally optimal solution for each control horizon. With the
same terminal SoC constraint in each receding horizon, the
SoC trajectories of RBF-100 MPC and DMPC are overlaid.
The SoC trajectory for the 5-stage MC approach differs
noticeably from the DMPC benchmark, due to poor velocity
prediction.
3) Computation Time: From Table IV, it is obvious that
the Markov-chain method is computationally heavier than the
NN-based and exponentially varying methods. This is because
the process of generating a stochastic Markov emission needs
additional calculations to form the probability distributions.
4) Constraint Violation: The exponentially-varying and
Markov-chain approaches do not violate the velocity and
acceleration constraints. In the RBF-100 case, the constraints
are violated 24 times (0.18%) in the UDDS cycle. The
constraint violation pertaining to the RBF-100 predictor is thus
negligible.
Fig. 5. Normalized fuel consumption and final SoC comparison. SE is simple
ECMS, OE is optimal ECMS, and DM is DMPC.
E. Summarized Simulation Results
Similar outcomes are achieved for the remaining seven tests.
Fig. 5 shows the average values and standard deviations
of normalized fuel consumption and final SoC. The DP
and OECMS produce the optimal/near-optimal results if the
driving profiles are known a prior. The NN-based predictors
generally maintain the average fuel optimality over 92%,
which is considerably better than the other two predictors and
fairly close to DMPC, both in terms of average and standard
deviation. Compared with simple ECMS, MPC with NN-based
velocity predictors saves more than 7% fuel, and guarantees
the final SoC constrained.
Furthermore, the 4% gap between DP and DMPC fuel
economy is because DP is a global approach, whereas
DMPC conservatively optimizes the fuel consumption locally.
Compared with DMPC, the extra fuel consumption caused
from velocity predicting errors varies from 2% to 21%
based on different methods. Note these results are evaluated
across both emission certification cycles and real-world drive
1204 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 23, NO. 3, MAY 2015
cycle data. Consequently, we conclude the NNs provide a
promising blend of prediction capability and computational
efficiency for MPC energy management in HEVs.
V. CONCLUSION
This brief presents a comparative study of three classes
of velocity predictors for MPC-based energy management
in a power-split HEV: 1) generalized exponentially varying;
2) Markov chains; and 3) artificial NNs. The NN-based
horizon velocity predictor is proposed, and the sensitivity of
its prediction precision on different NN structures is examined
to elucidate a successful template. Generalized exponentially
varying and Markov-chain velocity predictors are systemat-
ically described and compared with the NN-based predictor
in terms of the prediction precision, computational cost, and
fuel economy. Results demonstrate that NN-based velocity
predictors provide the best overall performance across a range
of certification and real-world drive cycles. This brief provides
better understanding for predictive HEV energy management.
ACKNOWLEDGMENT
The authors would like to thank Prof. J. K. Hedrick at
the University of California at Berkeley, Berkeley, CA, USA,
for substantial help and enlightening discussions on model
predictive control theory and implementation.
REFERENCES
[1] A. Sciarretta, M. Back, and L. Guzzella, “Optimal control of parallel
hybrid electric vehicles,” IEEE Trans. Control Syst. Technol., vol. 12,
no. 3, pp. 352–363, May 2004.
[2] A. Sciarretta and L. Guzzella, “Control of hybrid electric vehicles,” IEEE
Control Syst., vol. 27, no. 2, pp. 60–70, Apr. 2007.
[3] C.-C. Lin, H. Peng, J. W. Grizzle, and J.-M. Kang, “Power management
strategy for a parallel hybrid electric truck,” IEEE Trans. Control Syst.
Technol., vol. 11, no. 6, pp. 839–849, Nov. 2003.
[4] L. Serrao, S. Onori, and G. Rizzoni, “ECMS as a realization of
Pontryagin’s minimum principle for HEV control,” in Proc. Amer.
Control Conf. (ACC), Jun. 2009, pp. 3964–3969.
[5] N. Kim, S. Cha, and H. Peng, “Optimal control of hybrid electric vehi-
cles based on Pontryagin’s minimum principle,” IEEE Trans. Control
Syst. Technol., vol. 19, no. 5, pp. 1279–1287, Sep. 2011.
[6] P. Pisu and G. Rizzoni, “A comparative study of supervisory control
strategies for hybrid electric vehicles,” IEEE Trans. Control Syst. Tech-
nol., vol. 15, no. 3, pp. 506–518, May 2007.
[7] A. Sciarretta, L. Guzzella, and M. Back, “A real-time optimal control
strategy for parallel hybrid vehicles with on-board estimation of the
control parameters,” in Proc. IFAC Symp. Adv. Autom. Control, 2004,
pp. 19–23.
[8] C. Musardo, G. Rizzoni, Y. Guezennec, and B. Staccia, “A-ECMS:
An adaptive algorithm for hybrid electric vehicle energy management,”
Eur. J. Control, vol. 11, nos. 4–5, pp. 509–524, 2005.
[9] H. Borhan, A. Vahidi, A. M. Phillips, M. L. Kuang, I. V. Kolmanovsky,
and S. Di Cairano, “MPC-based energy management of a power-split
hybrid electric vehicle,” IEEE Trans. Control Syst. Technol., vol. 20,
no. 3, pp. 593–603, May 2012.
[10] D. Rotenberg, A. Vahidi, and I. Kolmanovsky, “Ultracapacitor assisted
powertrains: Modeling, control, sizing, and the impact on fuel economy,”
IEEE Trans. Control Syst. Technol., vol. 19, no. 3, pp. 576–589,
May 2011.
[11] V. Ngo, T. Hofman, M. Steinbuch, and A. Serrarens, “Predictive gear
shift control for a parallel hybrid electric vehicle,” in Proc. IEEE Veh.
Power Propuls. Conf. (VPPC), Sep. 2011, pp. 1–6.
[12] L. Johannesson, M. Asbogard, and B. Egardt, Assessing the potential
of predictive control for hybrid vehicle powertrains using stochastic
dynamic programming,” IEEE Trans. Intell. Transp. Syst., vol. 8, no. 1,
pp. 71–83, Mar. 2007.
[13] S. J. Moura, H. K. Fathy, D. S. Callaway, and J. L. Stein, “A stochastic
optimal control approach for power management in plug-in hybrid
electric vehicles,” IEEE Trans. Control Syst. Technol., vol. 19, no. 3,
pp. 545–555, May 2011.
[14] J. Liu and H. Peng, “Modeling and control of a power-split
hybrid vehicle,” IEEE Trans. Control Syst. Technol., vol. 16, no. 6,
pp. 1242–1251, Nov. 2008.
[15] G. Ripaccioli, D. Bernardini, S. Di Cairano, A. Bemporad, and
I. V. Kolmanovsky, A stochastic model predictive control approach
for series hybrid electric vehicle power management,” in Proc. Amer.
Control Conf. (ACC), Jun./Jul. 2010, pp. 5844–5849.
[16] R. Langari and J.-S. Won, “Intelligent energy management agent
for a parallel hybrid vehicle—Part I: System architecture and design of
the driving situation identification process,” IEEE Trans. Veh. Technol.,
vol. 54, no. 3, pp. 925–934, May 2005.
[17] G. P. Zhang, “Time series forecasting using a hybrid ARIMA and neural
network model,” Neurocomputing, vol. 50, pp. 159–175, Jan. 2003.
[18] P. Santos, A. G. Martins, and A. J. Pires, “Designing the input vector to
ANN-based models for short-term load forecast in electricity distribution
systems, Int. J. Elect. Power Energy Syst., vol. 29, no. 4, pp. 338–347,
2007.
[19] Y. Lin, P. Tang, W. J. Zhang, and Q. Yu, Artificial neural network
modelling of driver handling behaviour in a driver-vehicle-environment
system, Int. J. Veh. Design, vol. 37, no. 1, pp. 24–45, 2005.
[20] E. I. Vlahogianni, J. C. Golias, and M. G. Karlaftis, “Short-term traffic
forecasting: Overview of objectives and methods,” Transp. Rev., vol. 24,
no. 5, pp. 533–557, 2004.
[21] J. Park et al., “Intelligent vehicle power control based on machine
learning of optimal control parameters and prediction of road type
and traffic congestion,” IEEE Trans. Veh. Technol., vol. 58, no. 9,
pp. 4741–4756, Nov. 2009.
[22] S. Piller, M. Perrin, and A. Jossen, “Methods for state-of-charge
determination and their applications,” J. Power Sour., vol. 96, no. 1,
pp. 113–120, 2001.
[23] X. Hu, S. Li, and H. Peng, “A comparative study of equivalent circuit
models for Li-ion batteries,” J. Power Sour., vol. 198, pp. 359–367,
Jan. 2012.
[24] X. Hu, S. Li, H. Peng, and F. Sun, “Robustness analysis of state-of-
charge estimation methods for two types of Li-ion batteries, J. Power
Sour., vol. 217, pp. 209–219, Nov. 2012.
[25] L. Guzzella and A. Amstutz, “CAE tools for quasi-static modeling and
optimization of hybrid powertrains,” IEEE Trans. Veh. Technol., vol. 48,
no. 6, pp. 1762–1769, Nov. 1999.
[26] P. Falcone, F. Borrelli, H. E. Tsengz, J. Asgari, and D. Hrovat,
“A hierarchical model predictive control framework for autonomous
ground vehicles,” in Proc. Amer. Control Conf. (ACC), Jun. 2008,
pp. 3719–3724.
[27] E. F. Camacho and C. Bordons, Model Predictive Control,vol.2.
London, U.K.: Springer-Verlag, 2004.
[28] H. A. Borhan and A. Vahidi, “Model predictive control of a power-
split hybrid electric vehicle with combined battery and ultracapacitor
energy storage,” in Proc. Amer. Control Conf. (ACC), Jun./Jul. 2010,
pp. 5031–5036.
[29] M. Bichi, G. Ripaccioli, S. Di Cairano, D. Bernardini, A. Bemporad,
and I. V. Kolmanovsky, “Stochastic model predictive control with driver
behavior learning for improved powertrain control, in Proc. 49th IEEE
Conf. Decision Control (CDC), Dec. 2010, pp. 6077–6082.
[30] G. Bolch, S. Greiner, H. de Meer, and K. S. Trivedi, Queueing Net-
works and Markov Chains: Modeling and Performance Evaluation With
Computer Science Applications. New York, NY, USA: Wiley, 2006.
[31] M. T. Hagan, H. B. Demuth, and M. H. Beale, Neural Network Design.
Boston, MA, USA: PWS-Kent, 1996.
[32] A. Graves, M. Liwicki, S. Fernández, R. Bertolami, H. Bunke, and
J. Schmidhuber, “A novel connectionist system for unconstrained hand-
writing recognition,” IEEE Trans. Pattern Anal. Mach. Intell., vol. 31,
no. 5, pp. 855–868, May 2009.
[33] J. C. Herrera, D. B. Work, R. Herring, X. J. Ban, Q. Jacobson, and
A. M. Bayen, “Evaluation of traffic data obtained via GPS-enabled
mobile phones: The Mobile Century field experiment,” Transp. Res. C,
Emerg. Technol., vol. 18, no. 4, pp. 568–583, 2010.
[34] D. LeBlanc et al., “Road departure crash warning system field oper-
ational test: Methodology and results,” Univ. Michigan Transportation
Research Institute, Ann Arbor, MI, USA, Tech. Rep. UMTRI-2006-9-1,
Jun. 2006.
[35] R. V. Gopal and A. Rousseau, “System analysis using multiple expert
tools,” SAE Technical Paper 2011-01-0754, Jan. 2011.
... Therefore, it is considered to be the most valuable control strategy in actual deployment. [22][23][24] So far, MPC has been widely studied by enterprises and scholars, and a series of research results have been achieved. It has been widely used in the energy management of many different types of new energy vehicles, such as FCVs, HEVs, and EVs. 25 Depending on whether the uncertainty of future driving information is taken into account, MPC can be further divided into two categories: deterministic MPC and stochastic MPC. ...
... DMPC: DMPC with ANN based predictor. DMPC: DMPC with the improved deterministic predictor described in equation (22). SMPC: SMPC with the stochastic predictor described in equation (28). ...
Article
Full-text available
The performance of model predictive control strategies for hybrid electric vehicles (HEVs) highly depends on the accuracy of future speed predictions. This paper proposes improved prediction models for deterministic model predictive control (DMPC) and stochastic model predictive control (SMPC), respectively. For DMPC, the neural network-based predictor is first introduced and taken as the benchmark predictor. A novel deterministic predictor considering historical prediction errors is proposed, which relies on the assumption that the offset between the prediction and measurement at current instant is a good estimate of the offset in the short future. Based on the proposed deterministic predictor, a stochastic predictor that considers the distribution law of historical data at different locations is further proposed for SMPC. Simulation results show that the controller using the proposed deterministic prediction model improves fuel economy by 2.89%, and the controller using the proposed stochastic prediction model improves fuel economy by 4.5% compared with the benchmark.
... A Markov chain (MC) is the most representative stochastic VVP algorithm [3]. A multi-level MC model is considered to be an effective measure for improving the prediction accuracy of an MC; however, the size of the transition probability matrix grows exponentially with additional model inputs, which leads to a high computational burden when using this algorithm, making it impossible to cover all potential Markov states [4,5]. Deterministic VVP algorithms can be further divided into parametric methods and non-parametric methods [6]. ...
... (3) The acceleration parameter a was the most crucial feature parameter to enhancing the model prediction accuracy, while the number and type of feature parameters, as well as the distribution of the feature parameters between the training and testing data sets, had a significant impact on VVP model performance. (4) In terms of the sliding window length, the three VVP models achieved a relatively higher prediction accuracy when the sliding window was between 15 and 20 s. ...
Article
Full-text available
Vehicle velocity prediction (VVP) plays a pivotal role in determining the power demand of hybrid electric vehicles, which is crucial for establishing effective energy management strategies and, subsequently, improving the fuel economy. Neural networks (NNs) have emerged as a powerful tool for VVP, due to their robustness and non-linear mapping capabilities. This paper describes a comprehensive exploration of NN-based VVP methods employing both qualitative theory analysis and quantitative numerical simulations. The used methodology involved the extraction of key feature parameters for model inputs through the utilization of Pearson correlation coefficients and the random forest (RF) method. Subsequently, three distinct NN-based VVP models were constructed comprising the following: a backpropagation neural network (BPNN) model, a long short-term memory (LSTM) model, and a generative pre-training (GPT) model. Simulation experiments were conducted to investigate various factors, such as the feature parameters, sliding window length, and prediction horizon, and the prediction accuracy and computation time were identified as key performance metrics for VVP. Finally, the relationship between the model inputs and velocity prediction performance was revealed through various comparative analyses. This study not only facilitated the identification of an optimal NN model configuration to balance prediction accuracy and computation time, but also serves as a foundational step toward enhancing the energy efficiency of hybrid electric vehicles.
... If the prediction accuracy is improved, the exponential coefficient needs to change according to the different working conditions [4]. Sun et al. [5] studied the prediction accuracy and hybrid electric vehicle (HEV) fuel consumption of the exponential function prediction model. However, this prediction method relies on the assumption of exponential changes in future driving information, and if this assumption is violated, its wide application will be hindered. ...
Article
Full-text available
Accurate vehicle velocity prediction is of great significance in vehicle energy distribution and road traffic management. In light of the high time variability of vehicle velocity itself and the limitation of single model prediction, a velocity prediction method based on K-means-QPSO-LSTM with kinematic segment recognition is proposed in this paper. Firstly, the K-means algorithm was used to cluster samples with similar characteristics together, extract kinematic fragment samples in typical driving conditions, calculate their feature parameters, and carry out principal component analysis on the feature parameters to achieve dimensionality reduction transformation of information. Then, the vehicle velocity prediction sub-neural network models based on long short-term memory (LSTM) with the QPSO algorithm optimized were trained under different driving condition datasets. Furthermore, the kinematic segment recognition and traditional vehicle velocity prediction were integrated to form an adaptive vehicle velocity prediction method based on driving condition identification. Finally, the current driving condition type was identified and updated in real-time during vehicle velocity prediction, and then the corresponding sub-LSTM model was used for vehicle velocity prediction. The simulation experiment demonstrated a significant enhancement in both the velocity and accuracy of prediction through the proposed method. The proposed hybrid method has the potential to improve the accuracy and reliability of vehicle velocity prediction, making it applicable in various fields such as autonomous driving, traffic management, and energy management strategies for hybrid electric vehicles.
... Traditional statistical methods of machine learning that operate on recurrent neural networks, such as simple linear regression, direct method, and recursive methods often have difficulty making accurate predictions when dealing with large data and widely dispersed [8]. In this paper, recurrent neural network (RNN) is the preferable prediction model because it can accommodate the temporal sequence formed by state of charge (SOC), distance, and vehicle speed. ...
Article
Full-text available
The goal is to develop an adaptive energy management method that can predict driving circumstances, reducing stress on the primary storage battery in a hybrid electric vehicle (HEV). To improve battery lifespan and overall efficiency, power allocation among multiple energy storage sources (ESS) must go beyond traditional rule-based schemes. This paper offers a promising solution to address these challenges. Standardized driving cycles, such as the Urban Dynamometer Driving Schedule (UDDS), are used as load torque to evaluate the power demand in hybrid electric vehicles (HEV). First, feature extraction techniques improve the extraction of important features from historical driving cycle data, thereby facilitating the long short-term memory (LSTM) based on multistep velocity predictors using a warm-up algorithm. This predicted future velocity profile serves as input to the HEV system. Second, within the hybrid electric vehicle (HEV) system, the battery is linked to a modified interleaved DC-DC boost converter. The results show that this converter affects factors such as ripple, converter loss of 29%, state of charge (SOC) level, and efficiency improvement to 98%. Finally, the energy management system (EMS) with the modified rule-based scheme is implemented with neural networks for optimization. This EMS is responsible for continuously adjusting the power allocation. Simulation results show a significant average reduction of about 33.33% at peak battery level, thereby extending battery life. Compared with conventional rule-based EMS methods, the proposed strategy significantly improves efficiency by reducing the total energy loss by about 10.36%. These results emphasize the reliability and robustness of the proposed strategy.
Article
In-wheel-motor-driven electric vehicles (IWM-EVs) provide more potential to enhance vehicle stability performance. However, traditional stability control relies on the current status fed back by sensors for stability judgment and control, only taking effect after the vehicle has already become unstable. In response to this issue, this paper proposes a pre-stability control strategy based on a hybrid dynamic state prediction method to predict dangerous driving conditions and intervene in vehicle stability control in advance. First, a driver-vehicle model is established to characterize the driver's driving intention and obtain the vehicle's ideal motion responses. Then, the methodology for implementing vehicle pre-stability control is introduced, which mainly includes sideslip angle estimation utilizing the extended Kalman filter, a hybrid dynamic state prediction approach based on vehicle model and data trends, and a vehicle pre-stability judgment method. Subsequently, a vehicle hierarchical controller is designed to achieve pre-stability control. The upper-level controller focuses on calculating the required additional yaw moment, and the lower-level controller aims to optimize torque distribution among the four wheels. Finally, the proposed pre-stability control strategy is validated by the hardware-in-the-loop test bench. The results show that the proposed control strategy can intervene in dangerous driving conditions in advance, and its mean errors of the yaw rate and sideslip angle are reduced by over 17.1% and 23.5%, respectively, compared with the traditional method, which significantly enhances vehicle stability and driving safety.
Article
Full-text available
Many of today"s advanced simulation tools are suitable for modeling specific systems; however, they provide rather limited support for model building and management. Setting up a detailed vehicle simulation model requires more than writing down state equations and running them on a computer. In this paper, we describe how modern software techniques can be used to support modeling and design activities, with the objective of providing better system models more quickly by assembling these system models in a "plug-and-play" architecture. Instead of developing detailed models specifically for Argonne National Laboratory"s Autonomie modeling tool, we have chosen to place emphasis on integrating and re-using the system models, regardless of the environment in which they were initially developed. By way of example, this paper describes a vehicle model composed of a detailed engine model from GT Power, a transmission from AMESim, and with vehicle dynamics from CarSim. The paper will explain the different options available for the interface and how each of these options can be implemented. It will use a simple case study to show how the detailed expert simulation models can be used with simpler Simulink models to address different vehicle design issues.
Article
Full-text available
This paper presents a comparative study of twelve equivalent circuit models for Li-ion batteries. These twelve models were selected from state-of-the-art lumped models reported in the literature. The test data used is obtained from a battery test system with a climate chamber. The test schedule is designed to measure key cell attributes under highly dynamical excitations. The datasets were collected from two types of Li-ion cells under three different temperatures. The multi-swarm particle swarm optimization algorithm is used to identify the optimal model parameters for the two types of Li-ion cells. The usefulness of these models is then studied through a comprehensive evaluation by examining model complexity, model accuracy, and robustness of the model by applying the model to datasets obtained from other cells of the same chemistry type.
Article
Full-text available
Modelling driver handling behaviour in a driver-vehicle-environment (DVE) system is essentially useful for the design of vehicle systems and transport systems in the light of the safety and efficiency of human mobility. Driver handling behaviour is reflected in two aspects: the mental workload and the performance. Further, this behaviour is exposed through the interactions between driver-vehicle and driver-environment. There is generally a lack of the first principle with which to develop a model for human behaviour. In this study, several more sophisticated artificial neural network architectures for developing models for human drivers in a DVE system were used. The vehicle dynamics are modelled by a 3-d.o.f. model derived from the first principle. The experiment was performed and compared with a DVE simulation system in which the developed human driver behaviour model was included, together with the vehicle dynamics model. The comparative study showed that the simulation result is in good agreement with the experimental result, which further justifies the effectiveness of the developed driver behaviour model.
Article
The paper presents a novel approach for hybrid powertrain control, based on a real-time minimization of the equivalent fuel consumption. This approach is non-predictive, thus the control strategy developed requires only information on the current status of the powertrain. The control parameters are continuously estimated on-board using information deriving from static route mapping and telemetry. Simulations for a prototypical hybrid car under development show the very promising benefits in terms of fuel consumption reduction and charge sustaining that can be obtained with the controller presented.
Article
In the last two decades, the growing need for short-term prediction of traffic parameters embedded in a real-time intelligent transportation systems environment has led to the development of a vast number of forecasting algorithms. Despite this, there is still not a clear view about the various requirements involved in modelling. This field of research was examined by disaggregating the process of developing short-term traffic forecasting algorithms into three essential clusters: the determination of the scope, the conceptual process of specifying the output and the process of modelling, which includes several decisions concerning the selection of the proper methodological approach, the type of input and output data used, and the quality of the data. A critical discussion clarifies several interactions between the above and results in a logical flow that can be used as a framework for developing short-term traffic forecasting models.
Article
A power-split hybrid electric vehicle (HEV) combines the advantages of both series and parallel hybrid vehicle architectures by utilizing a planetary gear set to split and combine the power produced by electric machines and a combustion engine. Because of the different modes of operation, devising a near optimal energy management strategy is quite challenging and essential for these vehicles. To improve the fuel economy of a power-split HEV, we first formulate the energy management problem as a nonlinear and constrained optimal control problem. Then two different cost functions are defined and model predictive control (MPC) strategies are utilized to obtain the power split between the combustion engine and electrical machines and the system operating points at each sample time. Simulation results on a closed-loop high-fidelity model of a power-split HEV over multiple standard drive cycles and with different controllers are presented. The results of a nonlinear MPC strategy show a noticeable improvement in fuel economy with respect to those of an available controller in the commercial Powertrain System Analysis Toolkit (PSAT) software and the other proposed methodology by the authors based on a linear time-varying MPC.
Article
In this paper, Model Predictive Control (MPC) framework is exploited to synthesize a predictive controller for a parallel Hybrid Electric Vehicle (HEV) equipped with an Automated Manual Transmission. The algorithm also controls the gear shift command, together with the power split between the engine and electric machine and the engine on-off state using the route information ahead. A non-predictive controller based on a combination of Dynamic Programming (DP) and Pontryagin's Minimum Principle (PMP) is described and taken as a benchmark control solution for optimizing the gear shift problem of the parallel HEV in terms of computational efficiency. This so-called DP-PMP control approach is then utilized in the MPC framework to realize the predictive controller for a gear shift problem in a receding horizon mode. Simulation results show that the non-predictive controller improves the fuel economy up to 35.9% and 43.5% on NEDC and FTP75 respectively when compared with a conventional vehicle. Even with a short horizon, fuel saving of the predictive controller is very close to that of the non-predictive controller with a relative difference of 0.3%. Moreover, the predictive controller can be seen as a suitable realtime implementable control candidate with a fast computation property.