ullman-38162
ull75741_fm
December 18, 2008
16:19
The Mechanical Design Process
ullman-38162
ull75741_fm
December 18, 2008
16:19
McGraw-Hill Series in Mechanical Engineering
Alciatore/Histand Introduction to Mechatronics and Measurement System Anderson Fundamentals of Aerodynamics Anderson Introduction to Flight Anderson Modern Compressible Flow Barber Intermediate Mechanics of Materials Beer/Johnston Vector Mechanics for Engineers Beer/Johnston Mechanics of Materials Budynas Advanced Strength and Applied Stress Analysis Budynas/Nisbett Shigley’s Mechanical Engineering Design Cengel Heat Transfer: A Practical Approach Cengel Introduction to Thermodynamics & Heat Transfer Cengel/Boles Thermodynamics: An Engineering Approach Cengel/Clmbala Fluid Mechanics: Fundamentals and Applications Cengel/Turner Fundamentals of Thermal-Fluid Sciences Dieter Engineering Design: A Materials & Processing Approach Doebelin Measurement Systems: Application & Design Dorl/Byers Technology Ventures: From Idea to Enterprise Dunn Measurement & Data Analysis for Engineering and Science Fianemore/Franzial Fluid Mechanics with Engineering Applications Hamrock/Schmid/Jacobson Fundamentals of Machine Elements
Heywood Internal Combustion Engine Fundamentals Holman Experimental Methods for Engineers Holman Heat Transfer Hutton Fundamental of Finite Element Analysis Kays/Crawford/Welgand Convective Heat and Mass Transfer Meirovioeh Fundamentals of Vibrations Norton Design of Machinery Palm System Dynamics Reddy An Introduction to Finite Element Method Schey Introduction to Manufacturing Processes Shames Mechanics of Fluids Smith/Hashemi Foundations of Materials Science & Engineering Turns An Introduction to Combustion: Concepts and Applications Ugural Mechanical Design: An Integrated Approach Ullman The Mechanical Design Process White Fluid Mechanics White Viscous Fluid Flow Zeid CAD/CAM Theory and Practice Zeid Mastering CAD/CAM
ullman-38162
ull75741_fm
December 18, 2008
16:19
The Mechanical Design Process Fourth Edition
David G. Ullman Professor Emeritus, Oregon State University
ullman-38162
ull75741_FM
December 30, 2008
9:25
THE MECHANICAL DESIGN PROCESS, FOURTH EDITION Published by McGraw-Hill, a business unit of The McGraw-Hill Companies, Inc., 1221 Avenue of the Americas, New York, NY 10020. Copyright © 2010 by The McGraw-Hill Companies, Inc. All rights reserved. Previous editions © 2003, 1997, and 1992. No part of this publication may be reproduced or distributed in any form or by any means, or stored in a database or retrieval system, without the prior written consent of The McGraw-Hill Companies, Inc., including, but not limited to, in any network or other electronic storage or transmission, or broadcast for distance learning. Some ancillaries, including electronic and print components, may not be available to customers outside the United States. This book is printed on acid-free paper. 1 2 3 4 5 6 7 8 9 0 DOC/DOC 0 9 ISBN 978–0–07–297574–1 MHID 0–07–297574–1 Global Publisher: Raghothaman Srinivasan Senior Sponsoring Editor: Bill Stenquist Director of Development: Kristine Tibbetts Senior Marketing Manager: Curt Reynolds Senior Project Manager: Kay J. Brimeyer Senior Production Supervisor: Sherry L. Kane Lead Media Project Manager: Stacy A. Patch Associate Design Coordinator: Brenda A. Rolwes Cover Designer: Studio Montage, St. Louis, Missouri Cover Image: Irwin clamp: © Irwin Industrial Tools; Marin bike: © Marin Bicycles; MER: © NASA/JPL. Senior Photo Research Coordinator: John C. Leland Compositor: S4Carlisle Publishing Services Typeface: 10.5/12 Times Roman Printer: R. R. Donnelley Crawfordsville, IN Library of Congress Cataloging-in-Publication Data Ullman, David G., 1944The mechanical design process / David G. Ullman.—4th ed. p. cm.—(McGraw-Hill series in mechanical engineering) Includes index. ISBN 978–0–07–297574–1—ISBN 0–07–297574–1 (alk. paper) 1. Machine design. I. Title. TJ230.U54 2010 621.80002 15—dc22 www.mhhe.com
2008049434
ullman-38162
ull75741_fm
December 18, 2008
16:19
ABOUT THE AUTHOR
David G. Ullman is an active product designer who has taught, researched, and
written about design for over thirty years. He is president of Robust Decisions, Inc., a supplier of software products and training for product development and decision support. He is Emeritus Professor of Mechanical Design at Oregon State University. He has professionally designed fluid/thermal, control, and transportation systems. He has published over twenty papers focused on understanding the mechanical product design process and the development of tools to support it. He is founder of the American Society Mechanical Engineers (ASME)—Design Theory and Methodology Committee and is a Fellow in the ASME. He holds a Ph.D. in Mechanical Engineering from the Ohio State University.
ullman-38162
ull75741_fm
December 18, 2008
16:19
ullman-38162
ull75741_fm
December 18, 2008
16:19
CONTENTS
Preface
CHAPTER
xi
1
2.8 Sources 44 2.9 Exercises 45 2.10 On the Web 45
Why Study the Design Process? 1 1.1 1.2
Introduction 1 Measuring the Design Process with Product Cost, Quality, and Time to Market 3 1.3 The History of the Design Process 8 1.4 The Life of a Product 10 1.5 The Many Solutions for Design Problems 15 1.6 The Basic Actions of Problem Solving 17 1.7 Knowledge and Learning During Design 19 1.8 Design for Sustainability 20 1.9 Summary 21 1.10 Sources 22 1.11 Exercises 22 CHAPTER
2.3 2.4 2.5 2.6 2.7
3
Designers and Design Teams 47 3.1 3.2
Introduction 47 The Individual Designer: A Model of Human Information Processing 48 3.3 Mental Processes That Occur During Design 56 3.4 Characteristics of Creators 64 3.5 The Structure of Design Teams 66 3.6 Building Design Team Performance 72 3.7 Summary 78 3.8 Sources 78 3.9 Exercises 79 3.10 On the Web 80
2
Understanding Mechanical Design 25 2.1 2.2
CHAPTER
Introduction 25 Importance of Product Function, Behavior, and Performance 28 Mechanical Design Languages and Abstraction 30 Different Types of Mechanical Design Problems 33 Constraints, Goals, and Design Decisions 40 Product Decomposition 41 Summary 44
CHAPTER
4
The Design Process and Product Discovery 81 4.1 4.2 4.3 4.4 4.5 4.6 4.7 4.8 4.9
Introduction 81 Overview of the Design Process 81 Designing Quality into Products 92 Product Discovery 95 Choosing a Project 101 Summary 109 Sources 110 Exercises 110 On the Web 110
vii
ullman-38162
ull75741_fm
December 18, 2008
16:19
Contents
viii
CHAPTER
5
6.9
Planning for Design 111 5.1 5.2 5.3
Introduction 111 Types of Project Plans 113 Planning for Deliverables— The Development of Information 5.4 Building a Plan 126 5.5 Design Plan Examples 134 5.6 Communication During the Design Process 137 5.7 Summary 141 5.8 Sources 141 5.9 Exercises 142 5.10 On the Web 142 CHAPTER
117
6
Understanding the Problem and the Development of Engineering Specifications 143 6.1 6.2 6.3
6.4 6.5 6.6
6.7
6.8
Introduction 143 Step 1: Identify the Customers: Who Are They? 151 Step 2: Determine the Customers’ Requirements: What Do the Customers Want? 151 Step 3: Determine Relative Importance of the Requirements: Who Versus What 155 Step 4: Identify and Evaluate the Competition: How Satisfied Are the Customers Now ? 157 Step 5: Generate Engineering Specifications: How Will the Customers’ Requirement Be Met? 158 Step 6: Relate Customers’ Requirements to Engineering Specifications: How to Measure What? 163 Step 7: Set Engineering Specification Targets and Importance: How Much Is Good Enough? 164
6.10 6.11 6.12 6.13 6.14
Step 8: Identify Relationships Between Engineering Specifications: How Are the Hows Dependent on Each Other? 166 Further Comments on QFD 168 Summary 169 Sources 169 Exercises 169 On the Web 170
CHAPTER
7
Concept Generation 171 7.1 7.2 7.3 7.4 7.5 7.6 7.7 7.8 7.9 7.10 7.11 7.12 7.13
Introduction 171 Understanding the Function of Existing Devices 176 A Technique for Designing with Function 181 Basic Methods of Generating Concepts 189 Patents as a Source of Ideas 194 Using Contradictions to Generate Ideas 197 The Theory of Inventive Machines, TRIZ 201 Building a Morphology 204 Other Important Concerns During Concept Generation 208 Summary 209 Sources 209 Exercises 211 On the Web 211
CHAPTER
8
Concept Evaluation and Selection 213 8.1 8.2 8.3 8.4 8.5 8.6
Introduction 213 Concept Evaluation Information 215 Feasibility Evaluations 218 Technology Readiness 219 The Decision Matrix—Pugh’s Method 221 Product, Project, and Decision Risk 226
ullman-38162
ull75741_fm
December 18, 2008
16:19
Contents
8.7 8.8 8.9 8.10 8.11
Robust Decision Making Summary 239 Sources 239 Exercises 240 On the Web 240
CHAPTER
233
9
Product Generation 241 9.1 9.2 9.3 9.4 9.5 9.6
Introduction 241 BOMs 245 Form Generation 246 Materials and Process Selection 264 Vendor Development 266 Generating a Suspension Design for the Marin 2008 Mount Vision Pro Bicycle 269 9.7 Summary 276 9.8 Sources 276 9.9 Exercises 277 9.10 On the Web 278 CHAPTER
10
Product Evaluation for Performance and the Effects of Variation 279 10.1 10.2 10.3 10.4 10.5 10.6 10.7 10.8 10.9 10.10 10.11
Introduction 279 Monitoring Functional Change 280 The Goals of Performance Evaluation 281 Trade-Off Management 284 Accuracy, Variation, and Noise 286 Modeling for Performance Evaluation 292 Tolerance Analysis 296 Sensitivity Analysis 302 Robust Design by Analysis 305 Robust Design Through Testing 308 Summary 313
10.12 Sources 313 10.13 Exercises 314
CHAPTER
11
Product Evaluation: Design For Cost, Manufacture, Assembly, and Other Measures 315 11.1 11.2 11.3 11.4 11.5
Introduction 315 DFC—Design For Cost 315 DFV—Design For Value 325 DFM—Design For Manufacture 328 DFA—Design-For-Assembly Evaluation 329 11.6 DFR—Design For Reliability 350 11.7 DFT and DFM—Design For Test and Maintenance 357 11.8 DFE—Design For the Environment 358 11.9 Summary 360 11.10 Sources 361 11.11 Exercises 361 11.12 On the Web 362
CHAPTER
12
Wrapping Up the Design Process and Supporting the Product 363 12.1 12.2 12.3 12.4 12.5 12.6 12.7 12.8
Introduction 363 Design Documentation and Communication 366 Support 368 Engineering Changes 370 Patent Applications 371 Design for End of Life 375 Sources 378 On the Web 378
ix
ullman-38162
ull75741_fm
16:19
Contents
x
APPENDIX
December 18, 2008
A
Properties of 25 Materials Most Commonly Used in Mechanical Design 379 A.1 Introduction 379 A.2 Properties of the Most Commonly Used Materials 380 A.3 Materials Used in Common Items 393 A.4 Sources 394 APPENDIX
B
Normal Probability 397 B.1 Introduction 397 B.2 Other Measures 401 APPENDIX
APPENDIX
D
Human Factors in Design 415 D.1 D.2 D.3 D.4
Introduction 415 The Human in the Workspace 416 The Human as Source of Power 419 The Human as Sensor and Controller 419 D.5 Sources 426
Index 427
C
The Factor of Safety as a Design Variable 403 C.1 Introduction
C.2 The Classical Rule-of-Thumb Factor of Safety 405 C.3 The Statistical, Reliability-Based, Factor of Safety 406 C.4 Sources 414
403
ullman-38162
ull75741_fm
December 18, 2008
16:19
PREFACE
have been a designer all my life. I have designed bicycles, medical equipment, furniture, and sculpture, both static and dynamic. Designing objects has come easy for me. I have been fortunate in having whatever talents are necessary to be a successful designer. However, after a number of years of teaching mechanical design courses, I came to the realization that I didn’t know how to teach what I knew so well. I could show students examples of good-quality design and poorquality design. I could give them case histories of designers in action. I could suggest design ideas. But I could not tell them what to do to solve a design problem. Additionally, I realized from talking with other mechanical design teachers that I was not alone. This situation reminded me of an experience I had once had on ice skates. As a novice skater I could stand up and go forward, lamely. A friend (a teacher by trade) could easily skate forward and backward as well. He had been skating since he was a young boy, and it was second nature to him. One day while we were skating together, I asked him to teach me how to skate backward. He said it was easy, told me to watch, and skated off backward. But when I tried to do what he did, I immediately fell down. As he helped me up, I asked him to tell me exactly what to do, not just show me. After a moment’s thought, he concluded that he couldn’t actually describe the feat to me. I still can’t skate backward, and I suppose he still can’t explain the skills involved in skating backward. The frustration that I felt falling down as my friend skated with ease must have been the same emotion felt by my design students when I failed to tell them exactly what to do to solve a design problem. This realization led me to study the process of mechanical design, and it eventually led to this book. Part has been original research, part studying U.S. industry, part studying foreign design techniques, and part trying different teaching approaches on design classes. I came to four basic conclusions about mechanical design as a result of these studies:
I
1. The only way to learn about design is to do design. 2. In engineering design, the designer uses three types of knowledge: knowledge to generate ideas, knowledge to evaluate ideas and make decisions, and knowledge to structure the design process. Idea generation comes from experience and natural ability. Idea evaluation comes partially from experience and partially from formal training, and is the focus of most engineering education. Generative and evaluative knowledge are forms of domain-specific knowledge. Knowledge about the design process and decision making is largely independent of domain-specific knowledge. 3. A design process that results in a quality product can be learned, provided there is enough ability and experience to generate ideas and enough experience and training to evaluate them. xi
ullman-38162
xii
ull75741_fm
December 18, 2008
16:19
Preface
4. A design process should be learned in a dual setting: in an academic environment and, at the same time, in an environment that simulates industrial realities. I have incorporated these concepts into this book, which is organized so that readers can learn about the design process at the same time they are developing a product. Chaps. 1–3 present background on mechanical design, define the terms that are basic to the study of the design process, and discuss the human element of product design. Chaps. 4–12, the body of the book, present a step-by-step development of a design method that leads the reader from the realization that there is a design problem to a solution ready for manufacture and assembly. This material is presented in a manner independent of the exact problem being solved. The techniques discussed are used in industry, and their names have become buzzwords in mechanical design: quality function deployment, decision-making methods, concurrent engineering, design for assembly, and Taguchi’s method for robust design. These techniques have all been brought together in this book. Although they are presented sequentially as step-by-step methods, the overall process is highly iterative, and the steps are merely a guide to be used when needed. As mentioned earlier, domain knowledge is somewhat distinct from process knowledge. Because of this independence, a successful product can result from the design process regardless of the knowledge of the designer or the type of design problem. Even students at the freshman level could take a course using this text and learn most of the process. However, to produce any reasonably realistic design, substantial domain knowledge is required, and it is assumed throughout the book that the reader has a background in basic engineering science, material science, manufacturing processes, and engineering economics. Thus, this book is intended for upper-level undergraduate students, graduate students, and professional engineers who have never had a formal course in the mechanical design process.
ADDITIONS TO THE FOURTH EDITION Knowledge about the design process is increasing rapidly. A goal in writing the fourth edition was to incorporate this knowledge into the unified structure—one of the strong points of the first three editions. Throughout the new edition, topics have been updated and integrated with other best practices in the book. Some specific additions to the new edition include: 1. Improved material to ensure team success. 2. Over twenty blank templates are available for download from the book’s website (www.mhhe.com/ullman4e) to support activities throughout the design process. The text includes many of them filled out for student reference. 3. Improved material on project planning.
ullman-38162
ull75741_fm
December 18, 2008
16:19
Preface
4. Improved sections on Design for the Environment and Design for Sustainability. 5. Improved material on making design decisions. 6. A new section on using contradictions to generate ideas. 7. New examples from the industry, with new photos and diagrams to illustrate the examples throughout. Beyond these, many small changes have been made to keep the book current and useful.
ELECTRONIC TEXTBOOK CourseSmart is a new way for faculty to find and review eTextbooks. It’s also a great option for students who are interested in accessing their course materials digitally and saving money. CourseSmart offers thousands of the most commonly adopted textbooks across hundreds of courses from a wide variety of higher education publishers. It is the only place for faculty to review and compare the full text of a textbook online, providing immediate access without the environmental impact of requesting a print exam copy. At CourseSmart, students can save up to 50% off the cost of a print book, reduce their impact on the environment, and gain access to powerful Web tools for learning including full text search, notes and highlighting, and email tools for sharing notes between classmates. www.CourseSmart.com
ACKNOWLEDGMENTS I would like to thank these reviewers for their helpful comments: Patricia Brackin, Rose-Hulman Institute of Technology William Callen, Georgia Institute of Technology Xiaoping Du, University of Missouri-Rolla Ian Grosse, University of Massachusetts–Amherst Karl-Heinrich Grote, Otto-von-Guericke University, Magdeburg, Germany Mica Grujicic, Clemson University John Halloran, University of Michigan Peter Jones, Auburn University Mary Kasarda, Virginia Technical College Jesa Kreiner, California State University–Fullerton Yuyi Lin, University of Missouri–Columbia Ron Lumia, University of New Mexico Spencer Magleby, Brigham Young University Lorin Maletsky, University of Kansas
xiii
ullman-38162
xiv
ull75741_fm
December 18, 2008
16:19
Preface
Make McDermott, Texas A&M University Joel Ness, University of North Dakota Charles Pezeshki, Washington State University John Renaud, University of Notre Dame Keith Rouch, University of Kentucky Ali Sadegh, The City College of The City University of New York Shin-Min Song, Northern Illinois University Mark Steiner, Rensselaer Polytechnic Institute Joshua Summers, Clemson University Meenakshi Sundaram, Tennessee Technical University Shih-Hsi Tong, University of California–Los Angeles Kristin Wood, University of Texas Additionally, I would like to thank Bill Stenquist, senior sponsoring editor for mechanical engineering of McGraw-Hill, Robin Reed, developmental editor, Kay Brimeyer, project manager, and Lynn Steines, project editor, for their interest and encouragement in this project. Also, thanks to the following who helped with examples in the book: Wayne Collier, UGS Jason Faircloth, Marin Bicycles Marci Lackovic, Autodesk Samir Mesihovic, Volvo Trucks Professor Bob Paasch, Oregon State University Matt Popik, Irwin Tools Cary Rogers, GE Medical Professor Tim Simpson, Penn State University Ralf Strauss, Irwin Tools Christopher Voorhees, Jet Propulsion Laboratory Professor Joe Zaworski, Oregon State University Last and most important my thanks to my wife, Adele, for her never questioning confidence that I could finish this project.
ullman-38162
ull75741_01
December 17, 2008
11:15
C
H
1 A
P
T
E
R
Why Study the Design Process? KEY QUESTIONS ■ ■ ■ ■ ■ ■
What can be done to design quality mechanical products on time and within budget? What are the ten key features of design best practice that will lead to better products? What are the phases of a product’s life cycle? How are design problems different from analysis problems? Why is it during design, the more you know, the less design freedom you have? What are the Hanover Principles?
1.1 INTRODUCTION Beginning with the simple potter’s wheel and evolving to complex consumer products and transportation systems, humans have been designing mechanical objects for nearly five thousand years. Each of these objects is the end result of a long and often difficult design process. This book is about that process. Regardless of whether we are designing gearboxes, heat exchangers, satellites, or doorknobs, there are certain techniques that can be used during the design process to help ensure successful results. Since this book is about the process of mechanical design, it focuses not on the design of any one type of object but on techniques that apply to the design of all types of mechanical objects. If people have been designing for five thousand years and there are literally millions of mechanical objects that work and work well, why study the design process? The answer, simply put, is that there is a continuous need for new, cost-effective, high-quality products. Today’s products have become so complex that most require a team of people from diverse areas of expertise to develop an idea into hardware. The more people involved in a project, the greater is the need for assistance in communication and structure to ensure nothing important 1
ullman-38162
ull75741_01
December 17, 2008
2
CHAPTER 1
11:15
Why Study the Design Process?
is overlooked and customers will be satisfied. In addition, the global marketplace has fostered the need to develop new products at a very rapid and accelerating pace. To compete in this market, a company must be very efficient in the design of its products. It is the process that will be studied here that determines the efficiency of new product development. Finally, it has been estimated that 85% of the problems with new products not working as they should, taking too long to bring to market, or costing too much are the result of a poor design process. The goal of this book is to give you the tools to develop an efficient design process regardless of the product being developed. In this chapter the important features of design problems and the processes for solving them will be introduced. These features apply to any type of design problem, whether for mechanical, electrical, software, or construction projects. Subsequent chapters will focus more on mechanical design, but even these can be applied to a broader range of problems. Consider the important factors that determine the success or failure of a product (Fig. 1.1). These factors are organized into three ovals representing those factors important to product design, business, and production. Product design factors focus on the product’s function, which is a description of what the object does. The importance of function to the designer is a major topic of this book. Related to the function are the product’s form, materials, and manufacturing processes. Form includes the product’s architecture, its shape, its color, its texture, and other factors relating to its structure. Of equal importance to form are the materials and manufacturing processes used to produce the product. These four variables—function, form, materials, and manufacturing processes—
Business Target market
Promotion
Sales forecast Product form
Distribution coverage
Price
Product function Manufacturing processes
Materials
Cost/risk Production system
Product design
Production planning/ sourcing
Facilities Production
Figure 1.1 Controllable variables in product development.
ullman-38162
ull75741_01
1.2
December 17, 2008
11:15
Measuring the Design Process with Product Cost, Quality, and Time to Market
are of major concern to the designer. This product design oval is further refined in Fig. 9.3. The product form and function is also important to the business because the customers in the target market judge a product primarily on what it does (its function) and how it looks (its form). The target market is one factor important to the business, as shown in Fig. 1.1. The goal of a business is to make money— to meet its sales forecasts. Sales are also affected by the company’s ability to promote the product, distribute the product, and price the product, as shown in Fig. 1.1. The business is dependent not only on the product form and function, but also on the company’s ability to produce the product. As shown in the production oval in Fig. 1.1, the production system is the central factor. Notice how product design and production are both concerned with manufacturing processes. The choice of form and materials that give the product function affects the manufacturing processes that can be used. These processes, in turn, affect the cost and hence the price of the product. This is just one example of how intertwined product design, production, and businesses truly are. In this book we focus on the product design oval. But, we will also pay much attention to the business and production variables that are related to design. As shown in the upcoming sections, the design process has a great effect on product cost, quality, and time to market.
1.2 MEASURING THE DESIGN PROCESS WITH PRODUCT COST, QUALITY, AND TIME TO MARKET The three measures of the effectiveness of the design process are product cost, quality, and time to market. Regardless of the product being designed—whether it is an entire system, some small subpart of a larger product, or just a small change in an existing product—the customer and management always want it cheaper (lower cost), better (higher quality), and faster (less time). The actual cost of designing a product is usually a small part of the manufacturing cost of a product, as can be seen in Fig. 1.2, which is based on data from Ford Motor Company. The data show that only 5% of the manufacturing cost of a car (the cost to produce the car but not to distribute or sell it) is for design activities that were needed to develop it. This number varies with industry and product, but for most products the cost of design is a small part of the manufacturing cost. However, the effect of the quality of the design on the manufacturing cost is much greater than 5%. This is most accurately shown from the results of a detailed study of 18 different automatic coffeemakers. Each coffeemaker had the same function—to make coffee. The results of this study are shown in Fig. 1.3. Here the effects of changes in manufacturing efficiency, such as material cost, labor wages, and cost of equipment, have been separated from the effects of the design process. Note that manufacturing efficiency and design have about the same influence on the cost of manufacturing a product. The figure shows that
3
ullman-38162
4
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
15% Labor
30% Overhead
Design 5%
50% Material
Figure 1.2 Design cost as fraction of manufacturing cost.
$4.98 Good design Efficient manufacturing
$9.72 Good design Inefficient manufacturing
$8.17 Average design Average manufacturing
$8.06 Poor design Efficient manufacturing
$14.34 Poor design Inefficient manufacturing
Figure 1.3 The effect of design on manufacturing cost. (Source: Data reduced from “Assessing the Importance of Design through Product Archaeology,” Management Science, Vol. 44, No. 3, pp. 352–369, March 1998, by K. Ulrich and S. A. Pearson.)
Designers cost little, their impact on product cost, great. good design, regardless of manufacturing efficiency, cuts the cost by about 35%. In some industries this effect is as high as 75%. Thus, comparing Fig. 1.2 to Fig. 1.3, we can conclude that the decisions made during the design process have a great effect on the cost of a product but cost very little. Design decisions directly determine the materials used, the goods
ull75741_01
December 17, 2008
1.2
11:15
Measuring the Design Process with Product Cost, Quality, and Time to Market
Product cost is committed early in the design process and spent late in the process.
80
mitted
Conceptual design
100
Specification development
purchased, the parts, the shape of those parts, the product sold, the price of the product, and the sales. Another example of the relationship of the design process to cost comes from Xerox. In the 1960s and early 1970s, Xerox controlled the copier market. However, by 1980 there were over 40 different manufacturers of copiers in the marketplace and Xerox’s share of the market had fallen significantly. Part of the problem was the cost of Xerox’s products. In fact, in 1980 Xerox realized that some producers were able to sell a copier for less than Xerox was able to manufacture one of similar functionality. In one study of the problem, Xerox focused on the cost of individual parts. Comparing plastic parts from their machines and ones that performed a similar function in Japanese and European machines, they found that Japanese firms could produce a part for 50% less than American or European firms. Xerox attributed the cost difference to three factors: materials costs were 10% less in Japan, tooling and processing costs were 15% less, and the remaining 25% (half of the difference) was attributable to how the parts were designed. Not only is much of the product cost committed during the design process, it is committed early in the design process. As shown in Fig. 1.4, about 75% of the manufacturing cost of a typical product is committed by the end of the conceptual phase process. This means that decisions made after this time can influence only 25% of the product’s manufacturing cost. Also shown in the figure is the amount of cost incurred, which is the amount of money spent on the design of the product.
Cost com
60
40
Product design
Percentage of product cost committed
ullman-38162
Cost incurred
20
0
Time
Figure 1.4 Manufacturing cost commitment during design.
5
ullman-38162
6
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
Table 1.1 What determines quality
Works as it should Lasts a long time Is easy to maintain Looks attractive Incorporates latest technology/features
1989
2002
4.99 (1) 4.75 (2) 4.65 (3) 2.95 (4–5) 2.95 (4–5)
4.58 (1) 3.93 (5) 3.29 (5) 3.58 (3–4) 3.58 (3–4)
Scale: 5 = very important, 1 = not important at all, brackets denote rank. Sources: Based on a survey of consumers published in Time, Nov. 13, 1989, and a survey based on quality professional, R. Sebastianelli and N. Tamimi, “How Product Quality Dimensions Relate to Defining Quality,” International Journal of Quality and Reliability Management, Vol. 19, No. 4, pp. 442–453, 2002.
It is not until money is committed for production that large amounts of capital are spent. The results of the design process also have a great effect on product quality. In a survey taken in 1989, American consumers were asked, “What determines quality?” Their responses, shown in Table 1.1, indicate that “quality” is a composite of factors that are the responsibility of the design engineer. In a 2002 survey of engineers responsible for quality, what is important to “quality” is little changed. Although the surveys were of different groups, it is interesting to note that in the thirteen years between surveys, the importance of being easy to maintain has dropped, but the main measures of quality have remained unchanged. Note that the most important quality measure is “works as it should.” This, and “incorporates latest technology/features,” are both measures of product function. “Lasts a long time” and most of the other quality measures are dependent on the form designed and on the materials and the manufacturing process selected. What is evident is that the decisions made during the design process determine the product’s quality. Besides affecting cost and quality, the design process also affects the time it takes to produce a new product. Consider Fig. 1.5, which shows the number of design changes made by two automobile companies with different design philosophies. The data points for Company B are actual for a U.S. automobile manufacturer, and the dashed line for Company A is what is typical for Toyota. Iteration, or change, is an essential part of the design process. However, changes occurring late in the design process are more expensive than those occurring earlier, as prior work is scrapped. The curve for Company B shows that the company was still making changes after the design had been released for production. In fact, over 35% of the cost of the product occurred after it was in production. In essence, Company B was still designing the automobile as it was being sold as a product. This causes tooling and assembly-line changes during production and the possibility of recalling cars for retrofit, both of which would necessitate significant expense, to say nothing about the loss of customer confidence. Company A, on the other hand, made many changes early in the design process and finished the design of the car before it went into production. Early design changes require more engineering time and effort but do not require changes in hardware or documentation. A change that would cost $1000 in engineering time if made
1.2
December 17, 2008
11:15
Measuring the Design Process with Product Cost, Quality, and Time to Market
Start production
ull75741_01
Changes
ullman-38162
Company A Ideal effort
Company B Actual project hours
Time
Figure 1.5 Engineering changes during automobile development. (Source: Data from Tom Judd, Cognition Corp., “Taking DFSS to the Next Level,” WCBF, Design for Six Sigma Conference, Las Vegas, June 2005.)
Fail early; fail often. early in the design process may cost $10,000 later during product refinement and $1,000,000 or more in tooling, sales, and goodwill expenses if made after production has begun. Figure 1.5 also indicates that Company A took less time to design the automobile than Company B. This is due to differences in the design philosophies of the companies. Company A assigns a large engineering staff to the project early in product development and encourages these engineers to utilize the latest in design techniques and to explore all the options early to preclude the need for changes later on. Company B, on the other hand, assigns a small staff and pressures them for quick results, in the form of hardware, discouraging the engineers from exploring all options (the region in the oval in the figure). The design axiom, fail early, fail often, applies to this example. Changes are required in order to find a good design, and early changes are easier and less expensive than changes made later. The engineers in Company B spend much time “firefighting” after the product is in production. In fact, many engineers spend as much as 50% of their time firefighting for companies similar to Company B. An additional way that the design process affects product development time is in how long it takes to bring a product to market. Prior to the 1980s there was little emphasis on the length of time to develop new products, Since then competition has forced new products to be introduced at a faster and faster rate. During the 1990s development time in most industries was cut by half. This trend
7
ullman-38162
8
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
has continued into the twenty-first century. More on how the design process has played a major role in this reduction is in Chap. 4. Finally, for many years it was believed that there was a trade-off between high-quality products and low costs or time—namely, that it costs more and takes more time to develop and produce high-quality products. However, recent experience has shown that increasing quality and lowering costs and time can go hand in hand. Some of the examples we have discussed and ones throughout the rest of the book reinforce this point.
1.3 THE HISTORY OF THE DESIGN PROCESS During design activities, ideas are developed into hardware that is usable as a product. Whether this piece of hardware is a bookshelf or a space station, it is the result of a process that combines people and their knowledge, tools, and skills to develop a new creation. This task requires their time and costs money, and if the people are good at what they do and the environment they work in is well structured, they can do it efficiently. Further, if they are skilled, the final product will be well liked by those who use it and work with it—the customers will see it as a quality product. The design process, then, is the organization and management of people and the information they develop in the evolution of a product. In simpler times, one person could design and manufacture an entire product. Even for a large project such as the design of a ship or a bridge, one person had sufficient knowledge of the physics, materials, and manufacturing processes to manage all aspects of the design and construction of the project. By the middle of the twentieth century, products and manufacturing processes had become so complex that one person no longer had sufficient knowledge or time to focus on all the aspects of the evolving product. Different groups of people became responsible for marketing, design, manufacturing, and overall management. This evolution led to what is commonly known as the “over-thewall” design process (Fig. 1.6). In the structure shown in Fig. 1.6, the engineering design process is walled off from the other product development functions. Basically, people in marketing communicate a perceived market need to engineering either as a simple, written request or, in many instances, orally. This is effectively a one-way communication and is thus represented as information that is “thrown over the wall.”
Customers
Marketing
Figure 1.6 The over-the-wall design method.
Engineering design
Production
ullman-38162
ull75741_01
December 17, 2008
11:15
1.3
The History of the Design Process
Engineering interprets the request, develops concepts, and refines the best concept into manufacturing specifications (i.e., drawings, bills of materials, and assembly instructions). These manufacturing specifications are thrown over the wall to be produced. Manufacturing then interprets the information passed to it and builds what it thinks engineering wanted. Unfortunately, often what is manufactured by a company using the over-thewall process is not what the customer had in mind. This is because of the many weaknesses in this product development process. First, marketing may not be able to communicate to engineering a clear picture of what the customers want. Since the design engineers have no contact with the customers and limited communication with marketing, there is much room for poor understanding of the design problem. Second, design engineers do not know as much about the manufacturing processes as manufacturing specialists, and therefore some parts may not be able to be manufactured as drawn or manufactured on existing equipment. Further, manufacturing experts may know less-expensive methods to produce the product. Thus, this single-direction over-the-wall approach is inefficient and costly and may result in poor-quality products. Although many companies still use this method, most are realizing its weaknesses and are moving away from its use. In the late 1970s and early 1980s, the concept of simultaneous engineering began to break down the walls. This philosophy emphasized the simultaneous development of the manufacturing process with the evolution of the product. Simultaneous engineering was accomplished by assigning manufacturing representatives to be members of design teams so that they could interact with the design engineers throughout the design process. The goal was the simultaneous development of the product and the manufacturing process. In the 1980s the simultaneous design philosophy was broadened and called concurrent engineering, which, in the 1990s, became Integrated Product and Process Design (IPPD). Although the terms simultaneous, concurrent, and integrated are basically synonymous, the change in terms implies a greater refinement in thought about what it takes to efficiently develop a product. Throughout the rest of this text, the term concurrent engineering will be used to express this refinement. In the 1990s the concepts of Lean and Six Sigma became popular in manufacturing and began to have an influence on design. Lean manufacturing concepts were based on studies of the Toyota manufacturing system and introduced in the United States in the early 1990s. Lean manufacturing seeks to eliminate waste in all parts of the system, principally through teamwork. This means eliminating products nobody wants, unneeded steps, many different materials, and people waiting downstream because upstream activities haven’t been delivered on time. In design and manufacturing, the term “lean” has become synonymous with minimizing the time to do a task and the material to make a product. The Lean philosophy will be refined in later chapters. Where Lean focuses on time, Six Sigma focuses on quality. Six Sigma, sometimes written as (6σ) was developed at Motorola in the 1980s and popularized in the 1990s as a way to help ensure that products were manufactured to the highest
9
ullman-38162
10
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
Table 1.2 The ten key features of design best practice
1. 2. 3. 4. 5. 6. 7. 8. 9. 10.
Focus on the entire product life (Chap. 1) Use and support of design teams (Chap. 3) Realization that the processes are as important as the product (Chaps. 1 and 4) Attention to planning for information-centered tasks (Chap. 4) Careful product requirements development (Chap. 5) Encouragement of multiple concept generation and evaluation (Chaps. 6 and 7) Awareness of the decision-making process (Chap. 8) Attention to designing in quality during every phase of the design process (throughout) Concurrent development of product and manufacturing process (Chaps. 9–12) Emphasis on communication of the right information to the right people at the right time (throughout and in Section 1.4.)
standards of quality. Six Sigma uses statistical methods to account for and manage product manufacturing uncertainty and variation. Key to Six Sigma methodology is the five-step DMAIC process (Define, Measure, Analyze, Improve, and Control). Six Sigma brought improved quality to manufactured products. However, quality begins in the design of products, and processes, not in their manufacture. Recognizing this, the Six Sigma community began to emphasize quality earlier in the product development cycle, evolving DFSS (Design for Six Sigma) in the late 1990s. Essentially DFSS is a collection of design best practices similar to those introduced in this book. DFSS is still an emerging discipline. Beyond these formal methodologies, during the 1980s and 1990s many design process techniques were introduced and became popular. They are essential building blocks of the design philosophy introduced throughout the book. All of these methodologies and best practices are built around a concern for the ten key features listed in Table 1.2. These ten features are covered in the chapters shown and are integrated into the philosophy covered in this book. The primary focus is on the integration of teams of people, design tools and techniques, and information about the product and the processes used to develop and manufacture it. The use of teams, including all the “stakeholders” (people who have a concern for the product), eliminates many of the problems with the over-the-wall method. During each phase in the development of a product, different people will be important and will be included in the product development team. This mix of people with different views will also help the team address the entire life cycle of the product. Tools and techniques connect the teams with the information. Although many of the tools are computer-based, much design work is still done with pencil and paper. Thus, the emphasis in this book is not on computer-aided design but on the techniques that affect the culture of design and the tools used to support them.
1.4 THE LIFE OF A PRODUCT Regardless of the design process followed, every product has a life history, as described in Fig. 1.7. Here, each box represents a phase in the product’s life.
ull75741_01
December 17, 2008
11:15
1.4
Product development
The Life of a Product
Production and delivery
Identify need Plan for the design process Develop engineering specifications
Manufacture Assemble Distribute
Develop concepts Install Develop product
End of life
Use Use Operate in sequence 1
Retire
..
ullman-38162
Operate in sequence N Disassemble Reuse or recycle
Clean Maintain Diagnose Test Repair
Figure 1.7 The life of a product.
These phases are grouped into four broad areas. The first area concerns the development of the product, the focus of this book. The second group of phases includes the production and delivery of the product. The third group contains all the considerations important to the product’s use. And the final group focuses on what happens to the product after it is no longer useful. Each phase will be introduced in this section, and all are detailed later in the book. Note that designers, responsible for the first five phases, must fully understand all the subsequent phases if they are to develop a quality product. The design phases are: Identify need. Design projects are initiated either by a market requirement, the development of a new technology, or the desire to improve an existing product.
11
ullman-38162
12
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
The design process not only gives birth to a product but is also responsible for its life and death.
Plan for the design process. Efficient product development requires planning for the process to be followed. Planning for the design process is the topic of Chap. 4. Develop engineering requirements. The importance of developing a good set of specifications has become one of the key points in concurrent engineering. It has recently been realized that the time spent evolving complete specifications prior to developing concepts saves time and money and improves quality. A technique to help in developing specifications is covered in Chap. 6. Develop concepts. Chapters 7 and 8 focus on techniques for generating and evaluating new concepts. This is an important phase in the development of a product, as decisions made here affect all the downstream phases. Develop product. Turning a concept into a manufacturable product is a major engineering challenge. Chapters 9–12 present techniques to make this a more reliable process. This phase ends with manufacturing specifications and release to production. These first five phases all must take into account what will happen to the product in the remainder of its lifetime. When the design work is completed, the product is released for production, and except for engineering changes, the design engineers will have no further involvement with it. The production and delivery phases include: Manufacture. Some products are just assemblies of existing components. For most products, unique components need to be formed from raw materials and thus require some manufacturing. In the over-the-wall design philosophy, design engineers sometimes consider manufacturing issues, but since they are not experts, they sometimes do not make good decisions. Concurrent engineering encourages having manufacturing experts on the design team to ensure that the product can be produced and can meet cost requirements. The specific consideration of design for manufacturing and product cost estimation is covered in Chap. 11. Assemble. How a product is to be assembled is a major consideration during the product design phase. Part of Chap. 11. is devoted to a technique called design for assembly, which focuses on making a product easy to assemble. Distribute. Although distribution may not seem like a concern for the design engineer, each product must be delivered to the customer in a safe and costeffective manner. Design requirements may include the need for the product to be shipped in a prespecified container or on a standard pallet. Thus, the
ullman-38162
ull75741_01
December 17, 2008
11:15
1.4
The Life of a Product
design engineers may need to alter their product just to satisfy distribution needs. Install. Some products require installation before the customer can use them. This is especially true for manufacturing equipment and building industry products. Additionally, concern for installation can also mean concern for how customers will react to the statement, “Some assembly required.” The goal of product development, production, and delivery is the use of the product. The “Use” phases are: Operate. Most design requirements are aimed at specifying the use of the product. Products may have many different operating sequences that describe their use. Consider as an example a common hammer that can be used to put in nails or take them out. Each use involves a different sequence of operations, and both must be considered during the design of a hammer. Clean. Another aspect of a product’s use is keeping it clean. This can range from frequent need (e.g., public bathroom fixtures) to never. Every consumer has experienced the frustration of not being able to clean a product. This inability is seldom designed into the product on purpose; rather, it is usually simply the result of poor design. Maintain. As shown in Fig. 1.7, to maintain a product requires that problems must be diagnosed, the diagnosis may require tests, and the product must be repaired. Finally, every product has a finite life. End-of-life concerns have become increasingly important. Retire. The final phase in a product’s life is its retirement. In past years designers did not worry about a product beyond its use. However, during the 1980s increased concern for the environment forced designers to begin considering the entire life of their products. In the 1990s the European Union enacted legislation that makes the original manufacturer responsible for collecting and reusing or recycling its products when their usefulness is finished. This topic will be further discussed in Section 12.8. Disassemble. Before the 1970s, consumer products could be easily disassembled for repair, but now we live in a “throwaway” society, where disassembly of consumer goods is difficult and often impossible. However, due to legislation requiring us to recycle or reuse products, the need to design for disassembling a product is returning. Reuse or recycle. After a product has been disassembled, its parts can either be reused in other products or recycled—reduced to a more basic form and used again (e.g., metals can be melted, paper reduced to pulp again). This emphasis on the life of a product has resulted in the concept of Product Life-cycle Management (PLM). The term PLM was coined in the fall of 2001 as a blanket term for computer systems that support both the definition or authoring of product information from cradle to grave. PLM enables management
13
ullman-38162
14
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
of this information in forms and languages understandable by each constituency in the product life cycle—namely, the words and representations that the engineers understand are not the same as what manufacturing or service people understand. A predecessor to PLM was Product Data Management (PDM), which evolved in the 1980s to help control and share the product data. The change from “data” in PDM to life cycle in PLM reflects the realization that there is more to a product than the description of its geometry and function—the processes are also important. As shown in Fig. 1.8, PLM integrates six different major types of information. In the past these were separate, and communications between the communities Customer Needs
Environment Regulations
Systems Engineering
Features Functions Architecture Signals and connections Simulations Solid models Layout
Product Life-cycle Management (PLM)
Design Automation
Drawings
MCAD
Assembly ECAD Software
Bills of Materials
Manufacturing Engineering
Service, Diagnosis, Warrantee
Portfolio Planning
Figure 1.8 Product Life-cycle Management.
Detail
DFA DFM
ullman-38162
ull75741_01
December 17, 2008
11:15
1.5
The Many Solutions for Design Problems
was poor (think of the over-the-wall method, Fig. 1.6). Whereas Fig. 1.7 focuses on the activities that happen during a product’s life, PLM, Fig. 1.8 focuses on the information that must be managed to support that life. What PLM calls “Systems Engineering” is support for the technical development of the function of the product. The topics listed under Systems Engineering are all covered in this book. What historically was called CAD (Computer-Aided Design) is now often referred to as MCAD for Mechanical CAD to differentiate it from Electronic CAD (ECAD). These two, along with software are all part of design automation. Like most of PLM, this structure grew from the twigs to the root of the tree. Traditional drawings included layout and detailed and assembly drawings. The advent of solid models made them a part of an MCAD system. Bills Of Materials (BOMs) are effectively parts lists. BOMs are fundamental documents for manufacturing. However, as product is evolving in systems engineering so does the BOM; early on there may be no parts to list. In manufacturing, PLM manages information about Design For Manufacturing (DFM) and Assembly (DFA). Once the product is launched and in use, there is a need to maintain it, or as shown in Fig. 1.7, diagnose, test, and repair it. These activities are supported by service, diagnosis, and warrantee information in a PLM system. Finally, there is need to manage the product portfolio—namely, of the products that could be offered, which ones are chosen to be offered (the organization’s portfolio). Portfolio decisions are the part of doing business that determines which products will be developed and sold. This description of the life of a product and systems to manage it, gives a good basic understanding of the issues that will be addressed in this book. The rest of this chapter details the unique features of design problems and their solution processes.
1.5 THE MANY SOLUTIONS FOR DESIGN PROBLEMS Consider this problem from a textbook on the design of machine components (see Fig. 1.9): What size SAE grade 5 bolt should be used to fasten together two pieces of 1045 sheet steel, each 4 mm thick and 6 cm wide, which are lapped over each other and loaded with 100 N?
Figure 1.9 A simple lap joint.
15
ullman-38162
16
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
Design problems have many satisfactory solutions but no clear best solution.
In this problem the need is very clear, and if we know the methods for analyzing shear stress in bolts, the problem is easily understood. There is no necessity to design the joint because a design solution is already given, namely, a grade 5 bolt, with one parameter to be determined—its diameter. The product evaluation is straight from textbook formulas, and the only decision made is in determining whether we did the problem correctly. In comparison, consider this, only slightly different, problem: Design a joint to fasten together two pieces of 1045 sheet steel, each 4 mm thick and 6 cm wide, which are lapped over each other and loaded with 100 N.
The only difference between these problems is in their opening clauses (shown in italics) and a period replacing the question mark (you might want to think about this change in punctuation). The second problem is even easier to understand than the first; we do not need to know how to design for shear failure in bolted joints. However, there is much more latitude in generating ideas for potential concepts here. It may be possible to use a bolted joint, a glued joint, a joint in which the two pieces are folded over each other, a welded joint, a joint held by magnets, a Velcro joint, or a bubble-gum joint. Which one is best depends on other, unstated factors. This problem is not as well defined as the first one. To evaluate proposed concepts, more information about the joint will be needed. In other words, the problem is not really understood at all. Some questions still need to be answered: Will the joint require disassembly? Will it be used at high temperatures? What tools are available to make the joint? What skill levels do the joint manufacturers have? The first problem statement describes an analysis problem. To solve it we need to find the correct formula and plug in the right values. The second statement describes a design problem, which is ill-defined in that the problem statement does not give all the information needed to find the solution. The potential solutions are not given and the constraints on the solution are incomplete. This problem requires us to fill in missing information in order to understand it fully. Another difference between the two problems is in the number of potential solutions. For the first problem there is only one correct answer. For the second there is no correct answer. In fact, there may be many good solutions to this problem, and it may be difficult if not impossible to define what is meant by the “best solution.” Just consider all the different cars, televisions, and other products that compete in the same market. In each case, all the different models solve essentially the same problem, yet there are many different solutions. The goal in design is to find a good solution that leads to a quality product with the least commitment of time and other resources. All design problems have a multitude of satisfactory solutions and no clear best solution. This is shown graphically
ullman-38162
ull75741_01
December 17, 2008
11:15
1.6
The Basic Actions of Problem Solving
Design process knowledge
Resulting products that meet the need
Design need
Design process paths
Physics Electric Materials motors science Engineering Thermodynamics science Manufacturing Engineering processes economics Welding design Pumps Domain Kinematics knowledge
Figure 1.10 The many results of the design process.
in Fig. 1.10 where the factors that affect exactly what solution is developed are noted. Domain knowledge is developed through the study of engineering physics and other technical areas and through the observation of existing products. It is the study of science and engineering science that provides the basis on which the design process is based. Design process knowledge is the subject of this book. For mechanical design problems in particular, there is an additional characteristic: the solution must be a piece of working hardware—a product. Thus, mechanical design problems begin with an ill-defined need and result in an object that behaves in a certain way, a way that the designers feel meets this need. This creates a paradox. A designer must develop a device that, by definition, has the capabilities to meet some need that is not fully defined.
1.6 THE BASIC ACTIONS OF PROBLEM SOLVING Regardless of what design problem we are solving, we always, consciously or unconsciously, take six basic actions: 1. Establish the need or realize that there is a problem to be solved. 2. Plan how to solve the problem.
17
ullman-38162
18
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
3. Understand the problem by developing requirements and uncovering existing solutions for similar problems. 4. Generate alternative solutions. 5. Evaluate the alternatives by comparing them to the design requirements and to each other. 6. Decide on acceptable solutions. This model fits design whether we are looking at the entire product (see the product life-cycle diagram, Fig. 1.7) or the smallest detail of it. These actions are not necessarily taken in 1-2-3 order. In fact they are often intermingled with solution generation and evaluation improving the understanding of the problem, enabling new, improved solutions to be generated. This iterative nature of design is another feature that separates it from analysis. The list of actions is not complete. If we want anyone else on the design team to make use of our results, a seventh action is also needed: 7. Communicate the results. The need that initiates the process may be very clearly defined or ill-defined. Consider the problem statements for the design of the simple lap joint of two pieces of metal given earlier (Fig. 1.9). The need was given by the problem statement in both cases. In the first statement, understanding is the knowledge of what parameters are needed to characterize a problem of this type and the equations that relate the parameters to each other (a model of the joint). There is no need to generate potential solutions, evaluate them, or make any decision, because this is an analysis problem. The second problem statement needs work to understand. The requirements for an acceptable solution must be developed, and then alternative solutions can be generated and evaluated. Some of the evaluation may be the same as the analysis problem, if one of the concepts is a bolt. Some important observations: ■
■ ■ ■ ■

New needs are established throughout the design effort because new design problems arise as the product evolves. Details not addressed early in the process must be dealt with as they arise; thus, the design of these details poses new subproblems. Planning occurs mainly at the beginning of a project. Plans are always updated because understanding is improved as the process progresses. Formal efforts to understand new design problems continue throughout the process. Each new subproblem requires new understanding. There are two distinct modes of generation: concept generation and product generation. The techniques used in these two actions differ. Evaluation techniques also depend on the design phase; there are differences between the evaluation techniques used for concepts and those used for products. It is difficult to make decisions, as each decision requires a commitment based on incomplete evaluation. Additionally, since most design problems
ullman-38162
ull75741_01
December 17, 2008
11:15
1.7

Knowledge and Learning During Design
are solved by teams, a decision requires consensus, which is often difficult to obtain. Communication of the information developed to others on the design team and to management is an essential part of concurrent engineering.
We will return to these observations as the design process is developed through this text.
1.7 KNOWLEDGE AND LEARNING DURING DESIGN When a new design problem is begun, very little may be known about the solution, especially if the problem is a new one for the designer. As work on the project progresses, the designer’s knowledge about the technologies involved and the alternative solutions increases, as shown in Fig. 1.11. Therefore, after completing a project, most designers want a chance to start all over in order to do the project properly now that they fully understand it. Unfortunately, few designers get the opportunity to redo their projects. Throughout the solution process knowledge about the problem and its potential solutions is gained and, conversely, design freedom is lost. This can also be seen in Fig. 1.11, where the time into the design process is equivalent to exposure to the problem. The curve representing knowledge about the problem is a learning curve; the steeper the slope, the more knowledge is gained per unit time. Throughout most of the design process the learning rate is high. The second curve in Fig. 1.11 illustrates the degree of design freedom. As design decisions are made, the ability to change the product becomes increasingly limited. At the beginning the designer has great freedom because few decisions have been made and little capital has been committed. But by the time the product is in production, 100 Knowledge about the design problem
Percentage
80
60
40 Design freedom
20
0
Time into design process
Figure 1.11 The design process paradox.
19
ullman-38162
20
ull75741_01
December 17, 2008
CHAPTER 1
11:15
Why Study the Design Process?
A design paradox: The more you learn the less freedom you have to use what you know.
any change requires great expense, which limits freedom to make changes. Thus, the goal during the design process is to learn as much about the evolving product as early as possible in the design process because during the early phases changes are least expensive.
1.8 DESIGN FOR SUSTAINABILITY It is important to realize that design engineers have much control over what products are designed and how they interact with the earth over their lifetime. The responsibility that goes with designing is well summarized in the Hannover Principles. These were developed for EXPO 2000, The World’s Fair in Hannover, Germany. These principles define the basics of Designing For Sustainability (DFS) or Design For the Environment (DFE). DFS requires awareness of the short- and long-term consequences of your design decisions. The Hannover Principles aim to provide a platform on which designers can consider how to adapt their work toward sustainable ends. According to the World Commission on Environment and Development, the high-level goal is “Meeting the needs of the present without compromising the ability of future generations to meet their own needs.” The Hannover Principles are: 1. Insist on rights of humanity and nature to coexist in a healthy, supportive, diverse, and sustainable condition. 2. Recognize interdependence. The elements of human design interact with and depend on the natural world, with broad and diverse implications at every scale. Expand design considerations to recognizing even distant effects. 3. Accept responsibility for the consequences of design decisions on human well-being, the viability of natural systems and their right to coexist. 4. Create safe objects of long-term value. Do not burden future generations with requirements for maintenance or vigilant administration of potential danger due to the careless creation of products, processes, or standards. 5. Eliminate the concept of waste. Evaluate and optimize the full life cycle of products and processes to approach the state of natural systems in which there is no waste. 6. Rely on natural energy flows. Human designs should, like the living world, derive their creative forces from perpetual solar income. Incorporate this energy efficiently and safely for responsible use. 7. Understand the limitations of design. No human creation lasts forever and design does not solve all problems. Those who create and plan should practice
ullman-38162
ull75741_01
December 17, 2008
11:15
1.9
Summary
You are responsible for the impact of your products on others.
humility in the face of nature. Treat nature as a model and mentor, not as an inconvenience to be evaded or controlled. 8. Seek constant improvement by the sharing of knowledge. Encourage direct and open communication between colleagues, patrons, manufacturers, and users to link long-term sustainable considerations with ethical responsibility, and reestablish the integral relationship between natural processes and human activity. 9. Respect relationships between spirit and matter. Consider all aspects of human settlement including community, dwelling, industry, and trade in terms of existing and evolving connections between spiritual and material consciousness. We will work to respect these principles in the chapters that follow. We introduced the concept of “lean” earlier in this chapter as the effort to reduce waste (Principle 5). We will revisit this and the other principles throughout the book. In Chap. 11, we will specifically revisit DFS as part of Design for the Environment. In Chap. 12, we focus on product retirement. Many products are retired to landfills, but in keeping with the first three principles, and focusing on the fifth principle, it is best to design products that can be reused and recycled.
1.9 SUMMARY The design process is the organization and management of people and the information they develop in the evolution of a product. ■
■ ■

The success of the design process can be measured in the cost of the design effort, the cost of the final product, the quality of the final product, and the time needed to develop the product. Cost is committed early in the design process, so it is important to pay particular attention to early phases. The process described in this book integrates all the stakeholders from the beginning of the design process and emphasizes both the design of the product and concern for all processes—the design process, the manufacturing process, the assembly process, and the distribution process. All products have a life cycle beginning with establishing a need and ending with retirement. Although this book is primarily concerned with planning for the design process, engineering requirements development, conceptual design, and product design phases, attention to all the other phases is important. PLM systems are designed to support life-cycle information and communication.
21
ullman-38162
22
ull75741_01
December 17, 2008
CHAPTER 1
■ ■ ■
11:15
Why Study the Design Process?
The mechanical design process is a problem-solving process that transforms an ill-defined problem into a final product. Design problems have more than one satisfactory solution. Design for Sustainability embodied in the Hannover Principles is becoming an increasingly important part of the design process.
1.10 SOURCES Creveling, C. M., Dave Antis, and Jeffrey Lee Slutsky: Design for Six Sigma in Technology and Product Development, Prentice Hall PTR, 2002. A good book on DFSS. Ginn, D., and E. Varner: The Design for Six Sigma Memory Jogger, Goal/QPC, 2004. A quick introduction to DFSS The Hannover Principles, Design for Sustainability. Prepared for EXPO 2000, Hannover, Germany, http://www.mcdonough.com/principles.pdf Product life-cycle management (PLM) description based on work at Siemens PLM supplied by Wayne Embry their PLM Functional Architect. http://www.plm.automation.siemens.com/en_us/products/teamcenter/index.shtml http://www.johnstark.com/epwl4.html PLM listing of over 100 vendors. Ulrich, K. T., and S. A. Pearson: “Assessing the Importance of Design through Product Archaeology,” Management Science, Vol. 44, No. 3, pp. 352–369, March 1998, or “Does Product Design Really Determine 80% of Manufacturing Cost?” working paper 3601–93, Sloan School of Management, MIT, Cambridge, Mass., 1993. In the first edition of The Mechanical Design Process it was stated that design determined 80% of the cost of a product. To confirm or deny that statement, researchers at MIT performed a study of automatic coffeemakers and wrote this paper. The results show that the number is closer to 50% on the average (see Fig. 1.3) but can range as high as 75%. Womack, James P., and Daniel T. Jones: Lean Thinking: Banish Waste and Create Wealth in Your Corporation, Simon and Schuster, New York, 1996.
1.11 EXERCISES 1.1 Change a problem from one of your engineering science classes into a design problem. Try changing as few words as possible. 1.2 Identify the basic problem-solving actions for a. Selecting a new car b. Finding an item in a grocery store c. Installing a wall-mounted bookshelf d. Placing a piece in a puzzle 1.3 Find examples of products that are very different yet solve exactly the same design problem. Different brands of automobiles, bikes, CD players, cheese slicers, wine bottle openers, and personal computers are examples. For each, list its features, cost, and perceived quality. 1.4 How well do the products in Exercise 1.3 meet the Hannover Principles? 1.5 To experience the limitations of the over-the-wall design method try this. With a group of four to six people, have one person write down the description of some object that is
ullman-38162
ull75741_01
December 17, 2008
11:15
1.11
Exercises
not familiar to the others. This description should contain at least six different nouns that describe different features of the object. Without showing the description to the others, describe the object to one other person in such a manner that the others can’t hear. This can be done by whispering or leaving the room. Limit the description to what was written down. The second person now conveys the information to the third person, and so on until the last person redescribes the object to the whole group and compares it to the original written description. The modification that occurs is magnified with more complex objects and poorer communication. (Professor Mark Costello of Georgia Institute of Technology originated this problem.)
23
ullman-38162
ull75741_01
December 17, 2008
11:15
ullman-38162
ull75741_02
December 17, 2008
16:16
C
H
2 A
P
T
E
R
Understanding Mechanical Design KEY QUESTIONS ■ ■ ■ ■ ■
What is the difference between function, behavior, and performance? Why does mechanical design flow from function to form? What are the languages of mechanical design? Are all design problems the same? What can you learn from dissecting products?
2.1 INTRODUCTION For most of history, the discipline of mechanical design required knowledge of only mechanical parts and assemblies. But early in the twentieth century, electrical components were introduced in mechanical devices. Then, during World War II, in the 1940s, electronic control systems became part of the mix. Since this change, designers have often had to choose between purely mechanical systems and systems that were a mix of mechanical and electronic components and systems. These electronic systems have matured from very simple functions and logic to the incorporation of computers and complex logic. Many electromechanical products now include microprocessors. Consider, for example, cameras, office copiers, cars, and just about everything else. Systems that have mechanical, electronic, and software components are often called mechatronic devices. What makes the design of these devices difficult is the necessity for domain and design process knowledge in three overlapping but clearly different disciplines. But, no matter how electronic or computer-centric devices become, nearly all products require mechanical functions and a mechanical interface with humans. Additionally, all products require mechanical machinery for manufacture and assembly 25
ullman-38162
26
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
and mechanical components for housing. Thus, no matter how “smart” products become, there will always be the need for mechanical design. To explore systems that have significant mechanical components consider two examples that will be used throughout the book, the Irwin Quick-Grip clamp (Fig. 2.1) and the drive wheel assembly for the NASA Mars Exploration Rover (MER) developed by Cal Tech’s Jet Propulsion Laboratory (JPL) (Fig. 2.2).
Figure 2.1 Irwin Quick-Grip clamp. (Reprinted with permission of Irwin Industrial Tools.)
Figure 2.2 The Mars Exploration Rover being tested by JPL engineers. (Reprinted with permission of NASA/JPL.)
ullman-38162
ull75741_02
December 17, 2008
16:16
2.1
Introduction
Irwin is one of the largest manufacturers of one-handed bar clamps. What makes the model shown in Fig 2.1 unique is that it can generate over 550 lb (250 kg) of force with the strength of only one hand. Irwin introduced this product in 2006 and sells many tens of thousands of them a month. In contrast to the purely mechanical, high-production-volume Quick-Grip, only two MERs were made and they are highly mechatronic. The two MERs were launched toward Mars on June 10 and July 7, 2003, in search of answers about the history of water on Mars. They landed on Mars January 3 and January 24, 2004. They were designed for 90 Sol (Martian days, about 40 min longer than an Earth day) and were still operating in 2008, over 1300 Sols (over 3.5 years) past their design life. One of the Rovers, Opportunity, had traveled over 11 km (7.1 mi) during its five years of life. Each Rover is a six-wheeled, solar-powered robot that stands 1.5 m (4.9 ft) high and is 2.3 m (7.5 ft) wide and 1.6 m (5.2 ft) long. They weigh 180 kg (400 lb) on Earth, 35 kg (80 lb) of which is the wheel and suspension system. Mars has only 38% the gravitational pull of Earth. So they weigh 68.4 kg (152 lb) on Mars. As shown in Fig. 2.3, a very simplified diagram of the MER’s systems, propulsion and steering are two of the subsystems. Later in this chapter, we delve further into the MER, and in later chapters we will detail the wheels. In general, during the design process the function of the system and its decomposition are considered first. After the function has been decomposed into the finest subsystems possible, assemblies and components are developed to provide these functions. For mechanical devices, the general decomposition is system– subsystem–assembly–component. Figure 2.3 shows the MER propulsion system, within which the motor and transmission are two subsystems. The wheel is a component. Systems, subsystems, and components all have features, specific attributes that are important, such as dimensions, material properties, shapes, or functional details. For the MER propulsion system, an important feature is that it can propel the MER at 5 cm/sec. For the transmission, a feature is that it has a 1500:1 reduction ratio. For the MER wheel, some of the important features are its diameter, tread pattern, and flexibility. We must also note that many systems have both electrical and mechanical subsystems and components. Electrical systems generally provide energy, sensing, and control functions. The function of these electrical systems is fulfilled by circuits (electrical assemblies) that can be decomposed into electrical components (e.g., switches, transistors, and ICs), much as with mechanical objects. Finally, some of the control functions are filled by microprocessors. Physically, these are electric circuits, but the actual control function is provided by software programs in the processor. These programs are assemblies of coding modules composed of individual coding statements. Note that the function of the microprocessor could be filled by an electrical or possibly even a purely mechanical system. During the early phases of the design process, when developing systems is the focus of the effort, it is often unclear whether the actual function will be met by mechanical assemblies, electrical circuits, software programs, or a mix of these elements.
27
ullman-38162
28
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
Solar collectors
IMU (Inertial Measurement Unit)
REM (Rover Electronics Module)
Battery Propulsion
Steering
Drive motor (1 of 6)
Steering motor (1 of 4)
Transmission (1 of 6)
Transmission (1 of 4)
Wheel (1 of 6)
Encoder (1 of 4)
Figure 2.3 The MER Propulsion System showing some of the sub-systems and
components.
2.2 IMPORTANCE OF PRODUCT FUNCTION, BEHAVIOR, AND PERFORMANCE What is the function of the Irwin clamp? How does it behave? Does it have good performance? These three questions revolve around the terms “function,” “behavior,” and “performance”—similar, but different attributes of the clamp. There are many synonyms for the word function. In mechanical engineering, we commonly use the terms function, operation, and purpose to describe what a device does. A common way of classifying mechanical devices is by their function. In fact, some devices having only one main function are named for
ullman-38162
ull75741_02
December 17, 2008
2.2
16:16
Importance of Product Function, Behavior, and Performance
Function determines form and form, in turn, enables function. that function. For example, a screwdriver has the function of enabling a person to insert or remove a screw. The terms drive, insert, and remove are all verbs that tell what the screwdriver does. In telling what the screwdriver does, we have given no indication of how the screwdriver accomplishes its function. To discover how, we must have some information on the form of the device. The term form relates to any aspect of physical shape, geometry, construction, material, or size. As we shall see in Chap. 3, one of the main ways engineers mentally index their knowledge about the mechanical world is by function. Now reread this paragraph and replace the screwdriver example with the Quick-Grip clamp. In Fig. 2.3, we physically decomposed the Mars Rover propulsion and steering systems into subsystems and components at its physical boundaries. Functional decomposition is often much more difficult than physical decomposition, as each function may use part of many components and each component may serve many functions. Consider the handlebar of a bicycle. The handlebar is a bent piece of tubing, a single component that serves many functions. It enables the rider to “steer the bicycle” (“steer” is a verb that tells what the device does), and the handlebar “supports the rider” (again, a function telling what the handlebar does). Further, it not only “supports the brake levers” but also “transforms (another function) the gripping force” to a pull on the brake cable. The shape of the handlebar and its relationship with other components determine how it provides all these different functions. The handlebar, however, is not the only component needed to steer the bike. Additional components necessary to perform this function are the front fork, the bearings between the fork and the frame, the front wheel, and miscellaneous fasteners. Actually, it can be argued that all the components on a bike contribute to steering, since a bike without a seat or rear wheel would be hard to steer. In any case, the handlebar performs many different functions, but in fulfilling these functions, the handlebar is only a part of various assemblies. Similarly, the steering on the MER cannot actually steer it without the wheels in the propulsion system. The coupling between form and function makes mechanical design challenging. Many common devices are cataloged by their function. If we want to specify a bearing, for example, we can search a bearing catalog and find many different styles of bearings (plain, ball, or tapered roller, for example). Each “style” has a different geometry—a different form—though all have the same primary function, namely, to reduce friction between a shaft and another object. Cataloging is possible in mechanical design as long as the primary function is clearly defined by a single piece of hardware, either a single component or an assembly. In other words, the form and function are decomposed along the same boundaries. This is true of many mechanical devices, such as pumps, valves, heat exchangers, gearboxes, and fan blades, and is especially true of many electrical circuits and components, such as resistors, capacitors, and amplifier circuits.
29
ullman-38162
30
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
Known input
?????? To be designed
Desired performance
(a) Function Known input
Physical properties of system
Actual performance
(b) Behavior
Figure 2.4 Function and behavior.
Two other terms often related to function are behavior and performance. Function and behavior are often used synonymously. However, there is a subtle difference, as shown in Fig. 2.4. In this figure there are two standard system blocks with an input represented by an arrow into the box, the system acted on by the input represented by the box, and the reaction of the system to the input represented by the arrow out of the box. The box in the upper part of the figure shows that function is the desired output from a system that is yet to be designed. When we begin to design a device, the device itself is unknown, but what we want it to do is known. If the system is known, as in the second part of the figure, then the behavior of the system can be found. Behavior is the actual output, the response of the system’s physical properties to the input energy or control. Thus, the behavior can be simulated or measured, whereas function is only a desire. Performance is the measure of function and behavior—how well the device does what it is designed to do. When we say that one function of the handlebar is to steer the bicycle, we say nothing about how well it serves this purpose. Before designing a handlebar, we must develop a clear picture of its desired performance. For example, one design functional goal is that the handlebar must “support 50 kg,” a measurable desired performance for the handlebar. The development of clear performance measures is the focus of Chap. 6. Further, after designing the handlebar we can simulate its strength analytically or measure the strength of a prototype to find the actual performance for comparison to that desired. This comparison is a major focus of Chap. 10.
2.3 MECHANICAL DESIGN LANGUAGES AND ABSTRACTION Many “languages” or representations can be used to describe a mechanical object. Consider for a moment the difference between a detailed drawing of a component and the actual hardware that is the component. Both the drawing and the hardware represent the same object; however, they each represent it in a different language.
ullman-38162
ull75741_02
December 17, 2008
16:16
2.3
Mechanical Design Languages and Abstraction
A skilled designer speaks many languages. Extending this example further, if the component we are discussing is a bolt, then the word bolt is a textual (semantic or word) description of the component, a third language. Additionally, the bolt can be represented through equations (the final language) that describe its functionality and possibly its form. For example, the ability of the bolt to “carry shear stress” (a function) is described by the equation τ = F /A; the shear stress τ is equal to the shear force F on the bolt divided by the stress area A of the bolt. Based on this, we can use four different representations or languages to describe the bolt. These four can be used to describe any mechanical object: Semantic. The verbal or textual representation of the object—for example, the word bolt, or the sentence, “The shear stress on the bolt is the shear force divided by the stress area.” Graphical. The drawings of the object—for example, scale representations such as solid models, orthogonal drawings, sketches, or artistic renderings. Analytical. The equations, rules, or procedures representing the form or function of the object—for example, τ = F/A. Physical. The hardware or a physical model of the object. In most mechanical design problems, the initial need is expressed in a semantic language as a written specification or a verbal request by a customer or supervisor. The result of the design process is a physical object. Although the designer produces a graphical representation of the product, not the hardware itself, all the languages will be used as the product is refined from its initial, abstract semantic representation to its final physical form. Further complicating how we refer to objects being designed, consider two drawings for a MER wheel, as shown in Fig. 2.5. Figure 2.5a is a rough sketch, which gives only abstract information about the component. It centers on the
(a)
Figure 2.5 Abstract sketch and solid model of a MER wheel.
31
ullman-38162
32
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
function of the wheel’s spokes to act like springs. Figure 2.5b is a solid model of the same component, focused on the final form of the wheel. In progressing from the sketch to the solid model, the level of abstraction of the device is refined. Some design process techniques are better suited for abstract levels and others for levels that are more concrete. There are no true levels of abstractions, but rather a continuum on which the form or function can be represented. Descriptions of three levels of abstraction in each of the four languages are given in Table 2.1. The object we call a bolt is used as an example in Table 2.2. Another term that is often used in describing the analytical row in Table 2.1 is simulation fidelity. As analytical models or simulations increase in fidelity, their representation of the actual object or system becomes a more accurate representation of reality. Simulation fidelity will be further refined in Chap. 10. The process of making an object less abstract (or more concrete) is called refinement. Mechanical design is a continuous process of refining the given needs Table 2.1 Levels of abstraction in different languages
Level of abstraction Language
Abstract
−−−−−−→
Concrete
Semantic
Qualitative words (e.g., long, fast, lightest) Rough sketches
Reference to specific parameters or components Scale drawings
Qualitative relations (e.g., left of) None
Back-of-the-envelope calculations Models of the product
Reference to the values of the specific parameters or components Solid models with tolerances Detailed analysis
Graphical Analytical Physical
Final hardware
Table 2.2 Levels of abstraction in describing a bolt
Level of abstraction Language Semantic
Abstract A bolt
−−−−−−→ A short bolt
Concrete A 100030003 1/4−20 UNC Grade 5 bolt
Length of bolt 5 8 -UNC-2A
Graphical Body Length of diameter thread
Analytical Physical
Right-hand rule —
τ = F/A —
τ = F/A
ullman-38162
ull75741_02
December 17, 2008
16:16
2.4
Different Types of Mechanical Design Problems
to the final hardware. The refinement of the bolt in Table 2.4 is illustrated on a left-to-right continuum. In most design situations, the beginning of the problem appears in the upper left corner and the final product in the lower right. The path connecting these is a mix of the other representations and levels of abstractions.
2.4 DIFFERENT TYPES OF MECHANICAL DESIGN PROBLEMS Traditionally, we decompose mechanical engineering by discipline: fluids, thermodynamics, mechanics, and so on. In categorizing the types of mechanical design problems, this discipline-oriented approach is not appropriate. Consider, for example, the simplest kind of design problem, a selection design problem. Selection design means picking one (maybe more) item from a list such that the chosen item meets certain requirements. Common examples are selecting the correct bearing from a bearings catalog, selecting the correct lenses for an optical device, selecting the proper fan for cooling equipment, or selecting the proper heat exchanger for a heating or cooling process. The design process for each of these problems is essentially the same, even though the disciplines are very different. The goal of this section is to describe different types of design problems independently of the discipline. Before beginning, we must realize that most design situations are a mix of various types of problems. For example, we might be designing a new type of consumer product that will accept a whole raw egg, break it, fry it, and deliver it on a plate. Since this is a new product, there will be a lot of original design work to be done. As the design process proceeds, we will configure the various parts. To determine the thickness of the frying surface we will analyze the heat conduction of the frying component, which is parametric design. And we will select a heating element and various fasteners to hold the components together. Further, if we are clever, we may be able to redesign an existing product to meet some or all of the requirements. Each of the italicized terms is a different type of design problem. It is rare to find a problem that is purely one type.
2.4.1
Selection Design
Selection design involves choosing one item (or maybe more) from a list of similar items. We do this type of design every time we choose an item from a catalog. It may sound simple, but if the catalog contains more than a few items and there are many different features to the items, the decision can be quite complex. To solve a selection problem we must start with a clear need. The catalog or the list of choices then effectively generates potential solutions for the problem. We must evaluate the potential solutions with respect to our specific requirements to make the right choice. Consider the following example. During the process of designing a product, an engineer must select a bearing to support a shaft. The known information is given in Fig. 2.6. The shaft has a diameter of 20 mm (0.787 in.). There is a radial force of 6675 N (1500 lb) on the shaft at the bearing,
33
34
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
Bearing
20 mm
ullman-38162
6675 N
Shaft
2000 rpm
Housing
Figure 2.6 Load on a shaft.
and the shaft rotates at a maximum of 2000 rpm. The housing to support the bearing is still to be designed. All we need to do is select a bearing to meet the needs. The information on shaft size, maximum radial force, and maximum rpm given in bearing catalogs enables us to quickly develop a list of potential bearings (Table 2.3). This is the simplest type of design problem we could have, but it is still incompletely defined. We do not have enough information to select among the five possible choices. Even if a short list is developed—the most likely candidates being the 42-mm-deep groove ball bearing and the 24-mm needle bearing—there is no way to make a good decision without more knowledge of the function of the bearing and of the engineering requirements on it.
2.4.2
Configuration Design
A slightly more complex type of design is called configuration or packaging design. In this type of problem, all the components have been designed and the problem is how to assemble them into the completed product. Essentially, this type of design is similar to playing with an Erector set or other construction toy, or arranging living-room furniture. Consider packaging of the assemblies in the MER. The body of the MER is made up of a Rover Equipment Deck (RED) where all the experiments are mounted, a Rover Electronics Module (REM), an Inertial Measurement Unit (IMU), a Warm Electronics Box (WEB), a battery, a UHF radio, an X-band telecom HW, and a Solid-State Power Amplifier (SSPA), as shown in Fig. 2.7. Each of these assemblies is of known size and has certain constraints on its position. For example, the RED must be on top and the WEB on the bottom, but
ullman-38162
ull75741_02
December 17, 2008
16:16
2.4
Different Types of Mechanical Design Problems
35
Table 2.3 Potential bearings for a shaft
Outside diameter (mm)
Width (mm)
Load rating (lb)
Speed limit (rpm)
Deep-groove ball bearing
42 47 52
8 14 15
1560 2900 3900
18,000 15,000 9000
6000 6204 6304
Angular-contact ball bearing
47 37
14 9
3000 1960
13,000 34,000
7204 71,904
Roller bearing
47 52
14 15
6200 7350
13,000 13,000
204 220
Needle bearing
24 26
20 12
1930 2800
13,000 13,000
206 208
Nylon bushing
23
Variable
290. . 8
10 . . 500
4930
Type
IMU
RED
UHF radio
REM
X-band telecom HW
Battery “Front” WEB SSPA
Figure 2.7 The major assemblies in the MER.
+X
+Y +Z
Catalog number
ullman-38162
36
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
many of the other major assemblies can be anywhere inside the envelop defined by these two. Configuration design answers the question, How do we fit all the assemblies in an envelop? or Where do we put what? One methodology for solving this type of problem is to randomly select one component from the list and position it so that all the constraints on that assembly are met. We could start with the REM in the middle, then we select and place a second component. This procedure is continued until either we run into a conflict or all the components are in the MER. If a conflict arises, we back up and try again. For many configuration problems, some of the components to be fit into the assembly can be altered in size, shape, or function, giving the designer more latitude to determine potential configurations and making the problem solution more difficult. There are other methods to configure assemblies. They will be covered in Chap. 11.
2.4.3
Parametric Design
Parametric design involves finding values for the features that characterize the object being studied. This may seem easy enough—just find some values that meet the requirements. However, consider a very simple example. We want to design a cylindrical storage tank that must hold 4 m3 of liquid. This tank is described by the parameters r, its radius, and l, its length and its volume is determined by V = π r2 l Given a volume equal to 4 m3 , then r 2 l = 1.273 We can see that an infinite number of values for the radius and length will satisfy this equation. To what values should the parameters be set? The answer is not obvious, nor even completely defined with the information given. (This problem will be readdressed in Chap. 10, where the accuracy to which the radius and the length can be manufactured will be used to help find the best values for the parameters.) Let us extend the concept further. It may be that instead of a simple equation, a whole set of equations and rules govern the design. Consider the instance in which a major manufacturer of copying machines had to design paper-feed mechanisms for each new copier. (A paper feed is a set of rollers, drive wheels, and baffles that move a piece of paper from one location to another in the machine.) Many parameters—the number of rollers, their positions, the shape of the baffles, and the like—characterize this particular design problem, but obviously there are certain similarities in paper feeders, regardless of the relative positions of the beginning and end points of the paper, the obstructions (other components in the machine) that must be cleared, and the size and weight of the paper. The company developed a set of equations and rules to aid designers in developing workable paper paths, and using this information, the designers could generate values for parameters in new products.
ullman-38162
ull75741_02
December 17, 2008
16:16
2.4
2.4.4
Different Types of Mechanical Design Problems
Original Design
Any time the design problem requires the development of a process, assembly, or component not previously in existence it calls for an original design. (It can be said that if we have never seen a wheel and we design one, then we have an original design.) Though most selection, configuration, and parametric problems are represented by equations, rules, or some other logical scheme, original design problems usually cannot be reduced to any algorithm. Each one represents something new and unique. In many ways the other types of design problems—selection, configuration, and parametric—are simply constrained subsets of an original design. The potential solutions are limited to a list, an arrangement of components, or a set of related characterizing values. Thus, if we have a clear methodology for performing original design, we should be able to solve any design problem with a more limited set of potential solutions.
2.4.5
Redesign
Most design problems solved in industry are for the redesign of an existing product. Suppose a manufacturer of hydraulic cylinders makes a product that is 0.25 m long. If the customer needs a cylinder 0.3 m long, the manufacturer might lengthen the outer cylinder and the piston rod to meet this special need. These changes may require only parameter changes, or they may require something more extensive. What if the materials are not available in the needed length, or cylinder fill time becomes too slow with the added length? Then the redesign effort may require much more than parameter changes. Regardless of the change, this is an example of redesign, the modification of an existing product to meet new requirements. Many redesign problems are routine; the design domain is so well understood that the method used can be put in a handbook as a series of formulas or rules. The parameter changes in the example of the hydraulic cylinder are probably routine for the manufacturer. The hydraulic cylinder can also be used as an example of a mature design, in that it has remained virtually unchanged over many years. There are many examples of mature designs in our everyday lives: pencil sharpeners, hole punches, and staplers are a few found on the average desk. For these products, knowledge about the design problem is high. There is little more to learn. However, consider the bicycle. The basic configuration of the bicycle—the two tensioned, spoked wheels of equal diameter, the diamond-shaped frame, and the chain drive—was fairly refined late in the nineteenth century. While the 1890 Humber shown in Fig. 2.8 looks much like a modern bicycle, not all bicycles of this era were of this configuration. The Otto dicycle, shown in Fig. 2.9, had two spoked wheels and a chain; stopping and steering this machine must have been a challenge. In fact, the technology of bicycle design was so well developed by the end of the nineteenth century that a major book on the subject, Bicycles and Tricycles: An Elementary Treatise on Their Design and Construction, was
37
ullman-38162
38
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
Figure 2.8 1890 Humber bicycle.
Figure 2.9 The Otto dicycle.
ullman-38162
ull75741_02
December 17, 2008
16:16
2.4
Different Types of Mechanical Design Problems
Figure 2.10 The Marin Mount Vision. (Reprinted with permission of Marin Bicycles.)
Most design problems are redesign problems since they are based on prior, similar solutions. Conversely, most design problems are original as they contain something new that makes prior solutions inadequate. published in 1896.1 The only major change in bicycle design since the publication of that book was the introduction of the derailleur in the 1930s. However, in the 1980s the traditional bicycle design began to change again. For example, the mountain bike shown in Fig. 2.10 no longer has a diamondshaped frame. Why did a mature design like a bicycle begin evolving again? First, customers are always looking for improved performance. Bicycles of the style shown in Fig. 2.10 are better able to handle rough terrain than traditional bikes. Second, there is improved understanding of human comfort, ergonomics, and suspensions. Third, customers are always looking for something new and exciting even if performance is not greatly improved. Fourth, materials and components have improved. The point is that even mature designs change to meet new needs, to attract new customers, or to take advantage of new materials. Part of the design of a new bicycle like the Marin Mount Vision is routine, and part is original. Additionally, 1 The book, written by Archibald Sharp, has recently (1977) been reprinted by the MIT Press, Cambridge, Mass.
39
ullman-38162
40
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
many subproblems were parametric problems, selection problems, and configuration problems. Thus, the redesign of a product, even a mature one, may require a wide range of design activity.
2.4.6
Variant Design
Sometimes companies will produce a large number of variants as their products. A variant is a customized product designed to meet the needs of the customer. For example, when you order a new computer from companies such as Dell, you can specify one of three graphics cards, two battery configurations, three communication options, and two levels of memory. Any combination of these is a variant that is specifically tuned to your needs. Also, Volvo trucks estimates that of the 50,000 parts it has in its inventory it annually supplies over 5000 variants, different truck models specifically assembled to meet the needs of the customer.
2.4.7
Conceptual Design and Product Design
Two other terms that will be used throughout the book are conceptual design and product design. These are catchall terms for two parts of the product development process. First, you must develop a concept and then refine the concept into a product. The activities during the conceptual and product development phases may make use of original, parametric, and selection design and redesign as needed.
2.5 CONSTRAINTS, GOALS, AND DESIGN DECISIONS The progression from the initial need (the design problem) to the final product is made in increments punctuated by design decisions. Each design decision changes the design state. The state of a product is a snapshot of all the information known about it at any given time during the process. In the beginning, the design state is just the problem statement. During the process, the design state is a collection of all the knowledge, drawings, models, analyses, and notes thus far generated. Two different views can be taken of how the design process progresses from one design state to the next. One view is that products evolve by a continuous comparison between the design state and the goal, that is, the requirements for the product given in the problem statement. This philosophy implies that all the requirements are known at the beginning of the design problem and that the difference between them and the current design state can be easily found. This difference controls the process. This philosophy is the basis for the methods in Chap. 6. Another view of the design process is that when a new problem is begun, the design requirements effectively constrain the possible solutions to a subset of all possible product designs. As the design process continues, other constraints are added to further reduce the potential solutions to the problem, and potential solutions are continually eliminated until there is only one final design. In other
ullman-38162
ull75741_02
December 17, 2008
16:16
2.6
Product Decomposition
Constraints are often opportunities in disguise. words, design is the successive development and application of constraints until only one unique product remains. Beyond the constraints in the original problem specifications, constraints added during the design process come from two sources. The first is from the designer’s knowledge of mechanical devices and the specific problem being solved. If a designer says, “I know bolted joints are good for fastening together sheet metal,” this piece of knowledge constrains the solution to bolted joints only. Since every designer has different knowledge, the constraints introduced into the design process make each designer’s solution to a given problem unique. The second type of constraint added during the design process is the result of design decisions. If a designer says, “I will use 1-cm-diameter bolts to fasten these two pieces of sheet metal together,” the solution is constrained to 1-cm-diameter bolts, a constraint that may affect many other decisions—clearance for tools to tighten the bolt, thickness of materials used, and the like. During the design process, a majority of the constraints are based on the results of design decisions. Thus, the individual designer’s ability to make well-informed decisions throughout the design process is essential. Decision-making techniques are emphasized in Chap. 8.
2.6 PRODUCT DECOMPOSITION We will conclude this chapter with a method that can is the basis for understanding existing products. As such, it can serve as a starting point whether doing redesign, original design, or some other type of design, whether at the system or subsystem level. This product decomposition or “benchmarking” method helps us understand how a product is built, its parts, its assembly, and its function. It cannot be overemphasized how important it is to do decomposition and how it is the starting place for all design. In this chapter, we will decompose to understand the parts and assembly. In Chap. 7, the decomposition begun here will be extended to understand function. Figure 2.11 shows a template that can be used to organize the decomposition. It is partially filled in for a pre-2003 version of the Irwin Quick-Grip. This version is the starting point for the redesign effort that resulted in the product shown in Fig. 2.1 The template begins with a brief description of the product and how it works— its function. This follows with a section showing each part. Only a selection of the parts is shown for the clamp in Fig. 2.11. Each part is given a name, the number required, its material, and the manufacturing process. Often it can be hard to determine the material and manufacturing process. For plastics, there is a set of simple experiments for rough identification. Over the last few years, handheld devices have been developed that can identify materials
41
ullman-38162
42
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
Product Decomposition Design Organization: Example for the Mechanical Design Process Date: Aug. 14, 2007 Product Decomposed: Irwin Quick Grip—pre 2007 Description: This is the Quick-Grip Product that has been on the market for many years
How it works: Squeeze the pistol grip repeatedly to move the jaws closer together and
increase the clamping force. Squeeze the release trigger to release the clamping force. The foot (the part on the left in the picture that holds the face that is clamped against) is reversible so the clamping force can be made to push apart rather than squeeze together. Parts: Part #
Part Name
# Req’d.
Material
Mfg. Process
1
Main body
1
PPO or PVC
Injection molded
2
Trigger
1
PVC
Injection molded
4
Face plate, left
1
Polyethylene
Injection molded
Image
ullman-38162
ull75741_02
Part #
8
December 17, 2008
Part Name
16:16
# Req’d.
Material
Mfg. Process
Pad
2
??
Injection molded
13
Power spring
1
Steel
Wound wire
14
Jam plates
2
Steel
Stamped sheet
Image
Disassembly: Step #
Procedure
Part #s removed
1
Take off left face plate
4
12
Remove jam plates and power spring from main body assembly
13, 14, 1
13
Remove trigger from main body assembly
2
14
Pry off pad from main body assembly
8
The Mechanical Design Process Copyright 2008, McGraw-Hill
Image
Designed by Professor David G. Ullman Form # 1.0
Figure 2.11 Product decomposition samples for an older version of the Irwin Quick-Grip.
(Photos reprinted with permission of Irwin Industrial Tools.) 43
ullman-38162
44
ull75741_02
December 17, 2008
CHAPTER 2
16:16
Understanding Mechanical Design
just by pointing the device at a sample of the material. While the main market for these devices is recycling, they are very useful when decomposing a product. Details on these are given in the Sources section at the end of this chapter. The final section of the template is for the disassembly of the product. To build this section of the Product Decomposition report, remove one part at a time. Document the procedure needed to remove the part and the part numbers for those parts removed. Document what was done with a photograph. Figure 2.11 shows only a couple of the steps. Usually disassembly and part naming occur at the same time. Disassembly step 1 shows the left face plate, Part #4, was removed from the product. The internal parts of the clamp can now be seen in the photo. As this is a digital image in the actual template, it can easily be rescaled and studied as needed. Steps 12–14 are shown using a single image. The first one shows the removal of two parts, #13 and #14, at the same time as they come out together. Note how each procedure begins with a verb or verb phrase to tell what has to be done to remove the parts. Make these as descriptive as possible.
2.7 SUMMARY ■
■ ■ ■ ■ ■ ■ ■
■ ■
A product can be divided into functionally oriented operating systems. These are made-up of mechanical assemblies, electronic circuits, and computer programs. Mechanical assemblies are built of various components. The important form and function aspects of mechanical devices are called features. Function and behavior tell what a device does; form describes how it is accomplished. Mechanical design moves from function to form. One component may play a role in many functions, and a single function may require many different components. There are many different types of mechanical design problems: selection, configuration, parametric, original, redesign, routine, and mature. Mechanical objects can be described semantically, graphically, analytically, or physically. The design process is a continuous constraining of the potential product designs until one final product evolves. This constraining of the design space is made through repeated decisions based on comparison of design alternatives with design requirements. Mechanical design is the refinement from abstract representations to a final physical artifact. Product dissection is a useful way to understand the structure of a product.
2.8 SOURCES Good books on designing new products Clausing, Don, and Victor Fey: Effective Innovation: Development of Winning Technologies, ASME Press, 2004.
ullman-38162
ull75741_02
December 17, 2008
16:16
2.10
On the Web
Cooper, Robert G.: Winning at New Products, 3rd ed., Perseus Publishing, 2001. Vogel, C.M., J. Cagan, and P. Boatwright: The Design of Things to Come, Wharton School Publishing, 2005.
Plastics identification The PHAZIR is a handheld, battery-powered, point-and-shoot plastic identifier. It weighs only 4 lb (1.8 kg) and takes 1–2 sec to determine the makeup of the sample. www.polychromix.com
Metals identification The iSort is a handheld, battery-powered, point-and-shoot spectrometer for on-site identification and analysis of all common metal alloys. Metal identification just requires pointing the gun-shaped iSort at a clean metal sample. The iSort is fairly expensive. http://www.spectro.com/pages/e/p010101.htm An inexpensive method uses the color of a chemical deposition to identify the metal. The process requires putting a drop of solution on the sample, then using a battery-powered electric charge through the solution to cause a chemical deposition on a piece of blotter paper. The color of the resulting deposit identifies the metal. http://www.alloyid.com
2.9 EXERCISES 2.1 2.2 2.3
Decompose a simple system such as a home appliance, bicycle, or toy into its assemblies, components, electrical circuits, and the like. Figures 2.3 and 2.11 will help. For the device decomposed, list all the important features of one component. Select a fastener from a catalog that meets these requirements: ■ ■ ■ ■
Can attach two pieces of 14-gauge sheet steel (0.075 in., 1.9 mm) together Is easy to fasten with a standard tool Can only be removed with special tools Can be removed without destroying either base materials or fastener
2.4
Sketch at least five ways to configure two passengers in a new four-wheeled commuter vehicle that you are designing. 2.5 You are a designer of diving boards. A simple model of your product is a cantilever beam. You want to design a new board so that a 150-lb (67-kg) woman deflects the board 3 in. (7.6 cm) when standing on the end. Parametrically vary the length, material, and thickness of the board to find five configurations that will meet the deflection criterion. 2.6 Find five examples of mature designs. Also, find one mature design that has been recently redesigned. What pressures or new developments led to the change? 2.7 Describe your chair in each of the four languages at the three levels of abstraction, as was done with the bolt in Table 2.2.
2.10 ON THE WEB A template for the following document is available on the book’s website: www.mhhe.com/Ullman4e ■
Product Decomposition
45
ullman-38162
ull75741_02
December 17, 2008
16:16
ullman-38162
ull75741_03
December 17, 2008
14:55
C
H
3 A
P
T
E
R
Designers and Design Teams KEY QUESTIONS ■ ■ ■ ■ ■ ■ ■
Why is it important to know how people do design? How is your ability to design dependent on your cognitive preferences? What are the characteristics of creators? How do individual cognitive abilities interact with the abilities of others during team activities? Why is a team more than a group of people? What can you do to help teams be successful? How can you measure team health?
3.1 INTRODUCTION Since the time of the early potter’s wheel, mechanical devices have become increasingly complex and sophisticated. This sophistication has evolved without much concern for how humans solve design problems. Throughout history people who were just naturally good at design were trained, through an apprentice program, to be masters in their art. The design methods they used and the knowledge of the domain in which they worked was refined through their personal experiences and passed, in turn, to their apprentices. Much of this experience was gained through experiments, through building prototypes and then going “back to the drawing board” to iterate toward the next product. The results of these experiments taught the designers what worked and what did not and pointed the way to the next refinement. With this methodology, products took many generations to be refined to the point of mature design. However, as systems grew more complex and the world community grew more competitive, this mode of design became too time-consuming and too expensive. Designers recognized the need to find ways to deal with larger, more complex systems; to speed the design process; and to ensure that the final design 47
ullman-38162
48
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
be reached with a minimum use of resources and time. In this book we discuss design techniques that meet these goals. To understand how these techniques help streamline the design process, it is important to understand how designers and design teams progress from abstract needs to final, detailed products. To put this chapter in context, it is important to realize that design is the confluence of technical processes, cognitive processes, and social processes. We begin our discussion of how humans design mechanical objects by describing a cognitive model of how memory is structured in the individual designer. The types of information that are processed in this structure are explored, and the term knowledge is defined. Once we understand the information flow in human memory, we develop the different types of operations that a designer must perform in memory during the design process, and we explore creativity. Based on this model of the individual’s cognitive process, the chapter moves to the social aspect of design—working in teams. First, the structure of design teams is developed. This includes descriptions of the members of teams and how they are managed. Further, beyond the formal titles that people have, there is a more subtle, cognitive role that people play on teams. Second, an entire section is devoted to building and maintaining a design team. This includes how to start a team, inventory its health, and resolve problems as they develop. Supporting this chapter is a series of templates available at the book’s website.
3.2 THE INDIVIDUAL DESIGNER: A MODEL OF HUMAN INFORMATION PROCESSING The study of human problem-solving abilities is called cognitive psychology. Although this science has not yet fully explained the problem-solving process, psychologists have developed models that give us a pretty good idea of what happens inside our heads during design activities. A simplification of a generally accepted model is shown in Fig. 3.1. This model, called the information-processing system and developed in the late 1950s, describes the mental system used in the solution of any type of problem. In discussing that system here, we give special emphasis to the solution of mechanical design problems. Information processing takes place through the interaction of two environments: the internal environment (information storage and processing inside the human brain) and the external environment. The external environment comprises paper and pencil, catalogs, computer output, and whatever else is used outside the human body to extend the internal environment. In the internal environment, that is, within the human mind, there are two different types of memory: short-term memory, which is similar to a computer’s operating memory (its random access memory or RAM), and long-term memory, which is like a computer’s disk storage. Bringing information into this system from the external environment are sensors, such as the eyes, ears, and hands. Taste and smell are less often used in design. Information is output from the body with the use of the hands and the voice. There are other means of output, such
ullman-38162
ull75741_03
December 17, 2008
3.2
14:55
The Individual Designer: A Model of Human Information Processing
External environment
Internal environment
Sensors Sight
Longterm memory
Shortterm memory
Touch
Controller
Recorders
Notes and drawings
Figure 3.1 The human problem solver.
as body position, that are less often used in design. Additionally, as part of the internal processing capability, there is a controller that manages the information flow from the sensors to the short-term memory, between the short-term and the long-term memory, and between the short-term memory and the means of output. Before describing short-term and long-term memory and the control of information flow, we need to describe the information that is processed in this system. In a computer the information is in terms of bits, or binary digits (0s and 1s), but in the human brain, information is much more complex. In recent experiments, an orthographic drawing of a power transmission system consisting of shafts, gears, and bearings was shown to mechanical engineering students and professional engineers. The students were lower-level undergraduates who had not studied power transmission systems. The drawing was shown briefly and then removed, and the subjects were asked to sketch what they had seen. The students tended to reconstruct the drawings from the line segments and simple shapes they had seen in the original drawing. Not understanding the complexities of geared transmissions, they could not remember anything more complicated. They remembered and drew only the basic form of the components. On the other hand, the professional engineers were able to remember components grouped together by their function. In recalling a gear set, for example, the experts knew that two meshed gears and their associated shafts and bearings provide the function of changing the rpm and torque in the system. They also knew
49
ullman-38162
50
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
what geometry or line segments were needed to represent the form of a gear set. Thus, the experienced engineers using functional groupings were able to include substantially more information than the students in their sketches. The line segments remembered by the students and the functional groupings remembered by the experienced designers are called chunks of information by cognitive psychologists. The greater the expertise of the designer, the more content there is in the chunks of information processed. Exactly what types of information are in these chunks, however, is not always clear. Types of knowledge that might be in a chunk include ■


General knowledge, information that most people know and apply without regard to a specific domain. For example, red is a color, the number 4 is bigger than the number 3, an applied force causes a mass to accelerate—all exemplify general knowledge. This knowledge is gained through everyday experiences and basic schooling. Domain-specific knowledge, information on the form or function of an individual object or a class of objects. For example, all bolts have a head, a threaded body, and a tip; bolts are used to carry shear or axial stresses; the proof stress of a grade 5 bolt is 85 kpsi. This knowledge comes from study and experience in the specific domain. It is estimated that it takes about ten years to gain enough specific knowledge to be considered an expert in a domain. Formal education sets the foundation for gaining this knowledge. Procedural knowledge, the knowledge of what to do next. For example, if there is no answer to problem X, then decomposing X into two independent easier-to-solve subproblems, X1 and X2, would illustrate procedural knowledge. This knowledge comes from experience, but some procedural knowledge is also based on general knowledge and some on domainspecific knowledge. We must often make use of procedural knowledge to solve mechanical design problems.
In mechanical engineering the term feature is synonymous with chunks of information. Since a design feature is some important aspect of a component, assembly, or function, the gear set discussed in the preceding example is both a chunk and a feature. The exact language in which chunks of information are encoded in the brain is unknown. They might be dealt with as semantic information (text), graphical information (visual images), or analytical information (equations or relationships). Psychologists believe that most mechanical designers process information in terms of visual images and that these images are three-dimensional and are readily manipulated in the short-term memory.
All design and decision making is limited by human cognitive capabilities.
ullman-38162
ull75741_03
December 17, 2008
3.2
3.2.1
14:55
The Individual Designer: A Model of Human Information Processing
Short-Term Memory
The short-term memory is the main information processor in the human brain. It has no known specific anatomic location, yet it is known to have very specific attributes. One important attribute of the short-term memory is its quickness. Information chunks can be processed in the short-term memory in about 0.1 second. The term processed implies such actions as comparing one chunk of information to another, modifying a chunk by decomposing it into smaller parts, combining two or more chunks into one new one, changing a chunk’s size or distorting its shape, and making a decision about the chunk. It is unknown how much of the short-term memory is actually used to process the information. We do know that the harder it is to solve the problem, the more short-term memory is used for processing. The capacity of the short-term memory was first described in a paper titled “The Magical Number Seven, Plus or Minus Two” (see Section 3.8), which reported that the short-term memory is effectively limited to seven chunks of information (plus or minus two). This is like having a computer RAM with only seven memory locations. These approximately seven chunks—these seven unique things—are all that a person can deal with at one time. For example, let us say we are working on a design problem and have an idea (a chunk of information, maybe just a word or maybe a visual image) that we want to compare to some constraints on the design (other chunks of information). How many constraints can we compare to the idea in our head? Only two or three at a time, since the idea itself takes one slot in the short-term memory and the constraints take two or three more. That does not leave much memory to do the processing necessary for comparison. Add any more constraints and the processing stops; the short-term memory is simply too full to make any progress on solving the problem. A couple of quick experiments are convincing about the limits of the short term memory. Open a phone book and randomly choose a phone number in which the seven digits are unrelated to each other. (A number such as 555-2000 is not acceptable because the last four digits can be lumped together as a single chunk—two thousand.) After looking at the number briefly, close the phone book, walk across the room, and dial the number. Most people can manage to do this task if they are not interrupted or do not think about anything else. The same experiment can be tried with two unrelated phone numbers. Few people are able to remember them long enough to dial them both since they require dialing 14 pieces of information, which is beyond the capacity of the short-term memory. Granted, these 14 digits can be memorized, or stored in long-term memory, but that would take some study time. Another example of the size limitations of short-term memory is more mechanical in nature. Consider the four-bar linkage of Fig. 3.2. It is made up of four elements: the driver A–B, the link B–C, the follower C–D, and the base D–A. It is not difficult for most engineers to visualize the follower C–D rocking back and forth as the driver A–B is rotated. Point B makes a circle, and point C moves in an arc about point D. An expert on linkages would only use a single
51
ullman-38162
52
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
? F
C ? E B
A
D
Figure 3.2 A four-bar linkage.
chunk to encode this mechanism. But a novice in the domain of four-bar linkages would need to visualize four line segments, using four chunks plus others for processing the motion. To make the task more difficult, trace the path of point E on the link. This requires more short-term memory. Harder still is tracing the path of point F. In fact, this requires so many different parameters to track that only a few linkage experts can visualize the path of point F. Another feature of the short-term memory is the fading of information stored there. The phone number remembered earlier is probably forgotten within a few minutes. To keep from forgetting short-term information, like the phone number, many people keep repeating the information over and over. With such continuous refreshing, it is possible to retain certain objects or parts of objects within the short-term memory and to let only the unimportant information fade to make room for the processing of new chunks of information. Last, it is impossible for us to be aware of what is happening in our shortterm memory while we are solving problems. To follow our own thoughts, we need to use some of that memory to monitor and understand the problem-solving process, making that space no longer available for problem solving. Thus, you can not really observe what you are doing during problem solving without affecting what you are trying to observe.
3.2.2
Long-Term Memory
The long-term memory was earlier compared with the disk storage in a computer; like disk storage, it is for permanent retention of information. Let us look at the four major characteristics of long-term memory. First, long-term memory has seemingly unlimited capacity. Despite the cartoon in Fig. 3.3, there is no
ullman-38162
ull75741_03
December 17, 2008
3.2
14:55
The Individual Designer: A Model of Human Information Processing
Figure 3.3 Long-term memory problems.
documented case of anybody’s brain becoming “full,” regardless of head size. It is hypothesized that as we learn more we unconsciously find more efficient ways to organize the information by reorganizing the chunks in storage. Reconsider the difference between the student’s and the expert’s ways of remembering information about the power transmission system. The expert’s information storage was more efficient than the student’s. The second characteristic of the long-term memory is that it is fairly slow in recording information. It takes 2 to 5 min to memorize a single chunk of information. This explains why studying new material takes so long. The third characteristic is the speedy recovery of information from long term memory. Retrieval is much quicker than storage, the time depending on the complexity of the information and the recentness of its use. It can be as fast as 0.1 sec per chunk of information.
53
ullman-38162
54
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
The fourth characteristic is that the information stored in the long-term memory can be retrieved at different levels of abstraction, in different languages, and with different features. For example, consider the knowledge an average engineer can retrieve about a car (Fig. 3.4). The sample data ranges from images of entire vehicles to semantic rules and equations for diagnosing problems. Human memory is very powerful in matching the form of the data retrieved to that which is needed for processing in the short-term memory.
Car=drive train+body +interior
The body of a Corvette is fiberglass.
Car
65 mph; 0 to 60 in 6 seconds
It’s fast. It handles well.
If engine is running rough, then spark plugs might be fouled.
If car won’t start, then 1. Check fuel level. 2. Check battery.
Figure 3.4 Knowledge stored in memory about cars.
Parts list: #80312—Floatbody #87426—32 jet
hp=
Tn 5252
Tire size= 195/60
ullman-38162
ull75741_03
December 17, 2008
3.2
3.2.3
14:55
The Individual Designer: A Model of Human Information Processing
Control of the Information-Processing System
During problem solving, the controller (Fig. 3.1) enables us to encode outside information obtained through our senses or retrieve information from long-term memory for processing in the short-term memory. Some of the information in the short-term memory is allowed to fade, and new information is input as it is needed and becomes available. Additionally, the controller can help extend the short-term memory by making notes and sketches; these need to be done quickly so that they do not bog down the problem-solving process. When we have completed manipulating the information, the controller can store the results in long-term memory, or in the external environment by describing it in text, verbally, or in graphic images.
3.2.4
External Environment
The external environment—paper and pencil, computers, books—plays a number of roles in the design process: it is a source of information; it is an analytical capability; it is a documentation/communication facility; and, most importantly for designers, it is an extension for the short-term memory. The first three of these roles seem evident; however, the last role, as an extension for the short-term memory, needs some discussion. Because the short-term memory is a space-limited central processor, human problem solvers utilize the external environment as a short-term memory extension, much as a computer extends RAM by using cache memory. This is accomplished by making notes and sketches of ideas and other information needed in problem solving. In order to be useful to the short-term memory, any extension must share the characteristics of being very fast and having high information content. Watch any design engineer trying to solve a problem. He or she will make sketches even when not trying to communicate. These sketches serve as aids in generating and evaluating the ideas by serving as additional chunks of information to be processed. Sketches are fast to make and are information-rich.
3.2.5
Implications of the Model
One of the implications of the information-processing model of human problem solving is that the size of the short-term memory is a major limiting factor in the ability to solve problems. To accommodate this limitation we break down problems into finer and finer subproblems until we can “get our mind around it”—in other words, manage the information in our short-term memory. Typically, these fine-grained subproblems are worked on for about 1 minute before going to the next one. Thus, design of even a simple problem is the solution of many thousands of subproblems. Further, our thinking process has evolved so that, as we solve problems, our expertise about the constraints and potential solutions increases and our configuration of chunks becomes more efficient. This helps offset the “magic number” seven, but human designers are still quite limited. It would almost seem that these limitations would preclude our ability to solve
55
ullman-38162
56
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
If you try to think about what you are doing while you are doing it, you stop doing it. If you don’t reflect on what you just did, you are doomed to repeat it.
complex problems. As discussed in the upcoming sections, processing speed and flexibility of information storage and recovery enable designers to develop very complex products.
3.3 MENTAL PROCESSES THAT OCCUR DURING DESIGN We can now describe what happens when a designer faces a new design problem. The problem may be the design of a large, complex system or of some small feature on a component. We will focus on how a designer understands new information such as the problem statement, how ideas are generated, and how they are evaluated. In Section 1.6 we introduced seven basic actions of problem solving. The core actions—understand, generate, evaluate, and decide—are refined here.
3.3.1
Understanding the Problem
Consider what happens when a new problem is broached. If we think of its design state as a blackboard on which is written or drawn everything known about the device being designed, then the blackboard is initially blank, i.e., the design state is empty. Let us return to the fastening problem presented in Chap. 1 (see Fig. 1.9): Design a joint to fasten together two pieces of 1045 sheet steel, each 4 mm thick and 6 cm wide, that are lapped over each other and loaded with 100 N.
Before any information about the problem is put on the design-state blackboard, the problem statement must be understood. If the problem is outside the realm of experience (the designer does not know what the term lapped means, for example), then the problem cannot be understood. But how do we “understand” a problem? Most likely in this way: As the problem is read, it is “chunked” into significant packets of information. This happens in the short-term memory, where we naturally parse the sentence into phrases like “design a joint,” “to fasten together,” and so on. These chunks are compared with long-term memory information to see if they make sense, and then most are allowed to fade. The goal of this first pass through the problem is to try and retain only the major functions of the needed device. Usually a problem will be read or sensed a number of times until the major function(s) is identified. Unfortunately there is no guarantee that, from the usually incomplete data that
ullman-38162
ull75741_03
December 17, 2008
14:55
3.3
Mental Processes That Occur During Design
exist at the beginning of a design problem, the most important functions will be identified. In our example there is no ambiguity. The prime function is to transfer a load from one sheet of steel to another through a lapped joint. What is important to realize is that a problem is “understood” by comparing the requirements on the desired function to information in the long-term memory. Thus, every designer’s understanding of the problem is different, because each designer has different information stored in the long-term memory. (In Chap. 6 we develop a method to ensure that the problem is fully understood with minimal bias from the designer’s own knowledge.)
3.3.2
Generating Solutions
We have seen that in trying to understand a design problem, we compare the problem to information from the long-term memory. In order to retrieve information from the long-term memory, we need a way to index the knowledge stored there. We can index that information in many ways (Fig. 3.4). As in the gearbox example at the beginning of this chapter, the most efficient indexing method is by function. What are recalled and downloaded to the short-term memory are specific (usually abstract) visual images from past experience. Thus, we search by function and recall form or graphical representations. This is not always true: we can also index our memory by shape, size, or some other form feature. However, in solving design problems, function is usually the primary index. For some problems the information recalled meets all the design requirements and the problem is solved. If, in understanding a problem, we must recall images of previous designs, we have a predisposition to use these designs. Some designers get stuck on these initially recalled images and have difficulty evaluating them objectively and generating other, potentially better ideas. Many of the techniques discussed in Chaps. 7 and 11 are specifically designed to overcome this tendency. On the other hand, what happens if the problem being solved is new and we find no solution to it in the long-term memory? We then use a three-step approach: decompose the problem into subproblems, try to find partial solutions to the subproblems, and finally recombine the subsolutions to fashion a total solution. The subproblems are generally functional decompositions of the total problem. The creative part of this activity is in knowing how to decompose and recombine cognitive chunks.
3.3.3
Evaluating Solutions
Often people generate ideas but have no ability to evaluate them. Evaluation requires comparison between generated ideas and the laws of nature, the capability of technology, and the requirements of the design problem itself. Comparison, then, necessitates modeling the concept to see how it performs with respect to these measures. The ability to model is usually a function of knowledge in the domain. We will address evaluation techniques in Chaps. 8, 10, and 11.
57
58
CHAPTER 3
3.3.4
Designers and Design Teams
Deciding
At the end of each problem-solving activity, a decision is made. It may be to accept an idea that was generated and evaluated, or more likely, it will be to address another topic that is related to the problem. The rationale for how decisions are made is not well understood, but Sections 3.3.5 and 3.3.6 should help clarify what is known.
3.3.5
Controlling the Design Process
To understand how designers progress through a design problem, subjects were videotaped as they worked. In the study of these videotapes, it became evident that the path from initial problem presentation to solution was not very straightforward. It seemed like an almost random process—efforts on a subproblem made the designer aware of another subproblem, and the designer then focused attention on this second problem without having solved the first. No model for the control of focus was found. However, it was clear that the process for some designers is so chaotic that they never find solutions to their problems, while other designers rapidly proceed through the design effort. The techniques discussed in this book are intended to give structure to the design process so that the path from problem statement to solution is as controlled and direct as possible.
3.3.6
Problem-Solving Behavior
Everybody has a unique manner of problem solving. A person’s problem-solving behavior affects how decisions are made individually and has a significant impact on team effectiveness. The following discussion is centered around five personal problem-solving dimensions. These five are useful for describing how an individual solves a design problem because they describe an individual’s information management and decision-making preferences. Since all the team members bring their individual problem-solving processes to team activities, it is the interaction of all the individuals’ solution processes that determines the team’s health. For each of the five dimensions, suggestions for how to counteract extreme behavior are given. Some of these are useful to the individual working alone, and all are important in team situations and will be referenced later in the chapter when we talk about team health. A template for easily evaluating your problem-solving behavior is available. The first personal problem-solving dimension describes an individual’s energy source or extraversion. It is a measure of whether you are an internal or external problem solver. For a rough estimate of your, or a colleague’s, energy source, answer five questions. If scoring a colleague, pretend you are that person. The five questions are shown in Figs. 3.5–3.9, screen shots from the template. In each of the five questions are shown with two potential responses. In Fig. 3.5 the top responses indicate an internal energy source and the bottom responses indicate an external energy source. For the example here, internal is
3.3
Mental Processes That Occur During Design
Figure 3.5 Energy source personal problem-solving dimension.
selected for the first and third questions and so the person is 2/5 or 40% internal and 60% external. In the template, the bar chart updates as you select the responses. If a person is reflective, is a good listener, thinks and then speaks, and enjoys solving problems alone, then she is an internal problem solver. If the person’s energy comes from outside through interactions with others (i.e., the person is sociable and tends to speak and then think) she is an external problem solver. About 75% of all Americans and 48% of engineering students and top executives are external problem solvers. There is no right or wrong style; this is merely the way people operate. They may show slightly different styles in different situations, but will generally not deviate very far from type. In team settings both internals and externals have characteristics that are essential to the team but may cause difficulty—the externals tend to overwhelm the internals, who are reluctant to share their ideas. Here are some suggestions to keep the externals productive but not domineering: ■ ■
■ ■
Externals need to allow others time to think. Point out to them that it is not necessary to fill in all the pauses with words. Externals need to practice listening to the ideas and suggestions of others and pausing before they react. Brainstorming or another creativity-support activity can help here (see Section 7.4). Encourage externals to recap what has been said to make sure they have heard the contributions of others. Externals need to realize that silence does not always mean consent. Sometimes an external will overwhelm the internals, who will become quiet rather than argue the point.
Here are some suggestions to assist internals in getting their ideas out for consideration: ■

Encourage internals to share more than their final response. There is value in thinking out loud, as even the most trivial idea may be part of a good solution. The process will judge the value of the ideas. Try suggesting techniques that enable internals to have an equal say in selecting ideas and plans, such as the techniques in Chaps. 5–12.
59
60
CHAPTER 3

■ ■
Designers and Design Teams
Encourage internals to develop some nonverbal, body-language signals that indicate assent or dissent. Make sure that these signals are understood by other team members. Encourage internals to restate their ideas. This restating signifies to the internal that his or her ideas count and forces the externals to listen. Get internals to push externals for more clarity and meaning.
The second dimension reflects your preference for an information management style or originality. It is a measure of whether you like working with facts or possibilities. For an estimate of your or a colleague’s information management style, answer the questions in Fig. 3.6. For the example shown, the individual operates on both facts and possibilities with a slight tendency for possibilities. People who prefer facts and details are literal, practical, and realistic; they appreciate the here and now. Those who think in terms of possibilities, patterns, concepts, and theories are looking for relationships between pieces of information and the meaning of the information. About 75% of Americans are fact-oriented, as are 66% of top executives; yet only 34% of all engineering students are factoriented. This is interesting in light of the heavy emphasis on math and science that is the focus of an engineering education. Other labels that could be placed on the scale are Preserver and Explorer, where the Preservers maintain the system, the Explorers are the boat rockers. To solve most problems it is important to have a balance between the two extremes. When solving a problem alone, fact-oriented people have trouble getting started, whereas possibility-oriented people have trouble doing the details.This problem-solving dimension is the cause of most miscommunication, misunderstanding, and team problems. Design requires working with both facts and possibilities. Thus, both types of thinking are essential on a design team. However, individuals with a strong tendency toward either extreme may need help in the team setting. Some suggestions for fact-oriented team members are as follows: ■

Encourage fact-oriented team members to fantasize, think wildly, and allow others to think wildly. Wild ideas can lead to good ideas. Brainstorming (Section 7.4) and thinking out loud (rambling) bring out such ideas. Encourage fact-oriented team members to allow the team to set goals rather than dive right into the problem and tackle the details.
Figure 3.6 Information management style problem-solving dimension.
3.3
Mental Processes That Occur During Design
Here are some suggestions for team members who think in terms of possibilities: ■


Encourage possibility-oriented team members to deal with details. The best idea will never reach maturity if the details are not attended to. It is frustrating to them but possibly worthwhile to have them take on the responsibility of a detail task. Force possibility-oriented team members to be specific and avoid generalities. They should be encouraged to try to enumerate the exact items they want to address instead of making sweeping general statements. Remind possibility-oriented team members to stick to the issues. Other team members can control the flow of the problem solving by clearly stating the issues being addressed. Other issues that arise during discussion should be recorded and then shelved for later consideration.
The third dimension measures which information language a person prefers to use, verbal or visual. For a rough idea of your or a colleague’s information language style, answer the questions in Fig. 3.7. The example individual is primarily a visual problem solver, but can work verbally. Visual information includes pictures, diagrams, graphs, and hardware. Verbal information includes written or spoken words and mathematical formulas. It is interesting to note that most people favor visual information, yet most classes in school are presented in a verbal language. This mismatch is especially striking in science and engineering classes. When you are working alone, the language you use is not an important consideration. In teams, however, the preferred languages greatly affect the development of a shared vision of the problem and alternative solutions. Some guidelines on how to manage the two types of communication language in team situations follow. ■ ■
Help identify information that needs to be communicated, regardless of language. Help identify differences in team members’ mental models, encouraging extra effort by both visual and verbal people to communicate clearly with other members to develop a shared understanding.
Figure 3.7 Information language personal problem-solving dimension.
61
62
CHAPTER 3
Designers and Design Teams
Figure 3.8 Deliberation style personal problem-solving dimension. ■
If words and equations aren’t working, try a diagram or picture. If the picture isn’t working, try words and equations.
The fourth dimension reflects the deliberation style or accommodation, the objectivity or subjectivity with which problems are solved. To get an estimate of your or a colleague’s deliberation style, answer the five questions in Fig. 3.8. In the example, the person is primarily an objective problem solver. Some team members take a subjective approach, others an objective one. People who rely on interpersonal involvement, circumstances, and the “right thing to do” take a subjective approach to design. These team members can be referred to as “adaptors.” Conversely, team members who are logical, detached, and analytical take an objective approach to problems. They challenge others when their logic tells them that they are right. About 51% of Americans are objective decision-makers, as are 68% of engineering students and 95% of top executives. As it is important to have a variety of information-collection approaches on a design team, it is equally important to have a range of deliberation styles.Although engineers are trained to make decisions based on objective measures, the greatest number of decisions faced in every design problem have incomplete, inconsistent, qualitative information requiring subjective evaluation. For objective designers the following may help in working with the team: ■



Encourage objective team members to pay attention to the feelings of others. Gut feelings are often right, and sometimes a lack of information forces one to rely on these feelings. Help objective team members understand that how the team functions is as important as what is accomplished. If there is acrimony, no decisions will be made. Remind objective team members that not everyone likes to discuss a topic merely for the sake of argument. Others may drop out from exhaustion and be taken to be conceding the point. Encourage objective team members to express how they feel about the outcome once in a while. Objective decision-makers may have trouble expressing feelings.
Subjective people are in a minority on most design teams. Thus, they must develop techniques to get their opinions heard and not get their sensitivities hurt.
3.3
Mental Processes That Occur During Design
Figure 3.9 Decision closure style personal problem-solving dimension.
Here are some ideas: ■ ■ ■
Help subjective team members to realize that it is all right to disagree and argue. Reassure subjective team members that while harmony is important, not every resolved issue will satisfy everyone even if consensus is reached. Reinforce to subjective team members that discussions about ideas are not personal attacks.
The fifth and final personality dimension relates to the need to actually come to a conclusion during decision making. Decision closure style ranges from flexible to decisive. For a rough estimate of your or a colleague’s decision closure style, answer the questions in Fig. 3.9. Some people are flexible and others are decisive. If a person goes with the flow; is flexible, adaptive, and spontaneous; and finds it difficult to make and stick with decisions, he is considered flexible. If, on the other hand, he makes decisions with a minimum of stress and likes an environment that is ordered, scheduled, controlled, and deliberate, then he is decisive. About half of all Americans are decisive, as are 64% of engineering students and 88% of top executives. One characteristic of flexible decision makers is that they have a tendency to procrastinate because they want to remain adaptive. This can make working with them difficult. The following are some suggestions for flexible decision-makers on the team: ■ ■ ■ ■ ■
Give flexible decision-makers plans in advance so that they can think about them in their own time. Acknowledge the flexible decision-maker’s contribution as a step toward moving to closure. Remind them that problems are solved one step at a time. Set clear decision deadlines in advance. Encourage feedback from flexible decision-makers so that they can think about the direction of their thoughts. Encourage flexible decision-makers to settle on something and live with it a while before redesigning. Encourage them to take a clear position and stick to it. This may be difficult for them to do.
A contrary characteristic of decisive people is that they tend to jump to conclusions. This too can adversely affect teamwork, as many ideas may be generated
63
64
CHAPTER 3
Designers and Design Teams
and consensus may be needed to reach a decision. Here are some suggestions for slowing down decisive people: ■ ■ ■ ■
Ask decisive people questions about their decision process. Remind them that most problems need to be subdivided into smaller problems to be solved. Let decisive people organize the data collection and review process. Utilize techniques, such as brainstorming, that suppress judgment. Do not let them settle on the first good idea they hear. Remind decisive people that they are not always right.
This discussion may seem like a lot of detail for an engineering book. Research has shown, however, that paying attention to the psychological makeup of a team is critical.
3.4 CHARACTERISTICS OF CREATORS Some people seem naturally more creative than others. Before describing the characteristics of a creative design engineer, let us clarify what we mean by “creative.” A creative solution to a problem must meet two criteria: it must solve the problem in question, and it must be original. Solving a problem involves understanding it, generating solutions for it, evaluating the solutions, deciding on the best one, and determining what to do next. Thus, creativity is more than just coming up with good ideas. The second criterion, originality, depends on the knowledge of the designer and of society as a whole. What is new and original to one person may be old hat to another. If someone who has never before experienced a wheel designs one, then it is original for that person. But it is society that assesses “originality” and labels a solution or a person “creative.” As discussed earlier, all humans have the same cognitive, or problem-solving, structure. Why is it, then, that some engineers can generate ingenious ideas while others, who may be brilliant at complex analysis, cannot come up with new concepts no matter how hard they try? There has been a lot of research on creativity, yet this trait is still not very well understood. The best way to understand the results of the research to date is in terms the relationship of creativity to other attributes. Creativity and intelligence. There appears to be little correlation between creativity and intelligence. Creativity and visualization ability. Creative engineers have good ability to visualize, to generate and manipulate visual images in their heads. We have seen before that people represent information in their minds in three ways: as semantic information (words), as graphical information (visual images), and as analytical information (equations or relationships). Words and equations convey serial information. They are generally understood on the basis of word order or the order of variables and constants. Pictures, or visual images, on the other hand, contain parallel information—you can see many different
3.4
Characteristics of Creators
The odds are greatly against you being immensely smarter than everyone else. —John R. Page, Rules of Engineering
things in a single image. Some people are very good at decomposing and manipulating visual images in their heads, whereas others are not. It appears, however, that the ability to manipulate complex images of mechanical devices can be improved with practice. This may be related to the formation of more information-rich chunks having functional information or to some other mechanism. Creativity and knowledge. The model of the information-processing system implies that all designers start with what they know and modify this to meet the specific problem at hand. At every step of the way, the process involves small movements away from the known, and even these small movements are anchored in past experience. Since creative people form their new ideas out of bits of old designs, they must retain a storehouse of images of existing mechanical devices in their long-term memory. Thus, in order to be a creative mechanical designer, a person must have knowledge of existing mechanical products. Additionally, part of being creative is being able to evaluate the viability of ideas. Without knowledge about the domain, the designer cannot evaluate the design. Knowledge about a domain is only gained through hard work in that domain. Thus, a firm foundation in engineering science is essential to being a creative designer of mechanical devices. For example, during World War II many people sent ideas for weapons to the Department of War. Some were very far-fetched ideas for death rays or for building 5-mile-high walls or domes over Europe to stop the bombers. These were very original but unworkable and were therefore not creative. The “inventors” had good intentions but lacked the knowledge to develop creative solutions to the war problems. Creativity and partial solution manipulation. Since new ideas are born from the combination of parts of existing knowledge, the ability to decompose and manipulate this knowledge seems to be an important attribute of a creative designer. This attribute, more than any other so far discussed, appears to become stronger with exercise. Although there is no scientific evidence to support this contention, anecdotal evidence does support it. Creativity and risk taking. Another attribute of creative engineers is the willingness to take an intellectual chance. Fear of making a mistake or of spending time on a design that in the end does not work is characteristic of a noncreative individual. Edison tried hundreds of different lightbulb designs before he found the carbon filament. Creativity and conformity. Creative people also tend to be nonconformists. There are two types of nonconformists: constructive nonconformists and
65
66
CHAPTER 3
Designers and Design Teams
obstructive nonconformists. Constructive nonconformists take a stand because they think they are right. Obstructive nonconformists take a stand just to have an opposing view. The constructive nonconformist might generate a good idea; the obstructive nonconformist will only slow down the design progress. Creative engineers are constructive nonconformists who may be hard to manage since they want to do things their own way. Creativity and technique. Creative designers have more than one approach to problem solving. If the process they initially follow is not yielding solutions, they turn to alternative techniques. A number of books listed in Section 3.7 give methods to enhance creativity. Many of the techniques covered in these are woven into the mechanical design techniques presented in the remainder of this book. This is especially true in the chapters on concept and product generation (Chaps. 7 and 9). Creativity and environment. If the work environment allows risk taking and nonconformity and encourages new ideas, creativity will be higher. Further, if teammates and other colleagues are creative, the environment for creativity is greatly enhanced. In the discussion of teams in Section 3.5, it is stated that, on a team, the sum is greater than the parts. This is especially true for creativity. Creativity and practice. Creativity comes with practice. Most designers find that they have creative phases in their careers—periods when they have many good ideas. During these times the environment is supportive and one good idea builds on another. However, even with a supportive environment, practice enhances the number and quality of ideas. To summarize, the creative designer is generally a visualizer, a hard worker, and a constructive nonconformist with knowledge about the domain and the ability to dissect things in his or her head. Even designers who do not have a strong natural ability can develop creative methods by using good problem-solving techniques to help decompose the problem in ways that maximize the potential for understanding it, for generating good solutions, for evaluating the solutions, for deciding which solution is best, and for deciding what to do next. One final comment: There are many design tasks that require talents very different from those used to describe a creative person. Design requires much attention to detail and convention and demands strong analytic skills. Therefore, there are many good designers who are not particularly creative individuals; a design project requires people with a variety of skills and talents.
3.5 THE STRUCTURE OF DESIGN TEAMS The material already covered describes an individual designer. However, because of the complexity of most products, design work is generally done by design teams. As shown in Fig. 3.10, the complexity of mechanical devices has grown rapidly over the last 200 years. Gone are the days when a single individual could design an entire product. Even Edison had a team of others that worked with
3.5
The Structure of Design Teams
107 Space shuttle
Number of components
106
Boeing 777
105
Boeing 747
104
Wright Brothers “Flyer”
103 100
V2 Automobile
Bicycle Sewing machine
10 1
DC 3
Musket (51 parts)
Springfield rifle (140 parts)
0 1800 1820 1840 1860 1880 1900 1920 1940 1960 1980 2000 Year
Figure 3.10 Increasing complexity in mechanical design.
him. For example, the Boeing 777 aircraft, which has over 5 million components, required over 10 thousand person-years of design time. Thousands of designers worked over a three-year period on the project. Obviously, a single designer could not approach this effort. Modern design problems require a design team—a small number of people with complementary skills who are committed to a common purpose, common performance goals, and a common approach for which they hold themselves mutually accountable. A team is a group of people with complementary skills who are committed to a common purpose, performance goals, and approach for which they hold themselves mutually accountable. A group is not necessarily a team. Groups that interact primarily to share information and to help each individual perform within his or her area of responsibility is not a team. An effective team is more than the sum of it parts. Important points about teams are the following: 1. Teamwork is central to success in engineering as most problems are made of many interdependent subparts, all of which must be solved concurrently. Teams bring together complementary skills and experiences, which are needed to solve many engineering problems. 2. Management takes risks in forming teams as a team must be empowered to make decisions, removing this responsibility from the management. 3. Teams establish communication to support real-time problem solving. 4. Teams develop decisions by consensus rather than by authority. This leads to more robust decisions. In the most basic sense, teams solve design problems in the same way an individual does—understanding, generating, evaluating, and decision making.
67
68
CHAPTER 3
Designers and Design Teams
A team is a group of people in search of a common understanding.
However, there are some important differences. ■



Team members must learn how to collaborate with each other. Collaboration means more than just working together—it means getting the most out of other team members. The suggestions that follow help develop a collaborative team. Teams are generally empowered to make decisions. Since these are team decisions, members must compromise to reach them. Empowering teams to make these decisions means that management takes a risk in giving up responsibility for them. Further, developing decisions by consensus rather than by authority leads to more robust decisions. Team members must establish communication to support real-time problem solving. Further, members need to ensure that the others have the same understanding of design ideas and evaluations that they have. It is very difficult for people with different areas of expertise to develop a shared vision of the problem and its potential solutions. Developing this shared vision requires the development of a rich understanding of the problem. It is important that team members and management be committed to the good of the team. If they are not, it will be difficult reaching the other team goals.
To address what is special about teams, in this chapter we first itemize the different technical roles people play on teams and then, in Section 3.6, we address building teams and maintaining team health.
3.5.1
Members of Design Teams
In this section, we list the individuals who might fill a role on a product design team. The roles on a design team will vary with product development phase and from product to product, and the titles will vary from company to company. Each position on the team is described as if filled by one person. In a large design project, there may be many persons filling that role, whereas in a small project one individual may fill many roles. Product design engineer. The major design responsibility is carried by the product design engineer (hereafter referred to as the design engineer). This individual must be sure that the needs for the product are clearly understood and that engineering requirements are developed and met by the product. This usually requires both creative and analytical skills. The design engineer must bring knowledge about the design process and knowledge about specific technologies to the project. The person who fills this position usually has a four-year engineering degree. In smaller companies he or she may be a nondegreed designer who has extensive experience in the product area. For most product design projects, more than one design engineer will be involved.
3.5
The Structure of Design Teams
Product manager. In many companies, this individual has the ultimate responsibility for the development of the product and represents the major link between the product and the customer. Because the product manager is accountable for the success of the product in the marketplace, he or she is also often referred to as the marketing manager or the product marketing manager. The product manager is often from the sales or customer service department. In order to initiate a design project, management must appoint the nucleus of a design team—at a minimum, a design engineer and a product manager. Manufacturing engineer. Design engineers generally do not have the necessary breadth or depth of knowledge about various manufacturing processes to fully support the design of most products. This knowledge is provided by the manufacturing or industrial engineer, who must have a grasp not only of in-house manufacturing capabilities but also of what the industry as a whole has to offer. Designer. In many companies, the design engineer is responsible for specification development, planning, conceptual design, and the early stages of product design. The project is then turned over to designers, who finish detailing the product and developing the manufacturing and assembly documentation. Designers are often CAD experts with two-year technology degrees. At some companies designers are the same as design engineers. Technician. The technician aids the design engineer in developing the test apparatus, performing experiments, and reducing data in the development of the product. The insights gained from the technician’s hands-on experience are usually invaluable. Materials specialist. In some products, the choice of materials is forced by availability. In others, materials may be designed to fit the needs of the product. The more a product moves away from the use of known, available materials, the more a materials specialist is needed as a member of the design team. This individual is usually a degreed materials engineer or a materials scientist. Often the materials specialist will be a vendor’s representative who has extensive knowledge about the design potential and limitations of the vendor’s materials. Many vendors actually provide design assistance as part of their service. Quality control/quality assurance specialist. A quality control (QC) specialist has training in techniques for measuring a statistically significant sample to determine how well it meets specifications. This inspection is done on incoming raw materials, incoming products from vendors, and products produced in-house. A quality assurance (QA) specialist makes sure that the product meets any pertinent codes or standards. For example, for medical products, there are many FDA (Food and Drug Administration) regulations that must be met. Often QC and QA are covered by one person. Analyst. Many engineers work as analysts. Analysts usually perform complex mathematical studies of design performance using finite-element
69
70
CHAPTER 3
Designers and Design Teams
methods, thermal system modeling, or other advanced software. They are generally specialists who focus on one type of system or method. Industrial designer. Industrial designers are responsible for how a product looks and how well it interacts with consumers; they are the stylists who have a background in fine arts and in human factors analysis. They often design the envelope within which the engineer has to work. Assembly manager. Where the manufacturing engineer is concerned with making the components from raw materials, the assembly manager is responsible for putting the product together. As you will see in Chap. 11, concern for the assembly process is an important aspect of product design. Vendor’s or supplier’s representatives. Very few products are made entirely in one factory. In fact, many manufacturers outsource (i.e., have suppliers provide) 70% or more of their product. Usually there will be many suppliers of both raw and finished goods. There are three types of relationships with suppliers: (1) partnership—the supplier takes part in the process beginning with requirements and concept development; (2) mature—the supplier relies on the parent company’s requirements and concepts to develop needed items; and (3) parental—the supplier builds only what the parent company specifies. Often it is important to have critical suppliers on the design team, as the success of the product may be highly dependent on them. As Fig. 3.11 illustrates, having a design team made up of people with varying views may create difficulties, but teams are essential to the success of a product.
Figure 3.11 The design team at work.
3.5
The Structure of Design Teams
The breadth of these views helps in developing a quality design. Part of the promise of PLM is to help all these different contributors communicate in a consistent and productive manner.
3.5.2
Design Team Management
Since projects require team members with different domains of expertise, it is valuable to look at the different structures of teams in an organization. This is important because product design requires coordination across the functions of the product and across the phases in the product’s development process. Listed next are the five types of project structures. The number in parentheses is the percentage of development projects that use that type. These results are from a study of 540 projects in a wide variety of industries. Functional organization (13%). Each project is assigned to a relevant functional area or group within a functional area. A functional area focuses on a single discipline. For aircraft manufacturers, Boeing, for example, the main functions are aerodynamics, structures, payload, propulsion, and the like. The project is coordinated by functional and upper levels of management. Functional matrix (26%). A project manager with limited authority is designated to coordinate the project across different functional areas or groups. The functional managers retain responsibility and authority for their specific segments of the project. Balanced matrix (16%). A project manager is assigned to oversee the project and shares with the functional managers the responsibility and authority for completing the project. Project and functional managers jointly direct many work-flow segments and jointly approve many decisions. Project matrix (28%). A project manager is assigned to oversee the project and has primary responsibility and authority for completing the project. Functional managers assign personnel as needed and provide technical expertise. Project team (16%). A project manager is put in charge of a project team composed of a core group of personnel from several functional areas or groups, assigned on a full-time basis. The functional managers have no formal involvement. Project teams are sometimes called “Tiger teams,” “SWAT teams,” or some other aggressive name, because this is a high-energy structure and the team is disbanded after the project is completed. What is important about these structures is that some of them are more successful than others. Structures focused on the project are more successful than those built around the functional areas in the company (Fig. 3.12). Here the balanced matrix, project matrix, and project teams resulted in a higher percentage of success across all measures. Thus, when planning for a design project, organize the talent around the project whenever possible.
71
CHAPTER 3
Designers and Design Teams
80 70 Success (percentage)
72
60 50 40 30 20
Schedule
Cost
Technical
Functional
Functional matrix
Project matrix
Project team
Overall Balanced matrix
Figure 3.12 Project successes versus team structure.
3.6 BUILDING DESIGN TEAM PERFORMANCE It can be very exciting being part of a team that is productive and is making good use of all the members. Conversely, it can be hellish working on a team that is not functioning very well. So the goal of this section is to help you build and maintain successful teams. To help ensure success, we will use Team Contracts, Team Meeting Minutes, and Team Health Assessments. Each of these encourages behavior that leads to a successful team experience. According to a leading book on teams, there are ten characteristics of a successful team. Included in the description of each of these characteristics is a guide to where this text presents material to help make teams successful. 1. Clarity in goals. The process developed in this book focuses on goals during process planning in Chap. 4 and for the product itself in Chap. 6. Further, the Team Contract suggested later in this section encourages documenting the immediate team goals. 2. Plan of action. Chapter 4 is all about project planning. 3. Clearly defined roles. We have already discussed roles, and documenting them is part of the Team Contract. 4. Clear communication. Team Contracts, Team Meeting Minutes, and Team HealthAssessments (all in this chapter) plus virtually all the process methods in this book are designed to help with communication. 5. Beneficial team behaviors. As with communication, the material in this book is designed to result in beneficial behaviors. 6. Well-defined decision process. The decision process is introduced in Chap. 4 and is the focus of Chap. 8.
3.6
Building Design Team Performance
7. Balanced participation. Equal division of work is very important for a successful team. This is further discussed later in this chapter. 8. Established ground rules. This is discussed later in this chapter. 9. Awareness of team process. This is what we are talking about in this entire chapter. 10. Use of sound generation/evaluation approach. As introduced in Chap. 1, the seven activities of the design process are: Establish the Need, Plan, Understand, Generate, Evaluate, Decide, and Document. Generate and Evaluate are covered in Chaps. 7–12. To set the foundation for future work, the remainder of this chapter covers Team Contracts, Team Meeting Minutes, and Team Health Assessments.
3.6.1
Team Contract
A good starting point for a team is with a team contract. Team contracts are seldom done in industry because the basics of it are assumed in the employment contract, and it is further assumed that people know how to work to make a team successful. Here we will use a contract as both a learning tool and as a way to increase the odds of team success. Figure 3.13 shows an example Team Contract. The first section is for the assignment of roles on the team and goals for the team. As suggested in the list of characteristics for a successful team, the goals and roles need to be known and agreed to. Roles can be developed from the list in Section 3.5 and make the goals be as specific as possible. Much in Chaps. 4 and 5 focuses on goals. In the second section the team members sign, indicating they agree to a list of performance expectations that are shown in the example. Additionally, the form has room for other expectations to which the team may want to agree. The final section on the form is for strategies for conflict resolution. Hopefully these won’t be needed, but, like any other contract, methods for problem resolution need to be addressed at the beginning to prevent difficulties later. Suggested strategies are shown in the example.
3.6.2
Team Meeting Minutes
Before a meeting begins, it is essential to have an agenda. Without an agenda, meetings wander and it is often not clear whether anything was accomplished. Thus, the purpose of the first section in the team meeting minutes (Fig. 3.14) is to itemize the agenda. Agendas should be written in terms of the goals of the meeting. Agenda items such as “Present the results of the stress analysis” are not sufficient. Why are the results being presented? What is to be accomplished by telling others the results? It is better to state this in terms of what is to be accomplished: “Decide how the stress affects the assembly’s performance” or “Determine if the stress is low enough to meet the requirements of the system.” The second section itemizes the high points of the discussion. To understand why taking notes about the high points is so important, consider the results of an
73
74
CHAPTER 3
Designers and Design Teams
Team Contract Design Organization: The B Team Team Member
Date: Jan. 2, 2009
Roles
Signature
Jason Smathers
Lead designer
Jason Smathers
Brittany Spars
Structural engineer
Brittany Spars
Deon Warner
Systems engineer
Deon Warner
Team Goals
Responsible Member
1. Develop layout and initial input to solid model.
JS
2. Analyze for fatigue and other failures.
BS
3. Detail latching mechanism.
JS
4. Develop wiring plan.
DW
5. Team Performance Expectations
Initial
• Strive to complete all assigned tasks before or by deadlines.
JS
BS
DW
• Complete all tasks to the best of ability.
JS
BS
DW
• Listen carefully and attentively to all comments at meetings.
JS
BS
DW
• Accept and give criticism in a professional manner.
JS
BS
DW
• Focus on results before the fact, rather than excuses after.
JS
BS
DW
• Provide as much notice as possible of commitment problems.
JS
BS
DW
• Attend and participate in all scheduled group meetings.
JS
BS
DW
Strategies for Conflict Resolution • Amend contract with deadlines for agreed to tasks. • Reward entire team for goals met with some treat or social gathering. • As a team, go to a higher authority for assistance with a team problem. • Don’t kill messengers. Seek to encourage the airing of problems.
The Mechanical Design Process Copyright 2008, McGraw-Hill
Figure 3.13 Example team contract.
Designed by Professor David G. Ullman Form # 2.0
ullman-38162
ull75741_03
December 17, 2008
14:55
3.6
Building Design Team Performance
Team Meeting Minutes Design Organization: The C Team
Date: Jan. 30, 2009
Agenda
1. Finalize the plan for the exotherm system. 2. Decide on the final shape for the housing. 3. Resolve how to complete task 3. 4. Plan the postproject party. 5. 6. Discussion: Jason, Brittany, and Deon attended. The meeting lasted an hour. The agen-
da was fully covered and new issues were added to the list for the next meeting. Decisions Made
1. Exotherm plan finalized. See Attachment A. 2. Housing alternative 3 was chosen. 3. Action Items
Person Responsible
Deadline
Jason details Housing alternative 3
JS
Thursday
Brittany to plan party
BS
2/10
Deon will assist Brittany to get Task 3 completed by Thursday
BS
Thursday
Team member: Jason Smathers
Date for next meeting: Thursday
Team member: Brittany Spars Team member: Deon Warner Team member:
The Mechanical Design Process Copyright 2008, McGraw-Hill
Figure 3.14 Team meeting minutes.
Designed by Professor David G. Ullman Form # 3.0
75
ullman-38162
76
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
experiment where a group was asked 2 weeks after a meeting to recall specific details of that meeting. In recounting the meeting they ■ ■ ■ ■ ■
Omitted 90% of the specific points that were discussed. Recalled half of what they did remember incorrectly. Remembered comments that were not made. Transformed casual remarks into lengthy orations. Converted implicit meanings into explicit comments.
Recording the decisions made is even more important. Often decisions are clear. For example, “Choose to use 5056-T6 aluminum for the brace” or “The potential difference on anode and cathode of the X-ray tube will be 140 keV.” However, if you listen carefully to unstructured meetings, you find that they wander from topic to topic. When one topic gets difficult because some of the parties disagree or more information is needed, the conversation moves to another topic with no resolution of the initial topic. If stuck, decide what to do to get unstuck and record that call for action. For example, “A decision was made to gather more information on material x” or “We will use Belief Maps to help the team work toward agreement.” These decisions lead directly to the most important item in the meeting minutes, the action items—an itemized list of what is to happen next. State each action item as a clear deliverable, assign the responsible party, and determine by when it is to be done.
3.6.3
Team Health Assessment
One of the most important activities is assessing the team’s heath. A form for assessing team health is shown in Fig. 3.15. This form includes 17 measures (with room for more) to be assessed periodically by the team to measure how it is doing. For each measure, the response ranges from strongly agree to strongly disagree, with attention needed to remedy problems in areas where at least one person does not agree with the measure. The team needs to devise remedies for these “problem areas.” Not doing so allows problems to fester and worsen. This assessment should be used periodically and especially when any team members experience one of the following: ■ ■ ■ ■ ■ ■ ■ ■
A loss of enthusiasm A sense of helplessness A lack of purpose or identity Meetings in which the agenda is more important than the outcome Cynicism and mistrust Interpersonal attacks made behind peoples backs Floundering Overbearing or reluctant team members
ullman-38162
ull75741_03
December 17, 2008
14:55
3.6
Building Design Team Performance
Team Health Assessment Team Assessed:
Date:
SA = Strongly Agree, A = Agree, N = Neutral, D = Disagree, SD = Strongly Disagree, NA = Not Applicable Measure
SA A
N
D SD NA
1 Team mission and purpose are clear, consistent and attainable. 2 I feel that I am part of a team. 3 I feel good about the team’s progress. 4 Respect has been built within the team for diverse points of view. 5 Team environment is characterized by honesty, trust, mutual respect, and team work. 6 The roles and work assignments are clear. 7 Team treats every member’s ideas as having potential value. 8 Team encourages individual differences. 9 Conflicts within the team are aired and worked to resolution. 10 Team takes time to develop consensus by discussing the concerns of all members to arrive at an acceptable solution. 11 Decisions are made with input from all in a collaborative environment. 12 The environment encourages communication and does not “kill the messenger” when the news is bad. 13 When one team member has a problem others jump in to help. 14 Dysfunctional behavior is dealt with in an appropriate manner. 15 When someone on the team says they are going to do something, the team can count on it being done. 16 There is no “them and us” on the team. 17 Our team cultivates a “what we can learn” attitude when things do not go as expected. 18 19 20 Remedies for improving the Neutral (N), Disagree (D) and Strongly Disagree(SD) responses: Assessor:
The Mechanical Design Process Copyright 2008, McGraw-Hill
Figure 3.15 Team health assessment.
Designed by Professor David G. Ullman Form # 3.0
77
ullman-38162
78
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
3.7 SUMMARY ■ ■ ■
■ ■
■ ■
■ ■
The human mind uses the long-term memory, the short-term memory, and a controller in the internal environment when problem solving. Knowledge can be considered to be composed of chunks of information that are general, domain-specific, or procedural in content. The short-term memory is a small (seven chunks, features, or parameters) and fast (0.1-sec) processor. Its properties determine how we solve problems. We use the external environment to augment the size of the short-term memory. The long-term memory is the permanent storage facility in the brain. It is slow to remember, it is fast to recall (sometimes), and it never gets full. Creative designers are people of average intelligence; they are visualizers, hard workers, and constructive nonconformists with knowledge about the problem domain. Creativity takes hard work and can be aided by a good environment, practice, and design procedures. Because of the size and complexity of most products, design work is usually accomplished by teams rather than by individuals. Working in teams requires attention to every team member’s problem-solving style (including yours)—introverted or extroverted, fact or possibility, verbal or visual, objective or subjective, or decisive or flexible. It is important to have team goals and roles, keep meeting minutes, and assess team health. Many activities can help build team health.
3.8 SOURCES Adams, J. L.: Conceptual Blockbusting, Norton, New York, 1976. A basic book for general problem solving that develops the idea of blocks that interfere with problem solving and explains methods to overcome these blocks; methods given are similar to some of the techniques in this book. Larson, E., and D. Gobeli: “Organizing for Product Development Projects,” Journal of Product Innovation Management, No. 5, pp. 180–190, 1988. The study in Section 3.5.2 on design team management is from this paper. Koberg, D., and J. Bagnall: The Universal Traveler: A Systems Guide to Creativity, Problem Solving and the Process of Reaching Goals, Kaufman, Los Altos, Calif., 1976. A general book on problem solving that is easy reading. Miller, G. A.: “The Magical Number Seven, Plus or Minus Two: Some Limits on Our Capacity for Processing Information,” Psychological Review, Vol. 63, pp. 81–97, 1956. The classic study of short-term memory size, and the paper with the best title ever. Newell, A., and H. Simon: Human Problem Solving, Prentice Hall, Englewood Cliffs, N.J., 1972. This is the major reference on the information processing system. A classic psychology book. Plous, S.: The Psychology of Judgment and Decision Making, McGraw-Hill, New York, 1993. The importance of meeting notes example is from this interesting book. Weisberg, R. W.: Creativity: Genius and Other Myths, Freeman, San Francisco, 1986. Demystifies creativity; the view taken is similar to the one in this book.
ullman-38162
ull75741_03
December 17, 2008
14:55
3.9
Exercises
The next five titles are all good books on developing and maintaining teams. Belbin, R. M.: Management Teams, Heinemann, New York, 1981. Cleland, D. I., and H. Kerzner: Engineering Team Management, Van Nostrand Reinhold, New York, 1986. Johansen, R., et al.: Leading Business Teams, Addison-Wesley, New York, 1991. Katzenbach, J. R., and D. Smith: The Wisdom of Teams, Harvard Business School Press, 1993. Scholtes, P. R., et al.: The Team Handbook, 3rd edition, Oriel Inc, 2003.
The problem-solving dimensions in Section 3.3.5 are based on the MyersBriggs Type Indicator. These titles give more details on this method. Keirsey, D., and M. Bates: Please Understand Me, 5th ed., Prometheus Nemesis, 1978. Kroeger, O., and J. M. Thuesen: Type Talk at Work, Delta, 1992. Kroeger, O., and J. M. Thuesen: Type Talk, Delta, 1989.
3.9 EXERCISES 3.1 3.2 3.3
3.4 3.5
Develop a simple experiment to convince a colleague that the short-term memory has a capacity of about seven chunks. Think of a simple object, write about it, and sketch it in as many ways as possible. Refer to Table 2.1 and Fig. 3.4 to encourage a range of language and abstraction. Describe a mechanical design problem to a colleague. Be sure to describe only its function. Have the colleague describe it back to you in different terms. Did your colleague understand the problem the same way as you? Was the response in terms of previous partial solutions? During work on a team, identify the secondary roles each person is playing. Can you identify who fills each role? For a new team begin with these team-building activities. a. Paired introductions. Get to know each other by asking questions such as ■ What is your name? ■ What is your job (class)? ■ Where did you grow up (go to school)? ■ What do you like best about your job (school)? ■ What do you like least about your job (school)? ■ What are your hobbies? ■ What is your family like? b. Third-party introductions. Have one member of the team tell another the information in (a). Then the second member introduces the first member to the rest of the team using all the information that he or she can remember. It makes no difference if the team heard the initial introduction. c. Talk about first job. Have each member of the team tell the others about his or her first job or other professional experience. Information such as this can be included: ■ What did you do? ■ How effective was your manager? ■ What did you learn about the real world?
79
ullman-38162
80
ull75741_03
December 17, 2008
CHAPTER 3
14:55
Designers and Design Teams
d.
3.6
“What I want for myself out of this.” Have each member of the team tell the others for 3 to 5 min what his or her goals are for participation in the project. What do they want to learn or do, and why? Consider personal goals such as getting to know other people, feeling good about oneself, learning new skills, and other nontask goals. e. Team name. Have each person write down as many potential team names as possible (at least five). Discuss the names in the team, and choose one. Try to observe who plays which secondary role. Pick an item from the team health assessment. For that item, one member of a four-person team checks “Strongly Disagrees.” Develop a list of actions you would take as a team leader or team member.
3.10 ON THE WEB Templates for the following documents are available on the book’s website: www.mhhe.com/Ullman4e ■ ■ ■ ■
Personal Problem Solving Dimensions Team Contract Team Meeting Minutes Team Health Assessment
ullman-38162
ull75741_04
December 17, 2008
15:4
C
H
4 A
P
T
E
R
The Design Process and Product Discovery KEY QUESTIONS ■ ■ ■ ■ ■ ■
What are the six phases of the mechanical design process? What are the three prime sources for new products? What does it mean for a product to be “mature”? How can a SWOT analysis help choose which products to develop? How did Benjamin Franklin contribute to decision making? What are the six basic decision-making activities?
4.1 INTRODUCTION In this chapter, we introduce the major phases in the design process and tackle the first of them, discovering the need. The six-phase design process established here sets the structure for the rest of this book. Since design is fundamentally the effort to fulfill a need, discovering the need is always the first phase in the process. Because there are always more needs than there are resources to meet them, key here is deciding which product ideas to develop. Thus, in this chapter we also introduce the basics of decision making. Making good decisions is probably the most important and least studied engineering skill. We will refine decision making when choosing a concept and again when making Product Development decisions.
4.2 OVERVIEW OF THE DESIGN PROCESS Regardless of the product being developed or changed, or the industry, there is a generic set of phases that must be accomplished for all projects. These are listed 81
ullman-38162
82
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Design is a process—not just building hardware. —Tim Carver, OSU student, 2000
in Fig. 4.1. They are a refinement of the phases in a product’s life cycle (Fig. 1.8) that are of concern to the designer. For each phase, there are a series of activities that need to be accomplished. The phases and activities are briefly introduced in this chapter and refined throughout the rest of the book. After this introduction, the first phase, Product Discovery, is explained in detail. This design process, as shown, applies to design of systems, subsystems, assemblies, and components. It applies to new, innovative products and to changes in existing products. Of course, the detail and emphasis will change with the level of decomposition and with the amount of change needed. To help introduce the phases and how they are used at all levels in a product’s decomposition consider the design of a General Electric CT Scanner.
Product Discovery
Project Planning
Product Definition
Conceptual Design
Product Development
Product Support
Figure 4.1 The mechanical design process.
ullman-38162
ull75741_04
January 1, 2009
9:53
4.2
Overview of the Design Process
General Electric designs and manufactures many different types of products including home appliances, lightbulbs, jet engines, and a host of medical products. One of the products developed by GE’s healthcare business is the CT Scanner shown in Fig, 4.2. The full name of the technology used in this scanner is X-ray Computed Tomography (CT). CT is a diagnostic imaging technique that can produce solid images of the organs inside patients. A CT system consists of a patient table that can be positioned and moved through the bore of the gantry. Beneath the sleek outer casing, the gantry houses a frame that holds an X-ray tube and a detector. The X-ray tube is on the top at the 1 o’clock position in Fig. 4.3 and the arc-shaped detector is on the bottom at the 7 o’clock position. The frame, X-ray tube, and detector rotate around the patient at 120 rpm. This means that there is a centrifugal acceleration on the components of more than 10gs. Thus, the X-ray tube components experience very large radial body loads and covey centrifugal loading to the gantry support of approximately 2000 N of radial force. In order to generate images of organs the tube emits rays that pass through the patient, are sensed by the detector, and are processed by a computer, as shown in Fig. 4.4. To accomplish this, the X-ray tube emits bursts of X-rays. During emission, the tube requires 60–100 kW of power. This power must be transmitted to the rotating tube, where the majority of the power is converted into waste heat that must be transferred out of the gantry. Making the design task even more
Figure 4.2 GE CT Scanner. (Source: Reprinted with permission of GE Medical.)
83
ullman-38162
84
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Figure 4.3 The insides of a CT gantry. (Source: Reprinted with permission of GE Medical.)
Monitor
X-ray tube
Detectors
Computer
Figure 4.4 How a CT works.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.2
Overview of the Design Process
difficult, the anode in the X-ray tube is rotating on an axis perpendicular to the plane of the gantry at 7000–10,000 rpm. The bearings for the anode are operating in a vacuum at a temperature of 450◦ C. The design of the X-ray tube is a tremendous undertaking requiring hundreds of design and manufacturing engineers, materials scientists, technicians, purchasing agents, drafters, and quality-control specialists, all working over several years. To recap: ■ ■ ■ ■ ■
The system is the CT Scanner. Major subsystems are the patient table and the gantry. A major assembly in the gantry is the frame with the X-ray tube and detector. The X-ray tube itself is a subsystem in the frame assembly. Two components in the X-ray tube are the anode and its bearings.
Regardless of which of these are being designed or changed, there will be a Plan, Product Definition, and Conceptual Design before there are products. These phases, itemized in Fig. 4.1, are common to the design of every system, subsystem, assembly, and component. Let’s expand each of the phases.
4.2.1
Product Discovery
Before the original design or redesign of a product can begin, the need for it must be established. As shown in Fig. 4.5, there are three primary sources for design projects: technology, market, and change. We will delve into these sources later in this chapter. Regardless of the source, a common activity at most companies is maintaining a list of potential projects. Since companies have limited people and money, the second activity, after identifying the products, is choosing which Technology push
Market pull
Itemize projects
Product change
Develop more product ideas
Choose project
To project planning
Figure 4.5 The Discovery phase of the mechanical design process.
85
ullman-38162
86
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
of them to work on. Sometimes this decision comes before Project Planning as structured in this book, and sometimes it is postponed until later, after Planning, Product Definition, and Conceptual Design has been done and more is known about each of the options. The ordering of work on these phases will be further discussed in Chap. 5. The GE CT Scanner is a mature product, but new products are advancing the state of the art at a rapid rate. For mature products, design changes focus on improved reliability, cost, and supply chain management. New products are pulled by new imaging applications and performance capability. For the X-ray tube itself, changes are usually in response to market pull for more detailed X-rays and faster times. These system-level needs are projected down as projects to redesign the X-ray tube. Within these projects, the needs are communicated as specifications for higher power, more rotational speed, better heat removal, and other technical changes.
4.2.2
Project Planning
The second phase is to plan so that the company’s resources of money, people, and equipment can be allocated and accounted for (Fig 4.6). Planning needs to precede any commitment of resources; however, as with much design activity, this requires speculating about the unknown—and that makes the planning for a product that is similar to an earlier product easier than planning for a totally new one. Since planning requires a commitment of people and resources from all parts of the company, part of the planning is forming the design team. As discussed in Chap. 3, few products or even subsystems of products are designed by one person. Additionally, much planning work goes into developing a schedule and estimating the costs. The final goal of the activities in this phase is generating a set of tasks that need to be performed and a sequence for them. Planning is covered in detail in Chap. 5. The plan for redesigning the X-ray tube is very complex as it is usually only a small part of the plan to redesign the entire CT Scanner to create the next model. Thus, the tasks, schedule, and budget must integrate with many other similar plans.
4.2.3
Product Definition
During the product definition phase (Fig. 4.7), the goal is to understand the problem and lay the foundation for the remainder of the design project. Understanding the problem may appear to be a simple task, but since most design problems are poorly defined, finding the definition can be a major undertaking. In Chap. 6, we will look at a technique to accomplish this. Using this technique, the first activity will be to identify the customers for the product. This activity serves as the basis to generate the customers’ requirements. These requirements are then used to evaluate the competition and to generate engineering specifications, measurable behaviors of the product-to-be that, later in the design process, will help in
ullman-38162
ull75741_04
December 17, 2008
15:4
4.2
Overview of the Design Process
Identify the tasks
Develop teams
Develop schedule
Estimate time
Sequence tasks
Refine plan
Approve plan
To product definition
Cancel project
Figure 4.6 The Project Planning phase of the mechanical design.
determining product quality. Finally, in order to measure the “quality” of the product, we set targets for its performance. Often, the results of the activities in this phase determine how the design problem is decomposed into smaller, more manageable design subproblems. Sometimes not enough information is yet known about the product, and decomposition occurs later in the design process. In redesigning the X-ray tube, the needs are translated into realizable targets for power, rotational speed, heat removal, and other technical specifications. These specifications are developed in concert with other design teams that need to supply the power, structurally support and power the rotating X-ray tube, and dispose of the waste heat.
4.2.4
Conceptual Design
Designers use the results of the Planning and Product Definition phases to generate and evaluate concepts for the product or product changes (Fig 4.8). When we generate concepts, the customer’s requirements serve as a basis for developing a
87
ullman-38162
88
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Identify customers
Generate customers’ requirements
Evaluate competition
Generate engineering specifications
Refine product definition
Set targets To conceptual design Approve specifications
Cancel project
Figure 4.7 The Product Definition phase of the mechanical design.
Developing a concept into a product without prior effort on the earlier phases of the design process is like building a house with no foundation.
functional model of the product. The understanding gained through this functional approach is essential for developing concepts that will eventually lead to a quality product. Techniques for concept generation are given in Chap. 7. After we evaluate concepts, the goal is to compare the concepts generated to the requirements developed during Product Definition and make decisions.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.2
Overview of the Design Process
Generate concepts
Evaluate concepts
Refine concepts
Make concept decisions
Document and communicate
Refine plan Refine specifications
To product design Approve concepts
Cancel project
Figure 4.8 The Conceptual Design phase of the mechanical design process.
Concept decisions are made with limited knowledge.As shown in Fig. 1.11 knowledge increases with time and effort. One goal in Conceptual Design is choosing the best alternatives with the least expenditure of time and other resources needed to gain knowledge. Techniques helpful in concept evaluation and decision making are in Chap. 8. During projects to redesign the X-ray tube, concepts are small changes to existing products, and the X-ray design team at GE uses detailed analytical models to evaluate them. However, for the Mars Rover, introduced in Chap. 2, new wheel concepts were dramatically different from those previously used and concept evaluation was much less analytical than at GE.
89
ullman-38162
90
ull75741_04
December 17, 2008
CHAPTER 4
4.2.5
15:4
The Design Process and Product Discovery
Product Development
After concepts have been generated and evaluated, it is time to refine the best of them into actual products (see Fig. 4.9). The Product Development phase is discussed in detail in Chaps. 9–11. Unfortunately, many design projects are begun here, without benefit of prior specification or concept development. This design approach often leads to poor-quality products and in many cases causes costly changes late in the design process. It cannot be overemphasized: Starting a project by developing product, without concern for the earlier phases, is poor design practice. At the end of the Product Development phase, the product is released for production. At this time, the technical documentation defining manufacturing,
Generate product
Evaluate product For performance and robustness
For cost
For other DFX
For production
Make product decisions
Document and communicate
Refine concept
Release for production approval
To product support
Cancel project
Figure 4.9 The Product Development phase of the mechanical design process.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.2
Overview of the Design Process
assembly, and quality control instructions must be complete and ready for the purchase, manufacture, and assembly of components. The GE design team used refined analytical models and component system testing during Product Development. Their final prototypes generally use actual production processes and production lines to fabricate final prototypes. This helps to ensure they capture the expected product quality and not be misled by “laboratory” produced prototypes.
4.2.6
Product Support
The design engineer’s responsibility may not end with release to production. Often there is continued need for manufacturing and assembly support, support for vendors, and help in introducing the product to the customer (see Fig. 4.10). Additionally, design engineers are usually involved in the engineering change process. This is the process where changes made to the product, for whatever reason, are managed and documented. This is one of the Product Support topics discussed in Chap. 12.
Develop design documentation
Support vendors, customers, and manufacturing and assembly
Maintain engineering changes
Apply for patents
Retire product
Figure 4.10 The Product Support
phase of the mechanical design process.
91
ullman-38162
92
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Finally, the designers may be involved in the retirement of the product. This is especially true for products that are designed for specialized short-term use and then decommissioning. But, as pointed out in the Hannover Principles, this should be a concern regardless of the product throughout the design process. Whereas the GE team must continue to support the X-ray tubes that are in use, a better example of postproduction support is the Mars Rover. The two Rovers were designed for 90 Mars days of operation. As of this writing, they have both lasted over 3.5 years. One of the Rovers is operating with only five of its original six legs providing power as one drive motor has ceased to function. The other has lost one of its four steering motors. In both cases, the engineers had to figure out how to change the Rovers from Earth to compensate for the failures. This is an extreme example of postproject Product Support. Before refining the first phase, Product Discovery, later in the chapter, some justification is in order for why a product needs to be developed carefully through these six phases.
4.3 DESIGNING QUALITY INTO PRODUCTS A good design process will support designing quality into the product. Traditionally, quality has been the concern of Quality Control (QC) or Quality Assurance (QA). QC/QA specialists inspect products as they are being manufactured and assembled. They check for conformance with the technical documentation (i.e., drawings, material properties, and other specifications) developed during design. They check dimensions, material properties, surface finishes, and other factors that are critical for form and function. This is often referred to as “inspecting quality into a product.” It is less expensive and much more effective to design quality into a product. This implies not only designing a product that works as it should, lasts a long time, and meets the other customer desires listed in Table 1.1, but it also means designing the components and assemblies so they are easy to make, they have few or no tightly toleranced dimensions, and they have few critical (i.e., prone to failure) features. Finally, designing quality into a product also implies designing the product so that it is easy and foolproof to assemble. Many engineering best practices help design quality into a product. Table 4.1 itemizes techniques generally considered as best practice and discussed in this text. They appear in the order in which they are generally applied to a typical design problem. However, each design problem is different, and some techniques may not be applicable to some problems. Additionally, even though the techniques are described in an order that reflects sequential and specific design phases, they
Quality cannot be manufactured or inspected into a product, it must be designed into it.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.3
Designing Quality into Products
Table 4.1 Best practices presented in this text
Project Planning (Chap. 5) Generating a product development plan Managing the project Specification Development (Chap. 6) Understanding the design problem Developing customer’s requirements Assessing the competition Generating engineering specifications Establishing engineering targets Conceptual Design Generating concepts (Chap. 7) Functional decomposition Generating concepts from functions Evaluating concepts (Chap. 8) Judging feasibility Assessing technology readiness Using the decision matrix Robust decision making
Product Development Product generation (Chap. 9) Form generation from function Form representation Materials and process selection Vendor development Product evaluation (Chaps. 10 and 11) Functional evaluation Evaluating performance Tolerance analysis Sensitivity analysis Robust design Design for cost Design for value Design for manufacture Design for assembly Design for reliability Design for test and maintenance Design for the environment Product Support (Chap. 12) Developing design documentation Maintaining engineering changes Applying for a patent Design for end of product life
are often used in different order and in different phases. Understanding the techniques and how they add quality to the product aids in selecting the best technique for each situation. The techniques described in this text comprise a design strategy that will help in the development of a quality product that meets the needs of the customer. Although these techniques will consume time early in the design process, they may eliminate expensive changes later. The importance of this design strategy is clearly shown in Fig. 4.11, a reprint of Fig. 1.5. Figure 4.11 shows that Company A structures its design process so that changes are made early, while Company B is still refining the product after it has been released to production. At this point, changes are expensive, and early users are subjected to a low-quality product. The goal of the design process is not to eliminate changes but to manage the evolution of the design so that most changes come through iterations early in the process. The techniques listed in Table 4.1 also help in developing creative solutions to design problems. This may sound paradoxical, as lists imply rigidity and creativity implies freedom, however, creativity does not spring from randomness. Thomas Edison, certainly one of the most creative designers in history, expressed it well: “Genius,” he said, “is 1% inspiration and 99% perspiration.” The inspiration for creativity can only occur if the perspiration is properly directed and focused. The techniques presented here
93
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Start production
94
ull75741_04
Changes
ullman-38162
Company A Ideal effort
Company B Actual project hours
Time
Figure 4.11 Engineering changes during automobile development.
help the perspiration occur early in the design process so that the inspiration does not occur when it is too late to have any influence on the product. Inspiration is still vital to good design. The techniques that make up the design process are only an attempt to organize the perspiration. These techniques also force documentation of the progress of the design, requiring the development of notes, sketches, informational tables and matrices, prototypes, and analyses—records of the design’s evolution that will be useful later in the design process. In the 1980s, it was realized that the process was as important as the product. One result of this realization is that Product Development is now often referred to as integrated product and process development or IPPD. Note that the term process is on equal footing with the product. Note also that IPPD implies that the product and process are under development. They are evolving. Another result of this awareness is the increasing use of the International Standard Organization’s ISO 9000, the quality management system. ISO 9000 was first issued in 1987 and now has been adopted by most countries. There are millions of companies with ISO-9000 certification worldwide. All major manufacturing companies are ISO-9000 certified and, regardless of size, any company involved in international Product Development or manufacturing is also. Prior to 2000 there were five standards numbered 9000 through 9005. In 2000, these were reduced to: ISO 9000, fundamentals and vocabulary; ISO 9001, requirements; and ISO 9004, guidance for performance improvement. ISO-9000 registration means that the company has a quality system that 1. Standardizes, organizes, and controls operations. 2. Provides for consistent dissemination of information.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.4
Product Discovery
3. Improves various aspects of the business-based use of statistical data and analysis. 4. Enhances customer responsiveness to products and service. 5. Encourages improvement. Companies decide to seek ISO-9000 certification because they feel the need to control the quality of their products and services, to reduce the costs associated with poor quality, or to become more competitive. Also, they may choose this path simply because their customers expect them to be certified or because a regulatory body has made it mandatory. In order to receive the certification, they must first develop a process that describes how they develop products, handle product problems, and interact with customers and vendors. Among the materials that must be prepared are written procedures that ■
■ ■
Describe how most work in the organization gets carried out (i.e., the design of new products, the manufacture of products, and the retirement of products). Control distribution and reissue of documents. Design and implement a corrective and preventive action system to prevent problems from recurring.
Once this material is developed the company invites an accredited external auditor (registrar) to evaluate the effectiveness of the process. If the auditors like what they see, they will certify that the quality system has met all of the ISO’s requirements. They will then issue an official certificate. The company can then announce to the world that the quality of their products and services is managed, controlled, and assured by a registered ISO-9000 quality system. The certification typically expires after three years. Also, the registration agency typically requires surveillance audits at six-month intervals to maintain the currency of the certificate. It must be made clear that ISO 9000 does not give a plan or process for developing products. It only requires a company to have a documented Product Development process on which the plan for a particular product can be based. The certification is not on the quality of the process itself, but that it exists, is maintained, and is used. Thus, a company can have a very poor methodology for developing products and still be certified. However, it is assumed that if a company is going to go to the trouble to get certified and wants to remain competitive in its markets, it will work to make this process and its Product Development plans as good as it can.
4.4 PRODUCT DISCOVERY The goal of Product Discovery (Fig. 4.12), the first phase in the design process, is to develop a list of design projects that includes new products and product changes, and to choose which projects to work on. The term “discovery” may sound odd,
95
ullman-38162
96
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Technology push
Market pull
Itemize projects
Product change
Develop more product ideas
Choose project
To project planning
Figure 4.12 The Product Discovery phase of the mechanical
design process.
but every design effort begins with the discovery of a need for product. There are three prime sources for new products: market pull, technology push, and product change. Market pull occurs when there is customer demand for new products or product features. About 80% of new product development is market-driven. Without a customer for the product, there is no way to recover the costs of design and manufacture. Conversely, technology push is when a new technology is developed before there is customer demand. Let us refine these two product sources. To manage market pull, the sales and marketing departments of most companies have a long list of new products or product improvements that they would like. When they see customers purchase a competitor’s product, they wish their products had the unique features found on that product. Further, if they are doing a good job, they project the customer demand into the future. If sales and marketing had their way, there would be a continuous flow of product improvements and new products so that all potential customers could be satisfied. In fact, this is the direction that product development has been taking for the last few years—near-custom products with short development time. At the same time, engineers and scientists have ideas for new products and product improvements based on technology. Rather than being driven by the customer, these ideas are driven by new technologies and what is learned during the design process. In fact, most product-producing companies spend from 2% to 10% of their revenue on research and development. And, since design is learning, by the time a designer finishes with a project, she or he knows enough to improve it. Most engineers would like to have a second chance at each project so they can, based on their new understanding, do it better the second time.
ull75741_04
December 17, 2008
15:4
4.4
Product Discovery
When a company wants to develop a product without market demand, utilizing a new technology, they are forced to commit capital investment and possibly years of scientific and engineering time. Even though the resulting ideas may be innovative and clever, they are useless unless they can be matched to a market need or a new market can be developed for them. Of course, devices such as sticky notes and many other products serve as examples of products that have been successfully introduced without an obvious market need. While these types of products have high financial risk, they can reap a large profit because of their uniqueness.
4.4.1
Product Maturity
Let’s explore the need for new products further by examining the technology maturity “S” curve shown in Fig. 4.13. This shows the stages a technology matures through as it goes from a new product to a mature product. Products are often introduced to the market while some of the technologies it uses are still in the “make it work properly” stage, some even sooner. Product changes and improvements occur as technologies mature over time. Think of each of these improvements as redesign projects—they are. By the time a technology begins to reach maturity, the market is saturated with competition and companies need to decide if they are going to continue to develop using the existing technologies or innovate, develop new technologies, and begin the “S” curve again, as shown in Fig. 4.14. If companies stay with the current technologies and further refine them, they probably have much competition and little room for improvement. If they innovate, they are taking a risk as the product matures.
4.4.2
Kano’s Model of Customer Satisfaction
Another way to look at the need for product development is to examine Kano’s Model of Customer Satisfaction. The Kano model was developed by Dr. Noriaki Kano in the early 1980s to describe customer satisfaction. This model will help us understand how and why features mature. Kano’s model plots customer
Mature product
Technology maturity
ullman-38162
MINIMIZE COST MAXIMIZE RELIABILITY MAXIMIZE EFFICIENCY MAXIMIZE PERFORMANCE MAKE IT WORK PROPERLY New product
MAKE IT WORK Time
Figure 4.13 Product maturity “S” curve.
97
98
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Continue to mature
Innovate
Figure 4.14 A decision point on the “S” curve.
Delighted T im
Customer satisfaction
ullman-38162
e
Ex
e c it
me
Absent
nt
ce
an
m or
rf
Pe
Fully implemented
si Ba
c
Disgusted
Product function Figure 4.15 The Kano diagram for customer
satisfaction.
satisfaction, from disgusted to delighted, versus product function, from absent to fully implemented, as shown in Fig. 4.15. This plot shows three lines representing basic features, performance features, and excitement features. Basic features refer to customers’ requirements that are not verbalized as they specify assumed functions of the device. The only time a customer will mention them is if they are missing. If they are absent in the final product, the customer will be disgusted with it. If they are included, the customer will be neutral. An example is the requirement that a car should have brakes. If there are no brakes, then the customer is going to be disgusted with the product (and may be injured also). Brakes are expected on cars and so, just being there is not a cause for delight, just a neutral reaction from the customer. However, how well the brakes perform is a concern.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.4
Product Discovery
Requirements for performance features are verbalized in the form that the better the performance, the better the product. For example, a requirement on brake stopping distance is clearly a performance requirement. Generally, the shorter the distance, the more delighted the customer. But what if your car would apply the brakes when you said so? What if you said, “Slow down” and the car gently decelerated, and when you yelled, “STOP!” it braked hard? This capability is unexpected. If it is absent, the customers are neutral because they don’t expect it anyway. However, if customers’ reactions to the final product are surprise and delight at the additional functions, then the product’s chance of success in the market is high. Requirements for excitement-quality features are often called “wow requirements.” If you went to a car showroom and testdrove a car with voice-activated brakes, this would be unexpected. Your reaction to the system would be “wow.” If the system worked well, you would be delighted, if it were not there at all, you wouldn’t know the difference and so would be neutral. Excitement-level features on a product generally require new technologies. Over time, excitement-level features become performance-level features and, ultimately, basic features. This is true for most features of home entertainment systems, cars, and other consumer products. When first introduced, a new feature is special in one brand and consumers are surprised and delighted. The next year, as the technology matures, every brand has the feature and some perform better than others. Companies then work the “S” curve to improve performance, efficiency, reliability, and cost. After a few years, the feature is not even mentioned in advertising because it is an expected feature of the product. The Kano model is just another view of technology maturity. Companies need to make decisions about whether to invest in innovation to “wow” customers or improve performance, efficiency, reliability, and cost and work their way farther up the “S” curve. In addition to market pull and technology push, the third source of design projects is in response to the need for a change. There are three major sources for product changes: ■


A vendor can no longer supply materials or components used in the product or has recommended improved ones. This may require the development of new plans, specifications, and concepts. Manufacturing, assembly, or another downstream phase in the product’s life cycle has identified a quality, time, or cost improvement that results in a cost-effective change in the product. The product fails in some way and the design needs to be changed. This type of change can be very costly. Reflect back to Fig. 4.11, where the automobile manufacturer was still making design changes after release for production. As discussed there, these changes are very expensive.
Change-driven projects are so important that an entire section will be devoted to them in Chap. 12.
99
ullman-38162
100
ull75741_04
December 17, 2008
CHAPTER 4
4.4.3
15:4
The Design Process and Product Discovery
Product Proposal
Regardless of the source, one deliverable from this phase of the design process is the product proposal. A template for developing such a proposal is available and is shown with a simplistic example in Fig. 4.16. Note in this example that there is sufficient information to at least initiate discussions about how much resources should be allocated to following up on this proposal. In a real situation, much more documentation would be needed on each of these items.
Product Proposal Design Organization: xxxxxxxx
Date: June 23, 2010
Proposed Product Name: The Toastalator Summary: Customers who live in small spaces and have the need in the morning to
both make coffee and cook toast. The concept here is for a device that combines these two products in a small space. Background of the Product: Observations of people living in small apartments have
revealed an opportunity to minimize the space used when preparing breakfast. Since we manufacture both coffee makers and toasters this seems like a reasonable opportunity to pursue. Market for the Product: Although there is no firm evidence, there is anecdotal demand for this product. Studies of space availability and market size are needed. An initial survey shows the potential for up to 10 million customers. Competition: There is no known product such as this on the market today. And an initial
patent survey has shown no recent activity with similar products. Manufacturing Capability: XXXXX currently manufactures similar products independently. Distribution Details: XXXX as distribution channels for similar products. Proposal Details:
Task 1: Develop better market numbers. Task 2: Develop project plans through the Conceptual Design phase. Task 3: Develop product definition. Task 4: Develop and evaluate a proof-of-concept prototype. Team member:
Prepared by:
Team member:
Checked by:
Team member:
Approved by:
Team member:
The Mechanical Design Process Copyright 2008, McGraw-Hill
Figure 4.16 The product proposal template.
Designed by Professor David G. Ullman Form # 8.0
ullman-38162
ull75741_04
December 17, 2008
15:4
4.5
Choosing a Project
4.5 CHOOSING A PROJECT The hard part of this phase of the design process is deciding which projects to undertake and which to leave for later. We all think we make good decisions. It has been estimated, however, that over half of all decisions fail! A failed decision is later remade or the results ignored altogether. Failed decisions result in lost time and cost money. Any time you revisit a decision, all the work, tooling, prototypes, and CAD models made in the interim have little value. During the Product Definition phase, there are usually more product ideas than there are time, people and money to do them all. The goal here is to choose which projects to undertake and which to leave for later or not attempt at all. This effort is commonly called project portfolio management, where a portfolio is a list of potential projects and the goal is to decide which of them to undertake. To choose the best we need to know how to make decisions. We will introduce good decision-making practice and then specifically address portfolio decisions, the key decision needed during discovery. We will revisit decision making in Conceptual Design and then again during Product Development, adding to what we learn here. In the remainder of this section, three methods will be presented that can help in choosing a project from the portfolio. The first two are simple, but somewhat limited. The third sets the foundation for decision-making processes that will be developed later in the book.
4.5.1
SWOT Analysis
The first decision support method we will use to help us choose a project is called a SWOT analysis. SWOT stands for Strengths, Weaknesses, Opportunities, and Threats. This method is commonly used in business, can be applied to the evaluation of single projects, and is easy to do. The basics of the method are to list the four SWOT items on a quadchart (each of four quadrants filled in with SWOT entries), as shown in Fig. 4.17 and then informally weigh the strengths versus the weaknesses and the opportunities versus the threats. As an example in the figure, a bicycle manufacturing company is considering adding a tandem bicycle to its product line. Filling out a SWOT analysis makes it easier to judge whether or not a single potential project should be undertaken. Although this method does lay out the major points to consider when for decision making, it does not actually help in making the decision. It is still not clear whether or not BURL should undertake building a tandem bicycle.
Design is the technical and social evolution of information punctuated by decision making.
101
ullman-38162
102
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
SWOT Analysis Design Organization: BURL Bicycles
Date: Nov. 11, 2007
Topic of SWOT Analysis: Explore the potential for adding a tandem bicycle to the
product line in 2008. Strengths:
Weaknesses: • Market for tandems is small, ⬍1% of
• BURL has the technology to design
a top quality tandem bicycle.
all bicycle sales.
• BURL’s engineers want to do this project.
• The profit margin may be smaller
• It will expand the product line.
than on traditional bikes.
• Market for tandems is growing,
• Cost to develop may exceed
$40,000.
although no exact market numbers have been collected.
• Pay back time is estimated at 3 years.
• For the most part, they can be made
• It will take 6 months to get to mar-
with current equipment and processes.
ket, missing the current sales season.
• We can use our patented suspension to
• A tandem is just different enough to
differentiate BURL’s tandem from the rest. Opportunities: • A tandem will open BURL into new
markets. • A tandem might allow bike shops that
carry BURL to expand business and order more bikes.
need unique marketing and shipping. Threats: • The product is not unique enough to
attract customers. • We can’t get bike shops to carry them. • It will cost more than $40,000 to
develop. • Engineering can’t get it to ride like a
CLIEN. Team member: Fred Flemer
Prepared by: Fred Flemer
Team member: Bob Ksaskins
Checked by: Bob Ksaskins
Team member:
Approved by: Betty Booper
The Mechanical Design Process Copyright 2008, McGraw-Hill
Designed by Professor David G. Ullman Form # 11.0
Figure 4.17 SWOT diagram example.
4.5.2
Pro-Con Analysis
To take the SWOT analysis one step further, consider a pro-con analysis. An early, recorded use of this type of analysis is by Ben Franklin. Besides being a statesman, he was a designer of stoves, bifocals and many other inventions. In a 1772 letter to Joseph Priestly (the discoverer of oxygen), Franklin explained how he analyzed his problems when intuition failed him.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.5
Choosing a Project
Dear Sir: In the affair of so much importance to you, where in you ask my advice, I cannot, for want of sufficient premises, advise you what to determine, but if you please I will tell you how. When those difficult cases occur, they are difficult, chiefly because while we have them under consideration, all the reasons pro and con are not present to the mind at the same time; but sometimes one set present themselves, and at other times another, the first being out of sight. Hence the various purposes or information that alternatively prevail, and the uncertainty that perplexes us. To get over this, my way is to divide a sheet of paper by a line into two columns; writing over the one Pro, and over the other Con. Then, during three or four days consideration, I put down under the different heads short hints of the different motives, that at different times occur to me, for or against the measure. When I have thus got them all together in one view, I endeavor to estimate their respective weights; and when I find two, one on each side, that seem equal, I strike them both out. If I find a reason pro equal to some two reasons con, I strike out the three. If I judge some two reasons con, equal to three reasons pro, I strike out the five; and thus proceeding I find at length where the balance lies; and if, after a day or two of further consideration, nothing new that is of importance occurs on either side, I come to a determination accordingly. And, though the weight of the reasons cannot be taken with the precision of algebraic quantities, yet when each is thus considered, separately and comparatively, and the whole lies before me, I think I can judge better, and am less liable to make a rash step, and in fact I have found great advantage from this kind of equation . . .
Franklin considers whether to accept or reject a single alternative. This is really a choice between two alternatives: do this or do something else (including nothing). Franklin advises five steps for making a decision: Step 1: Make two columns on a sheet of paper and label one “Pros” and the other “Cons.” Step 2: Fill in the columns with all the pros and cons of an alternative. Step 3: Estimate the importance of each pro and each con. Step 4: Eliminate pros and cons this way: a. When two are of about equal importance, cross them both out and b. Find other importance equalities of pros and cons—for example, the importance of two pros equals three cons—and then strike them out. Step 5: When one or the other column becomes dominant, then “come to the determination accordingly.” You can extend the idea of using pro-con lists to include more than one alternative, but the balancing step quickly becomes complex. Still, NASA frequently uses this approach to help organize experts when evaluating multiple project proposals. For each proposal, the experts list the pros and cons. They then informally balance the pros and cons to differentiate among the alternatives. This helps to tease out the good and bad points.
103
ullman-38162
104
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Pro-Con Analysis Design Organization: BURL Bicycles
Date:
Topic of Pro-Con Analysis: Should BURL market a tandem bicycle? Pro:
Con:
• BURL has the technology to design a
top-quality tandem bicycle. • BURL’s engineers want to do this project. • It will expand the product line. • Market for tandems is growing,
although no exact market numbers have been collected. • For the most part they can be made
with current equipment and processes. • We can use our patented suspension to
differentiate BURL’s tandem from the rest. A tandem will open BURL into new markets. • A tandem might allow bike shops that
carry BURL to expand business and order more bikes.
• Market for tandems is small, ⬍1% of
all bicycle sales. • The profit margin may be smaller than
for traditional bikes. • Cost to develop may exceed $40,000. • Pay-back time is estimated at 3 years. • It will take 6 months to get to market,
missing the current sales season. • A tandem is just different enough to
need unique marketing and shipping. • The product is not unique enough to
attract customers. • We can’t get bike shops to carry them. • It will cost more than $40,000 to
develop. • Engineering can’t get it to ride like a
BURL. Team member: Fred Flemer
Prepared by: Fred Flemer
Team member: Bob Ksaskins
Checked by: Bob Ksaskins
Team member:
Approved by: Betty Booper
The Mechanical Design Process Copyright 2008, McGraw-Hill
Designed by Professor David G. Ullman Form # 9.0
Figure 4.18 Pro-con analysis example.
If you look back at the SWOT analysis, the statements there are all an argument either for or against designing and marketing a tandem bicycle. In Fig. 4.18, these are reordered on Pro-Con Analysis Template. Thus, we have already completed steps 1 and 2 of Franklin’s method. Step 3 forces you to put a value on how important each of the pro and con statements is to the success of the project in preparation for step 4. For example, in looking down the list, it appears that Market for tandems is growing although no exact market numbers have been collected.
is about as important as A tandem is just different enough to need unique marketing and shipping.
ullman-38162
ull75741_04
December 17, 2008
15:4
4.5
Choosing a Project
So, according to step 4, they need to be crossed out. Then, BURL has the technology to design a top-quality tandem bicycle. and BURL’s engineers want to do this project.
is about as important as Engineering can’t get it to ride like a BURL.
So they too need to be crossed out. Continuing this way BURL ultimately sees that the cons outweigh the pros and decides not to undertake a tandem project.
4.5.3
Basics of Decision Making
Although the two methods just presented begin to get the information organized for good decision making, they are both limited to one alternative. In this section, we will formalize the entire decision-making process and make a protocol decision. The basic structure of decision making is the same, whether addressing discovery issues or concept selection or choosing product details. In each case, there are six basic activities. Let’s look at these activities in more detail: 1. 2. 3. 4. 5. 6.
Clarify the issue that needs a satisfactory solution. Generate alternatives—itemize the potential solutions for the issue. Develop criteria as they measure a satisfactory solution for the issue. Identify criteria importance of each criterion relative to the others. Evaluate the value of the alternatives by comparing them to the criteria. Based on the evaluation results, decide what to do next. This decision will direct the process to a. Add, eliminate, or refine alternatives. b. Refine criteria. c. Refine evaluation—work to gain consensus and reduce uncertainty. d. Choose an alternative—you’ve made a decision, document it and address other issues.
These are shown in a flow diagram in Fig. 4.19. We will reuse this list of activities and this diagram numerous times throughout the book. Comparing the SWOT analysis to this ideal flow, SWOT is limited to activities 1, 2, and 5. It addresses only a single alternative and never actually itemizes the criteria for evaluation, even though they are inherent in the SWOT statements. (As we shall see in a moment.) SWOT focuses informally on the evaluation and never really gets to “what to do next.” Thus, it is not really a decision-making method by our definition, even though it supports some of the activities. The pro-con method adds concern for the importance (activity 4) of the statements and gives a limited idea of “what to do” to the process (activity 6).
4.5.4
Making a Portfolio Decision
Here we will apply the activities listed in the previous section to the bicycle example.
105
ullman-38162
106
ull75741_04
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
1. Clarify the issue
2. Generate alternatives
3. Develop criteria
4. Identify criteria importance Add, eliminate or refine alternatives
Refine criteria
5. Evaluate alternatives relative to criteria
Refine evaluation
6. Decide what to do next
Choose an alternative
Move to next issue
Figure 4.19 The decision-making flow.
Activity 1. BURL clarifies the issue. This was already done earlier, but we will make it broader here: “Choose, from a list of alternative product development projects, which one should be undertaken first?” In general, an issue is a question that needs to be addressed with some object or course of action chosen to answer the question and resolve the issue. Activity 2. BURL itemizes the alternatives to be considered. This list can be as few as two items or in the hundreds, and spans all the way from minor product
ullman-38162
ull75741_04
December 17, 2008
15:4
4.5
Choosing a Project
changes to new models of existing products to innovative new products. For our bicycle company example, the options are ■ ■ ■
Upgrade the current road bike. Introduce a tandem (as already considered). Add a front suspension to a soft-tail mountain bike product.
Activity 3. BURL develops criteria that are the basis for evaluating the alternatives. This is such an important activity that all of Chap. 6 is devoted to developing engineering specifications, the criteria for evaluating concepts and products. For many types of issues, those that are commonly repeated, a generic set of criteria can be used, at least as a starting place. For portfolio issues, the following list of criteria have evolved over time and can be used here: ■
■ ■ ■ ■ ■ ■ ■ ■
Acceptable program complexity: The complexity of the effort is within the experience of the organization or vendors. People are available with the skill sets needed to do the work. Clear market need: There is an established need in a market. (If evaluating innovative products, this may not be important.) Acceptable competitive intensity: The competitive intensity is reasonable and the alternative is not so new to the organization to impede commercialization. Acceptable five-year cash flow: The cash needed or generated over a five-year period is within reason. Reasonable payback time: The payback period for the needed investment and costs is acceptable. Acceptable start-up time: The time to realize cash flow is within the means of the organization. Good company fit: The newness or impact on the organization is acceptable— the new product or improvement fits the organization’s image. Strong proprietary position: The ability to withstand the competition’s efforts to erode the unique features that discriminate is good. Good platform for growth: The effort leads to future products or services.
In the SWOT analysis, we can see that the strengths, weaknesses, opportunities, and threats listed are evaluations of the criteria just listed. For example, the SWOT statements: “Market for tandems is growing although no exact market numbers have been collected” and “Market for tandems is small, < 1% of all bicycle sales” are qualitative evaluation statements for “Clear market need.” One way to develop a list of criteria is to begin with a SWOT or pro-con analysis and then group the statements in categories. In fact, the list of protocol criteria was developed by examining many different protocol decisions, SWOT analyses and pro-con lists to find the common measures. Activity 4. BURL decides what is most important. Not all of the nine criteria listed above are of equal importance. Complicating this is that the importance is in the eye of the beholder. For example, BURL’s financial people think that
107
ullman-38162
ull75741_04
108
December 17, 2008
CHAPTER 4
15:4
The Design Process and Product Discovery
Table 4.2 The portfolio scoring by BURL
Alternatives Upgrade road bike Criteria Acceptable program complexity complexity Clear market need Acceptable competitive intensity Acceptable five-year cash flow Reasonable payback time Acceptable start-up time Good company fit Strong proprietary position Good platform for growth
Front suspension for mountain bike
Tandem
Agreement
Certainty
Agreement
Certainty
Agreement
Certainty
SA N A D N A A SD D
C VC C C C VC C C C
N D A D D A D SA A
C U N C U VC C C U
D SA N A A N N A A
VU VC N C U C C C C
“acceptable five-year cash flow” and “reasonable payback time” are most important, and marketing wants to see “Good company fit.” Engineering wants a “Strong proprietary position” and a “Good platform for growth.” For now, we will assume they are all equally important and address this activity further in Chap. 8. Activity 5. BURL evaluates the alternatives relative to the criterion. These evaluations can range from qualitative assessments to the results of analytical simulations. For now, we will work with the qualitative statements made in the SWOT analysis and use a very simplified method to evaluate and decide what to do next. This will be refined as the product matures and more numerical analyses and simulation become possible. To support this evaluation BURL used a Decision Matrix, a table with the alternatives in columns and the criteria in rows (Table 4.2). The cells of the matrix contain the evaluation results. For this qualitative assessment, BURL evaluated each alternative relative to each criterion using two measures. The first is how well the alternative meets the criterion in terms of level of agreement with the statement “I <X> that the has ” where <X> equals ■ ■ ■ ■ ■
Strongly agree (SA) Agree (A) Neutral (N) Disagree (D) Strongly disagree (SD)
For example “I that the
isu °v 3.3 3.0 2.7
70°
2.2 1.9 3 max. allow. 1.5 elbow rise
69.6 l.w. 64.8 av. 58.8 s.w. 65.4 60.9 55.3 _1.2 normal slump 57.4 53.3 48.7
50
50% tile
6'-8' min. ceiling ht. 6'-8'–7'-0' door ht. std. 6'-6' min. door ht.
keep lights above to minimize glare unimportant displays
al
73.1–57.2 range
opt. control zone
ullman-38162
3' ea. woman 36
D 60° 0-10°
36 av. oven door
18 max.
16 min. clear. for stool
noncritical or casual controls, manuals and 45 displays noncritical controls 6 (stooping is 39 feasible) writing shelf 9
6 FT 68' cannot see over ht. 64.5 highest shelf 63 min. cat walk ceiling 5 FT
55 opt. tel. dial_women
4 FT
3 FT reach R. 28.3 26.2 23.4
2 FT
20.8 19.4 17.8
58–60 thermostat ht. std. 58 mirror CL 58 tel. dial_m. & w. standing
50 push plates std. 48–54 grab bars_women 48 wall switch std.
42 bar ht. std., door pulls 41 opt. tel dial–m. & w. sitting 39.5 opt. tel. dial–women sit. 39 max. counter ht. 38 door knob_std. 36 sink rim std., hand rails 34 hand rails_stairs 33 panic bars 33.5–34.8 cleaning surfaces 32.3–34.3 food preparation 31.5–33.5 cooker hot plate 31 lavatories 30–32 ironing boards 27–28 table ht. 26.4 stool ht.—36' counter 25–26 typing table 24 min. knee clear_sitting
16 av. seat ht.—women 15.75 max. ht. W.C. 15–15.3 opt. seat ht.—public 12 max. rung spacing 4.8 3.6 3.4
7–7.5 opt. stair riser 6–8 bar rail
1.6
4–6
4 min. toe space 1.6 av. heel_women
0' datum 6 min. 34 min. gangway
Figure D.2 Anthropometric woman at control panel (Source: Adapted from H. Dreyfuss, The Measure of Man: Human Factors in Design, Whitney Library of Design, New York, 1967).
ullman-38162
ull75741_appD
December 17, 2008
11:30
D.4
The Human as Sensor and Controller
drawing is of a 50th-percentile woman. The dimensions on the control panel are such that a majority of women will feel comfortable looking at the displays and working the controls. Returning to our lawn mower: the handle should be at about elbow level, height 5 in Fig. D.1 and Table D.1. To fit all men and women between the 5th and 95th percentiles, the handle must be adjustable between 94.9 cm (37.4 in.) for the 5th-percentile woman and 117.8 cm (46.4 in.) for the 95th-percentile man. Anthropometric data from the references also show that the pull starter should be 69 cm (27 in.) off the ground for the average person. For this uncommon position, only an average value is given in the references. For positions even more unique, the engineer may have to develop measurements of a typical user community in order to get the data necessary for quality products.
D.3 THE HUMAN AS SOURCE OF POWER Humans often have to supply some force to power a product or actuate its controls. The lawn-mower operator must pull on the starter cord and push on the handle or move the steering wheel. Human force-generation data are often included with anthropometric data. This information comes from the study of biomechanics (the mechanics of the human body). Listed in Fig. D.3 is the average human strength for differing body positions. In the data for “arm forces standing,” we find that the average pushing force 40 in. off the ground (the average height of the mower handle) is 73 lb, with a note that hand forces of greater than 30 to 40 lb are fatiguing. Although only averages, these values do give some indication of the maximum forces that should be used as design requirements. More detailed information on biomechanics is available in MIL-HDBK (Military Handbook) 759A and The Human Body in Equipment Design (see Sources at the end of this appendix).
D.4 THE HUMAN AS SENSOR AND CONTROLLER Most interfaces between humans and machines require that humans sense the state of the device and, based on the data received, control it. Thus, products must be designed with important features readily apparent, and they must provide for easy control of these features. Consider the control panel from a clothes dryer (Fig. D.4). The panel has three controls, each of which is intended both to actuate the features and to relate the settings to the person using the dryer. On the left are two toggle switches. The top switch is a three-position switch that controls the temperature setting to either “Low,” “Permanent Press,” or “High.” The bottom switch is a two-position switch that is automatically toggled to off at the end of the cycle or when the dryer door is opened. This switch must be pushed to start the dryer. The dial on the right controls the time for either the no-heat cycle (air dry) on the top half of the dial or the heated cycle on the bottom half. The dryer controls must communicate two functions to the human: temperature setting and time. Unfortunately, the temperature settings on this panel are
419
ull75741_appD
December 17, 2008
420
11:30
Human Factors in Design
APPENDIX D
HUMAN STRENGTH (for short durations) strength correction factors; X 0.9 left hand and arm X 0.84 hand – age 60 X 0.5 arm & leg – age 60 Arm forces standing X 0.72 women floor 28 approx. opt. lever 17 30 Lb.
Arm forces sitting A
14 lb. r. hand 19 lb. r. hand
back support
52 50 R. 50 42 L.
15 16
x
56 42
17 R. 13 L.
42 30 20 18
40 50
58'
18'
Pos. X 61 73
18'
40' Arm forces sitting
28'
B
8 L.
hand forces > 30–40 Lb. are fatiguing
max. force
13 L.
0 8 15
16'
large force
20 R. 20 15
Lifting forces
14 R.
Pos. X
close to body
24'
Leg forces sitting
100
–30
50
–1
–11
0 120 ° ° 135 –155°
0–5
1
05
b.
2.5
'm a ank x. trav le el
25 Lb. max.
0–2 0L
ullman-38162
00
0
back lift is 40% leg lift.
30
5 Ft.
45
4 Ft.
70
3 Ft.
125
2 Ft.
145 Lb.
1 Ft.
max. hand squeeze: 85 Lb. R.H. 77 Lb. L.H.
Figure D.3 Average human strength for different tasks (Source: Adapted from H. Dreyfuss, The Measure of Man: Human Factors in Design, Whitney Library of Design, New York, 1967).
ullman-38162
ull75741_appD
December 17, 2008
11:30
The Human as Sensor and Controller
D.4
Air Fluff Cycle Temperature
Fabricare Cycle 50
Permanent Press Low
40 30
20
60
High
10 Off
70
90
No iron Cool down 20
Push to start
80 70 30 40 50
Bradford
60
Permanent Press Timed Dry Cycle
Figure D.4 Clothes dryer control panel (Source: Adapted from J. H. Burgess,
Designing for Humans: The Human Factor in Engineering, Petrocelli Books, Princeton, N.J., 1986).
hard to sense because the “Temperature” rocker switch does not clearly indicate the status of the setting and the air-dry setting for temperature is on the dial that can override the setting of the “Temperature” switch. There are two communication problems in the time setting also: the difference between the top half of the dial and the bottom half is not clear and the time scale is the reverse of the traditional clockwise dial. The user must not only sense the time and temperature but must regulate them through the controls. Additionally, there must be a control to turn the dryer on. For this dryer, the rocker switch does not appear to be the best choice for this function. Finally, the labeling is confusing. This control panel is typical of many that are seen every day. The user can figure out what to do and what information is available, but it takes some conjecturing. The more guessing required to understand the information and to control the action of the product, the lower the perceived quality of the product. If the controls and labeling were as unclear on a fire extinguisher, for example, it would be all but useless—and therefore dangerous. There are many ways to communicate the status of a product to a human. Usually the communication is visual; however, it can also be through tactile or audible signals. The basic types of visual displays are shown in Fig. D.5. When choosing which of these displays to use, it is important to consider the type of information that needs to be communicated. Figure D.6 relates five different types of information to the types of displays. Comparing the clothes-dryer control panel of Fig. D.4 to the information of Fig. D.6, the temperature controls require only discrete settings and the time control a continuous (but not accurate) numerical value. Since toggle switches are not very good at displaying information, the top switch on the panel of Fig. D.4 should be replaced by any of the displays recommended for discrete information. The use of the dial to communicate the time setting seems satisfactory. To input information into the product, there must be controls that readily interface with the human. Figure D.7 shows 18 common types of controls and
421
422
ull75741_appD
December 17, 2008
APPENDIX D
11:30
Human Factors in Design
4
3
0
7
0
Digital counter
Icon, symbol display
110
0 12
10 0
ullman-38162
Linear dial Curved dial Fixed pointer on moving scale
Indicator light
Linear dial Circular dial
Moving pointer on fixed scale
Graphical display
Dan
ger f d g f k f r k g r ej k d f d k j g r e j g f v j r o j o
Mechanical indicator
Pictorial display
Figure D.5 Types of virtual displays.
their use characteristics; it also gives dimensional, force, and recommended use information. Note that the rotary selector switch is recommended for more than two positions and is rated between “acceptable” and “recommended” for precise adjustment. Thus, the rotary switch is a good choice for the time control of the dryer. Also, for rotary switches with diameters between 30 and 70 mm, the torque to rotate them should be in the range from 0.3 to 0.6 N · m. This is important information when one is designing or selecting the timing switch mechanism. In addition, note that for the rocker switch, no more than two positions are recommended. Thus, the top switch on the dryer, Fig. D.4, is not a good choice for the temperature setting. An alternative design of the dryer control panel is shown in Fig. D.8. The functions of the dryer have been separated, with the temperature control on one rotary switch. The “Start” function, a discrete control action, is now a button, and the timer switch has been given a single scale and made to rotate clockwise. Additionally, the labeling is clear and the model number is displayed for easy reference in service calls. In general, when designing controls for interface with humans, it is always best to simplify the structure of the tasks required to operate the product. Recall
ullman-38162
ull75741_appD
December 17, 2008
11:30
D.4
Exact value
Rate of change
The Human as Sensor and Controller
Trend, direction of change
Discrete information
Adjusted to desired value
Digital counter Moving pointer on fixed scale Fixed pointer on moving scale Mechanical indicator Symbol display Indicator light Graphical display Pictorial display Not suitable
Acceptable
Recommended
Figure D.6 Appropriate uses of common visual displays.
the characteristics of the short-term memory discussed in Chap. 3. We learned there that humans can deal with only seven unrelated items at a time. Thus, it is important not to expect the user of any product to remember more than four or five steps. One way to overcome the need for numerous steps is to give the user mental aids. Office reproducing machines often have a clearly numbered sequence (symbol display) marked on the parts to show how to clear a paper jam, for example. In selecting the type of controller, it is important to make the actions required by the system match the intentions of the human. An obvious example of a mismatch would be to design the steering wheel of a car so that it rotates clockwise for a left turn—opposite to the intention of the driver and inconsistent with the effect on the system. This is an extreme example; the effect of controls is not always so obvious. It is important to make sure that people can easily determine the relationship between the intention and the action and the relationship between the action and the effect on the system. A product must be designed so that when a person interacts with it, there is only one obviously correct thing to do. If the action required is ambiguous, the person might or might not do the right thing. The odds are that many people will not do what was wanted, will make an error, and, as a result, will have a low opinion of the product.
423
December 17, 2008
Control
Human Factors in Design
Force F, N Moment M, N ⋅ m
Dimension, mm
d
Handwheel
D: 160–800 d: 30–40
D
M
160–800 mm 200–250 mm
2–40 N ⋅ m 4–60 N ⋅ m
D Crank
d h
Rotary knob Turning movement
Hand (finger) r: <250 (<100) l: 100 (30) d: 32 (16)
r l
Hand (finger) D: 25–100 (15–25) h: >20 (>15)
D Rotary selector switch
b l h
l: 30–70 h: >20 b: 10–25
r
M
<100 mm 100–250 mm
0.6–3 N ⋅ m 5–14 N ⋅ m
D
M
15–25 mm 25–100 mm
0.02–0.05 N ⋅ m 0.3–0.7 N ⋅ m
l
M
30 mm 30–70 mm
0.1–0.3 N ⋅ m 0.3–0.6 N ⋅ m
Setting visible Accidental actuation
APPENDIX D
Continuous adjustment Precise adjustment Quick adjustment Large force application Tactile feedback
424
11:30
>2 positions
ull75741_appD
2 positions
ullman-38162
Thumbwheel b: > 8
F = 0.4 –5 N
D: 60–120
F = 0.4 –5 N
d: 30–40 l: 100–120
F1 = 10–200 N F2 = 7–140 N
d: 30–40 b: 110–130
F = 10–200 N
b D
Rollball
Handle (slide)
d l
1
2
D-handle
b
d
*
Push button d
Finger: d > 15 Hand: d > 50 Foot: d > 50
Linear movement
Slide
Finger: F = 1–8 N Hand: F = 4–16 N Foot: F = 15–90 N
b l: >15 b: >15
F = 1–5 N (Touch grip)
b: >10 h: >15
F = 1–10 N (Thumb-finger grip)
l Slide
b h
Sensor key b
l: >14 b: >14
l *Recessed installation
Figure D.7 Appropriate uses of hand- and foot-operated controls (Source: Adapted from
G. Salvendy (ed.), Handbook of Human Factors, Wiley, 1987).
11:30
Lever
d
Force F, N Moment M, N⋅m
Setting visible Accidental actuation
Dimension, mm
425
Precise adjustment Quick adjustment Large force application Tactile feedback
Control
The Human as Sensor and Controller
Continuous adjustment
D.4
>2 positions
December 17, 2008
2 positions
ull75741_appD
l
Joystick
d: 30–40 l: 100–120
F = 10–200 N
s: 20–150 d: 10–20
F = 5–50 N
b: >10 l: >15
F = 2–10 N
b: >10 l: >15
F = 2–8 N
d: 12–15 D: 50–80
F = 1–2 N
d s
b l
Toggle switch Swiveling movement
ullman-38162
D Rocker switch
b
l
Rotary disk D d Pedal
l b
b: 50–100 l: 200–300 l: 50–100 (forefoot)
Sitting: F = 16–100 N Standing: F = 80–250 N
Figure D.7 (continued).
TIME SET START
Cool down
20
Remove to prevent wrinkling Low
Air fluff
Hi
30
Set
40
Perma press
50 90 80
60 70
BRADFORD Model 78345
Figure D.8 Redesign of the clothes dryer control panel of Fig. D.4.
ullman-38162
426
ull75741_appD
December 17, 2008
APPENDIX D
11:30
Human Factors in Design
D.5 SOURCES Burgess, J. H.: Designing for Humans: The Human Factor in Engineering, Petrocelli Books, Princeton, N.J., 1986. A good text on human factors written for use by engineers; the dryer example is from this book. Damon, A. et al., The Human Body in Equipment Design, Harvard University Press, Boston, 1966. Dreyfuss, H.: The Measure of Man: Human Factor in Design, Whitney Library of Design, New York, 1967. This is a loose-leaf book of 30 anthropometric and biomechanical charts suitable for mounting; two are life-size, showing a 50th-percentile man and woman. A classic. Human Engineering Design Criteria for Military Systems, Equipment, and Facilities, MIL-STD 1472F. http://hfetag.dtic.mil/docs-hfs/mil-std-1472f.pdf. Four hundred pages of human factors information. Human Engineering Design Data Digest, Department of Defense Human Factors Engineering Technical Advisory Group, April 2000, http://hfetag.dtic.mil/hfs_docs.html. Excellent online source. Human Factors Design Standard (HFDS), FAA, http://hf.tc.faa.gov/hfds/. Another excellent online source. Jones, J. V.: Engineering Design: Reliability, Maintainability and Testability, TAB Professional and Reference books, Blue Ridge Summit, Pa., 1988. This book considers engineering design from the view of military procurement, relying strongly on military specifications and handbooks. MIL-HDBK-759C, Human Engineering Design Guidelines, 1995. Norman, D.: The Psychology of Everyday Things, Basic Books, New York, 1988. Guidance for designing good interfaces for humans; light reading. Moggridge, B.: Designing Interactions, http://www.designinginteractions.com/. An online book for designing human interfaces for the 21st century. Salvendy, G. (ed.): Handbook of Human Factors, 3rd edition, Wiley, New York, 2006. Seventeen hundred pages of information on every aspect of human factors. System Safety Program Requirements, MIL-STD 882D. U.S. Government Printing Office, Washington, D.C. http://safetycenter.navy.mil/instructions/osh/milstd882d.pdf. The hazard assessment is from this standard. Tilly, A. R.: The Measure of Man and Woman, Whitney Library of Design, New York, 1993. An updated version of the preceding classic rewritten by one of Dreyfuss’s associates.
ullman-38162
ull75741_IND
December 23, 2008
15:18
INDEX
A abstraction, levels of, 32, 215 accuracy, modeling and, 286 additive tolerance stack-up, 299–301 aging/deterioration effects, 290 aisle chair, 147, 158 analogies, 191–192 analysis problems, 16 analytical models, 124–125, 294–295 assembly drawings, 122–123 efficiency, 331, 333 instructions, 367 manager, 70 requirements, in engineering specifications, 162 B behavior, human problem-solving, 58–64 behavior, product, 30 Belief Map, 235–239 benchmarking, 157–158 best practices for product evaluation, 279–280 bicycle product discovery phase and, 101, 102, 106–109 redesign, 37–39 Bill of Materials (BOM),15, 245–246 brainstorming, 190 brainwriting, 190–191 C CAD systems, 118–119, 123–124 chunks of information, 50, 51, 53 clamp (see Irwin) coefficient of variation, 409–413 cognitive psychology, 48 Commercial Off The Shelf (COTS) components, 267 communication, during design process, 137–141 competition benchmarking, 157–158 component assembly, 331 development, 253–260 handling, 331, 343–346
mating, 331, 347–349 retrieval, 331, 342–343 components, 27 configuring, 247–249, 271–273 cost of injection-molded, 325 cost of machined, 321–324 developing, 253–260 developing connections/interfaces between, 249–253, 274–275 from vendors, 266–269 Computed Tomography (CT) Scanner, 82–85, 86, 89 computer-aided design (CAD) systems, 119, 123–124 computer-generated solid models, 118–119 concept combining, 207–208 defined, 171 developed for each function, 206–207 concept evaluation and selection, 213–239 assessing risk and, 226–233 decision-matrix method, 221–226 feasibility, 218–219 level of abstraction and language for, 215–218 robust decision-making, 233–239 technology readiness, 219–221 concept generation, 171–209 amount of time spent on, 171–172 basic methods of, 189–194 clamp, 173–176 contradictions used for, 197–201 functional decomposition technique, 181–189 morphological method, 204–208 reverse engineering, 178–180 Theory of Inventive Machines (TRIZ), 201–204 through patent literature, 194–197 understanding function of existing designs and, 176–180 conceptual design, 40, 87–89. See also concept evaluation and selection; concept generation phases of, 213–214 simplicity and, 208–209 concurrent engineering, 9 configuration design, 34–36 configuration of components, 247–249, 271–273 conformity, creativity and, 65–66
427
ullman-38162
428
ull75741_IND
December 23, 2008
15:18
Index
connections, 249–253 constraints, design, 40–41 contradictions, to generate ideas, 197–201 cost estimates, 320–321 estimating product development, 133 of injection-molded components, 325 of machined components, 321–324 cost, product determining, 316–320 measuring design process with, 3–6 cost requirements, 161 creativity, in designers, 64–66 “creeping specifications,” 143–144 Critical Path Method (CPM), 131 CT Scanner, 82–85, 86, 89 finding overall function of, 183–184 subfunction description, 187–188 customer relationships, 370 customers determining requirements of, 151–155 evaluating importance of requirements of, 155–157 identifying, 151 relating engineering specifications to, 163–164 satisfaction of, Kano model of, 97–99 satisfaction with competition, 157–158 D decision-making basics of, 105–106 choosing a project, 101–109 concept selection, 216, 233–239 portfolio decision, 105–109 risk, 233 Decision Matrix, 108–109, 221–226, 234 decisive decision-makers, 62–63 decomposition, product, 40–44 functional, 184–188, 204–205 reverse engineering and, 178–180 deliverables, 118–124, 128 design best practices, key features, 10 design-build-test cycle, 217 design decisions, 40 design engineer, 68 designers. See also design teams creativity of, 64–66 generating solutions, 57 human information processing and, 48–56 mental processes of, during design process, 56–64 as part of design team, 69 problem-solving behaviors by, 58–64 understanding the design problem, 56–57
design evaluation. See concept evaluation and selection Design-For-Assembly (DFA), 329–349 design for cost (DFC), 315–325 Design for Manufacture (DFM), 12 328–329 Design for Reliability (DFR), 350–357 Design for Six Sigma (DFSS), 10 Design for the Environment (DFE), 20, 358–360, 375–376 Designing For Sustainability (DFS), 20 design notebooks, 137–138 design patents, 373 design problems. See also Quality Function Deployment (QFD) technique basic actions for solving, 17–19 configuration design, 34–36 documentation of, 140 knowledge and learning during design and, 19–20 many solutions for, 15–17 mechanical, 33–40 mental processes of designers and, 56–57 original design, 37 parametric design, 36 redesign, 37–40 selection design, 33–34 solutions for, 15–17 understanding, 143–144, 143–151 design process. See also designers; mechanical design; product discovery communication during, 137–141 conceptual design phase, See concept generation and concept selection “creeping specifications” and, 143–144 defined, 8 designing quality, 92–95 documentation and, 363, 366–368 end of, 363–365 history of, 8–10 human factors and, 415–425 measuring, 3–8 need for studying, 1–3 overview of, 81–85 product definition phase. See product generation and product evaluation product development phase. See product development product discovery phase. See product discovery product support phase, 91–92, 368–370 project planning phase. See project planning safety factor in, 403–414 design report, 139–141 design reviews, 113, 138–139
ullman-38162
ull75741_IND
December 23, 2008
15:18
Index
Design Structure Matrix (DSM), 132 design teams assessing health of, 76–77 building performance, 72–73 characteristics of successful, 72–73 contract, 73–79 management of, 71–72 meeting minutes for, 73, 75, 76 members of, 68–71 need for, 66–68 desktop prototyping, 118 detail drawings, 121–122 deterioration/aging effects, 290 DFA (Design-For-Assembly), 329–349 DFC (design for cost), 315–325 DFE (design for the environment). See Design for the Environment (DFE) DFM (Design for Manufacture), 328–329 DFR (design for reliability), 350–357 DFSS (Design for Six Sigma), 10 DFV (value engineering), 325–328 disassembly, of product, 13 disclosure, patent 373 documentation, communicating final design, 139–141 domain-specific knowledge, 50 drawings assembly, 122–123 detail, 121–122 layout, 120–121 Dreamliner, Boeing, 146–147 E efficiency, assembly, 331–333 end-of-life, product, 13 End-of-Life Vehicles (ELVs), 376–378 energy flows, 177, 180 Engineering Change Notice (ECN), 371 engineering changes, 370–371 engineering specifications determining importance of, 164–165 developing, 158–163 guidelines for good, 162–163 identifying relationships between, 166–167 measuring competitions’ products, 165 relating customer requirements to, 163–164 targets, 165–166 types of, 160–162 evaluation. See also product evaluation of concepts, 88–89 importance of customer requirements, 155–157 Evaporating Cloud (EC) method, 197–198 excitement-level features, 98–99
F factor of safety, 403–414 Failure Modes and Effects Analysis (FMEA), 232, 350–353 failure rate, 355 fasteners, minimizing use of, 335–338 Fault Tree Analysis (FTA), 352, 353–355 Feasibility evaluation, 218–219 features basic, 98 excitement-level, 98–99 definition, 27 performance, 98 fidelity, 124, 216–217, 293 flexible decision-makers, 62 flow of energy, information, and material, 177, 179–180 focus-group technique, 152, 153, 154 force flow visualization, 257–259 form generation, 246–264 form of the product, 2–3, 29, 243, 244 Franklin, Benjamin, 102–103 function, 2–3, 28–40, 243 behavior and, 30 defining, 177–178 developing concepts for each, 206–207 finding the overall, 181, 183–184 modeling 181–189 monitoring change in, 280–281 using reverse engineering, 178–180 functional decomposition, 29, 172, 181–194, 204–205 functional performance requirements, 160 function diagram, 130 G Gantt chart, 131, 140 General Electric CT Scanner. See CT Scanner generating concepts, 87–88 graphical models, 118–124 green design (Design for the Environment), 358–360 group technology, 260 H handling, component, 331, 343–346 Hannover Principles, 20–21, 209, 357 house of quality, See Quality Function Deployment (QFD) Technique human factor requirements, 160 human factors, 415–425 human information processing, 48–56
429
ullman-38162
430
ull75741_IND
December 23, 2008
15:18
Index I industrial designer, 70 information, human memory and, 49–50 information language, problem-solving behavior and, 61–62 information processing, human, 48–56 injection-molded components, costs of, 325 installation instructions, 367 installation, product, 13 instruction manuals, 367–368 Integrated Product and Process Design (IPPD), 9, 94 interfaces between components, 249–253 International Standard Organization’s ISO 9000 system, 94–95 Irwin Quick-Grip clamp, 26, 27 product decomposition, 41–44, 179 project planning and, 113–115 redesign of one-handed bar clamp, 173–176 reverse engineering, 178–180 subfunction description, 187 ISO 9000 quality management system, 94–95 J Jet Propulsion Laboratory (JPL) (Cal Tech), 26 K Kano Model of customer satisfaction, 97–99 Kano, Noriaki, 97 Key features of design best practice, 10 knowledge increase during design, 19 types of, 50 creativity and, 65 L language concept evaluation and, 215–218 encoding chunks of information, 50 mechanical design, 30–32 layout drawings, 120–121 Lean manufacturing, 9 level of abstraction, 32, 215 Level of Certainty, 235, 237 Level of Criterion Satisfaction, 235, 236 life cycle, product, 161 long-term memory, 52–54 M machined components, costs of, 321–324 maintainability, 357
maintenance instructions, 367 manufacturing cost, 3–4, 5–6, 317–324 engineer, 69 instructions, 366 processes, 2–3 requirements, in engineering specifications, 162 variance, 290, 297 Marin Mount Vision Pro bike, 39 product evaluation and, 291–292, 299–300 product generation for, 269–276 market pull, 96–97, 99 Mars Exploration Rover (MER), 26 Choosing a wheel for configuration design, 34–36 mechanical design language and, 31–32 planning for, 132 product support and, 92 safety factors, 40 sub-systems, 28 material costs, 317 materials, properties of the most commonly used, 380–392 selection of, 264–266 materials specialists, 69 mating, component, 331, 347–349 mature design, 37 Mean Time Between Failures (MTBF), 355–357 mean value, 398–399 measurement of the design process, 3–8 mechanical failure, 350 mechanical fuse, 358 mechatronic devices, 25 meeting minutes, design team, 73, 75, 76 memory, human, 48–50 long-term, 52–54 short-term, 51–52 MER. See Mars Exploration Rover (MER) milestone chart, 131 MIL-STD 882D (Standard Practice for System Safety), 230–231 modeling, 117–126, 286, 292–296 modularity, 248–249 morphological method, 204–208 N “NIH” (Not Invented Here) policy, 178, 218 noise, 290–294 nominal tolerances, 297 nonconformity, 65–66 normal distribution, 397–401
ullman-38162
ull75741_IND
December 23, 2008
15:18
Index O objective approach to problem-solving, 61 observation of customers, 152, 153, 154 obstructive nonconformists, 66 operation instructions, 367 ordering subfunctions, 186–188 original design, 37 originality, 60 overall assembly, evaluation of, 333–341 overall function, 181, 183–184 over-the-wall design method, 8–9, 10, 12 P packaging (configuration) design, 34–36 parallel tasks, 131–132 parametric design, 36 part numbers, 245 patching, 260–261, 263–264 patent applications, 371–375 searches, 194–197 P-diagram, 282–283, 291–292 Performance and function performance evaluation, 281–286, 292–296 performance features, 98 PERT (Program Evaluation and Review Technique) method, 130–131 physical models, 117–118, 217, 286, 295–296 physical requirements, 160 planning. See project planning portfolio decision, 105–109 preproduction run, 118 Priestly, Joseph, 102–103 probability, normal, 397–401 problem-solving behavior, 58–64 decision closure style, 63–64 deliberation style, 62–63 energy source, 58–60 information language, 61–62 information management style, 60–61 pro-con analysis, 102–105 product change, 96, 99 Product Data Management (PDM), 14 product decomposition, 41–44 product design, 40 product design engineer, 68 product development phase, 90–91 product discovery, 85–86, 95–100 choosing a project, 101–109 customer satisfaction and, 97–99 goal of, 95–96 market pull and technology push, 96–97
product maturity and, 97 product proposal, 99–100 product evaluation, 279–313, 315–360 accuracy, variation, and noise, 286–292 best practices for, 279–280 Design-For-Assembly (DFA), 329–349 Design for Cost (DFC), 315–325 Design for Manufacture (DFM), 328–329 Design for Reliability (DFR), 350–357 Design for test and maintenance, 357–358 Design for the Environment, 358–360, 375–376 goals of performance evaluation, 281–284 modeling for, 292–296 monitoring functional change, 280–281 sensitivity analysis, 302–305 tolerance analysis, 296–302 trade-off management, 284–286 value engineering, 325–328 product generation, 241–276 Bill of Materials, 245–246 developing components, 253–260 form generation, 246–264 for Marin Mount Vision Pro bicycle, 269–276 materials and process selection, 264–266 vendor development, 266–269 Product Life-cycle Management (PLM), 13–15, 245 product manager, 69 product maturity, “S” curve, 97–98 product proposal, 99–100 product quality. See quality, product product risk, 230–233 product function of, 28–29 liability, 229–230 life of, 10–15 safety of, 227–229 product support, 91–92, 368–370 project planning, 86, 111–141 activities of, 111–112 choosing best models and prototypes for, 125–126 design plan examples, 134–137 goal of, 111 physical models and prototypes used in, 117–118 plan template, 125–133, 128–133 types of plans, 113–117 project portfolio management, 101 project structures, 71
431
ullman-38162
432
ull75741_IND
December 23, 2008
15:18
Index
prototypes, 117–118 choosing, 125 proof-of-concept prototype, 118 proof-of-function prototype, 118 proof-of-process prototype, 118 proof-of-production prototype, 118 proof-of-product prototype, 118 Pugh’s method. See decision-matrix method purchased-parts cost, 317 Q QFD method. See Quality Function Deployment (QFD) technique Quality Assurance (QA), 366 Quality Assurance (QA) specialists, 69, 92 Quality Control (QC), 366 Quality Control (QC) specialists, 69, 92 Quality Function Deployment (QFD) technique, 145–169 determining what the customers want, 151–155 developing engineering specifications, 158–163 evaluating importance of customer requirements, 155–157 identifying and evaluating the competition, 157–158 identifying customers, 151 identifying relationships between engineering specifications, 166–167 relating customer requirements to engineering specifications, 163–164 reverse engineering and, 178 setting engineering specification targets and importance, 164–166 uses of, 168 quality, product design process and, 92–95 determinants of, 6 effect of variation on, 289–292 measuring design process with, 3, 6 Quick-grip clamp. See Irwin R rapid prototyping, 118 recycling, 13, 359, 360 redesign, 37–40 of clamp, 173–176 QFD method and, 145 refining products, 260–264 refining subfunctions, 188–189 reliability, 161, 350, 355–357
reliability-based factor of safety, 406–414 reparability, 357 resource concerns, in engineering specifications, 161–162 retirement, product, 13 retrieval, component, 331, 342–343 reuse, of product, 13 reverse engineering, 178–180 risk, 226–233 decision, 233 product, 230–233 project, 232–233 robust decision making, 233–239 robust design by analysis, 305–308 through testing, 308–313 Rover, Mars. See Mars Exploration Rover (MER) S safety, product, 227–229, 403–414 sample mean, 398–399 sample standard deviation, 398–399 sample variance, 399 “S” curve, product maturity, 97 selection design, 33–34 sensitivity analysis, 302–305 sequential tasks, 131 serviceability, 357 short-term memory, 48, 51–52, 55 simple design plan, 134–135 simultaneous engineering, 9 6-3-5 method, 190–191 Six Sigma philosophy, 9–10, 297 sketches, 119 solid models, 118–119, 123 spatial constraints, 247, 269–270 specification, patent, 373–374 spiral process, 115–117 standard deviation, 398–399 Standard Practice for System Safety (MIL-STD 882D), 230–231 standards, 161–162 Stage-Gate Process, 113 statistical stack-up analysis, 301–302 subfunction ordering, 186–188 refining, 188–189 descriptions, 184–186 subjective approach to problem-solving, 61–62 subsystems, 27 surveys, 152, 154 sustainability, design for, 20–21 SWOT analysis, 101–102, 105
ullman-38162
ull75741_IND
December 23, 2008
15:18
Index T Taguchi, Genichi, 305 Taguchi’s method, 305–306 tank problem, 283–284 targets, engineering specifications, 165–166 tasks planning, 126–128 sequence, 131–133 teams, design. See design teams technicians, 69 technology push, 96, 99 technology readiness, 219–221 Templates (All available on line BOM, 246 Change order, 372 Design for Assembly, 330 Design Report, 139–141 FMEA, 351 Machined Part Cost Calculator, 322 Meeting minutes, 75 Morphology, 205 Patent prospects, 375 Personal Problem Solving Dimensions, 59–63 Product Decomposition, 42–43 Project Plan, 127 Team contract, 74 Team health inventory, 77 Plastics Part Cost Calculator, 325 Product Proposal, 100 Pro/Con Analysis, 104 Reverse Engineering, 182 Swot Analysis, 102 Technology Readiness, 221 testability, 357 Theory of Inventive Machines (TRIZ), 201–204
time product development, 6–8 project planning and, 128–130 spent on developing concepts, 171–172 time requirements, in engineering specifications, 161 tolerance analysis, 296–302 trade-off management, 284–286 TRIZ. See Theory of Inventive Machines U UL standards, 162 uncertainties, 285–286 uncoupled tasks, 132 “use” phase of products, 13 utility patents, 373 V value engineering/analysis, 325–328 variant design, 40 variation, 286–292, 297 vendor development, 266–269 vendor relationships, 368–370 vendor representatives, 70 verbal problem-solvers, 61 W Waterfall model of project planning, 113 work breakdown structure, 131 worst-case analysis, 301 X X-Ray CT Scanner. See CT Scanner
433
  1. Fox Float Rp23 Manual
  2. Fox Float Rp23 Manual
  3. Download Fox Float Rp23 Manual Lawns

Installing Your Shock General Maintenance Before You Ride Measuring Sag Setting Sag Adjusting Rebound ProPedal Service Intervals Important Safety Information Stuck Down Shock Air Sleeve Maintenance


Fox Float Rl Rear Shock Service Manual. And are compatible with all the Fox Float (air sprung) rear shocks (RP23. RP3, RP2, R, RL. Rl, 2009 fox float rp2, fox rp3 rebuild, fox rp3 manual, fox. Previous Page Fox float ctd shock manual download on iubmb-2013-3.org free books. Fox float RP23 factory series rear mountain bike shock Center bolt measure 215mm Originally from a yeti SB66 Excellent condition. Recently serviced at local bike shop. Kashima coat. Propedal adjustable. Works fine, just got a new bike. DOWNLOAD PDF. Ullman-38162 ull75741_fm December 18, 2008 16:19 The Mechanical Design Process ullman-38162 ull75741_fm December 18, 2008 16:19 McGraw-Hill Series in.

features/adjustments

high volume standard air sleeve, boost valve, angled air valve, lightweight chassis, DOHC ProPedal with 2 positions, ProPedal tuning adjust with 3 positions, air spring pressure, rebound adjust

spring

air

intended use

freeride, all-mountain, cross-country

Installing Your Shock

Fox Float Rp23 Manual

If you are installing your shock on a bike in which the shock is not original equipment:

  1. Install the shock.
  2. Remove the main air chamber air cap and let all the air out of the main air chamber.
  3. Carefully cycle the suspension through its entire travel.
  4. Check that all parts of the shock are clear of the frame and swingarm as it cycles through the travel.
  5. Pressurize your main air chamber to a minimum of 50 psi and no more than 300 psi. You will tune to a more specific air pressure in the Setting Sag section below.
  6. Set sag.

General Maintenance

There may be a small amount of air sleeve lubricant residue on the body. This is normal. If this residual air sleeve lubricant is not present, this is an indication that the air sleeve should be re-lubricated. Some other things to consider for all shock models:

  • If you ride in extreme conditions, service your shock and air sleeve more frequently. Check the maintenance schedule for your shock.
  • Wash your shock with soap and water only.
  • Do not use a high pressure washer to clean your shock.
  • Internal service should be performed by FOX Racing Shox or an Authorized Service Center.

Before You Ride

  1. Clean the outside of your shock with soap and water and wipe dry with a soft dry rag. Do not use a high pressure washer on your shock.
  2. Inspect entire exterior of shock for damage. The shock should not be used if any of the exterior parts appear to be damaged. Please contact your local dealer or FOX Racing Shox for further inspection and repair.
  3. Check that quick-release levers (or thru-axle pinch bolts) are properly adjusted and tightened.
  4. Check headset adjustment. If loose, adjust according to manufacturer’s recommendations.
  5. Check that brake cables or hoses are properly fastened.
  6. Check that the front and rear brakes operate properly on flat land.

Setting Sag

You can also view a Flash video on Setting Sag.

To set sag:

  1. Measure sag, and compare it to the recommended sag setting shown in the Air Spring Setting Guidelines table below. Continue if the sag is not to specification.
  2. Locate the Schrader air valve on the shock and remove the air valve cap.
  3. Screw the FOX Racing Shox High Pressure Pump onto the air valve until the pump shows pressure on the gauge. Do not over-tighten.
  4. Add air pressure until desired pressure is shown on the gauge. Refer to the Air Spring Setting Guidelines table below for the proper sag setting.
  5. Unthread the pump from the air valve and measure sag.
  6. Repeat steps 2-5 until proper sag is achieved, then replace the air valve cap.

    Air Spring Setting Guidelines

    Shock Travel
    (in./mm)

    Sag
    (in./mm)

    1.00/25.4

    0.25/6.4

    1.25/31.7

    0.31/7.9

    1.50/38.1

    0.38/9.5

    Now we recommend you to Download first result Unnai Kanda Naal Video Salim Vijay Antony Tamil HD Songs MP3 which is uploaded by Saregama Tamil of size 6.12 MB, duration 4 minutes and 39 seconds and bitrate is 192 Kbps. For your search query Unnai Kanda Naal Muthal MP3 we have found 1000000 songs matching your query but showing only top 10 results. Kanda naal mudhalai.

    1.75/44.4

    0.44/11.1

    2.00/50.8

    0.50/12.7

    2.25/57.1

    0.56/14.2

Adjusting Rebound

Rebound controls the rate at which your shock returns after it has been compressed. The proper rebound setting is a personal preference, and changes with rider weight, riding style and conditions. A rule of thumb is that rebound should be as fast as possible without kicking back and pushing the rider off the saddle.

The rebound knob has 8-10 clicks of adjustment.

For slower rebound, turn the red adjuster knob clockwise.

For faster rebound, turn the red adjuster knob counterclockwise.

ProPedal

ProPedal Lever

The ProPedal lever allows for on-the-fly ProPedal adjustment. ProPedal damping reduces pedal-induced suspension bob. The two ProPedal lever settings are:

Fox Float Rp23 Manual

  • OPEN
  • PROPEDAL

Use each setting to adjust the shock for different riding conditions and situations. For example, use PROPEDAL for riding to the top of the mountain, and then switch to OPEN for the descent.

To determine which ProPedal position is better for your condition and situation, pedal the bicycle and monitor the shock movement. Switch between positions and select the one that reduces suspension movement most effectively while providing the desired amount of bump absorption.

Because suspension designs and riding skills vary, optimal settings can vary from bike to bike and rider to rider. For more precise ProPedal tuning and to further eliminate pedal-induced bob while maintaining bump compliance, adjust the ProPedal knob. As with the ProPedal lever, switch positions and select a setting that reduces suspension movement most effectively while providing the desired amount of bump absorption.

ProPedal Knob

The 3-position ProPedal knob (shown below) allows you to adjust ProPedal firmness when the ProPedal lever is in the PROPEDAL position. The ProPedal knob only changes damping when the ProPedal lever is in the PROPEDAL position.

The ProPedal knob settings are denoted by the numbers etched onto the ProPedal knob. The three ProPedal knob settings are:

  • (1) PROPEDAL Light
  • (2) PROPEDAL Medium
  • (3) PROPEDAL Firm

Download Fox Float Rp23 Manual Lawns

To adjust the ProPedal knob:

  1. Turn the ProPedal lever to the PROPEDAL position, as shown in graphic above.
  2. Lift the ProPedal knob (see frame #2 in the drawing below).
  3. Turn the ProPedal knob clockwise (relative to the ProPedal knob facing the user) until the selection you want—1, 2, or 3—is aligned with the ProPedal lever (#3). The ProPedal knob clicks twice per setting as it turns. The first click occurs as you exit the current setting; the second click as you engage the new setting.
  4. Push the ProPedal knob into its new position (#4).

    Caution!The ProPedal knob should NOT be adjusted on-the-fly. It should only be adjusted while in a stationary position.

Turning the ProPedal knob (animated clip):


Fox

Bushing Technology & InspectionSeals & Foam RingsSeal CleaningControl DirectionOil VolumesStructural InspectionDropout Thickness InspectionTorque ValuesUnit ConversionSuspension Tuning TipsUsing the PumpImportant Safety InformationService IntervalsContact FOX ServiceWarranty InformationFOXHelp Service Web Site


Copyright © 2009
FOX Factory Inc.

Coments are closed
© 2020 - d8cmb.netlify.com
Scroll to top