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

Only $11.99/month after trial. Cancel anytime.

Tribological Processes in the Valve Train Systems with Lightweight Valves: New Research and Modelling
Tribological Processes in the Valve Train Systems with Lightweight Valves: New Research and Modelling
Tribological Processes in the Valve Train Systems with Lightweight Valves: New Research and Modelling
Ebook594 pages11 hours

Tribological Processes in the Valve Train Systems with Lightweight Valves: New Research and Modelling

Rating: 1.5 out of 5 stars

1.5/5

()

Read preview

About this ebook

Tribological Processes in Valvetrain Systems with Lightweight Valves: New Research and Modelling provides readers with the latest methodologies to reduce friction and wear in valvetrain systems—a severe problem for designers and manufacturers. The solution is achieved by identifying the tribological processes and phenomena in the friction nodes of lightweight valves made of titanium alloys and ceramics, both cam and camless driven.

The book provides a set of structured information on the current tribological problems in modern internal combustion engines—from an introduction to the valvetrain operation to the processes that produce wear in the components of the valvetrain. A valuable resource for teachers and students of mechanical or automotive engineering, as well as automotive manufacturers, automotive designers, and tuning engineers.

  • Shows the tribological problems occurring in the guide-light valve-seat insert
  • Combines numerical and experimental solutions of wear and friction processes in valvetrain systems
  • Discusses various types of cam and camless drives the valves used in valve trains of internal combustion engines—both SI and CI
  • Examines the materials used, protective layers and geometric parameters of lightweight valves, as well as mating guides and seat inserts
LanguageEnglish
Release dateJun 17, 2016
ISBN9780081009734
Tribological Processes in the Valve Train Systems with Lightweight Valves: New Research and Modelling
Author

Krzysztof Jan Siczek

Dr. Siczek is a Master Engineer in Mechanical Engineering, with a specialization in Cars and Tractors at Technical University of Lodz, Poland. He teaches Automobile Mechatronics at the Lodz Centre of Excellence for Teacher Training and Practical Training. He is also a Lecturer in the Department of Machine Design and Exploatation/Department of Precise Design/Department of Vehicle and Fundamentals of Machine Design. Responsible for teaching of Descriptive Geometry, Technical Drawing, Informatics, CAD. His current research focuses on selfstarters, valvetrain elements, shock absorbers, loom mechanisms, and properties of composites.

Related to Tribological Processes in the Valve Train Systems with Lightweight Valves

Related ebooks

Mechanical Engineering For You

View More

Related articles

Reviews for Tribological Processes in the Valve Train Systems with Lightweight Valves

Rating: 1.5 out of 5 stars
1.5/5

2 ratings0 reviews

What did you think?

Tap to rate

Review must be at least 10 words

    Book preview

    Tribological Processes in the Valve Train Systems with Lightweight Valves - Krzysztof Jan Siczek

    Tribological Processes in the Valve Train Systems with Lightweight Valves

    New Research and Modeling

    Krzysztof Jan Siczek

    Department of Vehicles and Fundamentals of Machine Design, Lodz University of Technology, Poland

    Table of Contents

    Cover image

    Title page

    Copyright

    Preface

    Chapter 1. Introduction

    Abstract

    Chapter 2. Principles of valve train operation

    Abstract

    Withdrawal from the Basic Valve Timing

    Lead, Lag, and Overlap

    Valve Timing Diagrams

    Four-Stroke Cycle

    Two-Stroke Engines

    Rotary Port System for IC Engines

    Arrangement of Poppet Valves

    Valve Train System with Poppet Valves

    Classification of Variable Valve Actuation Technology

    Variable Valve Timing

    Deactivation of the Cylinder and Valve

    Chapter 3. Spark-ignition engine valve trains

    Abstract

    Effect of Variable Control on the Operation of Lightweight Valves

    Variable Valve Timing Systems

    Camshaft-Based Mechanisms for the Valve Variable Operation in the Produced Engines

    Variable Valve Timing via Shifting the Camshaft Phases

    Variable Valve Timing via Special Design

    Chapter 4. Compression-ignition engine valve trains

    Abstract

    Valve Timing Control Systems in Compression-Ignition Engines

    Course of Valve Lifts during Timing Phase Changes in CI Engines

    Hydraulic Systems of Lost Lift

    Profile Generation Systems

    Variable Speed Systems

    Use of Variable Valve Control Systems for Standard CI Engines

    Review of Cam Valve Drives

    Chapter 5. Valve train thermodynamic effects

    Abstract

    Effects of Changes to Outlet (Exhaust) Valve Opening Timing

    Effects of Changes to Outlet Valve Closing Timing

    Effect of Changes to Inlet Valve Opening Timing

    Effect of Changes to Inlet Valve Closing Timing

    Inlet (Intake) Cam Phasing

    Outlet (Exhaust) Cam Phasing

    Dual-Equal Cam Phasing

    Dual-Independent Cam Phasing

    Cold-Start Valve Phasing Strategies

    Effects of Valve Overlap

    Effect of Valve Stroke

    Exhaust Gas Recirculation

    The Effect of Valve Timing on Effective Compression Ratio

    The Effect of Valve Train on In-Cylinder Turbulence

    The Effect of Valve Train on the Exhaust Temperature

    The Effect of Valve Train on Overexpansion

    The Effect of Valve Train on Turbo Charging

    Two- and Three-Step Strategies of Variable Valve Actuation

    Strategies for Full Variable Valve Timing Control

    Chapter 6. Valve train kinetic effects

    Abstract

    Control Cycle of Valve Motion

    Operating Conditions of the Valve Train Components

    Valve Rotation

    Auxiliary Rotation System

    Misalignment of Seat Insert Relative to Valve Guide

    Forces Loading Elements of Valve Train

    Modeling of Valve Train

    Stiffness of Valve Train

    Contact Between Cam and Follower

    Cam Profile

    Spring

    Lash Adjuster and Hydraulic Chain Tensioners

    Friction Phenomena in the Nodes of the Cam Valve Train With Fixed Phases

    Criteria for the Tribological Quality of the System

    Chapter 7. Valve train tribology

    Abstract

    Tribological Problems in the Guide–Lightweight Valve–Seat Insert Subsystem

    Basic Concepts Related to Friction

    Guidelines for the Design of the Model for the Guide–Valve–Seat Insert Assembly Treated as a Tribological System

    The Wear Process of Friction Pairs in the HOPI–SOPG System

    Chapter 8. Mechanical component design and analysis

    Abstract

    Drive System of the Valve Train

    Camshafts

    Valve Springs

    Small Parts in the Valve Train

    Classical Valves

    Lightweight Valves

    Valve Guides

    Seat Inserts

    Chapter 9. Advanced mechanical valve train design and analysis

    Abstract

    Variable Valve Stroke by Switching the Cam Profile

    Systems with Continuous Change of Valve Stroke

    Variable Valve Lift, Connecting the Valve Timing Change, and Changing of the Profile

    Variable Control of the Valves via the Camshaft

    Summary of Cam Valve Drives

    Chapter 10. Future valve train systems

    Abstract

    Electromagnetic Valve Drive

    Electromechanical Valve Drive

    Electrohydraulic Valve Drive

    Electropneumatic Valve Drive

    Chapter 11. Research on valve trains

    Abstract

    Testing Methods

    Computer Simulation of Friction and Wear at the Nodes Valve–Guide and Valve–Seat Insert

    Effect of the Wear of Components of the HOPI–SOPG System on the Sum of Flows in the Gap Between the Valve Stem and Guide and Between the Seat Faces of the Valve and Its Insert

    Simulation Algorithm

    References

    Index

    Copyright

    Butterworth-Heinemann is an imprint of Elsevier

    The Boulevard, Langford Lane, Kidlington, Oxford OX5 1GB, UK

    50 Hampshire Street, 5th Floor, Cambridge, MA 02139, USA

    Copyright © 2016 Elsevier Ltd. 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: 978-0-08-100956-7

    For Information on all Butterworth-Heinemann publications visit our website at http://www.elsevier.com/

    Publisher: Joe Hayton

    Acquisition Editor & Editorial Project Manager: Carrie Bolger

    Production Project Manager: Anusha Sambamoorthy

    Designer: Matthew Limbert

    Typeset by MPS Limited, Chennai, India

    Preface

    Nowadays the goal of achieving the best engine performance and lowest fuel consumption and emissions drives the development of all engine assemblies. Because the valve train operation influences all these aspects, its development is of crucial importance.

    Valve trains in modern engines operate in complex conditions that change both from cycle to cycle and long term. Proper recognition of them should take into account not only courses of loading, temperature, and lubrication conditions during the operating cycle but also wear and friction for all mating elements. The introduction of new materials also changes the tribological properties for mating surfaces.

    Although there are many excellent articles and books about valve train design, the effects of wear and friction of all valve train components on its operation as a whole are rarely taken into account. Because modeling and control of valve trains have played key roles in the technical development of all assemblies in modern combustion engines, I decided to write this book.

    The material assembled in this book represents my work during the past decade on engine valve train research and development at Lodz University of Technology in Poland. This book is intended to provide a better understanding of engine valve trains by presenting the major aspects of valve train modeling, control, simulation, and design.

    This book consists of 11 chapters. Chapter 1 describes different solutions for obtaining increased fuel economy and lower emissions, including variable valve control, exhaust gas recirculation, direct injection, and hybridization of vehicles. The approximate criterion for classifying valves as lightweight is also presented.

    Chapter 2 describes the principles of valve train operation. These include engine types, lead, lag, overlap, scavenging, rotary port system, poppet valves arrangement, variable valve actuation, variable valve timing, and cylinder and valve deactivation.

    Chapter 3 presents the different spark-ignition engine valve trains. They can use camshaft phaser, adjustable timing, stepwise adjustable valve lifts, or stepless adjustable valve lifts.

    Chapter 4 discusses the different compression-ignition engine valve trains, including compression-ignition engine valve timing control, systems of direct action, hydraulic systems of lost lift, profile generation systems, and variable speed systems.

    Chapter 5 presents valve train thermodynamic effects and describes valve opening strategy, valve closing strategy, exhaust gas recirculation, cam phasing, cold-start valve phasing, the role of valve overlap, valve stroke, effective compression ratio, exhaust temperature, and turbocharging.

    Chapter 6 discusses the valve train kinetic effects. It is especially considers valve train operating conditions, valve rotation, seat insert - guide misalignment, cam profile, forces loading valve train, valve train stiffness, valve spring, lash adjuster, friction phenomena in valve train nodes, tribological quality criteria and quality indicators.

    Chapter 7 addresses the valve train tribology. In particular, the term tribology is explained. The chapter also discusses the tribological phenomena, friction models and compensation, lubrication, wear intensity, and models and the role of pollutants.

    Chapter 8 discusses the mechanical component design and analysis. It discusses the materials, design, and analysis methods for the valve train drive system, including gear, chain, and cogged belt drives for camshafts; valve springs; and small parts in the valve train, including spring accessories, rocker arms and cam followers, lifters, pushrods, and valve lash adjustment elements. Different aspects of the classical and lightweight valves, guides, and seat inserts are also presented.

    Chapter 9 elaborates on the advanced mechanical valve train design and analysis. It presents solutions for obtaining variable valve stroke by switching the cam profile. It also discusses systems with continuous change of valve stroke, variable valve lift, and variable control of the valves via the camshaft. In addition, it provides a review of the cam valve drives.

    Chapter 10 is concerned with the future of valve train systems. Camless drives (electromagnetic, electromechanic, electrohydraulic, and electropneumatic) are presented, and the role of valve settling speed is discussed.

    Finally, Chapter 11 discusses research on valve trains, including testing methods, testers for valve trains, computer simulations, and the role of sum of media flows. It also presents a simplified simulation algorithm.

    This book was written as an engineering reference book on the analysis and modeling of valve trains with lightweight valves. It can be useful for training courses on valve train development and design. It should enable design engineers to understand valve train control algorithm design and development. It can be useful for both undergraduate- and graduate-level valve train modeling and design courses. I hope that this book will succeed in helping the reader understand this interesting technology.

    I thank my colleague or, better, my mentor Krzysztof Zbierski, PhD. Eng., for his excellent cooperation and help during my research. I also thank my colleagues Maciej Kuchar, PhD. Eng., Zbigniew Kossowski, PhD. Eng., and Piotr Jozwiak, MSc. Eng., for their help with my research. In addition, I thank Prof. Krzysztof Wituszynski for his help and insightful comments. Also, I cordially thank the reviewers and the Elsevier team, especially Ms. Carrie Bolger, for their cooperation and assistance during preparation of this book.

    Chapter 1

    Introduction

    Abstract

    In the chapter the different solutions for obtaining the increased fuel economy and lower emissions are pointed, including variable valve control, exhaust gas recirculation, direct injection and hybridization of vehicles. The variable valve control system adds a few degrees of freedom to control the internal combustion engine. Increasing the speed of engine with cam or camless valve train requires a low weight of moving parts, like valves, to reduce inertia forces loading the timing and the power required to its drive. The change of valve material needs usually also changing the materials of guides and seat inserts, what influences the operation conditions, friction between mating elements and their wear intensity. Also the lubrication of contacts can be changed, and the possible solutions are shortly discussed. The approximate criterion for classifying valves as lightweight is also presented.

    Keywords

    solutions increasing fuel economy and lower emissions; variable valve control system; tribological problems related to lightweight valves; lubrication system changes; criterion for classifying valves as lightweight

    In the current worldwide population of several million vehicles equipped with internal combustion engines, different solutions are employed to obtain increased fuel economy and lower emissions, which are necessary due to increasingly stringent environmental standards [1]. Some are well known, whereas others are still in development. Examples of such solutions include variable valve actuation (VVA), exhaust gas recirculation, direct injection, and hybridization of vehicles. The VVA system adds a few degrees of freedom to control the internal combustion engine.

    Tribological processes that occur in the existing valve train with cam-driven valves are well known and described in the literature [2–4]. In current solutions of valve timing with cam drive, the steel valves are used in conjunction with seat inserts and guides of cast alloy. The operation is provided under conditions of mixed friction due to intentional limits on the amount of oil supplied to the contact zones of the valve stem, guide, and valve seats and seat insert. Extortions acting on elements of the guide–valve–seat insert set are repeatable and subject to duty cycle of the engine, applied geometry, and stiffness in the elements of the valve train. Variations in these conditions occur mainly during cold engine warm-up and are short-lived.

    Increasing the speed of engines with a cam or camless valve train requires the moving parts, such as valves, to be lightweight to reduce inertia forces loading the timing and the power required to drive it.

    A relatively new area of use of VVA engines is hybrid vehicles—electrical, with fuel cells, or pneumatic. In such vehicles, the engine can operate at the optimal operating point due to the load and speed. Due to the necessity for frequent engine shutdown, the VVA engine is best suited to operate in such conditions.

    The introduction of new systems of control valves, including the VVA system, changes waveforms of load, relative velocity, and temperature characterizing operation of components of the guide–valve–seat insert system. This results in changes in courses of the resistance of motion in the valve stems against guides and wear intensity for components of those systems. Operational conditions of each controlled system and the type of drive valve are specific to each system because each system has its own unique dynamics based on the algorithm used and the control and drive components. The requirements for increasing the accuracy of control algorithms for valve motion necessitate the consideration of changes in the resistance of motion between the valve stem and its guide and the introduction of their compensation.

    The use of new lightweight valves, matching seat inserts, and guides made of new materials changes the resistance of motion and wear intensity compared to those of the previously used valves made of steel. The resulting issues that arise have not been sufficiently recognized.

    One of the unresolved issues is lubrication. For camless drives, the elimination of some elements of the classic cam-driven timing changes the conditions for the supply of oil to the contact valve stem–guide. This may result in the need to increase oil pressure in the main oil circuit, resulting in more power to drive the oil pump. It may also lead to increased complexity of the oil system and increased resistance to flow because of additional channels supplying oil to bearings of valve drives. As a result, the reduction in power needed to drive the valves will be offset by the increase in power to drive the unit supplying the oil system.

    The preferable solution is to eliminate timing from the main lubrication system of the engine. This creates new tribological problems associated with organizing a new way of delivering lubricant to the contact area valve stem–guide or taking actions to prevent the reduction of valve life, despite the elimination of lubrication of moving parts in the timing.

    Then, lubrication of the contact valve stem–guide can be provided using, for example, additional oil storage tanks or self-lubricating bushings. Oil selection and design of such bushings require separate tests for each drive configuration. The best solution is to use engine oil and bushings geometry similar to the geometry of classic guides. Complete elimination of oil may be possible in engines of lower speed and power, and it requires careful association of materials for guides and valve stems.

    Weights and key dimensions, such as the maximum diameter of the valve head dg, diameter of valve stem dt, and total height hz for valves on the market that are made of steel and TiAl alloys and used in the same engines were measured. The results allow for the assumption of an approximate criterion for classifying valves as lightweight, involving the fulfillment of the following condition [5]:

    (1.1)

    Chapter 2

    Principles of valve train operation

    Abstract

    In the chapter it was explained, that the valve performance depends on the engine type: spark-ignition and compression ignition, the type and a method for delivering components needed to carry out the combustion process in the engine, especially the fuel and oxidizer. It was also mentioned engines that use variable cycles and the engines of a homogeneous charge compression ignition (HCCI). In the chapter the departure from the basic engine valve timing was described. It was also explained concepts of lead, lag and overlap. It was presented valve timing diagrams for four–stoke and two–stroke engines both of the SI and CI type. It was explained the concept of the scavenging. It was also described the rotary port system for IC engines, poppet valves’ arrangement, types of valve train systems with poppet valves. It was discussed the classification of Variable Valve Actuation Technology, the role of Variable Valve Timing, and the deactivation of cylinder and valve.

    Keywords

    engine types; lead; lag; overlap; scavenging; rotary port system; poppet valves arrangement; variable valve actuation; variable valve timing; cylinder deactivation; valve deactivation

    The operation of valve train elements occurs under conditions of the repetitive operating cycle of the engine and depends on its course and parameters. Therefore, the engine type is one of the principal determinants of valve performance. Most cases of valve trains are seen in four-stroke cycle engines, and only a small portion of cases concern two-stroke engines.

    There are two main engine types: spark ignition (SI), operating in a version of the Otto cycle, and compression ignition (CI), which operates in a version of the diesel cycle.

    The valve performance is also determined by the type and the method of delivery of the components necessary to carry out the combustion process in the engine, especially fuel and the oxidizer. Both of these and interactions between them have an effect on pressure, temperature, the course of the combustion, and the produced atmosphere in which the valves operate. In SI engines, petrol is the common fuel; however, these engines may be powered with other fuels, such as autogas (LPG), methanol, ethanol, bioethanol, compressed natural gas, hydrogen, and nitromethane [6]. In most cases, CI engines are fuelled with gas oil.

    There are also engines that use variable cycles. An example is the Ricardo engine [7], in which the low-speed range of the two-stroke cycle is used and the four-stroke cycle is used at higher speeds. This involves the need to ensure greater efficiency throughout the engine speed range. This engine enables fuel savings of 27%.

    Relatively recently, engines with a homogeneous charge compression ignition (HCCI) have been developed that are hybrids of SI engines based on CI engine processes. The HCCI engine combines the high performance of the CI engine with the low NOx and particulate matter emissions of the SI engine. In the HCCI engine, fuel and air are mixed before combustion, as in the SI engine, and compression of the mixture causes self-ignition in the same way as in the CI engine. There are various methods of HCCI ignition control: inlet air temperature control [8], variable compression ratio [9], dual fuel injection [10], variable valve timing [11], and exhaust gas recirculation [12].

    Withdrawal from the Basic Valve Timing

    As explained in Ref. [13], the opening and closing of inlet and outlet valves are timed to match the beginning and the end of the induction and exhaust strokes, respectively. In the case of a variable-speed motor vehicle engine, such an orderly approach to valve timing would result in highly inefficient operation. In practice, it is necessary to change the basic valve timing implied by the four-stroke or less than two-stroke principle.

    The change in timing can be based on the factors involved, such as the following:

    1. Inertia effects of the incoming and outgoing cylinder gases

    2. The flexible nature of incoming and outgoing cylinder gases

    3. Mechanical stresses imposed by rapidly opening and closing valves

    To accommodate the previously mentioned effects, the basic valve timing of the four-stroke principle can be modified by providing for the lead (advanced time) and lag (delay time) of the inlet and outlet valve periods of opening.

    Lead, Lag, and Overlap

    The concept of lead, lag, and overlap is explained in Ref. [13].

    The inlet valve is given a lead in opening before the piston reaches top dead center on the exhaust stroke (Fig. 2.1A) so that least resistance is offered to the incoming flow of air and petrol mixture as the piston begins its induction stroke. It is also provided with a lag in closing after the piston reaches bottom dead center and begins the compression stroke (Fig. 2.1A) so as to take advantage of the reluctance of the incoming mixture to cease flowing as the piston ends its induction stroke. The maximum amount of air and petrol mixture is therefore induced to enter the cylinder, which directly affects the power developed by the engine.

    Figure 2.1 Valve timing: (A) exhaust and compression, (B) power and induction, and (C) timing diagram.

    The outlet valve is given a lead in opening before the piston reaches bottom dead center on the power stroke (Fig. 2.1B); thus the burnt gases are already leaving the cylinder under their own pressure as the piston begins its exhaust stroke. As a result, the engine expends less energy on expelling the exhaust gases than would otherwise be the case. The outlet valve is also provided with a lag in closing after the piston reaches top dead center and begins the induction stroke (Fig. 2.1B). This better scavenges the combustion chamber of exhaust gases and lowers cylinder pressure to facilitate flow of the incoming air and petrol mixture.

    The opening of the inlet valve before top dead center on the exhaust stroke and the closing of the outlet valve after top dead center on the induction stroke result in a period during which both valves are partially or fully open. This period when the inlet valve opens before the outlet valve closes is termed the valve overlap (Fig. 2.1C).

    Valve Timing Diagrams

    The opening and closing points of the valves are often shown in the form of a valve timing diagram (Fig. 2.1C), although these data can be arranged in a table or the number of degrees before top dead center when the inlet valve begins to open can be reported [13]. For tabulating valve timing information, the commonly used abbreviations BTDC and ATDC denote before top dead center and after top dead center, respectively, and refer to the positions of the crankshaft as the piston is respectively advancing toward and retreating from the combustion chamber. Similarly, BBDC and ABDC denote before bottom dead center and after bottom dead center, respectively, and relate to the opposite sense of piston movement.

    Four-Stroke Cycle

    In four-stroke cycle engines, both SI and CI, there are four strokes completing two rotations of the crankshaft. These are respectively the suction or charging, compression, power/work or expansion, and exhaust strokes. The important variable characterizing operational conditions in each engine is the brake mean effective pressure (bMEP), which is the mean effective pressure calculated from measured brake torque. It is defined by Eq. (2.1):

    (2.1)

    where

    iMEP is the indicated mean effective pressure, which is the mean effective pressure calculated from in-cylinder pressure—the average in-cylinder pressure over the engine cycle (720° in a four-stroke and 360° in a two-stroke). Direct iMEP measurement requires combustion pressure-sensing equipment.

    pMEP is the pumping mean effective pressure, which is the mean effective pressure calculated from work moving air in and out of the cylinder due to inlet throttling losses and residual gases in outlet.

    fMEP is the friction mean effective pressure, which is the theoretical mean effective pressure required to overcome engine friction. It can be thought of as mean effective pressure lost due to friction.

    Mean effective pressure is correlated with the peak pressure of gas in engine cylinders; however, such dependency is highly nonlinear and can be obtained from the measurement for a narrow class of engines or estimated using simulation models, which are very complex.

    The mean effective pressure and peak pressure affect the force loading the seat faces, the valves, and their inserts, which determines friction between seat faces and their wear rate.

    According to Ref. [14], for naturally aspirated SI engines, the maximum bMEP is within the range 850–1050 kPa, at speed at which maximum torque is obtained. At rated power, bMEP values are 10–15% lower. For boosted SI engines, the maximum bMEP falls within the range from 1.25 to 1.7 MPa. For four-stroke CI engines, the maximum bMEP is within the range 700–900 kPa for the naturally aspirated and 1.4–1.8 MPa for the boosted, respectively.

    Four-Stroke Cycle SI Engine

    In the SI engine, ignition is induced by sparks generated by spark plugs, where the operation cycle is adjusted to the engine speed and load using mechanical or computer-controlled ignition systems. Such adjustment is directly related to TDC positions and thus indirectly to the valve timing. In the SI engine, fuel is mixed with air, broken up into a mist, and partially vaporized. The compression ratio varies from 4:1 to 8:1, and the air–fuel mixture ratio varies from 10:1 to 20:1.

    The four strokes of a petrol engine sucking fuel–air mixture are shown in Fig. 2.2.

    Figure 2.2 The four strokes of an SI engine.

    Four-Stroke Cycle CI Engine

    In the CI engine, ignition takes place due to the heat produced in the engine cylinder at the end of the compression stroke. The four strokes of a CI engine sucking pure air are shown in Fig. 2.3. The compression ratio varies from 14:1 to 22:1. The pressure at the end of the compression stroke ranges from 30 to 45 kg/cm². The temperature near the end of the compression stroke is 650–800°C.

    Figure 2.3 The four strokes of a CI engine.

    The typical timing diagram for a CI engine is shown in Fig. 2.4.

    Figure 2.4 The timing diagram for a CI engine. From Ref. [15].

    Two-Stroke Engines

    Simple two-stroke SI engines fueled with a gasoline–oil mixture are often used in high-power, handheld applications, such as string trimmers and chainsaws. Such engines are preferred for small, portable, or specialized machines such as outboard motors, high-performance, small-capacity motorcycles, mopeds, underbones, scooters, tuk-tuks, snowmobiles, karts, ultralights, model airplanes and model vehicles, lawnmowers, chainsaws, weed trimmers, and dirt bikes.

    The two-stroke cycle is also used in CI large industrial and marine engines and in CI engines of some trucks and heavy machinery.

    Many modern two-stroke engines employ a power valve system to achieve better low-speed power without sacrificing high-speed power. The valves are normally in or around the exhaust ports. They operate in one of two modes:

    • By altering the exhaust port by way of closing off the top part of the port, which alters port timing, such as in Ski-Doo R.A.V.E., Yamaha YPVS, Honda RC-Valve, Kawasaki K.I.P.S., Cagiva C.T.S., or Suzuki AETC systems

    • By altering the volume of the exhaust, which changes the resonant frequency of the expansion chamber, such as in the Suzuki SAEC and Honda V-TACS systems

    In the case of CI loop-scavenged engines, intake and exhaust occur via piston-controlled ports. A uniflow CI engine takes in air via scavenge ports, whereas exhaust gases exit through an overhead poppet valve. Two-stroke CI engines are all scavenged by forced induction. Some designs use a mechanically driven Roots blower, whereas marine diesel engines normally use exhaust-driven turbochargers with power-driven auxiliary blowers for low-speed operation when exhaust turbochargers are unable to deliver enough air.

    Marine two-stroke CI engines directly coupled to the propeller can start and run in either direction as required. Fuel injection and valve timing are mechanically readjusted by using a different set of cams on the camshaft. Thus, the engine can be run in reverse.

    The two-stroke engine can use variable outlet valve closing; this is easily achieved with an electronically controlled camshaft-less engine or by involving hydraulic valves as in the case of the modified Sulzer RTA.

    Two-stroke CI engines have bMEP values similar to those of four-stroke CI engines. Very large low-speed CI engines such as the Wärtsilä-Sulzer RTA96-C run at bMEPs of up to 1.9 MPa [14].

    Two-Stroke Cycle SI Engine

    Port timing

    The two-stroke spark-ignition engine completes the cycle of actions—induction, compression, power, and exhaust—in one rotation of the crankshaft or two complete piston strokes. The port timing must take into account the time lapse before the ports are either fully uncovered or fully covered and also the inertia effects of the incoming and outgoing flows of the crankcase and cylinder gases. Some compromise is inevitable because the ports are necessarily uncovered and covered by the piston at equal angles on either side of the crankshaft dead centers. The port timing diagram of a conventional two-stroke engine is symmetrical.

    Scavenging

    The burnt gases in the IC engine cylinder are not completely exhausted before the suction stroke. A portion of the gases remain inside the cylinder and mix with the fresh charge, which gets diluted and its strength is reduced. The process of removing burnt gases from the combustion chamber of the engine cylinder is known as scavenging.

    In a four-stroke cycle engine, scavenging is very effective because during the exhaust stroke the piston pushes out the burnt gases from the engine cylinder. Note that a small quantity of burnt gases remains in the engine cylinder in the clearance space.

    In a two-stroke cycle engine, scavenging is less effective because the exhaust port is open for a small fraction of the crank rotation. As the transfer and exhaust port arc open simultaneously during a part of the crank rotation, the fresh charge also escapes out along with the burnt gases. This is overcome by way of designing the piston crown of a particular shape.

    Specific engine output is largely determined by the efficiency of the scavenging system and is directly related to bMEP. Scavenging efficiency varies with the delivery ratio and the type of scavenging. In this respect, cross-scavenging is the least efficient and produces the lowest bMEP. This is because the scavenging air flows through the cylinder but does not expel the exhaust residual gases effectively. The loop scavenging method is better than the cross-flow scavenging method. Even with a delivery ratio of 1.0, in all cases the scavenging efficiencies are approximately 53%, 67%, and 80% for cross-, loop, and uniflow scavenging systems with corresponding values of bMEPs of 3.5, 4.5, and 5.8 bar.

    The delivery ratio Rdel (Eq. 2.2) compares the actual scavenging air mass (or mixture mass) to that required in an ideal charging process [17]:

    (2.2)

    The scavenging efficiency ηsc (Eq. 2.3) indicates to what extent residual gases in the cylinder have been replaced with fresh air [17]:

    (2.3)

    Two-Stroke Cycle CI Engine

    The two-stroke cycle engine is sometimes called the Clerk engine. Uniflow scavenging occurs with fresh charge entering the combustion chamber above the piston while the exhaust outflow goes through ports uncovered by the piston at its outermost position.

    Low- and medium-speed two-stroke marine CI engines continue to use this system, but high-speed two-stroke CI engines reverse the scavenging flow by blowing fresh charge through the bottom inlet ports, sweeping up through the cylinder and out of the exhaust ports in the cylinder head. The three characteristic phases of the two-stroke CI engine are [16]: scavenging phase, compression phase, and power phase.

    With the two-stroke diesel engine, intake and exhaust phases take place during part of the compression and power stroke, respectively, so that a cycle of operation is completed in one crankshaft rotation or two piston strokes. Because there are no separate intake and exhaust strokes, a blower is necessary to pump air into the cylinder to push out exhaust gases and to supply the cylinder with fresh air for combustion.

    Rotary Port System for IC Engines

    To decrease pumping losses and hence increase overall engine output, the reciprocating motion of the most common poppet valve system can be replaced with a rotational porting system [18]. The main design advantage of the latter is the absence of reciprocating components and thus the elimination of springs. It can create far less noise (vibrations) in the system. Such a porting system also can be easily adapted to current engine configurations because its thermal cycle is theoretically identical to that of a poppet valve system.

    A horizontal-type solution was introduced by Coates International [19–21]. Instead of using cylindrical shafts, spherical lobes with discrete passageways were used to ensure fluid passage into and out of the cylinders.

    Another horizontal-type solution is a cross-type rotary port valve system elaborated by Mercedes-Ilmor and Bishop for their Formula V10 engine [22]. The bore tolerances are calculated so that inherent mechanical and thermal distortions of the valves never interfere with the cylinder bore lining. Close tolerance between the shaft and the bore prevents leakage between the exhaust and the intake port.

    The cross/Bishop–type valve allows for continuous feed or exhaust via lateral positioned ports that remain open through the entire rotation of the valve [18]. In contrast, the Ritter/Coates–type valves have

    Enjoying the preview?
    Page 1 of 1