Discover millions of ebooks, audiobooks, and so much more with a free trial

Only $11.99/month after trial. Cancel anytime.

Modeling, Dynamics, and Control of Electrified Vehicles
Modeling, Dynamics, and Control of Electrified Vehicles
Modeling, Dynamics, and Control of Electrified Vehicles
Ebook874 pages78 hours

Modeling, Dynamics, and Control of Electrified Vehicles

Rating: 5 out of 5 stars

5/5

()

Read preview

About this ebook

Modelling, Dynamics and Control of Electrified Vehicles provides a systematic overview of EV-related key components, including batteries, electric motors, ultracapacitors and system-level approaches, such as energy management systems, multi-source energy optimization, transmission design and control, braking system control and vehicle dynamics control. In addition, the book covers selected advanced topics, including Smart Grid and connected vehicles. This book shows how EV work, how to design them, how to save energy with them, and how to maintain their safety.

The book aims to be an all-in-one reference for readers who are interested in EVs, or those trying to understand its state-of-the-art technologies and future trends.

  • Offers a comprehensive knowledge of the multidisciplinary research related to EVs and a system-level understanding of technologies
  • Provides the state-of-the-art technologies and future trends
  • Covers the fundamentals of EVs and their methodologies
  • Written by successful researchers that show the deep understanding of EVs
LanguageEnglish
Release dateOct 19, 2017
ISBN9780128131091
Modeling, Dynamics, and Control of Electrified Vehicles

Read more from Haiping Du

Related to Modeling, Dynamics, and Control of Electrified Vehicles

Related ebooks

Mechanical Engineering For You

View More

Related articles

Reviews for Modeling, Dynamics, and Control of Electrified Vehicles

Rating: 5 out of 5 stars
5/5

1 rating0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Modeling, Dynamics, and Control of Electrified Vehicles - Haiping Du

    Modeling, Dynamics, and Control of Electrified Vehicles

    Edited by

    Hui Zhang

    Beihang University, Beijing, China

    Dongpu Cao

    Cranfield University, Bedford, United Kingdom

    Haiping Du

    University of Wollongong, Wollongong, NSW, Australia

    Table of Contents

    Cover image

    Title page

    Copyright

    List of Contributors

    Chapter 1. Modeling, Evaluation, and State Estimation for Batteries

    Abstract

    1.1 Introduction

    1.2 Battery Modeling

    1.3 Evaluation of Model Accuracy

    1.4 State Estimation

    1.5 Conclusions

    References

    Chapter 2. High-Power Energy Storage: Ultracapacitors

    Abstract

    2.1 Introduction

    2.2 Modeling

    2.3 UC State Estimation

    2.4 Conclusions

    Further Reading

    Chapter 3. HESS and Its Application in Series Hybrid Electric Vehicles

    Abstract

    3.1 Introduction

    3.2 Modeling and Application of HESS

    3.3 Conclusion

    References

    Chapter 4. Transmission Architecture and Topology Design of EVs and HEVs

    Abstract

    4.1 Introduction

    4.2 EV and HEV Architecture Representation

    4.3 Topology Design of Power-Split HEV

    4.4 Topology Design of Transmission for Parallel Hybrid EVs

    4.5 Conclusion

    Reference

    Chapter 5. Energy Management of Hybrid Electric Vehicles

    Abstract

    5.1 Introduction

    5.2 Energy Management of HEVs

    5.3 Case Study

    5.4 Model Predictive Control Strategy

    5.5 Results

    5.6 Conclusions

    References

    Chapter 6. Structure Optimization and Generalized Dynamics Control of Hybrid Electric Vehicles

    Abstract

    6.1 Introduction

    6.2 Generalized Dynamics Models

    6.3 Extended High-Efficiency Area Model

    6.4 Typicals Applications

    6.5 Conclusions

    References

    Chapter 7. Transmission Design and Control of EVs

    Abstract

    7.1 Introduction

    7.2 EVs Equipped with IMT Powertrain System

    7.3 Problem Formulation

    7.4 Oscillation Damping Controller Design

    7.5 Simulation Results

    7.6 Conclusion

    Funding

    References

    Further Reading

    Chapter 8. Brake-Blending Control of EVs

    Abstract

    8.1 Introduction

    8.2 Brake-Blending System Modeling

    8.3 Regenerative Braking Energy-Management Strategy

    8.4 Dynamic Brake-Blending Control Algorithm

    8.5 Conclusion

    References

    Further Reading

    Chapter 9. Dynamics Control for EVs

    Abstract

    9.1 Introduction

    9.2 Modeling and Control of EVs

    9.3 Sensing and Estimation

    9.4 Active Safety Control

    9.5 Riding and Energy Efficiency Control

    9.6 Conclusions

    References

    Chapter 10. Robust Gain-Scheduling Control of Vehicle Lateral Dynamics Through AFS/DYC

    Abstract

    10.1 Introduction

    10.2 Development of Uncertain Vehicle Dynamics Model

    10.3 Main Results

    10.4 Simulation Results

    10.5 Conclusions

    Acknowledgments

    References

    Chapter 11. State and Parameter Estimation of EVs

    Abstract

    11.1 Introduction

    11.2 Velocity Estimation (Longitudinal, and Total, Preferred Method and Alternatives)

    11.3 Slip-Angle Estimation

    11.4 Tire-Force and Tire–Road Friction Coefficient Estimation

    11.5 Vehicle Mass- and Road Slope-Estimation Method

    11.6 Conclusions

    References

    Further Reading

    Chapter 12. Modeling and Fault-Tolerant-Control of Four-Wheel-Independent-Drive EVs

    Abstract

    12.1 Introduction

    12.2 System Modeling and Problem Formulation

    12.3 Fault-Tolerant Tracking Controller Design

    12.4 Simulation Investigations

    12.5 Conclusions

    References

    Chapter 13. Integrated System Design and Energy Management of Plug-In Hybrid Electric Vehicles

    Abstract

    13.1 Introduction

    13.2 Powertrain Modeling

    13.3 Heuristic Scenarios

    13.4 Emission Mitigation via Renewable Energy Integration

    13.5 Optimal Scenario With Integrated System Design and Energy Management

    13.6 Battery-Health Implication

    13.7 Conclusions

    References

    Appendix

    Chapter 14. Integration of EVs With a Smart Grid

    Abstract

    14.1 Introduction

    14.2 Powertrain Modeling

    14.3 Formulation of Cost-Optimal Control Problem

    14.4 Results and Discussion

    14.5 Conclusions

    References

    Index

    Copyright

    Woodhead Publishing is an imprint of Elsevier

    The Officers’ Mess Business Centre, Royston Road, Duxford, CB22 4QH, United Kingdom

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States

    The Boulevard, Langford Lane, Kidlington, OX5 1GB, United Kingdom

    Copyright © 2018 Elsevier Inc. All rights reserved.

    No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/permissions.

    This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein).

    Notices

    Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary.

    Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility.

    To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein.

    British Library Cataloguing-in-Publication Data

    A catalogue record for this book is available from the British Library

    Library of Congress Cataloging-in-Publication Data

    A catalog record for this book is available from the Library of Congress

    ISBN: 978-0-12-812786-5 (print)

    ISBN: 978-0-12-813109-1 (online)

    For information on all Woodhead Publishing publications visit our website at https://www.elsevier.com/books-and-journals

    Publisher: Joe Hayton

    Acquisition Editor: Sonnini R. Yura

    Editorial Project Manager: Ana Claudia Garcia

    Production Project Manager: Omer Mukthar

    Cover Designer: Victoria Pearson

    Typeset by MPS Limited, Chennai, India

    List of Contributors

    Dongpu Cao,     Cranfield University, Bedford, United Kingdom

    Philip Commins,     University of Wollongong, Wollongong, NSW, Australia

    Haiping Du,     University of Wollongong, Wollongong, NSW, Australia

    Hiroshi Fujimoto,     The University of Tokyo, Tokyo, Japan

    Hongwen He,     Beijing Institute of Technology, Beijing, China

    Jibin Hu,     Beijing Institute of Technology, Beijing, China

    Xiaosong Hu,     Chongqing University, Chongqing, China

    Yanjun Huang,     University of Waterloo, ON, Canada

    Amir Khajepour,     University of Waterloo, ON, Canada

    Boyuan Li,     University of Wollongong, Wollongong, NSW, Australia

    Liang Li,     Tsinghua University, Beijing, China

    Wei Liu,     Beijing Institute of Technology, Beijing, China

    Chen Lv

    Cranfield University, Cranfield, United Kingdom

    Cranfield University, Bedford, United Kingdom

    Brett McAulay,     University of Wollongong, Wollongong, NSW, Australia

    Fei Meng,     Shanghai Maritime University, Shanghai, China

    Hao Mu,     Beijing Institute of Technology, Beijing, China

    Jun Ni,     Beijing Institute of Technology, Beijing, China

    Zengxiong Peng,     Beijing Institute of Technology, Beijing, China

    Hong Wang,     University of Waterloo, ON, Canada

    Junmin Wang,     The Ohio State University, Columbus, OH, United States

    Rongrong Wang,     Shanghai Jiao Tong University, Minhang, China

    Xiangyu Wang,     Tsinghua University, Beijing, China

    Yafei Wang,     Shanghai Jiao Tong University, Shanghai, China

    Rui Xiong,     Beijing Institute of Technology, Beijing, China

    Chao Yang,     Tsinghua University, Beijing, China

    Sixiong You,     Tsinghua University, Beijing, China

    Hui Zhang,     Beihang University, Beijing, China

    Lei Zhang,     Beijing Institute of Technology, Beijing, China

    Shuo Zhang,     Beijing Institute of Technology, Beijing, China

    Xiaoyuan Zhu,     Shanghai Maritime University, Shanghai, China

    Chapter 1

    Modeling, Evaluation, and State Estimation for Batteries

    Hao Mu and Rui Xiong,    Beijing Institute of Technology, Beijing, China

    Abstract

    State estimation of the lithium-ion battery has been the focus of many researchers, and the consensus is that the model-based method is an effective tool for state of charge (SoC) estimation. In this chapter, we start with battery modeling. Several modeling approaches are presented and the their advantages and disadvantages are discussed. Moreover, the balance problem between model accuracy and complexity of an nth order RC networks model is tackled using an evaluation index of terminal voltages. Finally, the adaptive extended Kalman filter algorithm is proposed to estimate the SoC and its validity is confirmed.

    Keywords

    Electric vehicles; evaluation of models; lithium-ion battery; modeling; state of charge; dual timescales; AEKF algorithm

    1.1 Introduction

    Currently, hybrid electric vehicles (HEVs) and electric vehicles (EVs) promise a future of green travel in which fuel-consuming engines are replaced with electric motors, thus reducing our dependence on fossil energy and ultimately producing less harmful emissions. Such vehicles can be plugged in at home overnight or at the office or in a parking space during the day, using electricity that is generated at a centralized power station or even by renewable sources. The key component to the achievement of these electrical systems is the energy storage system, namely, the battery technology.

    The lithium-ion (Li-ion) battery, as depicted in Fig. 1.1, is the most common choice for phone communication and portable appliances because of its many advantages, such as high energy-to-weight and power-to-weight ratios (180 Wh/kg and 1500 W/kg, respectively) and low self-discharge rate (Linden and Reddy, 2002; Capasso and Veneri, 2014). In addition, among all rechargeable electrochemical systems, Li-ion technology is the first-choice candidate as a power source for HEVs/EVs. However, this technology is still delicate and affected by numerous limitations, such as issues of safety (Doughty and Roth, 2012), cost (Lajunen and Suomela, 2012), recycling (Gaines, 2011), and charging infrastructure (Veneri et al., 2012).

    Figure 1.1 Different types of Li-ion batteries.

    To ensure the power battery works safely and reliably, which is a function of the battery management system (BMS), the temperature, voltage, and current of the batteries should be monitored and the states of the batteries should be estimated precisely in real time (Junping et al., 2009; He et al., 2010; Camus et al., 2011). However, it is hard to measure the states of batteries, like the state of charge (SoC), state of health (SoH), and state of function (SoF) directly due to the complicated electrochemical process and various factors in practical applications. Thus estimation methods based on battery models are developed broadly.

    The remainder of this chapter is organized as follows: Section 1.2 introduces several kinds of modeling approaches for Li-ion batteries, such as physical-based models, equivalent circuit models (ECMs), etc. In Section 1.3, some regular battery tests are presented, which are indispensable for battery research. Then, considering the popularity of different models, the ECMs are selected to illustrate parameter identification methods, which can be divided into offline and online ones according to real-time capability. Due to the balance problem between model accuracy and the computation burden of the BMS, an evaluation criterion is introduced to determine the optimal number of RC networks in the models. Section 1.4 is the core part of this chapter and covers state estimation of batteries, in particular about SoC estimation. Many SoC estimation methods will be classified systematically and the multiscale adaptive extended Kalman filter (MAEKF) algorithm for state and parameter collaborative estimation will be elaborated on since it is not only provides satisfactory estimation accuracy, but also low computation burden. Some conclusions are drawn in Section 1.5 and references are listed in references section.

    1.2 Battery Modeling

    Many battery models, which are lumped models with relatively few parameters, have been put forward especially for the purpose of vehicle power management control and BMS development. The most commonly used models can be categorized as electrochemical models and ECMs (Plett, 2004a; He et al., 2011a, 2011b; Vasebi et al., 2007; Zhu et al., 2011; Hussein and Batarseh, 2011; Hu et al., 2012). Electrochemical models utilize a set of coupled nonlinear differential equations to describe the pertinent transport, thermodynamic, and kinetic phenomena occurring in the cell. They can translate the distributions into easily measurable quantities such as cell current and voltage and build a relationship between the microscopic quantities, such as electrode and interfacial microstructure and the fundamental electrochemical studies and cell performance. However, they typically deploy partial differential equations (PDEs) with a large number of unknown parameters, which often leads to large memory requirements and heavy computation burdens, so the electrochemical battery models are not desirable for BMSs (Smith et al., 2010). The simplified electrochemical models, which ignore the thermodynamic and quantum effects, are proposed to simulate the electrochemical and voltage performance. The Shepherd model, the Unnewehr universal model, the Nernst model, and the combined model are the typical choices. The equivalent circuit battery models are developed by using resistors, capacitors, and voltage sources to form a circuit network. Typically, a big capacitor or an ideal voltage source is selected to describe the open-circuit voltage (OCV); the remainder of the circuit simulates the battery’s internal resistance and relaxation effects such as dynamic terminal voltage. The Rint model, the Thevenin model, the DP model, and their revisions are widely used.

    1.2.1 Physical-Based Models

    Electrochemical models usually use coupled nonlinear PDEs to describe ion transport phenomena and electrochemical reactions to achieve high accuracy, but incur heavy computation load. For instance, a pseudo two-dimensional (P2D) model, developed by Doyle et al. (1993), is one of the most popular variants and can take seconds to minutes to simulate (Ramadesigan et al., 2012). For simplicity, Atlung et al. (1979) developed a single particle model (SPM) that assumes electrodes are represented by two single spherical particles. To improve the accuracy of the SPM under high C-rate, several extended single particle models (E-SPMs) have been proposed (Luo et al., 2013; Schmidt et al., 2010; Khaleghi Rahimian et al., 2013), where Li-ion concentration and potential distribution in electrolyte are taken into account. In general, electrochemical models such as P2Ds, SPMs, and E-SPMs are more accurate than ECMs, but require a large number of immeasurable parameters, leading to overfitting in parametric identification. Therefore the pursuit for battery models with high accuracy and computational efficiency still remains a challenge.

    Although electrochemical battery models are suitable for understanding the electrochemical reactions inside the battery, their complexity often leads to the need for more memory and computational effort. Thus they may not be practical in the fast computation and real-time implementations needed for EV BMS. This problem has been addressed by many researchers by investigating reduced-order models (ROMs) that predict the battery behavior with varying degrees of fidelity (Smith et al., 2008, 2010). To reduce the order of an electrochemical battery model, discretization techniques can be applied to retain only the most significant dynamics of the full-order model (Tanim et al., 2015). Various discretization techniques are utilized to simplify the full model’s PDEs into a set of ODEs of the ROM while keeping the fundamental governing electrochemical equations. In Shi et al., 2011, six different discretization methods (listed in Table 3) are addressed and compared for battery system modeling.

    1.2.1.1 Single Particle Model

    The SPM assumes a single electrode particle in each electrode and negligible electrolyte diffusion. Conservation of Li+ species in a single spherical active material particle is described by Fick’s law of diffusion:

    (1.1)

    where r∈(0,Rs) is the radial coordinate, Rs is the particle radius, cs(r,t) is the concentration of Li+ ions in the particle as a function of radial position r and time t, and Ds is the solid-phase diffusion coefficient. We use the subscripts s and s, e to indicate the solid-phase and solid/electrolyte, interface, respectively. The boundary conditions are

    (1.2)

    (1.3)

    where j(x, t) is the rate of electrochemical reaction at the particle surface (with j>0 indicating ion discharge), F is Faraday’s constant (96,487 C/mol), and as is the specific interfacial surface area. For the spherical active material particles occupying electrode volume fraction εs, as=3εs/Rs. The linearized Butlere–Volmer electrochemical kinetics and is given by

    (1.4)

    where the overpotential η drives the current flow across the electrode/electrolyte interface and Rct denotes the charge transfer resistance, which can expressed as follows:

    (1.5)

    where i0 is the exchange current density, R is the universal gas constant, T is the temperature, and αa and αc are the anodic and cathodic transfer coefficients, respectively.

    1.2.1.2 Pseudo Two-Dimensional Model

    The P2D model, as depicted in Fig. 1.2, is constructed based on the assumption that electrodes are seen as an aggregation of spherical particles (2D representation) in which the Li+ ions are inserted. The first spatial dimension of this model, represented by variable x, is the horizontal axis. The second spatial dimension is the particle radius r. The cell is comprised of three regions that imply four distinct boundaries. The specific descriptions of this model can be found in Sabatier et al. (2015).

    Figure 1.2 Systematic chart of P2D model.

    1.2.2 Lumped Parameter Electric Model

    The complexity of the electrochemical models and limitations of the computers in the past led researchers to investigate another modeling approach called electrical circuit modeling or equivalent circuit (EC) modeling. Today, for many applications, it is important to strike a balance between model complexity and accuracy so that the models can be embedded in microprocessors and provide accurate results in real-time (Pattipati et al., 2011). In other words, it is important to have models that are accurate enough, and not unnecessarily complicated. EC modeling is one of the most common battery modeling approaches especially for EV applications. Having less complexity, these models have been used in a wide range of applications and for various types of batteries (Marc et al., 2008; Fotouhi et al., 2015; He et al., 2011). The EC models are constructed by putting resistors, capacitors, and voltage sources in a circuit. The simplest form of an EC battery model is the internal resistance model (Johnson, 2002). The model consists of an ideal voltage source Uoc and a resistance Ro. Adding one RC network to the internal resistance model can increase its accuracy by considering the polarization characteristics of a battery. Such models are called Thevenin models (Salameh et al., 1992) and are illustrated in Fig. 1.3; in this figure, Ut is the battery’s terminal voltage, Uoc is the OCV, IL is the load current, Ro is the internal resistance, Rp and Cp are equivalent polarization resistance and capacitance, respectively.

    Figure 1.3 Schematic of Thevenin model.

    Adding more RC networks to the battery model may improve its accuracy but it increases the complexity too. Thus a compromise is needed when computational effort and time are vital. This subject is discussed in more detail in the following sections.

    Recently, fractional order models (FOMs) have attracted increasing interest in the field of electrochemical energy storage systems. One of the earlier FOMs was proposed for NiMH batteries by Kuhn et al. (2004), in which the parameters were identified based on impedance data in the frequency domain. Wu et al. (2013) proposed an FOM and performed time-domain parametric identification with the Levenberge–Marquardt algorithm, but fixed the differentiation orders at 0.5 and 1 through the estimation study. Xu et al. (2013) presented a fractional Kalman filter for SoC estimation based on a FOM, where the differentiation order of the Warburg element was also fixed at 0.5, and the other model parameters were identified based on a single pulse response. The fixing of differentiation orders helps to reduce the difficulty of parametric identification, but also significantly limits the model accuracy.

    One common EC model used in EIS tests was proposed by John Edward Brough Randles in 1947. The model, called Randles circuit model, is illustrated in Fig. 1.4. In cell modeling using the EIS method, each component of the electrical circuit model is related to an electrochemical process in the cell.

    Figure 1.4 Randle circuit.

    In this model, Ro is the ohmic resistance, the pseudo RC network is used to simulate the charge transfer process and double layer effect, and the Warburg impedance is used to describe the diffusion phenomenon of ions in solid phase. In actual applications, due to the capacitance dispersion, the Warburg impedance can be expressed in s-domain as:

    (1.6)

    where ZW denotes the impedance, WD is the coefficient, α is the order to evaluate capacitance dispersion (0≤α≤1), and when α=0 is the resistance, α=1 is the capacitor.

    1.3 Evaluation of Model Accuracy

    1.3.1 Some Experiments

    1.3.1.1 OCV Test

    The OCV is a measure of the electromotive force (EMF) of the battery, which is known to have a monotonic relationship with the SoC of the battery.

    Existing OCV modeling approaches can be broadly classified into chemistry-based and current–voltage based approaches. In chemistry-based approaches, the OCV of each electrode (anode and cathode w.r.t. some reference) is expressed as a function of the utilization of the electrode (the lithium concentration in the electrode normalized by the maximum possible concentration) or the SoC of each electrode. It is generally assumed that this anode and cathode SoC varies linearly with the cell SoC. Subsequently, the difference between the OCV of the anode and cathode gives the OCV of the complete cell. High current rates (i.e., near the rated maximum) have been shown to affect the macroscopic processes in a way that the OCV hysteresis vanishes for Li-ion cells, which regularly show OCV hysteresis after low current application. Roscher et al. conducted OCV (full and partial charge-discharge cycle) tests on Li-ion phosphate (LiFePO4) batteries to characterize the hysteresis and recovery effects. The final OCV model is constructed by concatenating the actual SoC, the recovery factor, and the hysteresis factor.

    The current–voltage based OCV–SoC characterization can be summed up in two simple steps:

    1. Collect pairs of {OCV, SoC} values, spanning the entire range of SoC from 0 to 1.

    2. Use the above data to estimate the parameters of the function OCV=f (SoC) for a hypothesized function f.

    Some important factors will influence the OCV–SoC curve such as aging and temperature.

    On the left side of Fig. 1.5, the OCV–SoC characterization curves of new and aged batteries are shown. New battery curves are plotted in solid blue and aged battery curves are plotted in dashed red. Different curves of the same type correspond to temperatures ranging from 25 to 50. On the right, nominal OCV modeling uses Cnom=1.5 Ah in computing SoC at all temperatures.

    Figure 1.5 Aging and temperature will influence the OCV–SoC experiment results. Source: Pattipati, B., Balasingam, B., Avvari, G.V., Pattipati, K.R., Bar-Shalom, Y., 2014. Open circuit voltage charaterization of lithium-ion batteries. J. Power Sources 269, 317–333 (Pattipati et al., 2014).

    There are two main methods for OCV tests: low-current OCV tests and incremental OCV tests; Fig. 1.6 shown the latter.

    Figure 1.6 Results of OCV test with certain SoC intervals and rest periods.

    1.3.1.2 HPPC Test

    In order to acquire data to identify the model parameters, a hybrid pulse power characterization (HPPC) test procedure is conducted at certain SoC intervals (constant current C/3 discharge segments) starting from 1.0 to 0.1 and each interval follows by a 2-hour rest to allow the battery to get electrochemical and thermal equilibrium before applying the next. The HPPC current profile is shown in Fig. 1.7. The voltage, current, and SoC profiles of the HPPC test are shown in Fig. 1.7B–D. The sampling time is 1 second.

    Figure 1.7 (A) HPPC current profile; (B) current profiles of the HPPC test; (C) voltage profiles of the HPPC test; (D) calculated SoC profiles of the HPPC test.

    1.3.1.3 Driving Cycle Experiment

    The dynamic stress test (DST) and the federal urban dynamic schedule (FUDS) test are the commonly used test procedures given in battery test procedure manuals. The DST uses a 360 second sequence of power steps with seven discrete power levels. The DST is a typical driving cycle that is often used to evaluate various battery models and SoC estimation algorithms. The SoC profiles and zoomed current profiles of this test are plotted in Fig. 1.8. As the DST driving cycle, the FUDS is a standard time-velocity profile for urban driving vehicles as well, which can be seen in Fig. 1.9.

    Figure 1.8 The plots of the DST test: (A) the SoC versus time profiles and (B) the current versus time profiles.

    Figure 1.9 The plots of the FUDS test: (A) the SoC versus time profiles and (B) the sample current versus time of one FUDS cycle.

    1.3.2 Parameter Identification Methods

    1.3.2.1 Offline Methods

    To identify the parameters in different models, a least squares (LS) method and genetic algorithm are presented. The LS method can be applied to identify parameters in different SoCs of the battery via the HPPC test mentioned above. Taking the Thevenin model as an example, the state-space equations can be formulated as follows:

    (1.7)

    The parameters are the same as those in Fig. 1.3. The OCV–SoC curve can be fitted by the model:

    (1.8)

    where z denotes as the SoC. The discrete form of this equation can be achieved by using the first-order backward difference:

    (1.9)

    where the coefficients ci are:

    (1.10)

    This equation can be reformulated as the matrix form:

    (1.11)

    where Y is the vector of terminal voltage, Y=[Ut(1) Ut(2) … Ut(N)]T, H is the coefficient matrix, H=[φ(1) φ(2) … φ(N)]T, φ(k)=[1 Ut(k−1) ln(z(k)) ln(1−z(k)) IL(k) IL(k−1)]T, and θ=[c1 c2 c3 c4 c5 c6]T. Hence, the parameter vector θ can be calculated by the LS method, and the concrete equation is shown as follows:

    (1.12)

    Moreover, in practical use, the battery cannot work under the certain and regular driving cycle, like the HPPC test. Therefore the genetic algorithm is presented to deal with the parameter identification problems under some complex situation.

    To evaluate the accuracy of models, the root mean square error (RMSE) of terminal voltage is set as the indicator. Minimizing the index is the cost function for the optimization problem:

    (1.13)

    is the predicted terminal voltage from the model.

    1.3.2.2 Online Methods

    In order to improve the prediction precision of the battery model, we use the recursive least squares (RLS) method with an optimal forgetting factor to carry out online parameter identification.

    A model-based method can provide a cheap alternative in estimation or it can be used along with a sensor-based scheme to provide some redundancy. The RLS algorithm is based on the minimization of the sum of squared prediction errors, where estimated process model parameters are improved progressively with each new process data acquired. The RLS method with an optimal forgetting factor (RLSF) has been widely used in estimation and tracking of time varying parameters in various fields of engineering. Many successful implementations of RLSF-based adaptive control for time varying parameters estimation are available in the literature.

    Consider a single-input single-output (SISO) process described by the general higher-order autoregressive exogenous (ARX) model:

    (1.14)

    where y is measured system output, which denotes the terminal voltage Ut in this article. φ and θ are the information matrix and the unknown parameter matrix, respectively. The parameters in θ can either be constant or subject to infrequent jumps. ξ is a stochastic noise variable (random variable with normal distribution and zero mean), and k is a nonnegative integer, which denotes the sample interval, k=0, 1, 2, ….

    For the recursive function of Eq. (1.13), the system identification is realized as follows:

    (1.15)

    is the estimate of the parameter matrix, e(k) is the prediction error of the terminal voltage, K(k) is the algorithm gain, and P(k) is the covariance matrix; λ is the forgetting factor, typically λ=[0.95, 1], and is very important to obtain a good estimated parameter set with small error.

    To achieve an accurate SoC profile to evaluate OCV-based SoC estimates, we should build the SoC reference data as ture SoC first. Due to the hard determination of the exact SoC value, herein we determine the initial SoC and the terminal SoC of the lithium iron-phosphate cell according to the definition of SoC with a standard charging experiment and a further standard discharging experiment after finishing a test, so the initial SoC and the terminal SoC are accurate. The coulomb counting approach is used to calculate the SoC since it can keep track of the accurate SoC with accurate initial SoC, battery capacity and current. We also improve the SoC accuracy with a revision method based on the accurate terminal SoC. Considering all the battery experiments are carried out in a temperature chamber the SoC calculation method is feasible with an acceptable accuracy.

    After setting the initial value P0 and θ0, the online parameter identification model coded by Simulink/xPC Target can be used to get the model’s parameters. The online parameter identification results are shown in Fig. 1.10. Fig. 1.10A shows the online terminal voltage observer profiles, Fig. 1.10B shows the parameter estimation result for the first parameter of θ, Fig. 1.10C shows the parameter estimation result for the second parameter of θa1, and Fig. 1.10D shows the parameter estimation results for the third and fourth parameter of θa2,a3.

    Figure 1.10 The online parameter identification results: (A) terminal voltage; (B) the first parameter of θ; (C) the second parameter of θ; and (D) the third and fourth parameters of θ.

    On the one hand, we will evaluate the proposed online parameter identification performance by the terminal voltage estimation accuracy of the dynamic model. On the other hand, we will use the online OCV estimate to infer battery SoC with the OCV–SoC lookup table.

    The maximum error, minimum error, and mean error and RMSE are selected as evaluation indexes to evaluate the estimation accuracy of the terminal voltage observer. The estimation error can describe the difference between the experimental data and online estimated value directly. The RMSE is used to evaluate the deviation degree of estimated value and experiment value, which can describe the present error and the convergence of the estimation algorithm together.

    Fig. 1.11 describes the evaluation results for the online estimation accuracy of the terminal voltage, the voltage error between the sensor values. Fig. 1.11A indicates the online parameter identification method can estimate the terminal voltage accurately, especially with a bad initial value of P0 and θ0. In addition, the proposed method has robust performance for a bad initial value. Fig. 1.11A gives a direct reflection of the convergence characteristics for the proposed online parameters identification method, and the terminal voltage estimation result based on the RLSF algorithm with convergence to the true value within less than 30 sample intervals (the sample interval is 1 second). The detail error index is shown in Table 1.1. Although the absolute error is 1.2292 V, it converges to the true value quickly, so both the error and the RMSE (nominal voltage is 32 V of the battery module) are less than 1%.

    Figure 1.11 Evaluation of online terminal voltage estimation accuracy: (A) error and (B) RMSE performance.

    Table 1.1

    Statistic list of model errors

    The online estimation results for the OCV and the ohmic resistance can be deduced from Fig. 1.10, and are shown in Fig. 1.12. We can calculate the SoC based on the OCV estimates in Fig. 1.12B and the SoC–OCV data. The comparative profiles between true SoC and estimated SoC are shown in Fig. 1.13A, and the SoC estimated error profiles and RMSE performance profiles are shown in Fig. 1.13B. Fig. 1.13A shows that the OCV values provide an acceptable way to estimate the battery SoC with good robustness regardless of the initial value; its SoC estimate error is shown in Fig. 1.13B, which shows the fluctuation of its error is within 0.04 after several sampling times when the online parameter identification process is stable. Its RMSE performance shows that the SoC estimation error is getting smaller and converging to zero quickly. Therefore the online parameter identification method not only can ensure the dynamic voltage simulation accuracy of the Thevenin model with real-time model parameters, but also can provide an acceptable SoC estimation and a maximum error within 4%.

    Figure 1.12 Online parameter identification results for the Uoc and Ro: (A) Uoc and (B) Ro.

    Figure 1.13 OCV estimates based on SoC estimation accuracy: (A) SoC estimation results and (B) RMSE performance.

    1.3.3 Evaluation of n-RC Networks Model

    For Li-ion batteries, based on an analysis on the structure of the Rint model and the Thevenin model, an equivalent circuit model with n-RC networks, named the NP model hereafter, is built. Fig. 1.14 shows the NP model where IL is the load current with a positive value in the discharge process and a negative value in the charge process, UL is the terminal voltage, Uoc is the OCV, Ro is the equivalent ohmic resistance, Ci is the ith equivalent polarization capacitance, Ri is the ith equivalent polarization resistance simulating the transient response during a charge or discharge process, and Ui is the voltage across Ci. i=1, 2, 3, 4, …, n. The electrical behavior of the NP model can be expressed by Eq. (1.15) in the frequency domain:

    (1.16)

    where s is the frequency operator.

    Figure 1.14 Schematic diagram of the NP model.

    1.3.3.1 In Case of n=0

    The NP model is simplified as the Rint model and a discretization form of Eq. (1.16) is written as Eq. (1.17), where k denotes the discretization step with a sample interval of T, k=1, 2, 3, …:

    (1.17)

    , then,

    (1.18)

    1.3.3.2 In Case of n=1

    The NP model is simplified as the Thevenin model and then

    (1.19)

    Define EL=UL−Uoc, the transfer function G(s) of Eq. (1.19) can be written as Eq. (1.20):

    (1.20)

    A bilinear transformation method shown in Eq. (1.21) is employed for the discretization calculation of Eq. (1.20) and the result is shown in Eq. (1.22).

    (1.21)

    where z is the discretization operator.

    (1.22)

    , then Ro can be solved according to the united equations of a1, a2, and a3.

    (1.23)

    Eq. (1.19) is rewritten as Eq. (1.24) after discretization, where k=1, 2, 3, …

    (1.24)

    The OCV is greatly influenced by SoC, working temperature Tem, and working history H, which are all the functions of time t. Herein we define the OCV as a function of SoC, Tem, and H as shown in Eq. (1.25).

    (1.25)

    Differentiate Uoc in Eq. (1.24) with respect to time t:

    (1.26)

    definitely holds since H represents a long usage history.

    (1.27)

    (1.28)

    Then Eq. (1.24) is rewritten as Eq. (1.29):

    (1.29)

    , then,

    (1.30)

    In the case of online application, the UL(k) and IL(k) are sampled at constant period, the vector θ1 can be identified by a RLS algorithm according to Eq. (1.30), and then the parameters can be solved by the expressions of a1, a2, a3.

    1.3.3.3 In Case of n=2

    The NP model is simplified and then

    (1.31)

    The corresponding transfer function G(s) of Eq. (1.31) is

    (1.32)

    The bilinear transformation method shown in Eq. (1.21) is employed for the discretization calculation of Eq. (1.32), then

    (1.33)

    where b1, b2, b3, b4, and b5 are the coefficients solved from Eq. (1.32). Similar to the case of n=1, a discretization form of Eq. (1.31) is arranged as Eq. (1.34), where k=2, 3, 4, …:

    (1.34)

    Define

    , and

    , then

    (1.35)

    In the case of online application, the UL(k) and IL(k) are sampled at constant period, the vector θ2 can be identified by a RLS algorithm according to Eq. (1.35), and the model parameters can be solved by Eq. (1.36):

    (1.36)

    1.3.3.4 In Case of n=n

    The electrical behavior of the NP model is deduced and written as Eq. (1.37):

    (1.37)

    The bilinear transformation method shown in Eq. (1.21) is employed for the discretization of Eq. (1.37), and Eq. (1.38) is deduced according to Eq. (1.29) and Eq. (1.34), where k=n, n+1, n+2, …:

    (1.38)

    Similarly, a recursive function is built as :

    (1.39)

    (1.40)

    (1.41)

    We have compared different orders of RC network models under three dynamic driving cycles, and the corresponding evaluation results are given in Figs. 1.15–1.17. It can be seen that the model parameters will increase manifoldly with the increase of the number of RC networks, and the calculation burden will be heavier and a larger memory will be required to store the large amount of sample data. It is meaningful to properly select a minimum RC network with an acceptable accuracy.

    Figure 1.15 Evaluation results under the HPPC test: (A) the statistics results of the voltage error; (B) the statistics results of the RMSE; (C) the voltage profiles of the Shepherd model-based estimation, the Nernst model-based estimation, and the HPPC test; and (D) the voltage profiles of the DP model-based estimation and the HPPC test.

    Figure 1.16 Evaluation results under the DST test: (A) the statistics results of the voltage error and (B) the statistics results of the RMSE.

    Figure 1.17 Evaluation results under the FUDS test: (A) the statistics results of the voltage error; (B) the statistics results of the RMSE; (C) the RMSE profiles of four specific models; and (D) the voltage profiles of the DP model-based estimation and the FUDS test.

    Considering the model complexity, when the number of the RC networks n is more than 5, a big error will arise from the linear discrete method. However, if another nonlinear parameter identification method is used such as Kalman filters, huge computing costs will result due to the complex model structure. Thus the evaluation tests are only conducted for the General equivalent circuit model

    Enjoying the preview?
    Page 1 of 1