Вы находитесь на странице: 1из 353

Topic Subtopic

Master the principles of engineering through fascinating do-it-yourself projects Science & Mathematics Engineering
with a registered professional engineer.

Do-It-Yourself

Do-It-Yourself Engineering
“Pure intellectual stimulation that can be popped

Engineering
into the [audio or video player] anytime.”
—Harvard Magazine

“Passionate, erudite, living legend lecturers. Academia’s


best lecturers are being captured on tape.”
—The Los Angeles Times Course Guidebook
“A serious force in American education.”
—The Wall Street Journal Professor Stephen Ressler
United States Military Academy
at West Point

Stephen Ressler is Professor Emeritus from the United States Military Academy at West
Point, where he taught for 21 years. He holds M.S. and Ph.D. degrees in Civil Engineering
from Lehigh University and is a registered professional engineer in Virginia. Professor
Ressler served in a variety of military engineering assignments in the United States, Europe,
and Central Asia. He has focused his scholarly work and professional service in the area of
engineering education and has won numerous national awards for engineering education
and service.

THE GREAT COURSES ®


Corporate Headquarters
4840 Westfields Boulevard, Suite 500
Chantilly, VA 20151-2299
USA
Phone: 1-800-832-2412
www.thegreatcourses.com
Workbook

Professor Photo: © Jeff Mauritzen - inPhotograph.com.


Cover Image: © Dougal Waters/Digital Vision/Getty Images.

Course No. 1144 © 2017 The Teaching Company. OI1144A


PUBLISHED BY:

THE GREAT COURSES


Corporate Headquarters
4840 Westfields Boulevard, Suite 500
Chantilly, Virginia 20151-2299
Phone: 1-800-832-2412
Fax: 703-378-3819
www.thegreatcourses.com

Copyright © The Teaching Company, 2017

Printed in the United States of America

This book is in copyright. All rights reserved.

Without limiting the rights under copyright reserved above,


no part of this publication may be reproduced, stored in
or introduced into a retrieval system, or transmitted,
in any form, or by any means
(electronic, mechanical, photocopying, recording, or otherwise),
without the prior written permission of
The Teaching Company.
STEPHEN RESSLER,
PH.D., P.E.,
DIST.M.ASCE
Professor Emeritus
United States Military Academy at West Point

S tephen Ressler is Professor Emeritus from the United States Military Academy at West Point, where
he taught for 21 years. He earned a bachelor of science degree from West Point, master’s and Ph.D.
degrees in Civil Engineering from Lehigh University, and a master of strategic studies degree from the
U.S. Army War College. Professor Ressler is a registered professional engineer in the commonwealth of
Virginia. He is also an amateur artist and craftsman who created all of the physical models and most of
the computer models and graphics used in this course.

Professor Ressler served for 34 years as a commissioned officer in the U.S. Army Corps of Engineers
and retired at the rank of brigadier general in 2013. He served in a variety of military engineering
assignments in the United States, Europe, and Central Asia, including 21 years as a member of the
West Point faculty. At West Point, Professor Ressler taught courses in engineering mechanics, structural
analysis, structural design, construction management, professional practice, and civil engineering
history. In 2007, he deployed to Afghanistan to develop a civil engineering program for the newly
created National Military Academy of Afghanistan in Kabul. In that capacity, he designed the civil
engineering curriculum, hired the first cohort of Afghan faculty, and developed 2 laboratory facilities.

Professor Ressler has focused his scholarly work and professional service in the area of engineering
education. He has written more than 80 scholarly papers on teaching techniques, faculty development,
curriculum assessment, engineering outreach to primary and secondary schools, engineering
accreditation, and information technology. Professor Ressler’s work has earned 9 Best Paper Awards
from the American Society for Engineering Education (ASEE) and 1 from the American Society of
Civil Engineers (ASCE).

Professor Ressler was the creator and director of the West Point Bridge Design Contest, a nationwide
Internet‑based engineering competition that engaged more than 50,000 middle school and high school
students between 2001 and 2016. He is also a developer and principal instructor for the Excellence in
Civil Engineering Education (ExCEEd) Teaching Workshop, a landmark faculty development program
sponsored by ASCE. The workshop has provided rigorous teacher training to more than 500 civil
engineering faculty members from more than 200 colleges and universities over the past 14 years.

Professor Ressler has won numerous national awards for engineering education and service. From
ASCE, he received the President’s Medal, the ExCEEd Leadership Award, and the John I. Parcel‑Leif
J. Sverdrup Civil Engineering Management Award. ASCE also named him a distinguished member in
2005. From ASEE, Professor Ressler received the George K. Wadlin Distinguished Service Award, the

Professor Biography i
Distinguished Educator Award, and the Dow Outstanding New Faculty Award. In 2017, he was named
a fellow of ASEE. Professor Ressler also received the Society of American Military Engineers’ Bliss Medal
for Outstanding Contributions to Engineering Education, the American Association of Engineering
Societies’ Norm Augustine Award for Outstanding Achievement in Engineering Communications, the
Premier Award for Excellence in Engineering Education Courseware, and the EDUCOM Medal for
application of information technology in education.

Professor Ressler was named one of Engineering News‑Record’s Top 25 Newsmakers who served
construction in 2000. In 2011, he received ASCE’s highest award—the Outstanding Projects and
Leaders Award, which is presented to only 5 of ASCE’s 140,000 members each year.

Professor Ressler’s other Great Courses are Everyday Engineering: Understanding the Marvels of Daily
Life; Understanding Greek and Roman Technology: From Catapult to the Pantheon; and Understanding the
World’s Greatest Structures: Science and Innovation from Antiquity to Modernity.

ii 
TABLE OF CONTENTS

INTRODUCTION

Professor Biography . . . . . . . . . . . . . . . . . . . . . . . . . . . i
Disclaimer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v
Project Descriptions . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

GUIDES

Project 1 Golf Ball Launcher . . . . . . . . . . . . . . . . . . . . . . 3


Lesson 1 » Why DIY Engineering?

Project 2 Cardboard Tower . . . . . . . . . . . . . . . . . . . . . . 15


Lesson 2 » Exploring the Science of Structure
Lesson 3 » Design and Build a Cardboard Tower

Project 3 Beam Bridge . . . . . . . . . . . . . . . . . . . . . . . . . 38


Lesson 4 » Bridging with Beams

Project 4 Suspension Bridge . . . . . . . . . . . . . . . . . . . . . 55


Lesson 5 » Make a Suspension Bridge

Project 5 Concrete Sailboat . . . . . . . . . . . . . . . . . . . . . 76


Lesson 6 » Design a Concrete Sailboat
Lesson 7 » Set Sail!

Project 6 Radio-Controlled Blimp . . . . . . . . . . . . . . . . . 104


Lesson 8 » Make a Radio-Controlled Blimp

Project 7 Rubber-Powered Airplane . . . . . . . . . . . . . . . . . 121


Lesson 9 » Exploring Aerodynamics
Lesson 10 » Build a Model Airplane
Lesson 11 » Take Flight!

Project 8 Rubber-Powered Helicopter . . . . . . . . . . . . . . . . 147


Lesson 12 » Build a Model Helicopter

Table of Contents iii


Project 9 Camera-Equipped Rocket . . . . . . . . . . . . . . . . . 164
Lesson 13 » This Is Rocket Science
Lesson 14 » Build a Rocket
Lesson 16 » Let’s Do Launch!

Project 10 Electric Launch Controller . . . . . . . . . . . . . . 190


Lesson 15 » Make an Electric Launch Controller

Project 11 Ballista, Onager, and Trebuchet . . . . . . . . . . . 199


Lesson 17 » A Tale of Three Catapults
Lesson 18 » Build a Ballista, Onager, and Trebuchet

Project 12 Hydraulic Arm . . . . . . . . . . . . . . . . . . . . . . 228


Lesson 19 » Design a Hydraulic Arm

Project 13 Water Turbine . . . . . . . . . . . . . . . . . . . . . . 244


Lesson 20 » Make a Water Turbine

Project 14 Gear Train . . . . . . . . . . . . . . . . . . . . . . . . 257


Lesson 21 » Design a Gear Train

Project 15 Wooden Pendulum Clock . . . . . . . . . . . . . . . . . 268


Lesson 22 » Make a Mechanical Clock

Project 16 Electric-Powered Crane . . . . . . . . . . . . . . . . . 283


Lesson 23 » Design a Motor-Powered Crane

Project 17 Tribute to Rube Goldberg . . . . . . . . . . . . . . . . 297


Lesson 24 » Creative Design: A Tribute to Rube Goldberg

SUPPLEMENTARY MATERIAL

Glossary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 300
Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 317
Appendix A
Structural Testing Machine . . . . . . . . . . . . . . . . . . . . . 319
Appendix B
Cutting Patterns and Layouts . . . . . . . . . . . . . . . . . . . . . 322

iv 
DISCLAIMER

This series of lessons includes a variety of demonstrations and hands-on projects intended to increase
your understanding of math, science, and engineering. These demonstrations and projects are performed
by an experienced professional and involve extensive use of tools, equipment, and materials that can be
dangerous if not used with extreme caution.

WARNING: THE DEMONSTRATIONS AND PROJECTS PERFORMED IN THESE LESSONS


CAN BE DANGEROUS. ANY ATTEMPT TO PERFORM THESE DEMONSTRATIONS OR
PROJECTS ON YOUR OWN IS UNDERTAKEN AT YOUR OWN RISK.

The Teaching Company expressly DISCLAIMS LIABILITY for any DIRECT, INDIRECT,
INCIDENTAL, SPECIAL, OR CONSEQUENTIAL DAMAGES OR LOST PROFITS that result
directly or indirectly from the use of these lessons. In states that do not allow some or all of the above
limitations of liability, liability shall be limited to the greatest extent allowed by law.

Table of Contents v
vi 
PROJECT DESCRIPTIONS

This guidebook provides comprehensive descriptions for each of the 17 do‑it‑yourself projects presented
in this course. Each of these descriptions includes the following:

}} problem definition
}} references
}} design concept
}} detailed design (models, experimentation, and analysis calculations)
}} final design drawings
}} materials list
}} fabrication and assembly procedure
}} testing procedure

In addition, the PDF version of this guidebook includes full‑size cutting patterns and layout drawings
for all relevant projects (See Appendix B).

Throughout the course, Professor Ressler makes extensive use of SketchUp software for 3‑D modeling
and design. The free version of this software—called SketchUp Make—can be obtained from
http://www.sketchup.com/. If you would like to learn how to use SketchUp, it is strongly recommended
that you work through the tutorials provided at https://www.sketchup.com/learn.

Professor Ressler has provided his SketchUp 3‑D models for all 17 course projects on his personal
website at http://stephenjressler.com/diy-engineering/.

The projects in this course require basic‑level woodworking and metalworking skills. The following
major pieces of shop equipment and tools are used (though not all are used for all projects):

}} table saw
}} miter saw or radial arm saw
}} bandsaw or scroll saw
}} drill press (with sets of twist drill bits and Forstner bits)
}} sanding station with belt and disk sanders
}} spindle sander
}} shop vacuum
}} benchtop vise
}} clamps
}} various hand tools (see individual project descriptions for specific requirements)

If you attempt to do any of the projects presented in this course, please pay careful attention to the safety
overview presented in lesson 1 and all of the specific safety advisories presented throughout the course.

Project Descriptions 1
2 
PROJECT 1
GOLF BALL
LAUNCHER
DESIGN, BUILD, AND TEST A MECHANICAL
DEVICE CAPABLE OF LAUNCHING A GOLF BALL
AND HITTING A TARGET PLACED EXACTLY
10 FEET AWAY FROM THE LAUNCHER.
LESSON IN WHICH THIS PROJECT IS COVERED:
1 » Why DIY Engineering?

PROBLEM DEFINITION
Requirements
}} Must be capable of launching a standard‑size golf ball.
}} Must be capable of hitting an 8‑inch‑diameter target placed exactly 10 feet away from the launcher
and at the same elevation as the launcher.

Constraints
}} Project budget is $20.
}} Must be fabricated from commonly available off‑the‑shelf materials and components.
}} Combustion is strictly forbidden.

Project 1 » Golf Ball Launcher 3


Additional Information The Science of Springs
}} Weight of golf ball = 1.62 ounces = 0.101 pounds
}} Diameter of golf ball = 1.68 inches A spring is a device that stores energy
}} Acceleration of gravity = g = 32.2 feet per through the elastic deformation of a
material. There are many different
second squared
types, but the most common is the coil
spring, which you encounter every day in
common devices like ballpoint pens.

Coil springs come in two different


REFERENCES configurations: tension springs,
which are designed to be stretched,
and compression springs, which are
Hibbeler, Engineering Mechanics. designed to be shortened. The only
Kosky, Balmer, Keat, and Wise, Exploring Engineering. physical difference between the two is
that the compression spring has some
space between the coils, so there is room
for shortening, while the tension spring
has its coils packed tightly together,
to allow for the largest‑possible tensile
DESIGN CONCEPT elongation.

The mechanical behavior of both types


LAUNCH TUBE is essentially identical. When a spring is
deformed, it provides a resisting force.
PISTON
The magnitude of the resisting force
SPRING is directly proportional to the spring’s
STOCK
TRIGGER change in length.

The most important mechanical property


of a spring is its stiffness, defined as the
amount of resisting force developed in
the spring per unit change of length.

BASE The elastic energy stored in a spring


is equal to 1/2ks2, where k is the spring
stiffness and s is the spring’s change
in length.

Key Features
}} Spring‑driven piston is guided by a clear plastic launch tube.
}} Steel trigger pin engages with stem of piston.
}} Launch tube, piston, and trigger are mounted on a wooden stock.
}} Stock pivots on wooden base to set launch angle.

4 
DETAILED DESIGN
Model of the Mechanical System

Conservation of Energy

where:

General approach:

}} Use projectile range equation to determine the launch velocity, v, required for a range of 10 feet.
}} Assume s.
}} Solve the conservation of energy equation for k.

Project 1 » Golf Ball Launcher 5


Projectile Range Equation
Conservation
of Energy
The principle of conservation
of energy says that the total
energy of an isolated system must
where:
remain constant. An important
implication of this principle is that
energy can neither be created nor
destroyed, but can be converted
from one form to another.

In this course, the forms of energy


Solve for v:
we will use in applying the principle
of conservation of energy are:

kinetic energy = 1/2mv2 = the


energy associated with a mass,
Let θ = 30° to allow for range adjustments by varying the m, moving at velocity, v.
launch angle.
gravitational potential energy
= mgh = the energy associated
Substitute known values: with a mass, m, elevated at a
height, h, above a reference
elevation and subjected to the
acceleration of gravity, g.

spring energy = 1/2ks2 = the


elastic energy stored in a spring
with stiffness, k, that has
Substitute v into the conservation of energy equation: undergone a change in length, s.

Assume s = 2 inches:

Now let’s go shopping! Select a 3‑inch steel compression spring from McMaster‑Carr
(https://www.mcmaster.com; part number 9657K428):

6 
Solve for actual spring compression, s:

Therefore, we will use a 3‑inch steel compression spring from McMaster‑Carr (part number 9657K428),
and we will design the trigger mechanism such that the spring is compressed 1.8 inches.

FINAL DESIGN DRAWINGS

The SketchUp 3‑D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/.

Drawing 1.1 » Golf ball launcher,


side and front elevation views

6 7/32"
5 1⅓2" 2 ¾"

8 ¼"

4 ¼"

5" 1 ¾"
10" 3 ½"

SIDE ELEVATION FRONT ELEVATION

Project 1 » Golf Ball Launcher 7


Drawing 1.2 » Golf ball
launcher, perspective
view, angled at 30°,
projectile loaded

R
K P
H I P
G
M N O
F
E Q
J

C Drawing 1.3 » Golf ball


launcher, exploded view
L
B

8 
MATERIALS LIST
Wooden Components

NAME DESCRIPTION
A Lower base 10″ × 31/2″ × 3/4″ poplar
B Upper base 51/4″ × 11/2″ × 3/4″ poplar*
C Base pivot mount 31/2″ × 11/2″ × 3/4″ poplar; 1/4″‑diameter hole*
D Stock pivot mount 9″ × 11/2″ × 3/4″ poplar; 1/4″‑diameter hole*
E Stock base 91/4″ × 31/2″ × 3/4″ poplar
F Front launch tube holder 31/2″ × 31/4″ × 3/4″ poplar; 2″‑diameter hole*
G Piston guide 31/2″ × 31/4″ × 1/4″ poplar*
H Trigger mount 2″ × 23/4″ × 3/4″ poplar; 1/8″‑diameter hole*
I Rear launch tube holder 13/4″ diameter × 3/4″ poplar*
J Piston disk 111/16″ diameter × 1/4″ poplar*
K Launch angle protractor 21/4″ × 21/4″ × 1/16″ plywood*
*cutting pattern provided (see Appendix B)

Hardware and Other Components

NAME DESCRIPTION
L Pivot 1/4″ × 2″ machine bolt with nut and 2 washers
M Trigger pin 1/8″ steel wire
N Spring 3″ steel compression spring (McMaster‑Carr 9657K428)
O Piston shaft 1/4″‑diameter aluminum tube, 5½″ long
P Retainer pins (2) cut from 4d steel nails
Q Launch tube clear polycarbonate tube, 2″ outside diameter, 13/4″ inside diameter
(McMaster‑Carr 8585K26); 61/4″ long
R Launch angle indicator fishing sinker suspended from 4d nail by thread
— 11/4″ wood screws (10)

Project 1 » Golf Ball Launcher 9


FABRICATION AND ASSEMBLY PROCEDURE
1 Begin by prefabricating all of the wooden components listed above, using the procedures demonstrated
in lesson 1. Print the full‑size cutting patterns provided with the PDF version of this guidebook. Cut
them out with scissors, a hobby knife, or a razor blade, and then adhere them temporarily to the wood
with spray adhesive. For components with complex shapes, cut slightly oversize, using a scroll saw or
band saw, and then use a power sander to achieve the final shape.

2 The holes in components C, D, F, G, H, I, and J should be drilled with a drill press to ensure that each
hole is perfectly perpendicular to the surface of its associated workpiece. The 2‑inch hole in component
F can be cut with a hole saw, as demonstrated in lesson 1.

3 After each component has been fabricated, its paper cutting pattern can be removed with lacquer
thinner. However, the pattern for component K (the launch angle protractor) should not be removed
from the component.

4 Assemble the base by gluing components A, B, and C together with yellow wood glue, clamping them
together until the glue has set, and then reinforcing the glue joints with 11/4‑inch wood screws driven
into pilot holes, as shown below.

C
4

10 
5
5 Assemble the stock, first join
components D and E with glue
and screws.

I
6 N
 ext, glue components G, H, and I
H together as a subassembly.
G

7
G

H
I
7 Finally, attach subassembly
G/H/I and component F to F
the top of component E.
E

Project 1 » Golf Ball Launcher 11


8 To complete the stock, slide the 8
plastic launch tube (component Q)
into position and glue the launch
angle protractor (component K) to
the side of component G. The launch Q
angle indicator (component R) is G
a fishing sinker (or similar weight)
K
suspended from a 4d nail by a short
length of thread.
R

9 A
 ssemble the stock to the base with a 1/4‑inch steel
bolt (component L), which will serve as a pivot.

10 The piston shaft (component O) is cut from a piece of 1/4‑inch‑diameter aluminum tubing and drilled
with a 1/8‑inch hole for the trigger pin and two 5/64‑inch holes for the retainer pins (component P),
which are cut from steel 4d nails. The holes are drilled at the locations shown below. Note that the hole
for the trigger pin is drilled only halfway through the tube.

5 ½"
10 ¾" 1 ¾" 2 ⅝" ⅜"

P P

12 
11 The piston disk (component J) P 11
is glued to the front of the piston
shaft (component O) and the P
front retainer pin (component P) O
using epoxy glue.
J

12

P 12 A fter the glue has dried, the piston assembly is


inserted through the spring, then through the
rear of the stock, and then secured with the rear
retainer pin (component P).

13

H
13 The trigger pin (component M) is made from steel
M
wire bent into an L shape. It is inserted into the
1/8‑inch hole in component H.

The golf ball launcher is now complete and ready for testing!

Project 1 » Golf Ball Launcher 13


TESTING PROCEDURE

To test the golf ball launcher, clamp the base to a tabletop and set up a target 10 feet away at the same
elevation as the launcher. Initially, set the launch angle at 30°. Cock the launcher, set the trigger, and load
the projectile. Aim and pull the trigger. If the projectile falls short, increase the launch angle in small
increments until the range is correct.

14 
PROJECT 2
CARDBOARD
TOWER
DESIGN, BUILD, AND TEST A 3‑FOOT‑TALL TOWER
WITH A STRUCTURAL SYSTEM MADE ENTIRELY
OF CARDBOARD AND CAPABLE OF SUPPORTING
A 100‑POUND GRAVITY LOAD AND A 10‑POUND
LATERAL LOAD APPLIED AT ITS TOP LEVEL.
LESSONS IN WHICH THIS PROJECT IS COVERED:
2 » Exploring the Science of Structure
3 » Design and Build a Cardboard Tower

PROBLEM DEFINITION
Requirements
}} The tower must be 36 inches tall, from its base to the level at which the gravity load is applied.
}} The gravity load will consist of six 12‑inch square concrete pavers, which weigh 17 pounds each.
}} The lateral load will be applied by pulling horizontally with a luggage scale until a reading of 10
pounds is achieved.

Constraints
}} The only allowable materials are corrugated cardboard and manila file-folder cardboard.
}} Use only wood glue for adhesive.

Project 2 » Cardboard Tower 15


REFERENCES How Does a Braced
Hibbeler, Engineering Mechanics. Frame Work?
——— , Mechanics of Materials.
This simple frame is composed
of two columns and one beam
connected with a single pin or bolt
at each corner. In this rectangular
configuration, the structure is
DESIGN CONCEPT unstable and incapable of carrying
lateral load.

PAVERS But when a diagonal brace is


100
added, the rectangular panel is
LB
subdivided into two triangles.
10 LB
The resulting structure—called
a braced frame—is stable and
9” capable of carrying substantial
lateral loads.
COLUMN

9”
BEAM
36”

DIAGONAL 9”

9”

10”
Key Features
}} This type of structural system—called a braced
frame—consists of columns, beams, and
diagonal braces.
}} The columns are spaced as far apart as possible
to enhance the resistance of the structure to overturning, but they must also be sufficiently close
together that 12‑inch pavers can rest directly on top of the columns.
}} Beams are used to subdivide the 36‑inch columns into four stories. The resulting intermediate
bracing significantly increases columns’ resistance to buckling.
}} The diagonals provide stability and resistance to lateral load. These members cause the structural
system to be classified as a braced frame.

16 
B asic Structural Engineering Concepts
The engineers who design major structures, such as
One World Trade Center, must meet a daunting
set of challenges. They must create a design that
is safe, beyond a shadow of a doubt—not just
today, but over the building’s lifetime of perhaps
a hundred years or more. The design must also
be reasonably economical, because no structure
will get built if its owner cannot afford the cost of
construction. And most impressively, the design
must be completed in intricate detail before
construction ever begins.

How is this done? How can engineers predict the


behavior of such an incomprehensibly complex
structural system with such confidence before
the first shovelful of earth is turned? The answer
to this question lies in the science of structural
mechanics, which will be the focus of this and
the following two projects. In these projects, we
will see how structural engineers apply scientific
principles and mathematical models to design a
© ElisabethOstensvik/iStock/Thinkstock. variety of structures.

Three fundamental concepts of structural mechanics will guide every aspect of this project:

1 Loads acting on a structural system cause internal forces in all of the members constituting that
system.

2 Each member has a characteristic strength, which is based on the physical characteristics of the
member and its constituent materials.

3 The structural system is capable of carrying load safely if the internal force in each member is
less than the associated strength of that member.

A load is simply a force applied to a structure. For the purpose of structural design, loads are often
placed into two major categories:

†† Gravity loads are the downward forces associated with anything that has weight—the building
itself, occupants, furniture, etc.

†† Lateral loads are the predominantly horizontal forces associated with wind and earthquake
effects.

The structural system of One World Trade Center must simultaneously support gravity loads totaling in
excess of 1 billion pounds and lateral load due to wind alone on the order of 10 million pounds!

Project 2 » Cardboard Tower 17


DETAILED DESIGN

For this project, the goal of the detailed design phase is to determine the specific size and shape of every
member in the structural system of our tower such that the system will carry the required loads safely.

Forces, Vectors, and the Principle of Equilibrium


A force is simply a push or a pull applied to a body. According to the principle of equilibrium, if a
body is not moving or is moving at a constant velocity, then we can conclude that all of the forces acting
on the body are in balance. In short, the body is in equilibrium.

For the purpose of analysis, a system of forces is F = 5 lb


typically represented using a simple graphical tool
called a free body diagram. To construct this 37°
diagram, start by sketching the body of interest 4 lb
isolated from its surroundings. Then, draw all
forces acting on the body, representing each force y
as a vector—a mathematical entity that has
both magnitude and direction. Finally, add an
x‑y coordinate axis system, which will serve as a x 3 lb
reference for the analysis.

Now we can use the free body diagram as the basis for proving that this system of forces is in
equilibrium. Fy F = 5 lb

Restating the principle of equilibrium, if a body is in equilibrium, the vector sum of all forces acting on
37°
the body must equal zero. When we are working in two dimensions, this condition can be represented as
4 lb
two mathematical equations: F
x
y
†† The sum of all forces acting in the x‑direction equals zero, written as .

†† The sum of all forces acting in the y‑direction equals zero, written
x as 3 lb .

In applying these equations to our free body diagram, we see that the 4‑pound force is clearly oriented
in the x‑direction and the 3‑pound force is clearly oriented in the y‑direction, but what about the
diagonally oriented 5‑pound force?

To deal with this force, we must represent it as an equivalent pair of forces: one oriented in the
x‑direction and one oriented in the y‑direction. This is an extremely important vector math operation
that we will be using often throughout this course for a variety of purposes.

18 
37°
4 lb

x 3 lb

To perform this operation, first draw a rectangle


around the force, with the sides of the rectangle Fy F = 5 lb
parallel to the x‑ and y‑axes and with the original force
forming its diagonal. Now the two equivalent forces—
37°
which are called the x‑component (designated Fx )
4 lb
and y‑component (designated Fy ) of the original Fx
force—are represented by the rectangle’s horizontal y
and vertical sides, as shown here.

The magnitudes of these two components can be x 3 lb


calculated using trigonometry, as follows:

Now applying the principle of equilibrium to our revised free body diagram, we begin by summing forces
in the x‑direction. Assuming that forces oriented to the right are positive and those oriented to the left are
negative, the associated equilibrium equation is:

Next, summing forces in the y‑direction and assuming that forces oriented upward are positive and those
oriented downward are negative:

Because all forces add to zero in both the x‑ and y‑directions, we can conclude that the body is indeed
in equilibrium. This concept of equilibrium (and the associated equations of equilibrium) constitute an
extraordinarily powerful problem‑solving tool that will be applied throughout this course.

Project 2 » Cardboard Tower 19


Model of the Structural System

The design for our tower will be based on a structural analysis, the objective of which is to calculate the
internal forces in every element of the braced frame. To set the stage for this analysis, we must first model
the structure and its loads.

The structural model, shown here, is a free body


diagram of one of the tower’s two main frames. The 25 lb 25 lb
body of interest is the frame itself, annotated with all
relevant dimensions and with the letters A through J 5 lb I J
designating the joints at which the frame members are
interconnected. θ
8.75"
Because the 100‑pound gravity load will actually
rest directly on the tops of all four columns in G H
the three‑dimensional structural system, it can
be represented as four 25‑pound downward loads 8.75"
applied to the column tops. Two of these 25‑pound
loads will be applied to our two‑dimensional frame. E F
Similarly, half of the 10‑pound lateral load is applied
as a 5‑pound load at the top of the two‑dimensional 8.75"
model. At the bottoms of the columns, where the
structure will be supported on its foundation, three C D
unknown forces—called reactions—are applied.
8.75"
Reactions are forces developed at a structure’s Bx
supports to keep it in equilibrium with its applied y A
loads. In our model, the vertical reactions Ay and By B
represent restraints that prevent the structure from
Ay 10"
By
moving downward in response to the gravity load, x
and the horizontal reaction B x represents a restraint
that prevents the structure from sliding sideways in
response to the lateral load.

To perform our structural analysis, we will need to know θ, the angle at which the diagonal braces are
inclined. Applying the definition of tangent from trigonometry:

Therefore:

20 
Structural Analysis (Method of Joints)

Analysis of Joint J

We will begin the structural analysis with joint The Method of Joints
J, because only two members (IJ and HJ) are
connected at this joint; thus, there will only be When a structural member experiences an
two unknown forces (FIJ and FHJ) for which to internal force that causes it to stretch or elongate,
solve. This is important because we only have two it is said to be in tension. When a member
equations of equilibrium, and , experiences an internal force that causes it to
shorten, it is said to be in compression.
available at each joint.
For a structure in which all members are
Free body diagram: subjected only to pure tension or compression
(i.e., no members experience significant flexure
25 lb or bending), the structural analysis can be
performed using the method of joints.

The method of joints is implemented in four


steps, as follows:
FIJ
J 1 Isolate a joint.

2 Draw a free body diagram of the joint.


y
3 Apply the principle of equilibrium to
FHJ solve for unknown member forces and
reactions at the joint.
x

Equilibrium:

4 Repeat for all remaining joints.

When drawing the free body diagram of a


joint, it is advisable to assume that all unknown
internal forces are in tension. Then, if the
calculated force is positive, we can conclude
that the assumption was correct: The internal
force is indeed in tension. If the calculated result
is negative, then the internal force must be in
compression.

Project 2 » Cardboard Tower 21


Analysis of Joint I

Next, we will proceed to joint I. Although there are three members connected at this joint (IJ, HI, and GI),
we have already solved for the force in member IJ (FIJ = 0), so we will be able to solve for the remaining two
unknown forces using the two available equations of equilibrium.

Free body diagram: Equilibrium:

25 lb

5 lb I FIJ
FHIx
41.2°
y
FGI
FHIy FHI
x

Analysis of Joint H

From this point forward, we will step from joint to joint, taking advantage of previously calculated forces to
ensure that we can use our two equilibrium equations to solve for the remaining unknowns at each joint.

Free body diagram: Equilibrium:


FHI FHIy
FHJ

41.2°
FHIx
H
y FGH

x
FFH

22 
Analysis of Joint G

Free body diagram: Equilibrium:


FGI

FGH
G
FFGx
41.2°

y FEG
FFGy FFG

Analysis of Joint F

Free body diagram: Equilibrium:


FFG FFGy
FFH

41.2°
FFGx

F
y FEF

x
FDF

Project 2 » Cardboard Tower 23


Analysis of Joint E

Free body diagram: Equilibrium:


FEG

FEF

E
FDEx
41.2°

y FCE
FDEy FDE

Analysis of Joint D

Free body diagram: Equilibrium:


FDE FDEy
FDF

41.2°
FDEx
D
y FCD

x
FBD

24 
Analysis of Joint C

Free body diagram: Equilibrium:


FCE

FCD
C
FBCx
41.2°

y FAC
FBCy FBC

Analysis of Joint B

Free body diagram: Equilibrium:

FBC FBCy
FBD

41.2°
FBCx Bx
B

y (Note that the direction of a reaction force is


indicated with an arrow, not with compression
or tension.)
x By

Project 2 » Cardboard Tower 25


Analysis of Joint A

Free body diagram: Equilibrium:

FAC

y
Ay
x

Summary of Structural Analysis Results

Based on these results, we can make the


MEMBER INTERNAL FORCE
following design decisions:
AC 42.5 lb (compression)
CE 38.1 lb (compression) }} Given that the columns will experience a
EG 33.8 lb (compression) maximum compressive force in excess of 40
pounds, we will fabricate these members
GI 29.4 lb (compression)
Columns by forming corrugated cardboard into a
BD 11.9 lb (compression) hollow square cross‑section for maximum
DF 16.3 lb (compression) compressive strength. (See “Strength of
FH 20.6 lb (compression) Structural Members” sidebar.)
}} The subdivision of the tower’s 36‑inch
HJ 25.0 lb (compression)
height into four stories also increases the
CD 5.0 lb (compression) compressive strength of the columns by
EF 5.0 lb (compression) reducing the lengths of each individual
Beams column segment to approximately 9 inches.
GH 5.0 lb (compression)
}} The beams will experience a much smaller
IJ 0
compressive force of 5 pounds, so we can
BC 6.6 lb (tension) use significantly smaller tubes for these
DE 6.6 lb (tension) members.
Diagonals }} The diagonals will experience a tensile force
FG 6.6 lb (tension)
of 7 pounds, which can be carried quite
HI 6.6 lb (tension) safely with strips of file‑folder material.

26 
Experimental Determination
of Tensile Strength
Strength of
Structural Members
The tensile strength of file‑folder cardboard has been determined
experimentally by using a lever‑based testing machine to load In general, the strength of a
cardboard specimens of various widths until they rupture. structural member is the internal
Drawings and a materials list for the testing machine are force at which the member fails.
Most structural members have
provided in appendix A. Use of this machine is demonstrated
significantly different strengths in
in lesson 3. Results of the tests are shown in the graph of tensile tension and compression:
strength versus member width below.
†† When a member is loaded
25 in tension, it fails by
TENSILE STRENGTH (POUNDS)

physically breaking in two.


This failure mode is called
20
rupture, and the internal
force at which it occurs is
15 the tensile strength of the
member. Tensile strength
depends only on the
10 cross‑sectional area of the
member and the material
of which it is made.
5
†† When a member is loaded
0 in compression, it usually
fails by buckling—a form
0 0.2 0.4 0.6 0.8 1 1.2
of instability in which the
MEMBER WIDTH (INCHES) member suddenly bends
sideways. The compressive
strength (or buckling
strength) of a member
Design of Diagonal Members depends not only on
the type of material and
The general mathematical requirement for the design of any cross‑sectional area but
also on the cross‑section
structural member is as follows:
shape and, most
importantly, the member
length.

In general, a hollow tube has


For this project, this requirement is defined more specifically as: significantly greater buckling
strength than a solid cross‑section
with the same area and length (i.e.,
the same amount of material). And
shorter members have significantly
greater buckling strength than
Rearranging algebraically, we arrive at an inequality that will longer ones.
serve as the design basis for all of the members in our tower:

Project 2 » Cardboard Tower 27


From the structural analysis results, the internal force in all
diagonals is 6.6 pounds. Therefore, for these members:
The Factor of Safety
In engineering design, the factor
of safety is simply a number—
always greater than 1—that
provides a margin of error to
account for the many sources of
Based on our experimentally derived graph of tensile uncertainty inherent in structural
strength versus member width (above), the required design: unanticipated loads,
natural variability in material
strength of 11.6 pounds can be achieved with a member
properties, fabrication errors, etc.
width of 0.6 inches. We will round up to 5/8 inch for ease of
measurement. In engineering practice, factors
of safety are normally specified
in design codes and typically
Design of Columns range between 1.5 and 2, with
the actual number determined by
such considerations as the type of
In our structural analysis, we calculated a range of different structural member, material, type
internal forces in the various column segments; however, of loading, and consequences of
for the sake of simplicity, it is reasonable to use the absolute failure. For our project, we will use
largest force—43 pounds—as the basis for selecting a a factor of safety of 1.75, which
single member size that will be used for all columns in the gives us a reasonably comfortable
75% margin of error.
tower. Therefore, for the columns:

Using the lever‑based testing machine, the compressive strength of 9‑inch‑long corrugated cardboard tubes
is found to be:

}} 63 pounds for a 3/4‑inch‑square tube (formed over a 1/2‑inch‑square core)


}} 84 pounds for a 1‑inch‑square tube (formed over a 3/4‑inch‑square core)

Based on these results, we must use a 1‑inch‑square tube (formed over a ¾‑inch‑square core) to achieve the
required strength of 75 pounds.

Design of Connections and Beams

Before designing the beams, we must define the configuration of the actual three‑dimensional structural
system such that the beams, columns, and diagonals will all fit together in a way that transmits internal
forces safely between members.

28 
This configuration has the following key characteristics:

GUSSET PLATE
}} Each connection will use a pair
of gusset plates, which are
BEAM
sandwiched around the beam
and column to tie all connected
members together. DIAGONAL
}} To maintain the symmetry of the
frame, the diagonals are configured
in parallel pairs, with one attached COLUMN
to each gusset plate. Thus, while our
design calls for one 5/8‑inch‑wide strip
of file‑folder cardboard for each diagonal,
we will actually use two 5/16‑inch strips to
provide the required symmetry while also
maintaining the required strength.
}} To incorporate a beam into this connection, the
beam width must be the same as the column width
so that the beam will fit snugly between the two gusset plates.

For the beams, the required compressive strength is:

To meet this requirement, we will use a corrugated cardboard tube with a rectangular cross‑section
formed around a 1/4‑inch‑by‑3/4‑inch core. Tests of this member size indicate that its compressive strength is
approximately 25 pounds, significantly above the 8.8‑pound required strength.

FINAL DESIGN DRAWINGS

The SketchUp 3‑D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/.

A full‑size layout drawing for the structure and templates for the gusset plates are provided in the PDF
version of this guidebook (see Appendix B).

Project 2 » Cardboard Tower 29


Drawing 2.1 » Tower perspective view Drawing 2.2 » Tower elevation
view, showing three
gusset-plate designations

B B

C C

C C 36"

C C

A A
11 ¼"

30 
Two aspects of this design warrant further explanation:

}} On each main load‑carrying frame, a second set of diagonals has been added within each panel such
that the paired diagonals now form an X. The purpose of this second set of diagonals is to ensure that
the frame is able to carry lateral load applied in either direction.
}} The two main load‑carrying frames are connected together with both beams and diagonals. The
principal purpose of these members is to provide stability perpendicular to the plane of the main
frames. Without these diagonals, the entire frame would be susceptible to toppling sideways under
gravity loads, just like a house of cards.

MATERIALS LIST
NAME DESCRIPTION
Column (8) 1″ × 1″ × 18″ corrugated cardboard tube formed around a 3/4″ × 3/4″ core
Beam (16) 1″ × 1/2″ × 8 3/4″ corrugated cardboard tube formed around a 3/4″ × 1/4″ core
Diagonal (48) 5/16″ × 13 5/8″ manila file‑folder cardboard
Gusset plate A (12) See full‑size template.
Gusset plate B (12) See full‑size template.
Gusset plate C (36) See full‑size template.
Connecting plates (32) See full‑size template.
Lower footing (8) 2″ × 2″ corrugated cardboard
Upper footing (8) 3/4″ × 3/4″ corrugated cardboard
Base 14″ × 14″ × 3/4″ plywood or MDF

The following supplies and tools are used to construct the tower:

}} one 18″ cardboard moving carton


}} one legal‑size file folder
}} wooden base
}} wood glue
}} wax paper
}} sewing pins
}} 3/4″ × 3/4″ metal or wooden core for forming columns
}} 3/4″ × 1/4″ metal or wooden core for forming beams
}} scrap wood for tube‑forming jig
}} hobby knife or single‑edge razor blade
}} scissors
}} clamps
}} weights (to be used as clamps)

Project 2 » Cardboard Tower 31


FABRICATION AND ASSEMBLY PROCEDURE
1 Begin by fabricating the corrugated cardboard tubes: 8 for the columns and 16 for the beams. For each
member, cut out a cardboard rectangle measuring 18 inches by 5 inches for columns and 8 3/4 inches by
4 inches for beams. Then, using the layout drawing below, measure and mark the fold lines (indicated
by dotted lines in the drawing) and use a ballpoint pen to crease each line.

⅞"

1"

1" 5"

1 ⅛"

direction of corrugations 1"

18"
COLUMN LAYOUT (not to scale)

⅞"

½"

1"
4"
⅝"

direction of corrugations 1"

8.75"
BEAM LAYOUT (not to scale)

32 
2 The process of forming these tubes is best accomplished by using a simple jig to ensure both uniformity
and strength. Key dimensions of the jig are shown below.

2
CARDBOARD TUBE
UPPER JIG
⅞"

METAL CORE

1' 8"
1"

LOWER JIG

1"

3 To form each tube, fold the rectangle inward along each crease, wrap the cardboard around the core,
and then slide the tube and core downward into the jig, with the overlapping face oriented upward.
Run a bead of glue along the inside of the outer flap, then close the overlap, place the upper jig
on top of the fold, and clamp it in position until the glue sets. After about 5 minutes, remove the
cardboard tube and core from the jig and slide out the core to complete the member. (This procedure is
demonstrated in lesson 3.)

4 To fabricate the 48 diagonals, draw parallel lines 5/16 inch apart on a legal‑size file folder and then slice
along each line with a hobby knife or razor blade and a straightedge.

5 To fabricate the gusset plates and connecting plates, print the templates provided in the PDF version
of this guidebook. Use spray adhesive to attach these templates to sheets of cardboard and then cut out
each component with a knife or scissors. Be sure to remove the paper templates after these components
have been cut out.

Project 2 » Cardboard Tower 33


BEAM
6 Now glue a rectangular
connecting plate to each end of
each beam, as shown here, and
6 set them aside to dry.

CONNECTING PLATE

7 The frames are assembled directly


7
on a building board, prepared by
gluing a thin sheet of cork to a sheet
of ¾‑inch particleboard or plywood.
Print the full‑size layout drawings
(contained in the PDF version of
this guidebook) and place them
onto the building board, followed
by a sheet of wax paper, to ensure
that the cardboard components are
not accidentally glued to the paper
drawing.

8 To begin the assembly process,


pin the gusset plates in their
proper positions on both frame
drawings and then glue the
diagonals to the gussets, using
a pin to anchor each end of
each diagonal.

34 
9 On one of the two frames, glue the
columns to the gusset plates, with 9
the overlapped side of each tube
facing inward, and glue the beams
to both the gussets and columns.
Place a weight over each joint to
hold everything together while the
glue sets.

10
10 Finally, remove this frame from the
board. Apply glue at the gusset‑plate
locations, flip it around, position it on
the adjacent subassembly of gusset plates
and diagonals, and replace the weights.

11 Once the glue is completely dry, remove


this completed frame from the board
and build another one just like it.

12
12 Now, using the same procedure,
build the two subassemblies of
gusset plates, transverse beams,
and diagonals over the same set
of plans.

Project 2 » Cardboard Tower 35


13 Once these subassemblies are dry,
13
glue the two main frames onto one
of these subassemblies.

14
14 Finally, remove this entire assembly
from the board and glue it onto the
other subassembly of gusset plates,
transverse beams, and diagonals.
The frame is now complete.

15 To prepare a base for the tower, glue


15
these laminated cardboard squares—
called footings—at the columns
locations. These will transmit the
column compressive forces into the
base while also providing the required
horizontal reactions to resist the
applied lateral loading.

The cardboard tower is now complete and ready for testing!

36 
TESTING PROCEDURE
DIRECTION
To test the cardboard tower, place the base
on the floor or clamp it to the top of a sturdy OF LOADING
table. Place the tower onto its four footings.
Tie a loop of nylon string through holes in the top‑level gusset
plates on the “downwind” side of the tower such that the loop
passes behind the tops of the downwind columns, as shown
here. Hang a luggage scale from this loop.

Now carefully place six 12‑inch square concrete pavers


onto the tops of the columns, ensuring that each paver is
centered on the tower. When all six pavers are in place, pull
horizontally in the downwind direction until the reading on
the luggage scale is 10 pounds.

Project 2 » Cardboard Tower 37


PROJECT 3
BEAM BRIDGE
DESIGN, BUILD, AND TEST A BEAM BRIDGE
WITH A SPAN OF 8 FEET TO CARRY
PEDESTRIANS ACROSS A STREAM.
LESSON IN WHICH THIS PROJECT IS COVERED:
4 » Bridging with Beams

PROBLEM DEFINITION
Requirements
}} To accommodate the use of 8‑foot lumber while also allowing for some bearing surface at the
supports of the bridge, the actual span length from center to center of supports will be 7 feet 10
inches (or 94 inches).
}} The bridge deck will be 8 feet long and 2 feet wide.
}} The structure must carry the appropriate code‑specified pedestrian loading safely.

Constraints
}} Use only commonly available materials: pine lumber, plywood, and standard glue and fasteners.

REFERENCES
Hibbeler, Engineering Mechanics.
——— , Mechanics of Materials.

38 
Beam Structures
Beams are among our most
common structures. From
the fallen tree that serves as
an expedient bridge across
a woodland stream, to the
ubiquitous overpass bridges
of our modern interstate
highway system, to elegant
long‑span bridges over major
waterways, to the very floor
beneath your feet, beams are
everywhere!

DESIGN CONCEPTS

For this project, we will consider three alternative


configurations that will allow us to compare, CONCEPT A
in a rigorous way, the influence of a beam’s
cross‑section shape on its load‑carrying capacity.
To facilitate this comparison, all three alternatives
will use exactly the same 1/2‑inch plywood deck, and
all three will use two beams made of pine and
spaced 16 inches apart from center to center. CONCEPT B

1 Concept A will use a rectangular


cross‑section that is wider than it is tall.

2 Concept B will use a rectangular


cross‑section that is taller than it is wide. CONCEPT C

3 Concept C will use I‑shaped beams, each


consisting of a vertical element called a web
and two horizontal elements called flanges.

Project 3 » Beam Bridge 39


DETAILED DESIGN

For this project, the goal of the detailed design phase is to determine the required cross‑section dimensions
for all three alternative design concepts, using the same code‑specified loading condition for all three so
that we can make a rigorous quantitative comparison and selection. Once we have selected the best design
concept, we will complete the detailed design for the selected alternative only.

Model of the Structural System

The structural model on which our design will be based is shown here.

The model is a simple two‑dimensional w


representation of a single beam with
unknown reaction forces (R), representing
the supports at each end and subjected to a
uniform loading with magnitude designated
as w, representing the total weight of the R R
bridge itself and of the pedestrians crossing 94"
it—or, more precisely, the portion of this 96"
weight that will be carried by one beam.

When a load is represented graphically in this way—as a series of arrows connected by a line—it is called a
line loading and is considered to be distributed continuously and uniformly along the specified length. A
line loading is expressed in terms of force per length—for example, pounds per foot or pounds per inch.

This representation makes sense for the bridge’s self‑weight, which is indeed distributed continuously and
uniformly along the full length of the span.

Consider the self‑weight of the plywood deck. A 4‑by‑8‑foot sheet of common 1/2‑inch plywood weighs
about 40 pounds. Our deck will measure 2 feet by 8 feet—exactly half a sheet—so it will weigh 20 pounds.
Each of the two supporting beams will carry half of that weight, or 10 pounds. This weight will indeed be
distributed uniformly along the 8‑foot length of the structure; thus, the magnitude of this loading is:

Next, consider the pedestrian loading, which is more complex, more uncertain, and more variable. Logically,
the most severe pedestrian loading that is possible for our bridge would be a group of relatively heavy people
occupying the entire bridge deck. But how should the forces representing this loading be defined?

Research has shown that the complex loading associated with pedestrians crowded onto a bridge can
be modeled with reasonable accuracy as a simple uniform loading. The U.S. design code that governs
pedestrian bridges specifies this loading as 90 pounds per square foot, applied to the entire bridge deck.

40 
This is a surface loading—a pressure, specified 90 lb/ft2 = 90 lb applied to each
one-foot square
in force per area (specifically, pounds per square
foot). For our two‑dimensional model, this surface
loading must be converted to a line loading,
expressed in pounds per foot or pounds per inch.

To do this conversion, visualize our 2‑by‑8‑foot


bridge deck subdivided into 16 one‑foot squares, as
shown here. The code‑specified area loading of 90
pounds per square foot simply means that 90 pounds
are applied to each of these 16 squares.

Therefore:

Thus far, the only load we have not accounted for is the weight of the beam itself. But we cannot yet
calculate this weight, because we have not yet designed the beams. For now, the best we can do is make a
guess and then check it later, after the beams have been designed.

Assuming a beam self‑weight of 0.1 pounds/inch, we can now determine the total magnitude of the line
loading, w, as:

Having made this determination, we can also calculate the magnitude of the two reaction forces, R,
at the ends of the beam. According to the principle of equilibrium, the total downward load must be
counterbalanced by the total upward force. The structure is symmetrical, so the two reactions must be
equal in magnitude.

Therefore:

w =15.2 lb/in

Our complete structural model is shown here.

R = 730 lb R = 730 lb
94"
96"

Project 3 » Beam Bridge 41


Formulation of the Design Basis for Flexure

As we learned in lesson 3, the general


mathematical requirement for the design
of any structural member is as follows: Stress
Stress is a measure of the intensity of internal force in a
member. It is always expressed in units of force per area,
such as pounds per square inch (lb/in2 or psi) or newtons
For this project, we will define the actual per square meter.
and failure conditions in terms of stress.
Thus: If a structural member is subjected to pure tension or pure
compression, the internal stress is uniformly distributed on
the cross‑section. Thus, if we could cut a tension member
apart and see the stress inside, it would look like this.

For a wooden beam, the relevant failure F


F
stress is a material property called the σ=
A
modulus of rupture (σr), defined as the
stress at which a particular type of wood
fails in flexure.

Substituting the flexure formula (see


sidebar, next page) for the maximum stress
The magnitude of this stress—typically represented by the
and the modulus of rupture for the
Greek letter sigma (σ)—is equal to the applied force (F)
failure stress: divided by the cross‑sectional area (A).

The concept of stress is vitally important in structural


design, because the strengths of materials are always
defined in terms of stress. For example, a common type
of structural steel has a strength of 50,000 psi. It does not
If we rearrange this expression
matter whether we are analyzing a massive steel girder
algebraically and solve for the elastic or a tiny steel rod; if the stress in any portion of the steel
section modulus, we arrive at an inequality member reaches 50,000 psi, the material will begin to fail.
that will serve as the design basis for
determining the dimensions of our three
alternative beam cross‑sections:

Before we can proceed, we will need a reliable number for σr, the modulus of rupture for pine. Given the
huge variability in mechanical properties for various species of pine, we will need to determine this number
experimentally, using the same type of pine we will be using for the actual bridge.

42 
Flexure
The term flexure refers to the bending behavior of a beam.

When a structural member bends, it experiences compression on


the top and tension on the bottom, as shown here. Note that COMPRESSION
the transverse lines on this drawing get closer together on
the top and farther apart on the bottom.

Furthermore, because these lines remain straight, there


must be a linear variation from maximum compression
on the top to maximum tension on the bottom. Therefore,
there must also be a horizontal plane at mid‑height of the beam, TENSION
where there is no deformation. This plane is called the neutral surface,
and it corresponds to the dotted line in this drawing.

In this graphical representation of a bending beam, you are actually seeing compressive and tensile deformations.
But wherever load‑induced deformations occur, stresses are also occurring.

The diagram at right illustrates the mechanics of flexure.


FLEXURAL STRESS
If we could cut a bending beam apart and see
σmax
the flexural stresses inside, they would look like
this—with maximum compressive stress occurring
neutral surface
NEUTRAL SURFACE COMPRESSION
TENSION
at the top of the beam, maximum tensile stress
occurring at the bottom, and both decreasing σmax
linearly to zero at the neutral surface.
FLEXURE FORMULA
The magnitude of the maximum stress (σmax) is the
same at the top and bottom surfaces of the beam
and can be calculated using a simple equation σmax = M
called the flexure formula. S
where:
In this equation, M is the internal moment, which M = internal moment
can be conceptualized as the tendency of external S = elastic section modulus
loads to cause bending in a structural member.

These diagrams show the internal moments


INTERNAL MOMENT (M)
in simple beams subjected to two common
loading conditions: a single concentrated
CONCENTRATED LOAD DISTRIBUTED LOAD
load applied at mid‑span and a uniform
P w line loading. Note that, in both cases, the
internal moment varies along the length of
L L the beam and is highest at mid‑span.

PL wL2 The other variable in the flexure formula is


Mmax= Mmax=
4 8 S, the elastic section modulus—a measure
of the member’s resistance to bending.
A beam with a large section modulus
experiences less stress under a given load
than a beam with a smaller section modulus.
M M
Continued on next page…

Project 3 » Beam Bridge 43


Flexure (continued)
ELASTIC SECTION MODULUS (S)
RECTANGULAR SECTION I-SHAPED SECTION Pictured here are the
equations for the section
moduli of the two
cross‑sections we will be
analyzing in this project: a
rectangle and an I‑shape.
Note that S is a function
only of the cross‑section
tf shape and dimensions, not
h h
tw the material of which the
beam is made.

b b

Experimental Determination of σr

To determine σr experimentally, we load a specimen of pine in flexure until it fails; we then calculate the
stress at which the failure occurred.

Our test specimen is 18 inches long, and its cross‑section is 0.28 inches square. At failure, this beam is
loaded by a single concentrated force of 8.8 pounds applied at mid‑span. Therefore, the maximum internal
moment at the instant of failure is:

The elastic section modulus of the cross‑section is:

Substituting these numbers into the flexure formula, the maximum flexural stress is:

And because this flexural stress occurs at the instant of failure, it is, by definition, the modulus of rupture.
Thus:

44 
Designing the Beams

Returning to our design basis:

Given that each beam in our bridge will be subjected to a total uniform loading of w = 15.2 lb/in, the
maximum internal moment is:

Using a factor of safety of 2, we can calculate the required elastic section modulus as:

For concept A, we need to determine b and h, the dimensions of the rectangular cross‑section. For ease
of construction, we will plan on building this beam from two horizontal laminations of standard one‑by
lumber. Thus, the height, h, will be 2 × 3/4 inches = 1 1/2 inches.

The relevant equation for the elastic section modulus, S, of a rectangular cross‑section is:

Solving for b and substituting for S and h:

For concept B, we also need to determine b and h. Here we will use a single thickness of one‑by material
oriented vertically. Thus, b = 0.75 inches. Once again, we will start with the equation for the elastic section
modulus, S, of a rectangular cross‑section:

Solving for h and substituting for S and b:

Project 3 » Beam Bridge 45


For concept C—the I‑shaped cross‑section—we must determine the overall height (h), the width (b), the
flange thickness (tf ), and the web thickness (tw). We will plan on using 3/4‑inch lumber for the flanges and
1/4‑inch plywood for the web. Furthermore, to facilitate a direct comparison of designs B and C, we will
make the overall height, h, exactly equal to that of concept B: 5 inches. Now, with only one unknown
dimension remaining, we can use the expression for the section modulus of an I‑shape to calculate the
flange width, b, albeit after some extensive algebraic manipulations:

Solve for b, and substitute for S, h, tw, and tf :

Comparison of Alternative Design Concepts

We have finally reached the point at which we can


rigorously compare our three alternative
cross‑section configurations. They are
drawn exactly to scale at right.

A summary of
their associated
cross‑section
dimensions is CONCEPT A CONCEPT B CONCEPT C
provided below:

CONCEPT CROSS‑SECTION DIMENSIONS CROSS‑SECTIONAL AREA


(INCHES) (SQUARE INCHES)
b h tf tw
A 8.3 1.5 N/A N/A 12.5
B 0.75 5.0 N/A N/A 3.8
C 1.0 5.0 0.75 0.25 2.4

46 
Because all three beams are the same length, the cross‑sectional area is a direct measure of the amount of
material each beam will use.

In making this comparison, it is important to remember that all three cross‑sections have precisely the same
level of structural safety. All were designed so that the maximum stress in each beam will not exceed the
strength of pine, with a safety factor of exactly 2. Yet, despite this commonality, there are huge differences
in the economy our three alternatives. Concept B uses only 30% as much material as concept A, and
concept C uses less than 20% as much.

Clearly, the I‑shaped alternative—concept C—is the optimum choice, because of its efficiency in carrying
flexure. To complete the design, we must do five additional checks.

Check Beam Self‑Weight

In developing our design, we assumed a beam self‑weight of 0.1 pounds per inch. Now that we know the
actual cross‑section dimensions, we must check this assumption.

Given the unit weight of pine (γpine = 26.5 lb/ft3) and the cross‑sectional area of our I‑shaped beam (A = 2.4 in2):

Because this actual weight is significantly less than the assumed weight, we know that our design is safe.
We could now go back and redo our design calculations to account for this smaller load, but because the
beam self‑weight represents less than 1% of the total load, the effect of this tiny adjustment to our design
would be insignificant.

Check Shear

We need to check the effect of shear in the beams. The concept of shear and the associated calculations are
quite complex. Readers seeking in‑depth coverage of the topic should consult R. C. Hibbeler’s Mechanics of
Materials.

For the sake of completeness, the relevant calculations are provided below.

Maximum internal shear force, V:

Project 3 » Beam Bridge 47


Area moment of inertia, I:

First moment of outward area, Q, for an I‑shaped cross‑section:

Shear stress, τ:

The shear strength of pine parallel to the grain is approximately 1600 psi. Thus, the factor of safety of our
design with respect to shear is:

Thus, the beam is adequate with respect to shear.

Check Stability

When a beam bends, its top half is in compression. As a result, tall I‑shaped beams like the one we just
designed are susceptible to a form of instability called lateral‑torsion buckling. This failure occurs when
the compressive force in the top half of the beam causes it to fail by buckling sideways and twisting.

If lateral‑torsional buckling is allowed to occur, it can severely reduce a beam’s load‑carrying capacity.
Fortunately, this failure mode can be easily prevented in our bridge by providing a strong structural
connection between the beam’s top flange and the deck. With that connection in place, the top flange is
prevented from buckling sideways, and the full flexural load‑carrying capacity of the beam can be achieved.

48 
Check Resistance to Concentrated Reaction Forces

When our bridge is resting on its concrete foundations and is fully loaded, a concentrated reaction force of
730 pounds will be applied to the bottom of the flanges at each end of each beam. This large concentrated
force has two possible adverse effects: It could cause the ends of the beam to topple over sideways,
and it could crush the bottom flange at the supports because wood is particularly weak when loaded
perpendicular to the grain.

To prevent the beam flanges from END DIAPHRAGM


crushing, we will add hardwood
bearing pads at each support
location, and to prevent the beams
from bending or buckling sideways,
we will add this transverse element,
called a diaphragm, at the ends of
the bridge, between beams. We will
also install two intermediate diaphragms BEARING PAD
at the one‑third points of the span to keep
the two beams properly spaced and aligned.
INTERMEDIATE
DIAPHRAGM

Check Deflection

Deflection is the vertical distance a beam bends in response to its applied load. For a simple beam with
span length, L, subjected to a uniform loading, w, the maximum deflection, δ, can be calculated with the
following equation:

where E is a material property called the modulus of elasticity (1,480,000 psi for pine) and I is the area
moment of inertia for the cross‑section (7.8 in4, as calculated above). In applying this equation, the load, w,
includes only the pedestrian loading, not the self‑weight of the bridge. And because deflection only affects
the serviceability of a structure, not its structural safety, we can consider the load to be shared equally by
the two beams in our bridge. Thus:

Substituting:

Project 3 » Beam Bridge 49


Most current design codes specify that beams used in floor systems should not deflect more than the span
length divided by 360. For our 94‑inch span, this is:

By this standard, our design fails, because the predicted actual deflection (0.66 inches) is more than twice
the code‑specified standard (0.26 inches).

It’s important to recognize, however, that code‑specified limits on deflection have nothing to do with
structural safety. Rather, these limits are specified only because excessive deflections tend to make people
feel uncomfortable and because excessive deflections in buildings can cause serviceability problems, such as
cracking plaster and jamming doors.

If either of these factors were a concern for our bridge, we would have to modify our design, significantly
enlarging the beam cross‑section to meet the deflection standard. Fortunately, these factors are not a
concern for our bridge.

Moreover, when we build the structure, we will be fastening the deck rigidly to the top flanges of both
beams—to prevent lateral‑torsional buckling—and this will also tend to reduce the bridge’s deflection
significantly, because integrating the deck and beams into a single structural unit will greatly reduce its
tendency to bend.

FINAL DESIGN DRAWINGS

The SketchUp 3‑D computer model for this project


can be downloaded from http://stephenjressler.com/
24" diy-engineering/.

96"

16"
32½"
31"
32½"
Drawing 3.1 » Beam bridge, perspective view

50 
A

D
F
E
C F
B E C
C B
C
D
F

Drawing 3.2 » Beam bridge,


exploded view

MATERIALS LIST

NAME DESCRIPTION
A Deck 24″ × 96″ × 1/2″ plywood
B Beam web (2) 4″ × 96″ × 1/4″ plywood
C Beam flange (4) 1″ × 3/4″ × 96″ pine with 1/4″ × 1/4″ groove cut into one face
D End diaphragm (2) 5″ × 15 3/4″ × 3/4″ plywood, notched to accommodate top and bottom flanges
E Intermediate diaphragm (2) 4 1/4″ × 15 3/4″ × 3/4″ plywood, notched to accommodate top flanges
F Bearing pad (4) 3 1/2″ × 2 3/4″ × 3/4″ hardwood

The only other supplies needed for this project are 1 1/4‑inch wood screws and wood glue. If you plan
to install this bridge permanently outdoors, be sure to use galvanized screws and waterproof glue
(e.g., polyurethane glue).

Project 3 » Beam Bridge 51


FABRICATION AND ASSEMBLY PROCEDURE
1 Begin by prefabricating all of the wooden components listed above. The beam webs and flanges should
be cut precisely on a table saw, as demonstrated in lesson 4. The groove in each flange can be cut with
either a router or a dado blade mounted in a table saw.

2 Assemble each beam by gluing the edges of each web into the grooves in two flanges, as demonstrated
in lesson 4. Use plenty of glue, and clamp the joints securely until the glue has dried. If your bridge will
be exposed to moisture, these joints must be made with waterproof glue.

3 Connect the two beams together by gluing and screwing them to


the end diaphragms (component D). Ensure that the
entire assembly is square when viewed
from above.

4 Add the intermediate diaphragms (component E),


again making the necessary connections with glue
and wood screws.

52 
5
5 Attach the four bearing
pads (component F) to the
bottoms of the beam flanges
with wood screws.

6 Finally, set the deck


(component A) in position,
and fasten it to the top
flanges of the beams with
wood screws. Note that
these screws are used only in
the outer two thirds of the
span, so they will not reduce
the strength of the beams in 6
the center third, where they
are most heavily stressed.

The beam bridge is now complete and ready for testing!

Project 3 » Beam Bridge 53


TESTING PROCEDURE

At the testing site, construct foundations by placing four 12‑inch square concrete pavers according to the
layout plan shown here. The four concrete pavers should be level and at the same elevation.

CONCRETE PAVER

16"

96"

Then, place the bridge on its foundations, with the four bearing pads centered on the four pavers.

Load the bridge with people whose weight totals approximately 1440 pounds. Do not overload the structure!

54 
PROJECT 4
SUSPENSION
BRIDGE
DESIGN, BUILD, AND TEST A SUSPENSION
BRIDGE WITH A SPAN OF 8 FEET TO CARRY
PEDESTRIANS ACROSS A STREAM.
LESSON IN WHICH THIS PROJECT IS COVERED:
5 » Make a Suspension Bridge

S uspension Bridges
Since 1931, all of the world’s
longest‑spanning bridges have been
suspension bridges. Not only are these
structures extraordinarily efficient
engineered systems, but they are
also among our most beautiful and
majestic works of civil infrastructure.
The bridge we build in this project
will not be very majestic; nonetheless,
it will evoke these great structures
while also teaching us how they
© Ultima_Gaina/iStock/Thinkstock. function as structural systems.

Project 4 » Suspension Bridge 55


PROBLEM DEFINITION
Requirements
}} The span length from tower to tower will be 8 feet.
}} The bridge deck will be 18 inches wide.
}} The structure must safely carry a code‑specified pedestrian loading of 90 pounds per square foot.

Constraints
}} Use steel wire rope for main cables.
}} Use commonly available lumber, fasteners, and adhesive for all other components.

REFERENCES
Breyer, Fridley, Cobeen, and Pollock, Design of Wood Structures.
Hibbeler, Engineering Mechanics.
——— , Mechanics of Materials.

DESIGN CONCEPT

This project is quite complex; thus, the design concept will be presented not as a two‑dimensional sketch,
but rather as a three‑dimensional computer model so that we can better visualize the full three‑dimensional
structural system.

The key components of this structural system are as follows:

}} Two wooden towers are placed on concrete foundation blocks 8 feet apart on opposite banks of
the stream.
}} At a yet‑to‑be‑determined distance behind each tower are the anchorages—so named because they
anchor the ends of the main cables. To do their job, the anchorages must be below the surface of the
ground and must have substantial weight. For our bridge, each pair of anchorages will be constructed
as a wooden box, into which we will stack a yet‑to‑be‑determined number of concrete pavers.
}} The main cables run from anchorage to anchorage, across the tops of the towers.
}} Four transverse beams are suspended from the main cables by vertical wires called suspenders.
}} The bridge deck rests on the transverse beams, and its ends are fastened to the towers.

56 
TOWER

MAIN CABLE

SUSPENDER
DECK

BEAM

FOUNDATION
BLOCK

ANCHORAGE

Before we proceed with the detailed design, three key dimensions must be specified:

}} The sag of the main cables at mid‑span will be 30 inches.


}} The longitudinal spacing between suspenders will be 19 inches.
}} The lateral spacing between main cables will be 20 inches.

DETAILED DESIGN

Our design process will entail purposeful consideration of six different structural elements, in the following
sequence: deck, transverse beams, suspenders, main cables, towers, anchorages.

This design sequence is defined by the transmission of internal forces through the structural system. The
deck directly supports the pedestrian loading, the beams support the deck, the suspenders support the
beams, the main cables support the suspenders, and the towers and anchorages support the main cables.
This methodical approach—often called following the load path—is essential, because as internal forces are
transmitted through the structure, they accumulate. For example, the beams carry the pedestrian loading
plus the weight of the deck; the suspenders carry the pedestrians, deck, and beams; and so on. If we do not
design by following the load path, we are quite likely to overlook some of this accumulating load.

Project 4 » Suspension Bridge 57


Deck Design

Under its uniform loading, the deck


will bend in the complex configuration
shown here, and it must be designed
to prevent flexural failure at the points
of most severe bending—immediately
above the supporting deck beams and DECK BEAM
midway between them.

This is a very challenging structural analysis problem.


Fortunately, it is also one we can sidestep, thanks to the
standardized rating system used for structural plywood in
the United States. The stamped information on the sheet of
19/32‑inch plywood pictured at left includes a span rating
consisting of two numbers separated by a slash. The left
number—40—is the maximum allowable spacing between
supports, in inches, when this plywood is used for roof sheathing
in residential construction; the right number—20—is the
maximum spacing between supports for flooring.

Our bridge deck is equivalent to flooring, and our deck beams


will be spaced only 19 inches apart; thus, the span rating assures
us that this plywood will be structurally adequate. Thanks to
this ratings system, no design calculations are needed.

Beam Design

Having decided on an appropriate deck thickness, we can now move on to the next element in the load
path: the transverse beams that support the deck.
T w T
We can model one such beam, as shown
here, with a uniformly distributed load, w,
representing the sum of the pedestrian loading
and deck weight, plus two upward reactions, T, 18"
corresponding to the tension force in the two
20"
suspenders.

In theory, we should also consider the weight of the beam itself in this analysis; however, in practice, this
weight is so small in comparison with the other loads that we can safely ignore it.

To determine the magnitude of w, we must first determine how much of the pedestrian load is carried by
each of the four beams.

58 
19" TRIBUTARY AREA OF ONE BEAM
If we visualize the deck as five
equal‑sized panels, as shown here,
then each beam can be assumed
DECK 18"
to support the portion of the
deck extending from midpoint to
midpoint of the two adjacent panels.
This 18×19‑inch rectangle is called 19" 19" 19" 19" 19"
the tributary area of the beam.

The total pedestrian load applied to this tributary area is:

This number seems quite reasonable, as 214 pounds is about what we would expect for the maximum
weight of a person standing within this rather small area of the bridge deck.

But this number does not include the weight of the deck itself: 19/32‑inch plywood weighs 1.5 pounds
per square foot; thus, the deck weight associated with our 18×19‑inch tributary area can be calculated as
follows:

Thus, the total load that must be carried by one beam is:

Therefore, the magnitude of the distributed load, w, applied to one transverse beam is:

This beam can be analyzed using exactly the same structural mechanics tools we used for the beam bridge
project—most importantly, the design basis for flexure:

In this expression, M is the internal moment due to the distributed load, w, calculated as follows:

Project 4 » Suspension Bridge 59


Assuming that this beam will be made of the same pine we used in our previous bridge project, we can
use the same experimentally derived modulus of rupture, 10,800 psi, and the same factor of safety, 2.
Substituting these values into the design basis:

The elastic section modulus, S, for a rectangular cross‑section is given by:


h

We can now solve this expression for h and substitute appropriate values for S and b.
Assuming that we will use standard one‑by lumber, b = 0.75 inches. Therefore:

b = 0.75"

For simplicity of construction, we will round this result up to the next standard lumber size: 1 1/2 inches.
Thus, we will make our transverse beams from 1×2 pine (which actually measures 3/4 × 1 1/2 inches).

Suspender Design

With the beams designed, the next elements in the load path are the suspenders, which support the beams
in pure tension. The model we just used for the beam design included the suspender tension forces, T, and
we can calculate their magnitude quite easily:

But common sense should tell us that there is a problem with this number. Our structural model of the
transverse beam assumes that the pedestrian loading is uniformly distributed along its entire length, which
implies that all pedestrians are exactly centered on the deck. For this assumed load configuration, both
suspenders share the 217‑pound pedestrian load equally.

But in practice, on such a small bridge, we have no assurance that the pedestrian loading will actually be
centered on the deck. In the worst‑case circumstance, if a pedestrian stands at the very edge of the deck, his
or her entire weight will be carried by a single suspender.

60 
To ensure that this sort of asymmetrical loading will not cause a failure, we must design each suspender
to carry the full tributary pedestrian load of 214 pounds, plus half of the tributary beam and deck
weight—1.8 pounds—for a total of 216 pounds.

Our suspenders will be fabricated from common 7‑strand steel wire rope, which has a working load of 120
pounds. (Note that the term “working load” always indicates that an appropriate factor of safety has already
been applied to the rated capacity.)

Therefore, the required number of wire rope strands for each suspender is:

We will fabricate each suspender from two strands of 120‑pound wire rope.

Main Cable Design

In designing the main cables, our principal task is to predict the absolute maximum tension force that will
occur in these cables and then select a steel wire rope product that will do the job safely.

To calculate the maximum tension force, we will start by modeling one full cable as a series of line
segments, with points A and H designating the anchorages, B and G designating the tower supports, and C
through F designating points of attachment with the suspenders. We have just determined the maximum
tension in the suspenders to be 216 pounds; thus, we can add these forces as downward loads on the cable.
The unknown reactions at the towers and anchorages are designated RT and R A . The relevant dimensions
are shown. The unknown anchorage setback is designated as x.

B G

RT 30" RT
C F

D E 60"
216 lb 216 lb

216 lb 216 lb

A H

y RA x 19" 19" 19" 19" 19" x RA

Project 4 » Suspension Bridge 61


Note that we have not included the self‑weight of the cable or suspenders in this free body diagram, again
because the weights of these elements are so small in comparison with the other loads that we can safely
ignore them.

To calculate the internal tension force in the cable, we


TBC TBCy
will isolate a segment of the cable, from just to the right
B of point B to the midpoint of the span between points
TBCx D and E.

30" A free body diagram of this cable segment is shown


C
at left. Note that the two unknown tension forces in
D cable segments BC and DE are designated TBC and TDE ,
respectively. And to facilitate our analysis, the diagonal
216 lb TDE
force TBC is represented as a pair of equivalent x‑ and
y‑components, designated TBCx and TBCy.
216 lb
y
19" 19" With our free body diagram complete, we now can apply
the principle of equilibrium to calculate these three
x
unknown forces.

1 The sum of moments about point B, assuming that the counterclockwise direction is positive:

2 The sum of forces in the x‑direction, assuming that the positive direction is to the right:

3 The sum of forces in the y‑direction, assuming that the positive direction is upward:

62 
Now that we know the two perpendicular components of the tension TBC TBCy = 432 lb
in cable segment BC, we can calculate the total tension force using the
Pythagorean theorem:

An important characteristic of a cable structure that is subjected only to


B
gravity loads is that the absolute maximum cable tension always occurs at
TBCx = 410 lb
the point where the geometric slope of the cable is steepest. The steepest
slope of our cable occurs in segment BC; thus, 595 pounds is, in fact, y
the maximum tension.

x
This requirement can be satisfied with 3/16‑inch‑diameter wire rope from
McMaster‑Carr (part number 3440T55), which has a more‑than‑adequate
capacity of 750 pounds. As noted on the McMaster‑Carr website, this tensile strength already incorporates
a very generous safety factor of 5; thus, we can be very confident in the safety of this critical component.

Before moving on, we need to determine the unknown sag


of the cable at point C—an important dimension, because it TBC TBCy = 432 lb
will determine the required length of the suspender at C.

In the graphic shown here, note that two triangles—one


formed by the forces TBC and TBCy and one formed by cable
19"
segment BC and its vertical sag—are geometrically similar.
Their angles are identical; thus, the respective lengths of
their sides must be in the same proportion to each other. TBCx = 410 lb
B θ
Therefore, we can set up this mathematical proportion
based on the principle of similar triangles: sagC
y
C
x

But what about the backstays—the cable segments that extend from the towers back to the anchorages? To
understand how these elements work, we will focus on the top of one tower and draw a free body diagram
of the short segment of cable running across it. The three forces acting on this segment are the tension in
BC, which we have calculated as 595 pounds; the unknown tension in the backstay AB; and the upward
reaction exerted by the tower, designated RT.

Project 4 » Suspension Bridge 63


To keep the cable from slipping, TAB must equal TBC . And if these forces are equal, then the only way to
prevent the cable from imparting an unbalanced horizontal force at the top of the tower is to incline the
backstay AB at exactly the same angle as cable segment BC. Using trigonometry, this angle, designated as θ
in the previous diagram, is:

B
46.5° 46.5°
And if TAB and TBC have the same magnitude 410 lb 410 lb
and the same angle of inclination, then their
components must also be the same: 410 pounds RT
in the horizontal direction and 432 pounds
in the vertical direction. Thus, applying the
TAB = 595 lb 432 lb 432 lb TBC = 595 lb
principle of equilibrium:
y

Tower Design

If the tower is pushing upward on the cable with a force of 864 pounds, then each cable is also applying a
downward force of 864 pounds to the tower.

The principal structural elements of each tower are its two columns, each of which supports one cable. In
our design concept, we assumed that 2×4 boards would be used for these columns. Assuming that these
boards will be grade 2 spruce‑pine‑fir lumber, we must now check to ensure that each column is capable of
safely carrying a compressive force of 864 pounds.

The following equations are simplified versions of the column analysis provisions specified in the National
Design Specification for Wood Construction.

In general, the allowable compressive force in a wooden column, Pallow, is given by:

where:

F′C = allowable compressive stress


A = cross‑sectional area of the column

64 
Furthermore, F′C is specified as:

where:

Fc = tabulated allowable compressive stress for a particular wood species


Cp = column stability factor

The column stability factor is defined as:

In this equation:

where:

E = modulus of elasticity = 510,000 psi for the lumber we will be using


k = effective length factor = 2 for a column that is not laterally supported at its upper end
L = actual column length = 60 inches
d = least cross‑sectional dimension = 1.5 inches

Substituting these values:

The remaining values required for the calculation of Cp are:

Fc = tabulated allowable compressive stress = 1150 psi for grade 2 spruce‑pine‑fir lumber
c = bucking and crushing interaction factor = 0.8 for sawn lumber

Project 4 » Suspension Bridge 65


Substituting and solving for Cp:

SIDE
PLATE
The allowable compressive stress can now be calculated as:

And the allowable compressive force, Pallow, for a 2×4 cross‑section (which
actually measures 1.5 by 3.5 inches) is:

allow

Given that the actual maximum compressive force in these members is expected to
be 864 pounds, our planned 2×4 columns are clearly inadequate.

We could correct this problem by using heavier 4×4 columns, but we can achieve the same beneficial effect
more efficiently by using triangular side plates, as shown above. These plates significantly increase the
column’s resistance to buckling, and they look good, too!

B
Anchorage Design 46.5°

Before we can design the anchorages, we must first determine their setback—the
distance the anchorages must be placed behind the towers (assuming that the
AY
ST

ground between the anchorages and towers is level). 60"


CK
BA

We have already determined that the backstays are inclined at 46.5°, as


shown here, and the tower height is 60 inches. Thus, using trigonometry: TOWER

46.5°
A
y x

x
66 
In designing the anchorages, our most important task is to TABy = 432 lb TAB = 595 lb
determine how many concrete pavers are needed to anchor
the ends of the main cables. As this free body diagram
indicates, the force applied to each anchorage is TAB , the
SOIL
tension in the backstay. We already know that this force
RESISTANCE
can be represented as an x‑component of 410 pounds and a TABx= 410 lb
y‑component of 432 pounds.

The 410‑pound horizontal component will be resisted by the y


weight of pavers = 432 lb
wall of soil on the tower side of the anchorage. (This is why
the anchorages must be placed below ground level.) But the x
only way to resist the 432‑pound upward component is with
at least 432 pounds of weight. Each concrete paver weighs 17
pounds; thus, the required number of pavers per anchorage is:

Rounding up, we will need 26 pavers for each anchorage, a total of 104.

This surprisingly large anchorage weight for such a small bridge illustrates quite dramatically why the
anchorages of real‑world suspension bridges are so incredibly massive.

Our suspension bridge design is now complete.

HAER CAL,38‑SANFRA,140‑‑6

Project 4 » Suspension Bridge 67


FINAL DESIGN DRAWINGS

The SketchUp 3‑D computer model


for this project can be downloaded
from http://stephenjressler.com/
diy-engineering/.

Note: Pavers omitted for clarity

Drawing 4.1 » Suspension bridge,


overall perspective view

Drawing 4.2 » Tower elevations and exploded view

½" diameter ¾"


K
1 ½"

A
A
60 ¼"
E

B E

12" D
C
5"
C
8"
SIDE ELEVATION FRONT ELEVATION EXPLODED VIEW

68 
⅜" diameter
1 ⅜"
Drawing 4.3 » Anchorage elevation,
plan view, and exploded view

24 ¾"
P P
49 ½" 52 ½"

L
H L

M
G M

F H
PLAN VIEW

7 ¼" 8 ¾"

11 ⅞"
14 ½"
SIDE ELEVATION EXPLODED VIEW

P P
N N
Drawing 4.4 » Cables, suspenders,
beams, and deck, exploded view Q Q
and plan view Q Q R R
R
R
J

O
EXPLODED VIEW

PLAN VIEW
¼" diameter 1 ½" 1"
¾" 1 ¾"

18"

19" 19" 19" 19" 19"

Project 4 » Suspension Bridge 69


MATERIALS LIST
Wooden Components

NAME DESCRIPTION
A Tower column (4) 1 1/2″ × 3 1/2″ × 60 1/4″ lumber (2×4)
B Tower cross member (4) 1 1/2″ × 3 1/2″ × 21 1/2″ lumber (2×4)
C Tower footer (4) 1 1/2″ × 3 1/2″ × 8″ lumber (2×4)
D Tower diaphragm (2) 10 1/2″ × 14 1/2″ × 3/4″ plywood or oriented strand board
E Tower side plate (4) 60″ × 8″ × 1/2″ plywood
F Anchorage base (4) 1 1/2″ × 7 1/4″ × 52 1/2″ lumber (2×8)
G Anchorage back (2) 1 1/2″ × 7 1/4″ × 49 1/2″ lumber (2×8)
H Anchorage side (4) 1 1/2″ × 7 1/4″ × 14 1/2″ lumber (2×8)
I Beam 3/4″ × 1 1/2″ × 24″ pine (1×2)
J Deck 18″ × 96″ × 19/32″ plywood (span rating 40/20)

Hardware and Other Components

NAME DESCRIPTION
K Saddle (4) 1/2″ × 5″ steel hex bolt with nut and 2 washers
L Cable‑anchorage connector (4) 3/8″ steel U‑bolt, 1″ ID (McMaster‑Carr product number 3201T36)
M Cable clamp (12) 3/16″ cast wire rope clamp (McMaster‑Carr product number
30325T62)
N Suspender retainer (8) 3/16″‑diameter clamping shaft collar (McMaster‑Carr product number
6436K6)
O Suspender‑beam connector (8) 1/4″ steel U‑bolt, 3/4″ ID (McMaster‑Carr product number 3201T31)
P Main cable (2) 3/16″ 6×19 wire rope (McMaster‑Carr product number 3440T55)—
each cable 26 feet long
Q Outer suspender (4) 1/16″ 7‑strand wire rope (working load 120 lb)—each suspender
60″ long
R Inner suspender (4) 1/16″ 7‑strand wire rope (working load 120 lb)—each suspender
40″ long
S Concrete footing (4) 12″ × 12″ concrete paver
T Anchorage weight (104) 12″ × 12″ concrete paver

The only other supplies needed for this project are 2½‑inch and 1¼‑inch wood screws. If you plan to install
this bridge permanently outdoors, be sure to use galvanized screws.

70 
FABRICATION AND ASSEMBLY PROCEDURE
In the Shop

1 Begin by prefabricating all of the wooden components (A­–J listed above).

2 Assemble the two towers and anchorage boxes in accordance with drawings 4.1–4.4 above.
Use 2 1/2‑inch wood screws for all connections.

3 The saddles (component K) 3


support the main cables as they
pass over the tops of the towers.
These 1/2‑inch bolts are mounted
with two washers and a nut, as
shown here.

4 The cable‑anchorage connectors


(component L) are 3/8‑inch U‑bolts,
mounted in the anchorage bases.

5 The suspender‑beam
connectors (component
O) are 1/4‑inch U‑bolts,
mounted at both ends of
each beam. 5

Project 4 » Suspension Bridge 71


6 a) Fabricate the suspenders
from 1/16‑inch 7‑strand wire
rope, using the simple jig
pictured below. Note that
each suspender consists of a
doubled 7‑strand wire with
an eye splice formed at each
end. The inner suspenders
6a measure 18 inches long,
and the outer ones measure
28 inches long.

b) To form each eye splice,


begin by separating one
end of the 7‑strand wire
rope into two bundles, one
of three strands and one
of four.

c) Next, cross the bundles


6b 6c over each other to form
an eye and twist them
back together, effectively
reassembling them into
the original wire rope
configuration.

d) Continue twisting both


bundles together around
the circumference of
the eye.

6d 6e e) Beyond the base of the eye,


twist the free ends together.

f) Finally, bind these loose ends together with


soft wire.

All remaining assembly will be done in the field,


at the bridge site.

6f

72 
On Site

1 At the bridge site, construct the foundations by placing four 12‑inch square concrete pavers at the tower
locations and digging two holes for the anchorage boxes according to the site layout plan below. The
four concrete pavers must be level and at the same elevation. Ideally, the tops of the anchorage boxes
should also be at the same elevation as the tops of the tower foundations.

ANCHORAGE ANCHORAGE
HOLE HOLE
TOWER TOWER
FOUNDATIONS FOUNDATIONS

29 ¾" 54"

16" 56 ⅝" 95" 56 ⅝" 16"

2 If the site does not permit the anchorages to be set at this elevation (for example, if the ground
elevation at the anchorage locations is substantially higher than the ground elevation at the towers),
then the anchorages must be moved closer to the towers and the cables must be shortened to maintain
the correct angle of the backstay, as shown below.

A = ANCHORGE POSITION FOR LEVEL STREAM BANKS


B = ANCHORGE POSITION FOR SLOPING STREAM BANKS

B B

A A

Project 4 » Suspension Bridge 73


3 Once the layout and excavations are complete, place each tower on its foundation, and place each
anchorage box into its hole. Ensure that these assemblies are properly positioned and level before
proceeding.

4 Next, lay out the two main cables, precisely marking the
LOCATION LENGTH
locations of the anchorages, towers, and suspenders and
ON CABLE FROM END
leaving 1 foot of extra length at each end. If the tops of the
end 0
anchorage boxes are at the same elevation as the bases of
anchorage 12″
the towers, these markings should be at the locations along
the length of each cable indicated in the table at right. top of tower 93 7/8″
suspender 121 7/16″
5 To install the cables, first drape them over the towers and suspender 142 15/16″
hold them temporarily in the correct position with ties suspender 161 15/16″
of soft wire. Then, pull the backstays tight, pass them suspender 183 7/16″
through the cable‑anchorage connectors (U‑bolts) in the top of tower 211″
bases of the anchorage boxes, and then secure the free ends anchorage 292 7/8″
of the cable with three wire rope clamps (component M) end 304 7/8″
per connection.

6 Stack 52 concrete pavers into each anchorage box to properly anchor the cables.

7 Once the pavers are in place, the soil must be backfilled and compacted on the tower side of each
anchorage to provide the bearing strength necessary to resist the 410‑pound inward pull of each cable.

8 Now, at each suspender location, attach a steel collar (component N) to keep the suspender from
sliding down the cable. Each collar is secured with two Allen‑head screws.

9 With the collars in position, loop each suspender over the cable at its designated location, secure its two
eye splices to a transverse beam with a U‑bolt (component O), and bind the upper end with soft wire to
prevent it from slipping over the collar.

10 Once all four beams are hanging from the suspenders, add the deck, securing it to the beams and the
towers with 1¼‑inch wood screws.

The suspension bridge is now complete and ready for testing!

74 
Galloping Gertie
In July 1940, the Tacoma Narrows Bridge—also known as Galloping Gertie—famously shook itself to
pieces during a modest windstorm. The principal cause of this collapse was excessive flexibility of the
bridge deck, resulting from the designer’s misguided effort to give the bridge an elegant, slender profile.

Our bridge is sufficiently short that it is quite unlikely to fail as Galloping Gertie did. Nonetheless, our
bridge—as designed—is noticeably flexible when loaded.

This flexibility can be addressed very effectively by reinforcing the deck with two stiffening girders
made of 1×3 lumber. There is a striking difference in the flexural behavior of the bridge before and after
the addition of these girders.

TESTING PROCEDURE

Load the bridge with people whose weight totals approximately 1080 pounds. Do not overload the structure!

Project 4 » Suspension Bridge 75


PROJECT 5
CONCRETE
SAILBOAT
DESIGN, BUILD, AND TEST A CONCRETE SAILBOAT.
LESSONS IN WHICH THIS PROJECT IS COVERED:
6 » Design a Concrete Sailboat
7 » Set Sail!

PROBLEM DEFINITION
Requirements
}} The hull of the sailboat
A Tribute to Pier Nervi
will be no more than 12
inches long. Pier Luigi Nervi (1891–1979) was one of the 20th century’s most
}} The boat must be
innovative structural engineers. Nervi specialized in thin‑shell
concrete structures: spectacular domed and vaulted buildings
controllable. characterized by an extraordinary synthesis of beauty and
}} The boat must be capable structural efficiency, attained through the use of thin prefabricated
of operating safely in winds reinforced‑concrete modules, assembled into elegant structural
up to 10 miles per hour. systems that are both strong and astonishingly light.

Partly to demonstrate the viability of his method, Nervi also


Constraints designed and built a series of concrete boats—both transport
vessels and sailboats—that proved to be quite seaworthy and were
}} Use only concrete for significantly less expensive than equivalent wooden‑hulled vessels.
the hull.
}} Use any commonly This project was inspired by Nervi’s work. In designing and building
available materials for all a concrete boat of our own, we will honor the memory of this great
other components. engineer while also attempting to emulate his spirit of innovation.

76 
MAST
REFERENCES
SAIL
Brewer, Understanding Boat Design.
Hibbeler, Engineering Mechanics.
Mindess, Young, and Darwin, Concrete.

BOOM
DESIGN CONCEPT

Our design concept, shown here, is


characterized by the following key features:

}} The shape of the concrete hull will be


defined by a mold made from a 2‑liter
plastic beverage bottle sliced in half. We RUDDER
will square off the stern and add a keel and
prow to improve the vessel’s seaworthiness.
}} The boat will be steered by a movable rudder. KEEL
}} The boat will be propelled by a single triangular sail
supported on a mast and boom made from wooden dowels.

Why use a beverage bottle as a mold for our hull? One of the greatest advantages of concrete as a
construction material is its versatility of form—its ability to be cast into any shape, provided that a suitably
shaped mold can be provided. A plastic beverage bottle fulfills this need admirably, as it has a smooth,
nonstick surface; is cheap and disposable; and is at least approximately hull‑shaped.

DETAILED DESIGN The Principle of Buoyancy


First articulated by the ancient Greek mathematician
Buoyancy Analysis Archimedes, the principle of buoyancy says
that immersing an object in a fluid causes an
How can a concrete object be made to float upward force—called the buoyant force—with a
when concrete is so much denser than water? We magnitude equal to the weight of the displaced fluid.
Furthermore, the principle of equilibrium requires
can answer this question rigorously by applying
that, for a floating object, the buoyant force must
the principle of buoyancy—an important equal the weight of the object itself. Therefore, the
concept from the science of hydrostatics, which weight of a floating object must equal the weight of
examines the equilibrium and stability of objects the water displaced by that object.
immersed in fluids that are not moving.

Project 5 » Concrete Sailboat 77


To see how the principle of buoyancy can be applied SIMPLIFIED HULL MODEL
to our concrete boat hull, we must first develop a
simplified model of the hull geometry, just to
keep the math from getting too complicated.
For the purpose of this analysis, we will
assume that the complex half‑bottle shape of
our proposed hull can be approximated by a half
cylinder with a radius of 2.15 inches (the same as the
beverage bottle) and a length of 7.5 inches. This model 7.5"
does not accurately reflect the complex curves of the boat’s
bow, but it should yield analysis results that are sufficiently
accurate for analyzing buoyancy.

The weight, W, of any object can be calculated as follows: 2.15" radius

where:

Thus, given that the unit weight of concrete is approximately 0.081 pounds per cubic inch, the weight of
our idealized half‑cylindrical concrete hull is:

How much water would this hull displace? Assuming that the half cylinder is fully immersed, the displaced
water will have exactly the same semicylindrical shape as the hull itself; thus, it must have exactly the same
volume. Given that the unit weight of water is 0.036 pounds per cubic inch, the weight of water displaced
by this hull is:

78 
And because our hull is assumed to be fully immersed, this is the largest‑possible buoyant force it can develop.

The weight of the hull is 4.4 pounds, oriented downward; the buoyant force is only 2 pounds, oriented
upward. Therefore, this solid concrete hull is guaranteed to sink—hardly a surprising result!

But suppose that we hollow out the hull such SIMPLIFIED HOLLOW HULL MODEL
that its sides and ends are all the same
thickness (t).
t
As t gets smaller, the weight of
the hull will also get smaller. Thus,
there must be a critical value of t at
which the hull will weigh exactly 2 pounds.
Theoretically, this 2‑pound hull will float, with t
7.5"
its gunwales—the uppermost edges of the hull—at
exactly the same level as the surface of the water.

Why will this hull


GUNWALES
float? Note that, in this
2.15" radius
configuration, the volume of
displaced water is the same as it
was in our previous analysis, so the
buoyant force remains at 2 pounds. But
because the hull’s weight and the buoyant force
are now equal, the hull is in equilibrium, and it
floats—though quite precariously, because just a slight
disturbance will cause water to pour over the gunwales.

Clearly, however, we can improve this situation further by making the hull even thinner. Doing so will
reduce its weight below 2 pounds, resulting in a new equilibrium position higher in the water.

It is clear, then, that a


boat made of concrete
can be made to float, as
long as its hull can be
made sufficiently thin.
Furthermore, it is also FREEBOARD
clear that we can control
the freeboard—the
vertical distance from
the water level to the
gunwales—by controlling
the thickness of the hull.
The thinner the hull, the
greater the freeboard.

Project 5 » Concrete Sailboat 79


To facilitate our design, then, we must work out a mathematical relationship between hull thickness and
freeboard for our semicylindrical hull model. We will attempt to achieve this end in three steps:

1 Estimate the hull’s weight for a given thickness.

2 Estimate the weight of displaced water for a given amount of freeboard.

3 Consistent with the principle of buoyancy, set these two expressions equal to each other, and use
algebra to solve for freeboard in terms of thickness.

As before, our equation for the weight of the hull is: SIMPLIFIED HOLLOW HULL MODEL

SUBTRACTE
D
But now, we must account for the hollowing‑out of the hull VOLUME
by subtracting this smaller half‑cylindrical volume from
t
the larger one.

The result is:

L = 7.5"
r = 2.15"
where:

The equation for the weight of displaced water is:

where:
AREA OF DISPLACED WATER

r f
WATER a
LEVEL Adisplaced

80 
In geometric terms, the area, Adisplaced, is called a segment of a circle, defined as a shape bounded by an arc
of less than 180° and a line segment connecting the endpoints of the arc. The area of a segment of a circle
can be calculated with the following formula, defined in terms of the radius, r, and the angle, a:

Therefore, the weight of water displaced by our idealized hull is:

Now we have a mathematical expression for the weight of displaced water in terms of the angle a;
however, we need this equation to be expressed in terms of the freeboard, f. Fortunately, there is a simple
trigonometric relationship between a and f :

Consistent with the principle of buoyancy, we will now set our equations for hull weight and weight of
displaced water equal to each other and then solve for the freeboard, f. The resulting equation is as follows:

But here our mathematical luck


FREEBOARD VS. HULL THICKNESS
runs out! If we could solve this
1.2 equation for f, we could then
⅛" thickness
substitute various values of
1" freeboard
1.0 thickness, t, into the resulting
FREEBOARD (INCHES)

equation and calculate the


0.8 corresponding freeboard for each
thickness. But, unfortunately,
0.6 it is not algebraically possible to
solve this equation for f.
0.4
It is possible, however, to solve
0.2 this problem by trial and error,
⅜" thickness
using a computer spreadsheet.
zero freeboard
0
0.1 0.15 0.2 0.25 0.3 0.35 0.4 0.45 After performing this analysis for
HULL THICKNESS (INCHES) a range of hull thicknesses, the
results are shown as a graph here.

Project 5 » Concrete Sailboat 81


As this graph illustrates, the critical hull thickness
Concrete
at which the boat will just barely float—with zero
freeboard—is a little more than 3/8 inch. If we reduce Concrete is the world’s most commonly
used construction material. Its most
the thickness to about 1/8 inch, we can achieve a very important constituent material is
respectable freeboard of 1 inch, meaning that about half Portland cement, manufactured by
of the hull will project above the waterline. mixing limestone, clay, and other
ingredients and then baking this mixture
As expected, a thinner hull is better from the perspective at high temperatures to form a rocklike
of buoyancy. But, realistically, how thin can we actually substance, which is then ground into a
fine gray powder.
cast a concrete hull before it becomes too fragile?
To make concrete, Portland cement is
combined with three other materials in
Hull Thickness approximately the following proportions:

†† 1 part Portland cement


In this stage of the design process, we seek to determine
†† 2 parts sand
an optimum hull thickness—one that will achieve an †† 3 parts gravel
appropriate balance between high strength (which requires †† just enough water to combine with
a thicker hull) and greater freeboard (which requires a the cement to form a paste that
thinner hull). binds the sand and gravel together

To make this determination, we prepared concrete test Many people mistakenly believe that
concrete cures by drying out. In fact, it
specimens of various thicknesses and tested them in cures by a chemical reaction between
flexure, loading each specimen until it failed and then the cement and the water, which is
recording the applied load at which the failure occurred. why cured concrete becomes essentially
waterproof—a very advantageous
Based on these tests, we found that a 1/8‑inch‑thick hull characteristic for boatbuilding.
would be too fragile. It would also be quite difficult to
For our sailboat hull, rather than mixing
cast such a thin shell with adequate quality, given the all of these ingredients manually, we will
geometrically complex shape of our hull. use a premixed cement‑and‑sand patching
compound, available at most building
The 3/16‑inch option was found to be considerably supply stores. This material is perfect for
stronger. And as the graph on the previous page shows, the job because it is fast‑setting; includes
this thickness will result in a reasonably comfortable some chemical additives, which enhance
both strength and flexibility; and does not
freeboard of approximately 3/4 inch. include gravel, which would be too large
for the thin hull we will be casting.
A 1/4‑inch‑thick hull would be significantly more robust,
but perhaps more so than we really need. And the greater
weight of this alternative would reduce the freeboard to
under 1/2 inch.

Thus, we can conclude—based on both our buoyancy analysis and empirical tests of concrete strength—
that a 3/16‑inch hull thickness is a reasonable choice for our design.

82 
Heeling Analysis and Sail Design

But is 3/4 inch of freeboard enough? To answer this question, we must first recognize that the wind blowing
on a sail causes the vessel to roll sideways, a phenomenon called heeling.

For a given amount of freeboard, there is a


corresponding maximum allowable heel angle,
beyond which water will flow over the lower gunwale
HULL
θ f
and swamp the boat. This maximum heel angle is
entirely a function of the hull geometry and can be
WATER
r calculated as follows:
LEVEL

A sailboat’s tendency to heel depends primarily on its hull configuration and weight, the wind speed, and
the size and shape of the sail. Because we have specified a maximum wind speed of 10 miles per hour
and have already determined the hull configuration and weight, we can use the maximum allowable heel
angle—20°—as the basis for designing our sail.

The effect of wind blowing on a sail can be


idealized as an aerodynamic force, FA , acting
at the geometric center of the sail. This point
is called the center of effort (CE), and for a
triangular sail, it is located at one‑third of the
sail’s height.

This aerodynamic force is essentially the


same as the one produced by air flowing
across an airplane wing, and its magnitude FA
can be calculated using the same equation:
CE

where: ⅓ of sail height

R
CLR
Joe Mabel/Wikimedia Commons/CC BY‑SA 3.0.
Joe Mabel/Wikimedia Commons/CC BY-SA 3.0.

Project 5 » Concrete Sailboat 83


In response to this wind force, the water surrounding the submerged portion of the hull develops a lateral
resisting force, R, which is applied at the geometric center of the submerged portion of the hull—a point
called the center of lateral resistance (CLR). This force resists the tendency of the wind to push the
hull sideways through the water, in the downwind (or leeward) direction. The need to minimize this
phenomenon, called leeward drift, is the principal reason why every sailboat has a keel.

Viewing our boat from the stern (below), we can see why sailboats heel in the wind. The aerodynamic force
and lateral resisting force cause a moment, which tends to tip the vessel in the leeward direction. This is
called the heeling moment—and its magnitude, designated MH, is given by:

where:

UPRIGHT HEELING

WIND
DIRECTION

FA FA

y
W W
WATER
LEVEL
R R

B x B

84 
What prevents the boat from capsizing?

There are actually two more forces acting on this vessel: its weight, W, which acts downward through the
center of gravity (CG), and the buoyant force, B, which acts upward through the center of the displaced
water. When the boat is upright, these two opposing forces are aligned, so they offer no resistance to the
heeling moment.

But when the boat heels, its CG shifts to windward relative to the buoyant force. Consequently, these two
forces now cause a righting moment, which tends to rotate the hull back toward its upright position. Thus,
the righting moment counterbalances the heeling moment. The magnitude of the righting moment is:

where:

As the heel angle increases, the righting arm (and therefore the righting moment) also increases—right
up to the maximum heel angle of 20°. In other words, the more the hull heels, the greater its resistance to
heeling.

In the science of hydrostatics, this is, by definition, a stable hull. In response to any disturbance that causes
heeling—for example, a gust of wind or a wave—the hull will tend to right itself.

When a steady wind is blowing across a sail, the boat’s increasing heel angle causes a corresponding increase
in the righting moment. At any given wind speed, there is a characteristic heel angle at which the righting
moment equals the heeling moment. At this angle, equilibrium is achieved, and the boat heels no farther.

Our design goal is to ensure that a 10‑mile‑per‑hour wind will not cause a heel angle greater than 20°, the
angle at which water will start flowing over the gunwales. And we can achieve this end through the design
of our sail. In short, we want the sail to be as large as possible—to maximize its propulsive force—but not
so large that it causes the boat to heel more than 20° in a 10‑mile‑per‑hour wind.

To perform this analysis, we first need to estimate the


location of the boat’s center of gravity. Assuming that the hull
cross‑section can be represented as a semicircular arc, the CG r y
is located at a vertical distance, , measured downward from the CG
gunwales, as shown here.

Project 5 » Concrete Sailboat 85


Because our hull will have a finite thickness of 3/16 inch, the radius used in the equation above must be the
average of the hull’s inside and outside radii, calculated as follows:

As shown below, when the boat heels at an angle, θ, the horizontal movement of the CG, x, is given by:

Substituting:

y θ
θ
At the maximum heel angle of 20°, this distance is:

The estimated weight of our hull is approximately 1 pound. Thus, the


righting moment, MR , at the maximum heel angle of 20° is:

h
As noted above, the heeling moment is given by: CE
FA
h/3
y
Substituting the following equations: y’
CLR

The result is:

86 
Based on the principle of equilibrium, MH = MR . Therefore:

Now substitute the following numerical values:

After doing the necessary unit conversions and solving for the sail height, h, the result is:

It should be noted that the above analysis involves two substantial simplifications. It considers neither the
combined weight of the mast, boom, and sail (which tends to increase the heeling moment) nor the weight
of the keel (which tends to increase the righting moment). Because these two weights counteract each
other—at least in part—they can be safely ignored without introducing a large error into the analysis.

Keel Design
DI W
With the sail size defined, the remaining design RE IN
CT D DIRECTION
task is to determine the longitudinal position IO OF TRAVEL
N
of the keel. Here, our goal is to ensure that the
vessel can maintain a constant heading under
sail, with minimal tendency to turn either to
windward or to leeward. The associated design FF
criterion—common to all sailing vessels—is that
the CE should be located ahead of the CLR.
RL CLR
To understand why this criterion makes sense,
consider a sailboat for which the CE and CLR CE FL
are aligned longitudinally, as shown here.

For any combination of sail position and wind


RR
direction, the aerodynamic force on a sail can be
visualized as a pair of component forces: one in
the direction of travel—the heading—and one
perpendicular to it. Both forces are applied at d
the CE.

Project 5 » Concrete Sailboat 87


In response, the hull’s movement through the water causes approximately equal‑and‑opposite resisting
forces, applied at the CLR.

Note that the forward component of the aerodynamic force and the rearward component of the resisting
force are laterally offset by the distance, d. Thus, they cause a moment that tends to turn the vessel into the
wind. This effect is called weather helm, and it can only be counteracted with the application of opposite
rudder, which increases drag and thus reduces efficiency.

Note, also, that this lateral offset of the two forces is caused not only by the position of the sail, but also
by the leeward shift of the CE due to heeling. Thus, the phenomenon is guaranteed to occur, regardless of
sail position.

Now try shifting the CLR


toward the stern by changing
the keel position. Although the DI W
RE IN DIRECTION
magnitudes of the aerodynamic CT D OF TRAVEL
and resisting forces do not IO
change, the two lateral force N
components are now offset
longitudinally by a distance
called the lead. Thus, these
forces now cause a moment FF
that tends to turn the vessel in
the leeward direction, thereby
offsetting the weather helm and
allowing the vessel to maintain
its heading with minimal
application of rudder.
lead CE FL
As a rule of thumb, this
CLR
balanced condition is typically
achieved when the lead is about RL
20% of the total hull length
(measured along the waterline).
RR
With the keel position
determined, our design is
essentially complete. d

88 
FINAL DESIGN DRAWINGS:
SAILBOAT
MAST
The SketchUp 3‑D computer model for this project
can be downloaded from http://stephenjressler.com/ Drawing 5.1 » Concrete
sailboat, overall
diy-engineering/.
perspective view

Six full‑size cutting patterns are provided in the PDF


version of this guidebook (see Appendix B).

Key features of the design


SAIL
}} The concrete hull is cast integrally with the keel.
}} The wooden bow block is permanently
embedded in the bow, both to form the bow BOOM
shape and to provide a socket for the mast.
}} The wooden rudder is fixed to a steel wire HULL
rudder post, which pivots in an aluminum tube,
which is embedded in the concrete transom.
}} The top of the wire rudder post is bent
KEEL RUDDER
horizontally to form a tiller, which can be
clamped to the 1/32‑inch plywood deck to set the
rudder at any desired angle.

SCREW EYE PEG


CLEAT SCREW EYE
BOW BLOCK
TILLER
CLEAT
DECK TRANSOM

RUDDER
POST
Drawing 5.2 » Concrete
sailboat, hull details

RUDDER

Project 5 » Concrete Sailboat 89


}} The mast and boom are made of 3/16‑inch wooden dowels that are tapered at their ends.
}} The boom is attached to the mast with a screw eye, which serves as a pivot.
}} The sail is made of muslin, a lightweight cotton cloth.
}} Small screws are used as cleats, to which rigging lines will be secured.
}} At the stern end of the boom are a small peg, to which the sail will be secured, and another screw eye,
which will serve as a point of attachment for a rigging line called the sheet, which is used to control
the sail position.

FINAL DESIGN DRAWINGS:


HULL MOLDS

Drawing 5.3 » Hull molds, viewed from rear and above


(plastic shells omitted for clarity)

MALE MOLD

KEEL MOLD
(HALF) KEEL MOLD
(HALF)

FEMALE MOLD

90 
MALE MOLD

Drawing 5.4 » Hull molds,


viewed from front and
below (plastic shells
omitted for clarity)

KEEL MOLD
(HALF)

KEEL MOLD
(HALF)
FEMALE MOLD

L
J K
S H I
Drawing 5.5 » Hull molds,
U exploded view

G T
O E
T Q
F P
R D
B C
A
F G

Project 5 » Concrete Sailboat 91


MATERIALS LIST

NAME DESCRIPTION
Hull Portland cement patching compound (1 pint)
Bow block 2½″ × 1¼″ × 1¼″ balsa wood block*
Deck 4½″ × 1½″ × 1/32″ plywood*
Rudder 1½″ × 1″ × ¼″ poplar
Rudder post 1/32″ music wire, 6″ long*
Rudder pivot 1/32″ inside diameter aluminum or brass tubing, 2½″ long*
Rudder hardware 4‑40 × ½″ screw with wing nut
Mast 3/16″ dowel, 13¾″ long
Boom 3/16″ dowel, 9½″ long
Sail 10″ × 12″ muslin cloth
Fasteners Small screw eye (2)
Cleats Small panhead screw (5)
Rigging Nylon string, 2 feet
Hull mold (components A–N) ¾″ × 5½″ × 6′ poplar or plywood
Hull mold (components O–P) ¼″ × 5½″ × 2½″ poplar or plywood
Hull mold (components Q–R) 1/8″ × 2″ × 3″ plywood*
Mold shells (components T–U) 2‑liter plastic beverage bottle
Keel mold fasteners ¼″ × 2″ machine bolts
*These items are most likely to be available from a hobby shop that sells model airplane supplies.
An excellent online source is SIG Manufacturing Co. (http://www.sigmfg.com/).

The other supplies needed for this project are 1″ and 1¼″ wood screws, wood glue, epoxy glue, and
polyurethane varnish.

FABRICATION AND ASSEMBLY PROCEDURE

The first major phase of this project is to build the female and male molds that will be used to cast the
concrete hull. In assembling these molds, it is important to plan for the eventual removal of the completed
concrete casting. Because some degree of adhesion inevitably develops between the hardened concrete
and the mold, removing the casting typically requires complete disassembly of the molds. Thus, most
components of these molds are fastened together with screws or bolts, but no glue.

1 Begin by prefabricating all required wooden components of the molds. Print the full‑size cutting
patterns provided in the PDF version of this guidebook. Cut them out and adhere them to wood of the
appropriate thickness (as indicated in the materials list above); cut out each wooden component with

92 
a scroll saw, band saw, or miter saw, as appropriate; and sand smooth. Components B–D and G–L,
called formers, must be cut with particular precision, such that there is a consistent 3/16″ gap between
the male and female molds when they are assembled.

2 At the locations indicated on the cutting patterns, drill 1/4‑inch holes in components F, Q, and R;
1/2‑inch holes in both copies of component G; and 5/32‑inch pilot holes in components E, M, and N.

3 Assemble the female hull mold as follows:

}} Glue component B to component A with wood glue.


}} Glue component P to component E.
}} Glue the main formers—components C and D—to the base, M.
}} Fasten component O to component E with two 1‑inch wood screws inserted from the rear of E.
}} Fasten subassemblies AB and EOP to the base, M, with 1¼‑inch wood screws.

AB P E
C D
3

4 Build the two keel mold assemblies


by gluing components G, Q, and R to G
component F, as shown here. Q
F
R

Project 5 » Concrete Sailboat 93


5 To give the keel a streamlined shape, apply putty to the
ROUND OFF interior corners and round off the upper edge, to provide a
UPPER EDGE
smooth transition between the keel and hull.

APPLY PUTTY
TO CORNERS
KEEL MOLD 6
5
T
T

6 The two halves of the keel mold are fastened


together with 1/4‑inch‑by‑2‑inch machine bolts
(component T), and then the assembly is inserted
into the female mold between formers C and D
D
and fastened with 1 1/4‑inch wood screws driven C
up through the base (component M) from below. M

7 For the shell of the female mold, cut apart a 2‑liter


plastic beverage bottle, as shown here.

SLOT O
FOR KEEL
T

8 Remove component O from the wooden


framework of the female mold and then
fit the plastic shell (component T) into the
cradle formed by the framework. Screw
component O back into position to clamp
the shell in place. Then, use a sharp knife
to cut a slot into the bottom of the shell
immediately above the keel mold to allow
concrete to flow down into the keel.
8

94 
PROW

9 Use epoxy putty to form a more realistic prow shape within


the neck of the bottle. After the putty has cured, refine the
shape with a file or sandpaper.

10 To complete the female mold, set 9


this piece of 1/32‑inch inside‑diameter
aluminum or brass tubing vertically
into a 3/32‑inch hole in the base, just
ahead of the rear former. This tube
will eventually be encased in the
concrete transom and will serve as a
pivot for the rudder.

10
11 The male mold is assembled in much
RUDDER PIVOT the same way as the female mold,
with a series of wooden formers
(components H–L) screwed to a
central spine (component N). At the
bow, a carved balsa block (component
S) will become a permanent part of
the hull. The 3/16‑inch hole drilled
into its top surface will serve as a
socket for the mast, as well as a means
of temporarily attaching the block
to the male mold framework. This
framework supports another plastic
shell (component U), which is cut
from the other half of the 2‑liter
beverage bottle and trimmed to fit
over the male mold.

N 12 Now, with both the male and female


molds complete, give all of the
L wooden components a few coats of
K polyurethane varnish to waterproof
J
S H I them and help prevent them from
adhering to the concrete.
11 U
13 Once the varnish has dried, the two
plastic shells are attached to their
respective molds with masking tape,
as demonstrated in lesson 7.

Project 5 » Concrete Sailboat 95


Now we are ready to cast the hull. The required materials are Portland cement–based patching compound,
a cup of water, a plastic tub for mixing, some household oil, and the molds. The required tools are a
paintbrush, a putty knife, a screwdriver, and a hammer. The procedure for casting the concrete hull is
demonstrated in lesson 7.

14 To keep the concrete from adhering to the forms, brush a light coat of oil onto all surfaces that will
come into contact with concrete.

15 Combine about a pint of patching compound with just enough water to make the mixture flowable.
After mixing thoroughly, pour the concrete into the female mold until it fills the keel; then, consolidate
it with a thin piece of wire or wood to remove any air voids. Next, continue filling to a depth of about
1 inch; then, press the male mold down into position and screw it in place. Concrete should overflow
the gunwales around the full perimeter of the hull. If it does not, add concrete until the entire mold has
been filled to the brim.

16 Immediately tap the forms with a hammer to consolidate the concrete.

17 Finally, use a putty knife to remove any excess concrete and smooth the upper surface of the gunwales
and then set the casting aside to cure for at least a day.

18 Once the concrete has cured, carefully and methodically remove


the hull from the mold. First, cut away the tape just above the
gunwales. Then, disassemble the wooden framework of the
male mold, pull out the dowel that connects the balsa block
(component S) to the male mold, peel off the inner plastic
shell, pry both halves of the keel mold off the keel, and remove
the outer plastic shell. The result is a very big mess and a very
beautiful concrete hull!

1"
18
CUT LINE
FOLD LINE

19
9 ¾"
LUFF 19 To transform this concrete hull
LE

into a sailboat, we will need a


AC
H

mast, boom, and sail. The sail is


cut from a piece of lightweight
muslin cloth, using the
FOOT dimensions at left.
1"
½"

8"

96 
NYLON STRING
20 The muslin is folded over and glued along all
three edges—traditionally called the luff, leach,
and foot of the sail. The fold along the luff
20
forms a sleeve that will accommodate the mast. SLEEVE
FOR MAST
A similar sleeve along the foot will accommodate FOLD
the boom. And the fold along the leach encloses
a nylon string that will be used to secure the sail.
These folds can be secured either by gluing with
fabric cement or by sewing.

⁄16" HOLE
1

3
⁄16" DIAMETER MAST

SLEEVE FOR BOOM


21

21 The mast and boom are both fabricated from


SCREW EYE 3/16‑inch hardwood dowels tapered at their ends,
where 1/16‑inch‑diameter holes are drilled for
3
⁄16" DIAMETER BOOM
fastening the sail. The small peg mounted at the
⁄16" HOLE
1 end of the boom will also be used to secure the
PEG sail. A small screw eye is driven into the base
of the boom to serve as a pivot, and another is
mounted near the tip of the boom, where it will
SCREW EYE serve as a point of attachment for the sheet—the
rigging line that is used to control the sail position.

1
⁄32" PLYWOOD DECK

22 This 1/32‑inch plywood deck


is fastened to the stern with
epoxy glue.
22

Project 5 » Concrete Sailboat 97


23 The rudder is fabricated from a sheet of 1/4‑inch poplar, sanded to a 23
streamlined shape, and glued to an L‑shaped piece of 1/32‑inch‑diameter
steel music wire, which will serve as the rudder post.
1
⁄32" MUSIC WIRE
24 The rudder post is inserted from below into the aluminum RUDDER POST
tube embedded in the concrete transom of the hull. The top
of the wire rudder post is then bent forward to
form a tiller, which can be used to set the rudder
at any desired angle. The front end of the tiller is
bent into an eye that holds a screw and wing nut,
which can be clamped to the plywood deck to
hold the rudder in position. RUDDER

SCREW AND WING NUT


TILLER
TUBE
EYE
25

24 SAIL
TIED TO SAIL
CLEAT TIED
TO PEG

SHEET
TIED
TO CLEATS
25 The completed hull should be sealed with a few coats of
varnish to enhance its appearance and durability while also
reducing friction as the vessel moves through the water.

26 To prepare for sea trials, insert the boom into the sleeve at the foot
of the sail; then, insert the mast downward through the luff of the sail and then through the screw eye
at the end of the boom. The ends of the nylon string enclosed within the leach of the sail are inserted
through the holes at the upper end of the mast and the outer end of the boom. Now plug the mast into
its socket in the bow and secure the sail by tying off the loose ends of the nylon string—the upper end
to a cleat at the bow and the lower end to the peg at the tip of the boom. Finally, the sheet is looped
through the screw eye at the aft end of the boom, and both of its ends are wound around the cleats
mounted in the stern deck.

The concrete sailboat is now complete and ready for testing!

98 
The Physics of Wind and Sail
Before we can test our boat’s performance under sail, we must first understand how wind flowing across a sail propels a
boat forward, for a given combination of wind direction, heading, and sail position.

We begin with some basic nautical terminology.

A sailboat’s direction of movement with respect to the wind direction is defined using a set of terms called points of sail.

When the boat is heading in the same direction as the wind, it is said to be running. When it is heading more‑or‑less
perpendicular to the wind, it is said to be reaching or sailing on a beam reach. And when the boat is sailing as close
as possible to the wind while still making forward progress, it’s said to be close‑hauled. Beyond this point of sail is a
range of headings, called the no‑go zone, within which it is impossible to sail forward.

POINTS OF SAIL

running

reaching reaching

no-go
zone
close-hauled close-hauled

wind direction

Let us examine the interaction between SAILING ON A BEAM REACH


wind and sail by looking more closely
at a boat sailing on a beam reach. For FA FL
simplicity, we will assume that the vessel is
initially stationary in the water.
CE
FF
angle of attack = 20–30º

heading

wind direction

Continued on next page…

Project 5 » Concrete Sailboat 99


In this situation, the optimum position of the sail is as shown here, with the boom inclined at roughly 20° to 30° from
the wind direction. This angle is called the angle of attack—a term we will use again in our study of aerodynamics.

This is no coincidence, because most of the time the sail of a sailboat works in exactly the same way as the wing of an
airplane. As the wind flows across a properly positioned sail, it forms the sail into a curved shape, called an airfoil, and
this inclined, curved surface effectively deflects the mass of air moving across it. This change in the momentum of the
air can only occur if the sail is applying a force to the air. Based on Newton’s third law of motion, if the sail is applying
a force to the air, then the air must be applying an equal and opposite force to the sail.

This is the aerodynamic force we designated as FA in the design of our boat above. The magnitude of FA is a function of
the density of the air, the area of the sail, and the wind velocity squared. Its direction is approximately perpendicular to
the boom on which the sail is mounted.

This aerodynamic force can be visualized as a pair of force components that are parallel and perpendicular to the
vessel’s direction of travel. The parallel component of FA propels the vessel forward; thus, we will designate it FF. The
perpendicular component tends to push the hull sideways through the water, in the leeward (or downwind) direction,
so we will designate it FL .

This leeward force is generally detrimental and is counteracted by the lateral resistance of the keel as it moves sideways
through the water. Indeed, minimizing this adverse leeward drift, called leeway, is the principal reason why sailboats
always have very prominent keels.

To a large degree, a sailboat’s performance at a given point of sail is determined by the relative magnitudes of FF and FL .

Up to this point, we have been looking at the forces developed on a sail when the boat is at a standstill. But as soon as
the vessel starts moving, the mechanics of wind and sail will change, because forward motion causes a change in both
the magnitude and direction of the airflow across the sail.

Suppose that there is no wind and you are in a boat that is moving forward at 5 miles per hour. Under these conditions,
you will feel a 5‑mile‑per‑hour wind blowing from bow to stern, even though the velocity of the true wind is zero.
This phenomenon is called apparent wind, the airflow you actually feel from the perspective of your moving frame of
reference on board the boat.

Now assume that you are sailing on a beam


APPARENT WIND
reach at 5 miles per hour and the wind is
blowing at 10 miles per hour perpendicular to
boom repositioned
your heading, from port to starboard, as shown apparent to account for
at right. Under these conditions, you feel the wind true apparent wind
combined effect of the true wind and the boat’s wind
forward motion. Mathematically, the resulting
apparent wind—the airflow that you actually
effect of
feel—is the vector sum of these two velocities. forward motion
And the airflow that you feel is also the airflow
that the sail feels. heading

wind direction

100 
This concept of apparent wind has two important implications for the sailor.

First, as the boat accelerates from a standstill, the boom must be repositioned—rotated toward the stern—such that the
sail maintains its optimal angle of 20° to 30° with respect to the apparent wind.

Second, when sailing on a beam reach, the apparent wind speed is actually greater than the true wind speed. This
explains why many sailboats actually achieve their maximum speed when reaching, rather than running. In other
words, they are faster when sailing perpendicular to the wind than when sailing downwind.

Now let us consider the most interesting point of sail: the close‑hauled condition, when the boat is sailing diagonally
into the wind—or “working to windward,” in nautical parlance. The diagram below shows the vector corresponding to
the true wind, the effect of the boat’s forward motion, and the resulting apparent wind. Now when the sail is set at its
optimum angle with respect to the apparent wind, the boom is oriented almost directly aft.

In this orientation, because the apparent wind speed is so high, the sail produces a large aerodynamic force. The problem is
that only a very small component of this force is oriented in the forward direction and a very large component is leeward.
Thus, when close‑hauled, the vessel is subject to considerable heel and leeway, while making minimal forward progress.
Still, sailing slowly and obliquely into the wind is better than not sailing into the wind at all.

SAILING CLOSE-HAULED sail set at 20–30º


off the apparent

effect of
FL wind direction

forward
motion FA

apparent CE
true
wind wind
FF

heading

wind direction

Continued on next page…

Project 5 » Concrete Sailboat 101


The final point of sail we will consider is running, where the vessel’s direction of travel is essentially the same as the
direction of the true wind. Running is the exception to the rule in the sense that, here, the sail is set perpendicular to
the wind, rather than at the optimum angle of attack—20° to 30°. At this orientation, the sail’s aerodynamic force is
very nearly aligned with the direction of travel, so nearly all of it contributes to forward motion.

RUNNING
FA
heading

CE
effect of
forward
true motion
wind
apparent
wind

wind direction

For this reason, we might expect that a sailboat’s maximum possible speed would always be achieved when running
downwind. As we have seen, however, this is often not the case. At this point of sail, the apparent wind speed is
significantly less than the true wind speed. And the faster the boat moves downwind, the less propulsive force its sail is
able to produce.

Conversely, when a sailboat is reaching, its forward motion actually increases the apparent wind speed, resulting in
the potential for significantly faster speed through the water. Indeed, when reaching, it is theoretically possible for the
boat’s forward speed to exceed the wind speed. When running, a sailboat can never sail faster than the wind.

TESTING PROCEDURE

The sailboat should be tested at a pond or pool where the water is relatively still. A breeze is essential,
though it should not exceed 10 miles per hour.

Initially, place the boat into the water and check for leaks and adequate buoyancy. The hull should float
with approximately 3/4 inch of freeboard.

Next, check the vessel’s hydrostatic stability by tipping it sideways 10° to 15° and then releasing it. If the
hull is stable, it will quickly return to its normal upright orientation.

102 
Now determine the wind direction and choose the direction in which the boat will be sailing. Based on the
wind direction and your chosen heading, will the boat be running, reaching, or close‑hauled? Set the sail’s
angle of attack accordingly, using the sheet to secure the boom in the correct orientation.

Give the boat a gentle push in the intended direction and observe its tendency to turn. If the vessel does not
maintain a reasonably straight heading, retrieve it, adjust the rudder accordingly, and try again.

Project 5 » Concrete Sailboat 103


PROJECT 6
RADIO-
CONTROLLED
BLIMP
DESIGN, BUILD, AND TEST A RADIO-
CONTROLLED ELECTRIC-POWERED BLIMP
LESSON IN WHICH THIS PROJECT IS COVERED:
8 » Make a Radio-Controlled Blimp

PROBLEM DEFINITION
Requirements:
}} Must be capable of sustained flight in an indoor environment.
}} Must be capable of controlling both direction and altitude.

Constraints:
}} Use only commonly available materials and components.

REFERENCES
Hibbeler, Engineering Mechanics.
Simons, Model Aircraft Aerodynamics.

104 
DESIGN CONCEPT

Our design concept is shown in the sketch here. Like


all real-world blimps and hot-air balloons, the vehicle BALLOON
will be lighter than air, with buoyancy provided by
a common latex balloon inflated with helium. The
balloon will support a lightweight module—called
the gondola—which incorporates:
PROPELLER
TAIL BOOM
}} an adjustable mounting for two small
electric motors driving propellers;
}} an enclosure that houses a battery and the MOTOR FIN
GONDOLA
electronics required to control both the forward
motion and steering of the blimp;
}} an overhead framework to connect the gondola to the balloon; and
}} a vertical fin mounted on a tail boom to improve the aircraft’s
directional stability in flight.

Because the principal focus of this project is on aerodynamics and


buoyancy—not on electricity and electronics—we will obtain our
propulsion and radio-control systems by borrowing the motors and
electronics from an off-the-shelf toy tank, which was purchased from
a major online retailer for less than $20.

DETAILED DESIGN

Basic Aerodynamics Concepts


Any powered aircraft moving through the air is acted upon by four principal forces:

†† the weight of the aircraft, which acts vertically downward through the center of gravity;
†† lift, which acts upward;
†† thrust, which propels the craft forward; and
†† drag, which resists the forward movement of the craft through the air.

The flight of a powered aircraft is governed by the balance—and occasional imbalance—of these four forces.
According to the principle of equilibrium, if the aircraft is flying level at constant altitude, then lift must equal weight,
and if it is flying at constant speed, then thrust must equal drag.

Continued on next page…

Project 6 » Radio-Controlled Blimp 105


LIFT

THRUST DRAG
TGC SR

WEIGHT

For a conventional airplane, these two pairs of equal and opposite forces are both physically and mathematically
interrelated, because lift is created by air flowing across a wing. The balance between thrust and drag determines
the airplane’s airspeed; thus, thrust has a significant influence on lift. This interdependence creates some significant
challenges, not only for airplane design, but also for trimming—the process of adjusting an airplane’s balance and
controls to achieve optimum flight characteristics.

Fortunately, this will not be the case for our blimp. For any lighter-than-air vehicle, lift is provided primarily by
buoyancy, rather than airflow; thus, lift is largely independent of airspeed. A lighter-than-air vehicle will fly even if
it has no thrust, as we can readily see with hot-air balloons. And in still air, any thrust—no matter how small—will
result in forward motion.

Thanks to this fundamental characteristic of lighter-than-air flight, the design of our blimp in this project will be
considerably less complicated than the design of our model airplane in the next project.

Buoyancy Analysis

The design of our blimp requires just one science-based analysis, the purpose of which is to determine the
diameter of the helium-filled balloon required to provide the lift needed to support the vehicle’s total weight.

The relevant scientific concept is the principle of buoyancy, which says that an object immersed in a fluid
develops an upward buoyant force with magnitude equal to the weight of the fluid it displaces. For our
sailboat design, the relevant fluid was water. For our blimp, it is air. (The scientific definition of a fluid
encompasses both liquids and gases; thus, air is classified as a fluid.)

The critical criterion for this design is that the total weight of the blimp, Wblimp , cannot exceed the weight of
the air it displaces, Wair . Mathematically:

106 
We can calculate the weight of displaced air as follows:

where:

The weight of the blimp will be considered in three main components:

}} WG is the weight of the gondola, which includes both the gondola and its contents—motors,
propellers, batteries, electronics, and wiring. We will determine this weight simply and confidently by
building the gondola and weighing it.
}} W B is the weight of the latex balloon, not including its contents. This weight will depend on the
specific size and type of balloon we purchase for the project.
}} W H is the weight of the helium inside the balloon. Helium is only about one-seventh as dense as air,
which is why it is used in balloons and blimps; however, we cannot ignore this weight or we will
effectively introduce an error of one-seventh, or 14%, into our buoyancy calculation. In other words,
if we do not account for the weight of the helium, we will be designing for 14% less lift than is
required to keep the blimp aloft.

The weight of helium can be calculated as follows:

where:

We can now write the complete buoyancy equation for the blimp:

Project 6 » Radio-Controlled Blimp 107


Using algebra to solve for V:

Latex party balloons are sold in a variety of different diameters; thus, to select an appropriate balloon, we
will need to relate the volume to the diameter. These balloons generally are not perfectly spherical, but their
shape is sufficiently close to spherical that we can represent the volume in our buoyancy equation as the
volume of a sphere:


Making this substitution and solving for the radius, we get the following equation:

Doubling this equation gives us the required diameter of our balloon:

This equation, which derives directly from the principle of buoyancy, provides us with a very powerful
design tool. However, we cannot use this tool until we have determined the weight of the balloon and the
weight of the gondola.

A major party supply store sells latex balloons in two sizes: an 11-inch balloon that weighs 3 grams and a
24-inch balloon that weighs 22 grams. These will serve as our two principal design alternatives.

We will determine the weight of the gondola by designing, building, and weighing it.

Designing the Gondola

Before we can design and build the gondola, we must examine the functionality and inner workings of the
toy tank on which its design will be based. The key functional characteristics of this toy are as follows:

108 
}} Each track is independently driven by its own motor.
}} Each motor operates either at full speed forward, at full speed in reverse, or stopped. There are no
intermediate speeds.
}} Thus, the vehicle can move in only four discrete modes. When both tracks are spinning in the same
direction, the tank moves either straight forward or straight rearward, and when the tracks are
spinning in opposite directions, the vehicle pivots either right or left.

This propulsion and control system is entirely suitable for our blimp, except that the motors will no longer
be driving tracks but, rather, spinning propellers to create thrust.

These three graphics show the blimp viewed from below with one motor-propeller unit mounted on
either side of the gondola. Note that the blimp will operate in the same four modes as the tank. When
both propellers are producing forward thrust, the aircraft will move forward; when both are producing
rearward thrust, it will move backward; and when one propeller is producing forward thrust and the other
is producing rearward thrust, the aircraft will pivot, either to the left or right.

BOTH PROPELLERS PRODUCING FORWARD THRUST

FORWARD
MOTION

BLIMP VIEWED FROM BELOW

BOTH PROPELLERS PRODUCING REARWARD THRUST

BACKWARD
MOTION

BLIMP VIEWED FROM BELOW

Project 6 » Radio-Controlled Blimp 109


ONE PROPELLER PRODUCING FORWARD THRUST,
ONE PRODUCING REARWARD THRUST

PIVOTING MOTION

BLIMP VIEWED FROM BELOW

There is one critical difference between the operation of the tank and that of the blimp, however. There
is substantial friction between the tank’s tracks and the ground; thus, when the tank’s drive motors stop
spinning, the tank’s motion also stops immediately. But there is considerably less friction between our
blimp and the air, so once the vehicle is in motion—in whatever direction—it will tend to remain in
motion for a considerable period of time, even after the motors have stopped spinning.

In one sense, this behavior is advantageous, because it will allow us to use short pulses of forward thrust to
control the aircraft’s speed during forward flight. But this tendency can also be somewhat problematic when
executing turns, because even a short pulse of power with one propeller thrusting forward and the other
thrusting rearward will start the aircraft spinning (about its vertical axis). Then, it will just keep spinning for
an extended period of time, even after the motors have been switched off. This phenomenon tends to make
directional control somewhat more challenging for the blimp than it is for the tank. This is why the tail fin
has been included in the concept design for our blimp. The use of this fin—also called a vertical stabilizer—
greatly improves the directional control and stability of the aircraft, particularly during forward flight.

Blimp Propulsion and Control


The system of propulsion and
control we will be using for our
model is essentially the same as
the system used by many full-
size blimps and airships. Note,
in this photo, that the blimp
is propelled by two motors
mounted side by side and the
design includes tail fins for
© gobigpicture/iStock/Thinkstock. stability and directional control.

110 
This particular toy tank has a few additional features: an LED (light-emitting diode) light to indicate when
the power is switched on; a rotating turret (which requires a third motor); and the simulated sound of a diesel
engine (which is delivered through a small internal speaker). To save weight, we will disconnect the third
motor and speaker for our project; however, we will integrate the LED into our design as a navigation light.

Lesson 8 demonstrates the process of disassembling the toy tank to harvest the components we need:

}} two drive motors


}} circuit board
}} battery
}} battery-charging socket
}} switch
}} LED light
}} antenna

These components, which will provide our blimp with both its propulsion and control system, weigh a mere
9 grams—about one-third of an ounce.

In designing our gondola, we must note the size and shape of the motors so that we can design appropriate
mounts; the diameter of the motor shafts so that we can fit appropriately sized propellers; the length of the
wire leads that connect the motors to the circuit board so that we will know how far apart we can mount
the motors; and the sizes of all remaining components, which must be accommodated within the body of
the gondola.

Design drawings for the gondola and instructions for its fabrication and assembly are provided on the
following pages. When completed, it will weigh approximately 18 grams (0.04 pounds) with the propulsion
and control system installed.

Returning to the buoyancy equation we developed earlier, we must now make one small modification:

Note that an empirical correction factor, C, has been added to the equation. The purpose of this factor is to
account for several important discrepancies between our theoretical analysis and the real-world conditions
under which the blimp will operate. We will use C = 1.3, which effectively means that our calculated
diameter will provide 30% more lift than would have been provided by the theoretical solution.

Project 6 » Radio-Controlled Blimp 111


There are three reasons for building this rather generous margin of error into our design:

1 Commercial helium is often not 100% pure. The manufacturer’s specifications for helium we will be using
indicate that their product—advertised only as helium—is actually a mixture of 80% helium and 20% air.
Thus, 20% of the space inside our balloon will be occupied by a substance that provides no buoyancy.

2 Our calculations do not account for the fact that the stretch of the inflated latex balloon will compress
the helium somewhat. Consequently, the helium will be slightly denser than our calculations assume,
and the balloon’s lift will be reduced accordingly. This error is relatively small—on the order of 1%.

3 Even if our blimp flies perfectly on the first day, helium will slowly leak out of the balloon over time.
If we design for some excess buoyancy by making the balloon somewhat larger than the minimum
requirement, we can initially add ballast to establish the required balance between lift and weight.
Then, as the helium begins to leak out, we can gradually reduce the ballast to keep the blimp
operational for a longer period of time.

Returning to our two alternative balloon sizes—an 11-inch balloon that weighs 3 grams and a 24-inch
balloon that weighs 22 grams—it seems likely that the larger balloon will work. But it will be interesting to
see if the significantly lower self-weight of the smaller balloon might make this option feasible as well.

Substituting the densities of air and helium and the weights of the gondola and balloons into the buoyancy
equation:

Based on these calculations, we can conclude that the larger, heavier balloon only needs to be inflated to
an 18-inch diameter to meet our needs, while the smaller, lighter alternative would only work if we could
inflate it to 14.5 inches. Of course, this balloon has a specified maximum diameter of only 11 inches, so it
will not work.

With this determination, the design process is complete.

112 
FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/diy-
engineering/. A full-size layout drawing and cutting patterns are provided in the PDF version of this guidebook.

Drawing 6.1 » Blimp gondola,


perspective view from above

Drawing 6.2 » Blimp gondola,


perspective view from below,
with balloon mounted

Project 6 » Radio-Controlled Blimp 113


J
J
K L
H

I
I

C I
D I
A' C
F B
E
G A Drawing 6.3 »
Blimp gondola,
exploded view
D

The key features of this design are as follows:

}} The gondola body is an open-topped box made of balsa wood, except for the sides, which are thin
plywood. The sides serve as bearings for the pivoting motor mount assembly.
}} One side of the body is notched to accommodate the battery-charging socket.
}} Both the power switch and LED protrude through holes in the floor.
}} The motor mount assembly consists of a 1/8-inch wooden dowel with two mounting rings attached to
each end. One motor fits snugly within each pair of rings.
}} The two propellers—one right-handed and one left-handed (see “Right- and Left-Handed Propellers”
sidebar)—are hand-carved from balsa wood.
}} This overhead framework, made entirely of balsa, is used to fasten the balloon to the gondola and
provide the required clearance between the propellers and the balloon.
}} This framework also serves as the mounting point for the tail boom, on which the fin is mounted.

114 
Right- and Left-Handed Propellers
A right-handed propeller is one that produces forward thrust when rotating clockwise, viewed from
the perspective of the pilot looking forward. A left-handed propeller produces forward thrust when
rotating counterclockwise, viewed from the same perspective.

Our blimp uses one right-handed and one left-handed propeller, because our two electric motors are
wired to spin in opposite directions when the vehicle is moving forward. This situation is actually quite
advantageous, because of a phenomenon called torque effect—the tendency of an aircraft to roll in the
opposite direction of its propeller’s rotation.

When an aircraft has two propellers rotating in opposite directions, the torque effect of one cancels out
the torque effect of the other, and the tendency to roll is effectively eliminated. Thus, the use of one left-
handed and one right-handed propeller on our blimp will significantly improve its performance
and controllability.

MATERIALS LIST
NAME DESCRIPTION
A Gondola side 1/32″ plywood*
A’ Gondola side 1/32″ plywood*
B Gondola bottom 1/16″ balsa*
C Gondola bulkhead (2) 1/16″ balsa*
D Motor mount (4) 1/32″ plywood*
E Motor mount pivot 1/8″ dowel*
F Motor (2) Salvaged from radio-controlled toy tank
G Propeller (2) 3/16″ balsa*
H Balloon holder 1/16″ balsa*
I Post (4) 3/32″ × 3/16″ × 1 3/4″ balsa
J Balloon cradle (2) 3/32″ balsa*
K Tail boom 1/16″ bamboo or bass wood*
L Tail fin 1/32″ balsa*

* Full-size cutting patterns provided in PDF version of this guidebook.

Balsa wood and thin aircraft-grade plywood are available from hobby shops that sell model airplane
supplies. An excellent online source is the SIG Manufacturing Co. (http://www.sigmfg.com/).

The other supplies needed for this project are:

}} helium (obtained from any party supply store)

Project 6 » Radio-Controlled Blimp 115


}} circuit board, battery, battery-charging socket, switch, LED light, antenna, wiring, and transmitter,
all salvaged from a toy radio-controlled tank
}} screw eye and small steel washers, used for ballast
}} wood glue
}} cyanoacrylate glue

FABRICATION AND ASSEMBLY PROCEDURE

1 After printing the cutting


patterns, use spray adhesive to
attach them to plywood and
balsa sheets of the thicknesses
specified in the materials list
above.

2 Prefabricate all of the


required components.
These are very small and
delicate, so take your
time and work carefully.

116 
3 Begin the assembly process by gluing
components A, A’, B, and C together C
to create the gondola body.
A'

B
3 C
A

4 D 4 Glue one pair of mounting rings


(component D) to one end of the motor
mount pivot (component E).
D

5 Insert the motor mount pivot


(component E) through the
E
holes in the sides of the gondola
body and glue the second pair of
D
mounting rings (component D)
in place on the opposite end of
the dowel. D
In performing this step, it is
important to ensure that the
two pairs of rings are aligned so NO GLUE HERE
that the thrust lines of the two
motors will be parallel. Also,
ensure that no glue seeps into
the interface between the dowel
and the gondola sides so that the MOTOR
THRUST LINES
entire motor mount assembly PARALLEL
will be able to pivot freely.

Project 6 » Radio-Controlled Blimp 117


6 Construct the two sides of the
6
overhead frame directly over
the plans.

7
H

7 Glue these two subassemblies


to the outside faces of the
gondola body and add the
balloon holder (component H).

8 Glue the tail fin


(component L) to the tail 8
boom (component K) and
then glue this assembly to
the underside of the balloon
H
holder (component H).

K
L

118 
POWER SWITCH CIRCUIT
BOARD ANTENNA
LED MOTOR SHAFT
BATTERY PROPELLER
MOTOR CHARGING
SOCKET

9 10 11

9 Now, with the gondola essentially complete, add the motors and electronics. First, insert the power
switch and LED light into their respective holes in the floor of the gondola and secure them with a
drop of fast-drying cyanoacrylate glue.

10 Place the circuit board inside the gondola body and secure it in position with a toothpick and a rubber
band for easy removal. (For greater simplicity, the circuit board can be held in place with a piece of
electrical tape.) Finally, insert the motors into the motor mount rings, from the rear toward the front.

11 Cut the propellers from 3/16-inch-thick balsa, using the cutting pattern provided, and then carve them
using the procedure described in project 7. The hub of each propeller is drilled with a small hole to
match the diameter of the motor shaft, and then the propeller is secured to the shaft with a drop of
cyanoacrylate glue. Before gluing, be sure to check both motors’ directions of rotation; then, install
each propeller on the correct shaft so that the forward button on the radio transmitter causes both
propellers to produce forward thrust.

12

12 Inflate a 24-inch latex balloon to a


diameter of 18 inches with helium and
tie off its open end tightly. Then, insert
the neck of the balloon into the slot in INSERT BALLOON HERE
the balloon holder (component H).

The blimp is now complete and ready for testing!

Project 6 » Radio-Controlled Blimp 119


TESTING PROCEDURE

We will begin the flight-testing process by trimming the blimp, making adjustments to the aircraft’s
buoyancy to ensure that the desired flight characteristics are achieved.

Because we designed the blimp so that its buoyant force significantly exceeds its weight initially, it will have
a strong tendency to ascend. This tendency must be counteracted with ballast. Small metal washers serve
this purpose well. Mount a screw hook in the bottom of the gondola body and hang the washers from this
hook. Keep adding washers until the vehicle’s weight just slightly exceeds the buoyant force, resulting in a
slight tendency to descend when the motors are not running.

Next, use the adjustable motor mount to angle the motors upward so that the motors’ thrust will slightly
augment the balloon’s lift.

When trimmed in this way, the vehicle will ascend slightly when under power and descend slightly when
the power is shut off. This guarantees that we can recover the blimp if it loses power.

Now, with both motors running, release the blimp. It should climb gradually while maintaining a straight
heading. When the command to turn is given, the blimp should pivot in the appropriate direction and
will also descend slightly, because one of the two angled motors is now thrusting at a downward angle,
offsetting the upward thrust of the other motor. Renewed application of full forward thrust will arrest this
descent and initiate a shallow climb. Flying smoothly takes some practice!

Over time, the craft’s tendency to descend will increase as the balloon slowly leaks helium. To compensate,
remove some of the ballast.

120 
PROJECT 7
RUBBER-POWERED
AIRPLANE
DESIGN, BUILD, AND TEST A RUBBER-
POWERED FREE-FLIGHT MODEL AIRPLANE.
LESSONS IN WHICH THIS PROJECT IS COVERED:
9 » Exploring Aerodynamics
10 » Build a Model Airplane
11 » Take Flight!

PROBLEM DEFINITION
Requirements:
}} Must be capable of sustained stable flight for a duration of at least one minute.
}} Must be capable of varied flight characteristics through preflight adjustments.

Constraints:
}} Powered only by twisted rubber.
}} No form of remote control.
}} Use only readily available materials and components.

REFERENCES
Ross, Rubber Powered Model Airplanes.
Simons, Model Aircraft Aerodynamics.

Project 7 » Rubber-Powered Airplane 121


EXPLORING AERODYNAMICS
One of the greatest technological achievements in human history has been the design and construction of heavier-than-
air machines capable of carrying humans aloft in sustained, controlled flight. The Wright brothers did it first in 1903,
and the world has not been the same since.

In this project, we will follow in the footsteps of Orville and Wilbur as we design, build, and test a model airplane
constructed of wood, tissue paper, and just a bit of metal and capable of achieving surprisingly high levels of
performance, using only a few strands of rubber for power. This is a free-flight model, meaning that it must be capable
of flying on its own, with no means of remote control.

Achieving sustained stable flight will require the application of many important concepts from the science of
aerodynamics. Foremost among these are the four forces—lift, drag, thrust, and weight—that govern all forms of
powered flight. We will also consider the aerodynamic concepts of balance, stability, and control.

The tool most often used for studying aerodynamic lift and drag is the wind tunnel. Wind tunnels used by
organizations such as NASA, aerospace companies, and university research centers are highly sophisticated
technological systems, often costing millions of dollars and occupying entire buildings. Lacking any such equipment
for this course, we will use a simple DIY wind tunnel instead.

In a typical wind tunnel experiment, a test specimen—often a scale model of an airplane wing—is subjected to a high-
speed stream of air. Instruments are used to measure the lift and drag forces developed by the test specimen at a given
airspeed and angle of attack—defined as the wing’s inclination measured with respect to the airflow.

Two such experiments performed in our DIY wind tunnel produced the graph of lift versus angle of attack shown
below. The two test specimens were a simple flat plate and a Clark Y airfoil.

80
CLARK Y AIRFOIL
70
LIFT (GRAMS)

60

50

40 FLAT PLATE
30

20

10

0
-5 0 5 10 15 20 25 30
ANGLE OF ATTACK (DEGREES)

122 
An airfoil is a streamlined shape designed to produce a desired combination of lift and drag for a desired range
of airspeeds. Airfoils come in a nearly limitless variety of shapes and sizes that are carefully tailored to the desired
performance characteristics of the aircraft for which they are designed. But despite this variety, any given airfoil can be
fully defined in terms of just two geometric characteristics: camber and thickness form.

The camber of an airfoil is defined by the shape of an imaginary line drawn from the wing’s leading edge to its
trailing edge at exactly mid-height between the top and bottom of the airfoil, all along its length. If the camber line is
straight, then the airfoil is perfectly symmetrical; if the camber line is curved, then the airfoil is said to be “cambered”
and is asymmetrical, typically with its upper surface more sharply curved than its lower surface.

The second geometric characteristic—the thickness form—defines the thickness of an airfoil at every point along the
camber line. By definition, the thickness form always consists of pairs of equal offsets, one up and one down.

THICKNESS FORM

CAMBER
OFFSETS

THICKNESS FORM

CAMBER
OFFSETS

The Clark Y airfoil used in our wind tunnel experiment is a cambered airfoil. It was designed by U.S. Army officer
Virginius E. Clark in 1922 and used on many famous aircraft, such as the Spirit of Saint Louis, in which Charles
Lindberg piloted the first nonstop transatlantic flight; and the Lockheed Vega, in which Amelia Earhart earned the
distinction of becoming the first woman to fly solo across the Atlantic.

Comparing the two graphs produced by our wind tunnel experiments, we can see the clear advantage of the cambered
airfoil. It produces more lift at a given angle of attack, and it produces greater maximum lift.

But how does a cambered airfoil produce lift? It is quite surprising that today, more than a century after the Wright
brothers’ first successful flight, aerodynamicists still argue about the physical cause of lift. There are two common
explanations. The first is based on an important concept from fluid mechanics called Bernoulli’s principle. This
principle, first published by Daniel Bernoulli in 1738, states that when the speed of a moving fluid increases, its
pressure decreases—and vice versa.

Application of Bernoulli’s principle to the phenomenon of aerodynamic lift is best illustrated by the diagram below,
which shows a series of streamlines representing the flow of air around an airfoil. Note that, because of the airfoil’s
camber, the streamlines above the wing get closer together, indicating that the airfoil’s shape causes a constriction in
the flow above the wing. Constricting a flow causes its velocity to increase, and Bernoulli tells us that as the velocity of

Continued on next page…

Project 7 » Rubber-Powered Airplane 123


a flow increases, its pressure decreases. Consequently, the
flow of air across the wing creates a region of low pressure
above and relatively higher pressure below. The net effect
of this pressure difference is the upward force called lift.

The alternative explanation of lift says, simply, that the


curved top surface of an airfoil is particularly effective in
deflecting a large mass of air downward. This downward
change in the direction of the airflow tells us that the
wing is applying a large downward force to the air mass,
and therefore, according to Newton’s third law, the air
must be applying an equal but opposite upward force to
the wing. This upward force is lift.

This same explanation applies to the flat-plate test specimen, which also produced lift in our wind tunnel experiment.
The only difference between the flat plate and the cambered airfoil is that the curved shape of the airfoil bends the
airflow more effectively than the flat plate does.

Which of these two explanations is correct? From an engineering perspective, it really does not matter. What matters is
that effective mathematical tools for modeling aerodynamic phenomena such as lift and drag have been available since
the 1920s, primarily through the work of a brilliant German engineer named Ludwig Prandtl, who is widely regarded
as the father of modern aerodynamics. Thanks to the excellent predictive power of these mathematical models, today’s
aeronautical engineers can design efficient airfoils for every conceivable aeronautical purpose, even as the scientific
arguments about the fundamental cause of lift continue.

The basic mathematical equation for aerodynamic lift is:

where:

The lift coefficient, CL , is a non-dimensional parameter that is determined empirically through wind tunnel testing.

Both of the lift-versus-angle-of-attack curves that we developed in our wind tunnel experiment have a characteristic
shape. In general, lift increases with increasing angle of attack—but only to a point, after which the lift drops off rather
abruptly. This drop in lift at higher angles of attack is evidence of an aerodynamic phenomenon called the stall.

Stalling is caused by changes in the velocity and pressure of the air flowing over a wing. When air moves across the
top of an airfoil, it initially accelerates (over the front of the wing) and then decelerates (over the rear of the wing).
Consistent with Bernoulli’s principle, increasing velocity corresponds to decreasing pressure, and decreasing velocity
corresponds to increasing pressure. Thus, the pressure of the air moving across an airfoil initially decreases (over the

124 
front of the wing) and then increases (over the rear of the wing). In general, the movement of a fluid from a region of
high pressure to a region of lower pressure—as we see on the front portion of an upper wing surface—is favorable for
smooth, orderly flow. Conversely, movement from low pressure to higher pressure is unfavorable.

This latter condition—called an adverse pressure gradient—occurs over the rear of a wing’s top surface and tends to
get progressively more unfavorable with increasing angle of attack. At a critical combination of airspeed and angle of
attack, the adverse pressure gradient becomes so severe that the airflow in the immediate vicinity of the wing surface
actually reverses direction, causing the flow to separate from the wing. At the point of separation, the smooth, orderly
flow of air breaks down and is replaced by a turbulent wake, resulting in a sharp reduction in lift and a drastic increase
in drag.

And what is aerodynamic drag? As we have discovered in previous projects, drag is a force that resists the movement
of a body through a fluid. For an aircraft in subsonic flight, there are three principal types—skin friction drag, form
drag, and vortex drag—each with its own unique cause.

Skin friction drag is caused by the friction associated with a fluid moving across a surface. If the surface is smooth, skin
friction drag is low; if it is rough or bumpy, skin friction drag is higher.

Form drag is caused by variations of pressure associated with the flow of air around a body. A streamlined shape has
low form drag, while an angular shape typically has higher form drag. When an airplane wing stalls, the resulting drag
due to flow separation is primarily form drag.

Vortex drag (also called induced drag) is inextricably connected to lift. A lifting wing has relatively high-pressure air
below and low-pressure air above. Air naturally moves from high to low pressure by the most direct path available.
Thus, at the outer ends of a wing, the high-pressure air below spills up over the wing tip toward the low-pressure region
above, creating a spiraling wake called a vortex, which trails behind the airplane and resists its forward movement.
Because this type of drag originates in the pressure difference that causes lift, it is an inherent adverse consequence of
lift. In general, a wing with a higher lift coefficient also produces more vortex drag.

All three forms of drag contribute


to the total drag force acting on an
aircraft in flight, but they do not
contribute equally. The graph below
shows the relative contributions of
skin friction, form, and vortex drag as
DRAG FORCE

a function of airspeed. Note that both


skin friction and form drag increase
dramatically with airspeed, while TOTAL DRAG
vortex drag increases with decreasing
airspeed—and it dominates the total FORM DRAG
drag force at very low airspeeds.

This is quite important for our VORTEX DRAG


purposes, because our rubber-powered SKIN FRICTION DRAG
model airplane will be operating
primarily at low speed. Thus, our
AIRSPEED
design will need to address a particular
susceptibility to vortex drag.

Project 7 » Rubber-Powered Airplane 125


DESIGN CONCEPT

In developing our design concept for this project, we will be aided by many generations of aviation
enthusiasts, who have been building rubber-powered models for nearly a century and a half. The first model
airplanes capable of sustained flight and powered by twisted strands of rubber were built by French aviation
pioneer Alphonse Pénaud in 1871—more than three decades before the Wright brothers’ first successful
manned flight. In fact, Orville and Wilbur attributed their early interest in aviation to one of Pénaud’s
models, which was given to them by their father in 1878.

From Pénaud’s day to our own, aviation enthusiasts have been experimenting with the design of free-flight
models, determining the proportions and features that result in optimal performance and then passing
these observations along to future generations in the form of empirical design rules, like these:

s/15 0.25–0.4c c

α
CG
1.1 A

0.2–0.25s 0.35–0.4s 1 ¼"


min
0.75s
s 0.4–0.5s
wing aspect ratio (s:c) = 7:1–9:1
A = wing area = s × c
area of horizontal stabilizer = 0.35–0.4A
area of vertical stabilizer = 0.11A
α = incidence angle = 3–4º
dihedral angle = 12º
propeller pitch = 1.3 × diameter
length of rubber motor = 0.75–1.2s

Before we can apply these rules, we must choose a span length, s. From an aerodynamic perspective, larger is
generally better; however, from a practical perspective, a span larger than 3 feet will cause some significant
challenges for storage and transport. Thus, we will use s = 3 feet. Another practical advantage of this choice is
that the balsa wood we will be using to build the model is most commonly sold in 36-inch lengths.

Next, we will define the chord, the width of the wing. Doing so will also define an important aerodynamic
characteristic called the aspect ratio, the ratio of the wing’s span to its chord. The empirical design
rules suggest that the aspect ratio should be between 7:1 and 9:1, and we will use the upper end of this
range—9:1—resulting in a chord that is one-ninth of 36 inches, or 4 inches.

126 
Why this choice? Recall that vortex drag dominates aircraft performance for the low speeds at which our
model will be flying. Vortex drag is the inevitable adverse by-product of lift, so we cannot eliminate it. But
the science of aerodynamics tells us that we can reduce vortex drag somewhat by using a relatively high
aspect ratio. This strategy works, because the vortex phenomenon occurs primarily at the wing tips, and for
a high-aspect-ratio wing, the wing tips constitute a relatively smaller proportion of the total span.

Applying the remaining empirical design rules, the wing area, A, is:

The area of the horizontal stabilizer is specified as 35% to 40% of the wing area. Thus:

The vertical stabilizer should be 11% of the wing area:

The length of the fuselage—the body of the airplane—will be 75% of the span:

The portion of the fuselage extending from the nose rearward to the center of gravity will measure 25% of
the span:

The portion of the fuselage extending from the center of gravity rearward to the rubber peg—the dowel
that anchors the rear end of the rubber motor—will measure 35% of the span:

The height of the wing above the thrust line—an imaginary line corresponding to the axis of the
propeller—is given by:

Project 7 » Rubber-Powered Airplane 127


The angle at which the wing is mounted on the fuselage, called the
incidence angle, will be 3°.

And the angle at which the two halves of the wing are inclined,
called the dihedral angle, is specified as 12°.

Based on these calculated dimensions, the design concept for our


airplane is shown at right.

The physical dimensions of this design reflect the particular importance of aerodynamic stability in a
free-flight model airplane. (See “Aerodynamic Stability” sidebar.)

Aerodynamic Stability
The orientation of an aircraft flying in three-dimensional space is called its attitude and is defined with respect to the
three axes about which an aircraft can rotate: the pitch axis, the yaw axis, and the roll axis.
YAW Aerodynamic stability is the capacity of an aircraft to maintain its
attitude despite disturbances caused by wind gusts, turbulence,
and such. In general, aerodynamic stability is beneficial,
but it is possible to have too much of a good thing. Too
much stability can actually be detrimental to aircraft
performance, because this tendency to resist sudden
changes in attitude can reduce maneuverability. Fighter
planes and aerobatic aircraft are often designed to be
PITCH neutrally stable—meaning that they operate on the
ROLL borderline between stable and unstable—so they can
turn, climb, and dive more responsively. This is one
= AXIS reason why these types of aircraft require highly
= ROTATION ABOUT AXIS trained pilots.

But our model airplane will not have a pilot, so


it must have a high degree of stability. Indeed, many of the physical differences between our model and a full-size piloted
aircraft are directly related to the model’s critical need for more stability.

Aerodynamic stability is defined with respect to the aircraft’s pitch, yaw, and roll axes, and there is a separate
mechanism for achieving stability with respect to each axis.

†† Stability with respect to the pitch axis is provided by the horizontal stabilizer.
†† Stability with respect to the yaw axis is provided by the vertical stabilizer.
†† Stability with respect to the roll axis is provided by dihedral—the upward tilt of the airplane’s wings.

An aircraft has inherently greater roll stability if its wing is positioned above the thrust line. This is why training airplanes,
which must be very stable, either have a high wing with a moderate amount of dihedral or a low wing with much more
dihedral. And this is why highly maneuverable, neutrally stable aircraft typically have very little dihedral or none.

128 
DETAILED DESIGN
Airfoil Selection

In translating this concept into a detailed design, our first and most important task is to choose an airfoil
for the wing. We will use the Göttingen 795 airfoil, pictured below.

GÖTTINGEN 795 AIRFOIL

The two most important characteristics of this airfoil are

}} a flat bottom, which will facilitate ease of construction; and


}} a relatively thin profile, which is particularly suitable for a small wing flying at relatively low speed.

Propellers and Pitch


The purpose of a propeller is to produce the aerodynamic force called thrust.

In essence, a propeller is really just a rotating wing. Thus, the thrust generated by a propeller is really just lift oriented
in the forward direction.

This is a simple concept. However, in practice, the design of an efficient propeller is greatly complicated by the fact that,
unlike a wing, every point along the length of a propeller blade moves through the air at a different speed. To account
for this variation in speed, the incidence angle of a propeller blade must change continuously along its length. To ensure
that all parts of the blade contribute more or less equally to the development of thrust, the fast-moving tip must have a
relatively low angle, while the slower-moving sections closer to the hub must be set at correspondingly higher angles.

Consequently, each blade of a propeller has a distinct twist.

Surprisingly, the propeller’s complex geometry can be fully defined with a single number: the pitch, defined as the
forward distance through which the propeller would advance during one full rotation if it did not slip with respect to
the air. A high-pitch propeller has its blades set at a relatively high incidence angle. A low-pitch propeller has its blades
set at a relatively shallow angle.

The pitch of a propeller has a substantial influence on aircraft performance. In general, low pitch produces better low-
speed acceleration, while high pitch results in more efficient high-speed performance.

Project 7 » Rubber-Powered Airplane 129


Propeller Design

Our model will use a simple two-bladed propeller carved from a block of balsa wood.

Consistent with the empirical design rules defined previously, the propeller diameter will be:

And the pitch is specified as:

These characteristics—a large diameter and a relatively low pitch—reflect the low speed at which our
airplane will fly and the relatively low power output of its twisted rubber motor.

The incidence angle, θ, of the propeller blade tips can be determined from the following calculation:

Having selected an airfoil, designed the propeller, and determined all of the airplane’s key dimensions—
wingspan, aspect ratio, fuselage length, stabilizer dimensions, etc.—we can now complete the detailed
design by configuring its structural framework—the airframe—in a manner that provides an appropriate
balance of strength, serviceability, and light weight.

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering. A full-size layout drawing and cutting pattern are provided in the PDF version of this
guidebook. The layout drawing measures 36 inches by 24 inches and must be printed at exactly this size.
This can be accomplished by a commercial printing establishment for a few dollars. The cutting pattern is
designed to print on a single 8½-by-11-inch sheet and can be done on a home printer with the printing scale
set to 100%.

130 
Drawing 7.1 » Airframe, perspective
view from above left

Drawing 7.2 » Airframe, perspective


view from below right

PROPELLER

NOSE BLOCK

PROPELLER SHAFT

WASHERS
RATCHET
Drawing 7.3 » Propeller
shaft assembly, with THRUST BEARING
BEARING GLASS BEAD
exploded view

Key features:

}} The framework of the airplane is constructed primarily of balsa wood; however, a few components
that require additional strength are made of basswood or thin plywood.
}} The propeller shaft and landing-gear legs are made of 3/64-inch-diameter steel music wire. The rubber
peg is a short length of ¼-inch-diameter aluminum tubing.
}} The adjustable rudder is a piece of aluminum sheet salvaged from a beverage can.

Project 7 » Rubber-Powered Airplane 131


}} The airframe will be covered with lightweight Japanese tissue and sealed with nitrate dope mixed
with thinner.
}} The balsa propeller incorporates a bearing (made of brass tubing) and a ratchet assembly (made of
brass tubing and music wire) that allows the propeller shaft to drive the propeller when the rubber
is wound while also allowing the propeller to spin freely when the airplane is gliding. The latter
behavior—called freewheeling—substantially improves the airplane’s gliding performance by
reducing aerodynamic drag during the descent.
}} As illustrated in drawing 7.3 above, the propeller shaft assembly incorporates a hook at the rear, to hold
the rubber motor; a smaller hook up front, for winding the rubber motor; a nylon thrust bearing,
which allows the shaft to rotate freely in the nose block during powered flight; and a glass bead and two
brass washers, to reduce friction at the interface between the thrust bearing and the propeller.

MATERIALS LIST
SIG
COMPONENT DESCRIPTION PRODUCT
CODE*
Wing leading edge 3/32″ square bass SIGBA304
Wing trailing edge 3/32″ × 1/2″ balsa SIGB014
Wing ribs: R1 (14), R2 (2), R3 (2) 1/16″ balsa sheet SIGB143
Wing tips: W1 (2), W2 (2) 3/32″ balsa sheet SIGB144
Wing spars (2) 1/16″ square bass SIGBA004
Wing spar doublers 1/32″ birch plywood SIGPW004
Wing spar reinforcement 1/32″ balsa sheet SIGB141
All gussets 3/32″ balsa sheet SIGB144
Fuselage longitudinal frame members 3/32″ square bass SIGBA304
Fuselage transverse frame members 3/32″ square balsa SIGB009
Fuselage reinforcement: F4, F5, F7, F8 (2), 3/32″ balsa sheet SIGB144
F9 (2), F10 (2), F11 (2)
Fuselage reinforcement: F1, F6 (2) 1/32″ birch plywood SIGPW004
Wing hold-downs (2) 1/8″ hardwood dowel, 21/4″ long SIGBD026
Wing leading edge retainer 1/4″ square balsa, 11/4″ long SIGB031
Rubber peg 1/4″-diameter aluminum tubing, 11/2″ long **
Nose block: N2, N3, N5 1/8″ balsa sheet SIGB137
Nose block: N1, N4 1/16″ birch plywood SIGPW007
Horizontal stabilizer leading and trailing edges 3/32″ × 1/4″ balsa SIGB012
Horizontal stabilizer interior framework 3/32″ square balsa SIGB009
Horizontal stabilizer: S1, S2, S3 (2), S4 (2) 3/32″ balsa sheet SIGB144
Vertical stabilizer leading and trailing edges 3/32″ × 3/16″ balsa SIGB011

132 
Vertical stabilizer interior framework 3/32″ square balsa SIGB144
Vertical stabilizer: S5 3/32″ balsa sheet SIGB144
Rudder: R1 thin aluminum sheet
Landing gear 3/64″-diameter steel music wire SIGMW002
Wheels (2) 1/32″ birch plywood SIGPW004
Wheel hubs (2) 1/4″ hardwood dowel, 3/8″ long SIGBD029
Propeller 1″ × 3″ × 14″ balsa block SIGB358
Propeller shaft 3/64″ steel music wire SIGMW002
Thrust bearing large EBM thrust bearing **
Propeller bearing 1/16″-diameter brass tubing, 3/4″ long **
Propeller shaft washers Brass thrust washer SIGSH593
Glass bead Bead for propeller bearings SIGSH136
Ratchet 3/64″ steel music wire SIGMW002
Ratchet tube 1/16″-diameter brass tubing, 1/4″ long
Rubber motor Sport rubber, 3/16″ × 25' SIGSR325
Rubber lubricant Rubber lubricant, 4 oz. SIGRL001
Sealant Nitrate dope, pint SIGND001
Thinner Dope thinner, pint SIGSD106
Covering Lite-Flite Japanese tissue SIGPL001
Building board 3/4″ × 3' × 4' plywood or particleboard ***
Building board 1/4″ × 3' × 4' cork sheet ***

*Although these materials are available from many hobby shops, they can be obtained most reliably from online sources. A
particularly good one is the SIG Manufacturing Company (http://www.sigmfg.com/). All items for which a SIG Product
Code is listed can be obtained from this source.

**These items can be obtained from Easy Built Models (www.easybuiltmodels.com).

***These items can be obtained from most building supply stores.

The other supplies needed for this project are:

}} wood glue
}} cyanoacrylate glue
}} medium- and fine-grit sandpaper
}} wax paper
}} sewing pins
}} rubber bands
}} clothespins (for clamping)

The tools required for this project are:

}} single-edge razor blade or hobby knife


}} carving knife

Project 7 » Rubber-Powered Airplane 133


}} drill
}} razor saw or coping saw
}} assortment of small files
}} needle-nose pliers with wire cutter
}} metal straightedge

A power scroll saw and drill press are helpful—but not essential—for a few tasks in this project.

FABRICATION AND ASSEMBLY PROCEDURE

1 Begin by prefabricating all components made from balsa wood and plywood sheets. Print the full-size
cutting pattern provided in the PDF version of this guidebook. Use a light coat of spray adhesive to
adhere these patterns to wood of the correct type (balsa or plywood) and the correct thickness (1/16
inch, 3/32 inch, or 1/8 inch), as indicated on the pattern. Be sure that the pattern is properly aligned with
the grain direction, as indicated.

2 Very carefully cut out each piece with a hobby knife, razor blade, or razor saw; then, peel off the paper
pattern. If necessary, use a coping saw or scroll saw for the plywood pieces.

3 To fabricate the 16 identical balsa wing ribs (designated as R1 and R2), begin by cutting out two
identical wing rib templates from 1/16-inch plywood. Trace one of these templates 16 times onto a
1/16-inch balsa sheet; then, cut out the ribs slightly oversize. (See A below.) Sandwich these balsa ribs
between the two plywood
templates, using a light 3
coat of spray adhesive to
hold them together. (See
B.) Shape the entire stack
with a sheet of sandpaper
placed on a flat surface
until all of the balsa ribs
match the templates
(A) Cut ribs slightly oversize. (B) Sandwich ribs between templates.
exactly. (See C.) Use a
small file or razor blade
to cut the notches for the
wing spars and leading
edge. Finally, use lacquer
thinner to separate the
stack and clean off any
adhesive residue. (See D.)

(C) Sand ribs to shape and cut notches. (D) Separate the stack.

134 
4 To create the R2 ribs, cut off the rear half of
R1
two R1 ribs, as shown below.

R2

R1 4

R2

SANDING BLOCK 5 Shape the wing trailing


edges by sanding a
5 3/32-inch-by-½-inch balsa
strip to a triangular
cross-section using the
simple jig shown below
1⁄ " HARDWOOD DOWEL
8 and a piece of medium-
GLUED TO WOODEN BASE grit sandpaper wrapped
around a wooden block.
⁄32" × 1⁄2" BALSA TRAILING EDGE
3

6 The model will be assembled on a


building board consisting of a plywood WAX PAPER
or particleboard panel with a thin
sheet of cork glued to its top surface.
The full-size layout drawing is placed DRAWING
directly onto the building board,
followed by a sheet of wax paper to
prevent glue from sticking to the
CORK
plan. All wooden components of the
PLYWOOD OR
airframe should be assembled with PARTICLEBOARD
wood glue.

Project 7 » Rubber-Powered Airplane 135


7

RIB
L-SHAPED SPAR
BLOCK
TRAILING
EDGE
LEADING EDGE WINGTIP

7 The wing is constructed in three separate sections. For each outer wing panel, use a razor blade to cut
notches in the trailing edge at each wing rib location, and then pin the trailing edge to the board. Note
that the trailing edge need not be trimmed to its final length at this time. Next, glue and pin the wing
tip pieces (W1 and W2) in place, followed by the lower spar, and then the ribs (R1 and R2), and then
the leading edge. Rather than pinning through the leading edge (which would split the wood), use
L-shaped balsa blocks to hold it in position without doing any damage.

8 On each wing panel, glue the upper spar into the slots on top of the wing ribs. The top and bottom
spars are then reinforced with six 27/16-inch-by-5/16-inch rectangles cut from a 1/32-inch balsa sheet
and glued in place at the positions shown below. Finally, add small triangular pieces of 1/16-inch balsa
sheet—called gussets—to reinforce the connection of the leading edge to rib R2.

GUSSET

SPAR REINFORCEMENT

UPPER SPAR

136 
PLYWOOD DOUBLER

9 COMPONENT W3
PLYWOOD DOUBLER

GUSSET

9 Next, construct the center panel of the wing. Note that it incorporates two plywood doublers and
another pair of gussets, which will reinforce breaks in the continuity of the leading edge, spars, and
trailing edge at the points where the three wing panels are joined together. Note, also, that the two
wing ribs in the center panel are angled inward at precisely 12°, to accommodate the wing’s dihedral
angle. The trapezoidal balsa component W3 ensures that this angle is correct.

10 The two identical sides of the fuselage are also built directly on the board, using 3/32-inch-square
basswood strips for the longitudinal elements, 3/32-inch-square balsa for the remaining frame pieces,
sheet-balsa and plywood inserts (F5, F6, and F10) to strengthen the nose and rubber peg mount, and
additional inserts (F8 and F9) and gussets to increase the rigidity of the wing mount. When fastening
3/32-inch strips to the building board, do not pin through the wood; rather, use pairs of pins placed on
both sides of each strip, as shown below.

10
F5 & F6

F9

F8

F10

Project 7 » Rubber-Powered Airplane 137


11 The horizontal and vertical
stabilizers are flat balsa frameworks
that are constructed directly over
the plans, just like the fuselage.

11

12 The wing is assembled by gluing the two


outer panels to the center panel, with each
tip blocked up exactly 33/4 inches to provide
the required 12° dihedral angle.

12
3¾"

13 To assemble the fuselage, start by building the


two perpendicular frames, F2 and F3. Then, glue
the two fuselage halves to these frames. After
checking to ensure that the assembly is square,
set it aside to dry thoroughly.
F3

F2

13

138 
F7

14 Glue and clamp the two fuselage halves together at the tail,
and glue component F1 to the nose. Then, to complete the
framework, add the nose reinforcement (F11, both top
and bottom), the tail reinforcement (F7), and the
transverse 3/32-inch-square balsa frame pieces,
using a rubber band to clamp the joints 14
while the glue is setting up.
TRANSVERSE FRAME PIECES
F11

F1 ⁄8" DOWEL
1

WING RETAINER BLOCK


15 The wing hold-down
system includes two 1/8-inch ⁄8" DOWEL
1

hardwood dowels mounted in F9


components F8 and F9 and a
wing retainer block, which is
cut from ¼-inch-square balsa
carved to a triangular cross- 15
F8
section and glued in place
immediately above F8.

F4
F4

⁄64" MUSIC WIRE


3

16 The landing gear is a piece of 3/64-inch-


diameter steel music wire bent into the shape
shown on the plan and then sandwiched
16 between the two balsa mounts (F4). The
wheels (L1) are 1/32-inch plywood, with
hubs made from short lengths of ¼-inch
L1
hardwood dowel, each drilled with a 1/16-inch
⁄4" DOWEL HUB
1 hole to accommodate the landing-gear wire.

L1

Project 7 » Rubber-Powered Airplane 139


17 The complete landing-gear assembly is glued into
the fuselage and reinforced with a pair of gussets.

18 The nose block is built up from five prefabricated


layers: three balsa (N2, N3, and N5) and two
plywood (N1 and N4). After gluing these
laminations together and drilling a ¼-inch hole to 17
accommodate the propeller thrust bearing, sand
the block to a streamlined shape.

SLOT FOR RUBBER BAND


N1
1⁄4" HOLE
18
N2

N3

N4

N5

19 To prepare the airframe for covering, use sandpaper to round the edges of the wing tips, wing leading
edge, and stabilizers.

20 The airframe will be covered with lightweight water-shrinkable Japanese tissue adhered and sealed with
a mixture of 50% nitrate dope and 50% dope thinner. This substance is quite noxious, so be sure to do
this work in a well-ventilated space. The covering process is summarized below and is demonstrated in
lesson 10. For each component of the airframe:

}} Apply three coats of the 50-50 dope-thinner mixture to the outer edges, where the tissue will be adhered.
}} Cut a piece of tissue slightly larger than the surface to be covered. When covering the wing and
stabilizers, the grain of the tissue should be oriented parallel to the leading and trailing edges; when
covering the fuselage, the grain should be oriented longitudinally.
}} Lay the tissue in position. Brush dope thinner around its outer edge and press the tissue in place,
smoothing and stretching it to remove as many wrinkles as possible. As the thinner soaks in, it will
soften up the dope we previously applied to the framework and will bond the tissue in place.
}} After the thinner has dried, use an emery board or sanding block to trim off the excess tissue.

140 
}} For two-sided components (the wing and stabilizers), the opposite side of the frame is covered in the
same manner, except that the tissue should be trimmed about 1/8 inch oversize. The edge of the tissue
is then wrapped around the frame and adhered to the previously applied tissue with dope.
}} Use a spray bottle to apply a light mist of water to the tissue. As it dries, the tissue will shrink tight.

21 Once all components of the airframe have been covered, glue the horizontal
and vertical stabilizers to the fuselage. The adjustable rudder (component
R1) is cut from a piece of aluminum salvaged from a beverage can and RUDDER (R1)

glued into slots in the vertical stabilizer with VERTICAL


STABILIZER
cyanoacrylate glue.

22 Seal the tissue covering by painting the


entire model with two coats of the 50-50
dope-thinner mixture. You might also
HORIZONTAL
want to dress up the model by painting STABILIZER

the nose block and adding some


simulated windows. But don’t overdo it!
Any added weight will adversely affect
21
the performance of the airplane.

23 Prepare to fabricate the propeller by marking


23 a 1-inch-thick balsa block as shown below.
Its length corresponds to the diameter of the
propeller—13.2 inches—and its width is
1/16" hole based on the need for a 22.5° angle at the
blade tip, as calculated above. Once this
layout is complete, cut out the
½"
bow-tie shape and drill a
13 3/16"
1/16-inch-diameter hole
precisely at the center
1" to accommodate the
° propeller shaft.
22.5

2 7/16"

24
24 Use a sharp knife to carve away all of the material above
the diagonal along the length of one blade. This process is
demonstrated in lesson 11. Note that the original bow-tie
shape of the block guarantees the correct linear variation in
blade angle from 22.5° at the tip to 90° at the hub.

Project 7 » Rubber-Powered Airplane 141


25 Use the same procedure to carve the front side of the
opposite blade and the back of both blades, leaving a
uniform thickness of approximately 1/8 inch between the
two carved faces.

26 The blades are then tapered and sanded to an airfoil 25


shape, as demonstrated in lesson 11.

27 Fabricate the propeller bearing


RATCHET
from a piece of 1/16-inch-
diameter brass tubing and the
PROPELLER ratchet assembly from 1/16-inch
HUB brass tubing and 3/64-inch
music wire, as shown below.
Cut a slot for the ratchet into
the front face of the propeller
PROPELLER hub and then install both
BEARING
the bearing and ratchet with
cyanoacrylate glue. After
27 installation, ensure that the
ratchet can rotate freely.

28 Balance the propeller by inserting a piece of wire into the bearing and then lightly sanding the heavier
blade until the propeller remains in a horizontal orientation when released.

29 Fabricate the propeller shaft assembly from a 4-inch length of 3/64-inch music wire, a glass bead, two
brass washers, and a nylon thrust bearing, as illustrated in drawing 7.3 above. Use needle-nose pliers to
bend the rear hook in the propeller shaft; then, assemble all components, including the propeller and
nose block. Finally, bend the front hook and cut off any excess wire.

30 To prepare the rubber motor, drive two nails into a board 56 inches apart, wrap two complete loops of
3/16-inch rubber (4 strands) loosely around these nails, and tie the two free ends together tightly. These
four 56-inch strands will be doubled to form a rubber motor of eight 28-inch strands. (See “Designing
the Rubber Motor” sidebar for the basis of this design.)

142 
Designing the Rubber Motor
The rubber motor is designed after the airframe has been completely built, because the required amount
of rubber depends on the weight of the aircraft.

For optimal performance, the rubber motor should weigh about 35% to 40% as much as the airplane
itself (without rubber), and its length should be between 75% and 120% of the wingspan. A longer
motor can safely accommodate more twists and thus can provide longer flight duration. However, a
longer motor can also adversely affect the airplane’s glide performance, because the excess length of the
unwound rubber can shift to the front or rear of the fuselage and thus adversely affect the airplane’s
balance.

We will use a motor length of 28 inches (78% of the wingspan). Our completed airframe weighs 76
grams. High-quality synthetic rubber for DIY airplanes is sold in a variety of different widths, and the
standard 3/16-inch width is well suited for airplanes of this size. Thus, the amount of rubber required to
power our model can be calculated as follows:

We will use 8 strands of 3/16-inch rubber.

31 Before it is installed in the airplane, the rubber motor must be broken in and lubricated. Failure to do
so will significantly reduce both the longevity and performance of the motor. Rubber lubricant can be
purchased from model airplane supply vendors such as SIG Manufacturing, as noted in the materials
list above. The simple two-step procedure for break-in and lubrication is as follows:

}} Stretch the rubber motor to approximately three times its unstretched length. While holding it in
this position, apply rubber lubricant along the full length of the motor. After two minutes, release the
rubber and allow it to rest for 15 minutes.
}} Stretch the rubber to four times its unstretched length, hold it in this position for four minutes, and
then release.

Project 7 » Rubber-Powered Airplane 143


32 The rubber motor is installed in the airplane using a forked stick like the one pictured below. The
procedure is demonstrated in lesson 11.

RUBBER MOTOR
RUBBER PEG

32

33 The final step in the assembly process is to attach the airplane’s wing to the fuselage with rubber bands
to provide for further adjustability and improved crashworthiness.

The rubber-powered airplane is now complete and ready for testing!

TESTING PROCEDURE

Before the airplane can be flight tested, it must be balanced.

The term balance refers to the location of the airplane’s center of gravity, also called the CG or balance
point. Balance is closely related to stability. In general, the farther forward the CG is located, the more stable
the airplane will be about its pitch axis. An aircraft with its CG too far aft can be essentially uncontrollable.

A standard rule of thumb for free-flight models is that the CG (or balance point) should be located at
approximately one-third of the chord, measured rearward from the wing’s leading edge. If the airplane does
not hang level when suspended from this point, add weight to the nose block until it does.

To fly, this model safely requires a large open area that is free of obstructions, such as buildings, trees, and
power lines. For testing, in particular, it is preferable to choose a time with little or no wind. Early morning
and just before sunset are often the best times of day for calm, stable air.

Begin by trimming the aircraft for optimum glide performance. If there is any wind at all, aim directly into
it and then give the airplane a firm toss at a slight downward angle. Assess the test and make adjustments
as follows:

144 
}} The propeller should spin freely throughout the flight.
If not, adjust or lubricate the ratchet mechanism until it
Trimming
operates properly.
The process of adjusting the
}} The airplane should fly reasonably straight. If it exhibits
flight characteristics of an
a strong tendency to turn, adjust the rudder accordingly. aircraft is called trimming.
}} The airplane should descend at a slow, steady rate. If the Careful, methodical trimming is
descent is too steep, increase the wing’s incidence angle essential for successful free flight.
by adding a shim underneath the leading edge. If the Our model has four principal
airplane stalls, add a shim under the trailing edge. mechanisms for trimming:

†† The rudder is adjustable.


Repeat the glide test and continue making small adjustments Setting the rudder to the
until satisfactory gliding performance is achieved. left or right will cause the
airplane to turn (yaw and
Next, do a low-power test. Our rubber motor—8 strands of roll) in that direction.
3/16-inch rubber—should be able to hold about 1,200 twists †† Because the wing is held in
place with rubber bands,
before it reaches the breaking point. For this first test, use
its incidence angle can
approximately 20% of the maximum—about 240 turns—so be changed by inserting
that any problems with the airplane’s powered performance a balsa shim under the
can be corrected before they do serious damage. leading edge or trailing
edge. This is a pitch
Winding a rubber motor effectively and safely requires more adjustment. Increasing
than two hands. The easiest way to overcome this limitation the incidence angle will
cause the airplane’s rate of
is to have a friend hold the airplane, resisting the pull of the climb to increase or its rate
stretched rubber motor by gripping the rubber peg at the rear of descent to decrease—
of the fuselage. If no helpers are available, then the simple but only to a point. An
wooden cradle pictured below will do the job. Note that the excessively high incidence
airplane is connected to the cradle by a 1/8-inch steel pin that angle can cause stalling.
†† Because the nose block is
passes through the rubber peg. (This is why the rubber peg is
held in place with a rubber
an aluminum tube, rather than a solid wooden dowel.) Note band, the orientation of the
also that the cradle must be anchored with heavy nylon cord thrust line can be changed
to an immovable object, by shimming the nose
such as a car, tree, or block downward, upward,
fence post. or to one side. These
adjustments to the thrust
line are only effective
when the airplane is under
power. They do not affect
STEEL PIN
its gliding performance.
†† The power output and the
duration of the powered
flight can be controlled
NYLON CORD through the number of
twists applied to the rubber
motor prior to each flight.

Project 7 » Rubber-Powered Airplane 145


The procedure for winding the rubber motor is demonstrated in lesson 11. The motor is stretched to
approximately three times its unstretched length and then a mechanical device is used to apply the
appropriate number of turns to the hook at the front end of the propeller shaft.

Specially designed hand-operated rubber-winding devices can be purchased from model airplane supply
vendors; however, a common hand drill works just as well, as long as the winding hook (made from a piece
of heavy steel wire) is very firmly gripped in the drill chuck.

Regardless of which type of winding device is used, it is necessary to know the gear ratio of the machine.
The hand drill used for the flight tests demonstrated in lesson 11 has a gear ratio of 3.5 to 1, which means
that each turn of the crank produces 3½ turns of the hook. Thus, applying 240 twists to the rubber motor
requires only about 70 turns of the hand drill.

With the motor wound, the nose block is replaced in the fuselage, the ratchet is engaged with the propeller
shaft, the winder is removed, and the nose block is secured to the fuselage with a rubber band.

To initiate the test flight, aim directly into the wind, release the propeller and allow it to spin for a second
or two, and then launch the airplane at a slight upward angle.

Assess the test and make adjustments, as follows:

}} If the airplane stalls, angle the thrust line downward by adding a balsa shim at the top of the nose block.
}} If the airplane turns too sharply in one direction, angle the thrust line in the opposite direction by
adding a shim to the side of the nose block. (It is quite likely that the aircraft will turn sharply to the
left due to torque effect. See “Torque Effect” sidebar.)
}} If the airplane fails to climb, angle the thrust line upward by adding a balsa shim at the bottom of the
nose block.
}} To avoid losing the airplane on high-powered flights, it is generally desirable to trim for a gradual
turn while gliding. This result is best achieved with rudder adjustments.

After consistently satisfactory results have been attained in low-power tests, gradually increase the number
of twists in the rubber motor, making continued trimming adjustments as needed to achieve the desired
flight path and maximum flight duration. This airplane should be able to achieve flight durations of two
minutes or more.

Torque Effect
Torque effect is a manifestation of Newton’s third law: For every action, there is an equal and opposite
reaction. As the rubber motor applies a clockwise torque to a right-handed propeller, the propeller applies
a counterclockwise torque to the aircraft, resulting in an inherent tendency to turn left under power. As
the level of power increases, this turning tendency increases as well. If this tendency is not corrected by
adjusting the orientation of the thrust line, the dreaded “torque-induced death spiral” may result.

146 
PROJECT 8
RUBBER-
POWERED
HELICOPTER
BUILD AND TEST A FREE-FLIGHT, RUBBER-
POWERED MODEL HELICOPTER
LESSON IN WHICH THIS PROJECT IS COVERED:
12 » Build a Model Helicopter

PROBLEM DEFINITION
Requirements:
}} Must be capable of sustained, stable flight.
}} Must be capable of varied flight characteristics through initial rotor orientation.

Constraints:
}} Powered only by twisted rubber.
}} No form of remote control.
}} Use only readily available materials and components.

Project 8 » Rubber-Powered Helicopter 147


REFERENCES
Blattenberger, “Penni Helicopter.”
Ross, Rubber Powered Model Airplanes.
Simons, Model Aircraft Aerodynamics.

DESIGN CONCEPT

The format of this project differs somewhat from the other projects in this course. In this project, we will not
develop an original design; rather, we will explore helicopter aeronautics by building and flying a modified
version of a time-tested model helicopter design: the Penni helicopter, illustrated in drawing 8.1 on page 153.

Penni is a small free-flight, rubber-powered helicopter designed in the late 1960s by John Burkham, an
accomplished aeronautical engineer who worked for the Boeing Vertol company and was involved in
designing some of the world’s most advanced helicopters at that time. Throughout the 1960s, Burkham
also did pioneering work in the field of model helicopter design, and Penni is a unique product of that
work. Penni was the name of Burkham’s youngest daughter, and the design that bears her name was first
published in the January 1970 issue of American Aircraft Modeler magazine. This publication has long since
gone out of print, and Burkham died in 1999, yet the Penni helicopter lives on through the Internet and,
indeed, has become something of a classic in DIY circles. In this project, our adaptation of John Burkham’s
creation will serve as a tribute to this great engineer, who so generously shared his ingenuity and creative
spirit with the world.

Despite its small size (16-inch main-rotor diameter) and light weight (approximately 15 grams), Penni
replicates the mechanical functions and flight characteristics of a full-size helicopter with surprising
authenticity. To achieve this authenticity, however, the project demands a high level of care and precision
that is guaranteed to test your DIY skills.

Introduction to Helicopter Aeronautics


Real-world helicopters are designed in a variety of configurations, but for this project, we will focus on the most
common one—characterized by a single horizontally oriented main rotor on top and a small vertically oriented tail
rotor at the rear. This type of aircraft is often used for TV news and traffic reporting, law enforcement, medical
evacuation, and military troop transport.

Like all powered aircraft, a helicopter flies by achieving a balance of lift and weight, thrust and drag. The big difference
between a helicopter and a conventional airplane is that the helicopter uses one or more main rotors to generate both
lift and thrust.

148 
The main rotor is really just a rotating wing. This is why helicopters are often categorized as rotary-wing aircraft.
Like a conventional airplane wing, the rotor has an airfoil set at an angle of attack such that the desired amount of lift
is produced. Unlike an airplane wing, the airflow that generates lift is caused primarily by the rotor’s rotation rather
than by forward motion of the aircraft.

In this latter sense, a helicopter rotor is also a propeller that rotates on a vertical shaft rather than a horizontal one. Like
any propeller, the rotor develops an aerodynamic force, which can be represented as a vector oriented perpendicular to
the rotor disk, an imaginary circular disk defined by the sweep of the spinning blades.

A helicopter’s movement through the air is determined largely by the magnitude and direction of this aerodynamic
force. The pilot controls the magnitude of this force by adjusting the rotor blades’ angle of attack, more commonly
called the pitch. At a given rotor speed, higher pitch produces more lift.
LIFT
The pilot controls the direction of the main-rotor
force by tilting the rotor disk forward, rearward, or to
either side.

When the main-rotor disk is horizontal, its resulting


force vector is oriented straight upward. In this
orientation, the rotor produces all lift and zero thrust.
Consequently, the aircraft does not move horizontally
with respect to the air mass in which it is flying. If
the magnitude of the lift force is equal to the weight
of the aircraft, it hovers; if lift exceeds weight, it
ascends; and if lift is less than weight, it descends.

To move the aircraft forward, the pilot tilts the rotor WEIGHT
disk forward, thus reorienting its force vector as
shown below. If this vector is resolved into its horizontal and vertical components, the large vertical component still
supports the helicopter’s weight, but now the smaller horizontal component provides thrust, which accelerates the
aircraft forward until the drag force becomes large enough to restore equilibrium. Thereafter, the helicopter continues
to move forward at constant velocity. The system works the same way for rearward and sideways motion.

LIFT = vertical component of rotor force

horizontal component of rotor force = THRUST

DRAG

WEIGHT

Continued on next page…

Project 8 » Rubber-Powered Helicopter 149


The great advantage of a helicopter over a conventional airplane is that the helicopter’s main rotor is capable of
providing the required lift without the vehicle having to move forward through the air. But this advantage comes at
a cost: the need to compensate for torque effect. This phenomenon is a manifestation of Newton’s third law. As the
helicopter’s motor delivers torque (or twisting force) to the rotor, the rotor applies an equal and opposite torque to the
helicopter’s fuselage, causing it to rotate in the opposite direction.

All propeller-driven aircraft experience torque effect, but it has a much greater influence on helicopters than on fixed-wing
aircraft for several reasons. Most importantly, to generate enough lift to fly, a fixed-wing aircraft must move through the
air at a relatively high speed, thus allowing torque effect to be counteracted aerodynamically with very small adjustments
to the ailerons or rudder. Conversely, a helicopter must be capable of flying with zero forward speed; thus, torque effect
cannot be counterbalanced by the aerodynamic forces associated with air flowing across control surfaces.

The principal mechanism for counteracting torque effect on most single-main-rotor helicopters is the tail rotor, also
called the antitorque rotor for this reason. When the tail rotor spins, it generates an aerodynamic force that causes a
moment about the main-rotor shaft. This moment counteracts the torque produced by the main rotor.

In most real-world helicopters, the tail rotor is driven by a system of rigid shafts and gearboxes that links the rotor to
the engine. But our model will use a belt drive and pulley system, which is much lighter and less mechanically complex.

Main Rotor Operation and Control


The main rotor of a helicopter consists of two or more blades mounted on a hub. There are several different types of
helicopter rotors, but for our purposes in this project, we need only examine one: the relatively simple semirigid rotor,
commonly used on helicopters with just two rotor blades.

At the heart of any rotor is its hub, which transmits torque from the rotor shaft to the blades and plays a major role in
both the control and stability of the aircraft. The hub of a semirigid rotor is designed to facilitate the rotation of its two
blades in three different modes:

†† It guides the blades in their circular path around the main-rotor shaft.
†† It allows each blade to be rotated independently around its own longitudinal axis. This rotation, called
feathering, changes the blade’s pitch and therefore the amount of lift it produces. Feathering is controlled by
pitch linkages, which move up or down in response to the pilot’s control inputs.
†† The hub is connected to the shaft with a single horizontal hinge, which allows the entire rotor to tip like a
seesaw. This rotation is called teetering, and because it is uniquely associated with the semirigid rotor, this
configuration is often called a teetering rotor. Teetering is not directly controlled by the pilot. It simply
happens in response to aerodynamic forces, in a manner that facilitates both control and stability.

Some semirigid rotors also incorporate a stabilizer bar, consisting of two weights fixed to the ends of a metal bar
mounted perpendicular to the blades. The purpose of the stabilizer bar is to hold the rotor in its current plane of
rotation, resisting perturbations caused by wind gusts or other disturbances.

As noted above, a helicopter’s motion through the air is controlled principally by changing the magnitude and direction
of the aerodynamic force produced by the main rotor. The pilot uses two cockpit controls to make these changes.
The first, called the collective pitch, is a floor-mounted lever that is normally located on the pilot’s left side. True to

150 
its name, the collective pitch varies the pitch of both rotor blades collectively. As the lever is raised, the pitch of both
blades increases equally, thus causing the aircraft to ascend. Conversely, as the collective lever is lowered, the pitch of
both blades decreases, and the aircraft descends.

The second main-rotor control is the cyclic pitch, which usually takes the form of a joystick. Again, true to its name,
the cyclic pitch varies the pitch of each individual rotor blade cyclically such that each blade’s pitch depends on its
current rotational position. For example, when the pilot pushes the cyclic control forward, each blade increases pitch as
it passes through the rear sector of the rotor disk and then decreases pitch as it passes through the front sector. When
the blade’s pitch increases, the blade flies higher; when the pitch decreases, the blade flies lower, causing a teetering
rotation of the entire rotor. Consequently, the rotor disk tilts forward, and the resulting change in the direction of the
aerodynamic force vector causes a thrust component that propels the aircraft forward.

Similarly, if the pilot moves the cyclic control to the left, each blade increases its pitch as it flies through the right sector
of the rotor disk and decreases its pitch through the left sector. In response, the blade teeters, causing the rotor disk to
tilt leftward, and the aircraft moves to the left.

How are the pilot’s two control inputs—collective and cyclic pitch—actually translated into these very complex
feathering rotations of the blades? The answer lies with an ingenious mechanism called the swash plate. At the heart
of this device are two disks, mounted concentrically with the main-rotor shaft. The lower disk does not rotate and is
connected to the collective and cyclic pitch controls through a series of pushrods. The upper disk rotates along with the
rotor shaft and is connected to the blades by means of two pitch linkages.

The two disks of the swash plate are interconnected so that they move as a single unit. When the pilot moves the
collective pitch control, the entire swash plate slides up or down on the rotor shaft, thus causing the desired equal pitch
change in both blades. When the pilot moves the cyclic pitch control, the swash plate tilts in the same orientation as
the desired tilt of the rotor disk, and the pitch linkages follow along, causing the desired cyclic feathering rotations of
the blades.

Through this mechanism, the rotor hub controls both the magnitude and direction of the aerodynamic force produced
by the rotor. The hub must also compensate for an adverse phenomenon called the dissymmetry of lift. To understand
this phenomenon, assume that the tips of a hovering helicopter’s rotor blades are moving through the air at 500 miles
per hour. Then the aircraft begins moving forward, eventually reaching an airspeed of 100 miles per hour.

HOVERING FORWARD FLIGHT


500 MPH 400 MPH

RETREATING
BLADE

100 MPH

ADVANCING
BLADE

500 MPH 600 MPH

Continued on next page…

Project 8 » Rubber-Powered Helicopter 151


Note that, at any given time, there is always one rotor blade moving in the same direction as the aircraft and one moving in
the opposite direction. These are called the advancing blade and the retreating blade, respectively. Due to the combined
effects of rotor rotation and forward motion, the tip of the advancing blade reaches a maximum speed of 600 miles per
hour, while the tip of the retreating blade slows to 400 miles per hour. Consequently, the advancing blade develops more lift
than the retreating blade, causing a potentially strong tendency to roll in the direction of the retreating blade.

This is the dissymmetry of lift. It can be counteracted, in part, by the pilot applying cyclic pitch to compensate for the
rolling tendency. But the semirigid rotor also has an important feature that compensates for the dissymmetry of lift
automatically. This feature is teetering.

Because the advancing blade always generates more lift than the retreating blade, the rotor teeters upward on the advancing
side and downward on the retreating side. This vertical motion changes the relative wind experienced by each blade.

Relative wind is the airflow “felt” by the rotor blade as it moves through the air. As the advancing blade moves forward
through the air, it “feels” a component of relative wind flowing rearward across it. But because of the teetering motion,
the advancing blade is also moving rapidly upward; thus, this blade also “feels” a downward component of relative
wind. Added together, these two components result in a total relative wind that is angled downward, as shown below.
Consequently, the advancing blade has a significantly lower effective angle of attack than it would have if it did not
teeter. A lower angle of attack means reduced lift.

ADVANCING BLADE RETREATING BLADE


ACTUAL MOTION ACTUAL MOTION
TEETERING
EFFECTIVE
BLADE BLADE ANGLE OF
ROTATION ROTATION ATTACK

ROTOR BLADE EFFECTIVE


ANGLE OF
ATTACK
TEETERING

RELATIVE WIND RELATIVE WIND

EFFECT OF BLADE ROTATION TOT


AL
REL
EFFECT OF ATI
VE
D TEETERING WIN
EFFECT OF WIN D
TEETERING IVE
RELAT
TOTAL EFFECT OF BLADE ROTATION

On the other side of the rotor, the retreating blade is teetering downward, and the resulting upward component of
relative wind causes the effective angle of attack—and therefore the lift—to increase.

In this way, the teetering motion caused by the dissymmetry of lift actually compensates for the dissymmetry of lift,
at least in part, by increasing lift on the retreating blade and decreasing lift on the advancing blade. This corrective
mechanism happens automatically, with no intervention by the pilot.

The semirigid rotor and its associated control mechanisms constitute a marvelously sophisticated mechanical system,
many aspects of which will be replicated in our model.

152 
FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/. A full-size layout drawing, which includes all required cutting patterns, is provided in
the PDF version of this guidebook. This drawing measures 24 inches by 16 inches and must be printed at
exactly this size. Most commercial printing establishments will do this job for just a few dollars.

STABILIZER BAR
MAIN ROTOR HUB
MAIN ROTOR

REAR PULLEY
BELT DRIVE
MAIN ROTOR SHAFT

TAIL BOOM FORWARD PULLEY

TAIL ROTOR FUSELAGE


RUBBER MOTOR

Drawing 8.1 » Helicopter


model, major components
LANDING GEAR

EYE

HINGE PIN
MAIN ROTOR

MAIN ROTOR HUB Drawing 8.2 » Helicopter


model, Main rotor
MAIN ROTOR SHAFT shaft and hub
STABILIZER BAR

BELT DRIVE

FORWARD PULLEY

Project 8 » Rubber-Powered Helicopter 153


Key features:

}} The semirigid main rotor incorporates a stabilizer bar.


}} The rotor hub is connected to the main-rotor shaft by a hinge consisting of a hinge pin passing
loosely though a diamond-shaped eye at the top of the shaft. This hinge configuration allows the hub
to pivot freely about both the feathering and teetering axes.
}} These rotations work in conjunction with the stabilizer bar and a 5° forward sweep of the blades to
stabilize the rotor system and provide automatic compensation for the dissymmetry of lift.
}} The rotor blades are permanently attached to the hub; thus, there is no mechanism for collective pitch
adjustment. The fixed-blade pitch is designed to provide a slow ascent during the peak power output
and a steady descent as the rubber winds down.
}} The fixed-pitch tail rotor is coupled to the main rotor by two pulleys and a belt drive made from thread.
}} Power is provided by 4 to 6 strands of 1/8-inch rubber.

MATERIALS LIST
SIG
MATERIAL USED FOR PRODUCT
CODE*
3/32″ × 2″ × 36″ balsa sheet Pulleys, tail-rotor hub SIGB136
1/32″ × 2″ × 36″ balsa sheet Main-rotor blades, tail-rotor blades, gussets SIGB134
1/8″ × 2″ × 36″ balsa sheet Fuselage, main-rotor hub SIGB137
1/16″ square bamboo Landing-gear skids **
3/64″ steel music wire Main-rotor shaft SIGMW002
1/32″ steel music wire Fuselage front, landing-gear legs, lower rubber hook, SIGMW001
main-rotor hinge pin, tail-rotor shaft, stabilizer bar
Thin aluminum sheet Washers, bearings ***
(approx. 1 square inch)
0.02″ brass sheet (5/16″ × 5/16″) Washer KS16405
#22 copper wire (approx. 1″) Main-rotor pulley-to-shaft connection SIGSH330
3/32″-diameter solder Stabilizer bar ****
Glass bead Main-rotor bearing SIGSH136
Sport rubber (1/8″ × 25') Rubber motor SIGSR825
*These materials can be obtained most reliably from online sources. A particularly good one is the SIG Manufacturing
Company (http://www.sigmfg.com/). All items for which a SIG product code is listed can be obtained from this source.

**If bamboo is not available, use 1/32″ music wire.

***Can be salvaged from a used beverage can.

****Can be obtained from a building supply or hardware store.

154 
The other supplies needed for this project are:

}} wood glue
}} epoxy glue
}} medium- and fine-grit sandpaper
}} masking tape

The tools required for this project are:

}} single-edge razor blade or hobby knife


}} drill
}} metal shears or heavy-duty scissors
}} needle-nose pliers with wire cutter

FABRICATION AND ASSEMBLY PROCEDURE

1 Begin by building the


fuselage framework
from 1/8-inch-by-3/16-inch GUSSET
balsa wood strips, with
connections reinforced by FUSELAGE FRAMEWORK
1/32-inch-thick balsa gussets. NOSE
The landing-gear legs and
nose are bent from 1/32- 1
inch music wire, following
the patterns provided on
the project drawing. The
landing-gear skids are thin RUBBER HOOK
strips of bamboo salvaged
from an old leaf rake. If LANDING-GEAR LEG
bamboo is not available,
use 1/32-inch music wire for
the skids. The wire hook
at the base of the fuselage
LANDING-GEAR SKID
framework will provide the
lower anchorage for the
rubber motor.

Project 8 » Rubber-Powered Helicopter 155


2 2 The assembled fuselage is
shown below. All wood-to-metal
THREAD connections are reinforced by
wrapping with thread and coating
with glue.

3 Prefabricate the following washers and bearings:

NUMBER HOLE
SIZE MATERIAL PURPOSE
REQ’D DIAMETER
1 1/8″ × 1/8″ 3/64″ aluminum main-rotor-shaft washer
4 1/8″ × 1/8″ 1/32″ aluminum tail-rotor-shaft washers
2 1/8″ × 1/8″ 1/32″ aluminum main-rotor-hub washers
2 1/8″ × 1/4″ 3/64″ aluminum main-rotor bearings
1 5/16″ × 5/16″ 3/64″ brass forward-pulley mount

Begin this process by marking the outlines and hole locations on the appropriate metal sheet. Drill
all holes first; then, cut out each piece with a metal shears or heavy-duty scissors. If you do not have
drill bits of the correct size, the holes can be drilled with a short piece of sharpened music wire of the
appropriate diameter.

Mark and drill... ...then cut.

156 
3
⁄64" HOLE

4 Drill a 3/64-inch-diameter hole


vertically through the tail boom for
the main-rotor shaft.

⁄64" WIRE
3

⁄8" × 1⁄4" BEARING


1

5 Use epoxy to glue a 1/8-inch-by-1/4-inch


aluminum bearing on either side of the
hole, inserting a piece of 3/64-inch wire
temporarily to line up all three holes.

6 Use the same procedure to drill and


reinforce the 1/32-inch horizontal hole for
5
the tail-rotor shaft.

1
⁄8" × 1⁄8" BEARING

6
3
⁄64" WIRE

Project 8 » Rubber-Powered Helicopter 157


7 The main-rotor shaft is fabricated from a 3½-inch-
long piece of 3/64-inch wire. With a needle-nose pliers,
bend the diamond-shaped eye into the top end of the EYE
wire. Then, position the 5/16-inch-by-5/16-inch brass
washer 2¼ inches below the center of the eye, wrap a 7
few turns of copper wire tightly above it, and solder
both the wire and washer to the shaft, ensuring that 2 1⁄4"
the washer is exactly perpendicular to the shaft.
This joint will carry the full force of the fully wound ⁄64" WIRE
3
rubber motor, so it must be strong. (Note: The
procedure for soldering metal joints is covered in
project 10.) COPPER WIRE
8 Cut the two pulley disks very precisely from hard
3/32-inch balsa, using the patterns provided in the
project drawing. Then, using a sharp blade, cut a
V-groove into the edge of each disk around its full ⁄16" × 5⁄16" WASHER
5

circumference. Smooth the groove with a piece of


fine sandpaper folded into a V-shape.

REAR PULLEY FORWARD PULLEY


⁄8" DIAMETER DISK
5
1 13⁄16" DIAMETER DISK

MAIN
8 ROTOR SHAFT
V-GROOVE

SOLDER
EPOXY

9 Use epoxy to glue the


forward pulley onto
the brass washer we
previously soldered onto FORWARD
the main-rotor shaft. PULLEY
9

158 
REAR PULLEY
10 Glue the rear pulley onto the tail-rotor
shaft—a 1-inch-long piece of 1/32-inch
music wire with a short bend at one end.

TAIL ROTOR SHAFT

11 MAIN ROTOR SHAFT 10

WASHER
GLASS BEAD
11 Slide a washer and a glass bead onto the
BEARING bottom of the main-rotor shaft and insert the
shaft through its aluminum bearings.

MAIN
ROTOR SHAFT
12 With the shaft in position, bend a hook 12
for the rubber motor at its lower end.
Use two sets of pliers to avoid damaging
the balsa framework.

HOOK

1
⁄32" BLADE
13 Each main-rotor blade is laminated
LAMINATIONS
from two 1/32-inch layers. Use the
13 cutting patterns provided to cut out
the four laminations required for
two blades.

Project 8 » Rubber-Powered Helicopter 159


3" MAILING TUBE
14 For each blade, soak a pair of laminations in hot
LAMINATED BLADE water, glue them together with wood glue, and
then form them over a 3-inch-diameter mailing
tube or a similarly sized cylindrical object. Once
the laminated blade is in position, wrap it tightly
14 with masking tape to preserve its curved shape
and then set it aside to dry for at least 24 hours.

15 After removing the laminated blade from the tube,


sand it to a cambered airfoil shape.

15

16 Cut the rotor hub from hard


1/8-inch balsa, using the pattern
provided with the project drawings.
LAMINATED BLADE
Shape the ends of the hub to match
the undersides of the two blades.
AIRFOIL SHAPE
⁄8" BALSA HUB
1

16

SHAPED TO MATCH
UNDERSIDE OF BLADE

BLADE

BLADE

17 Glue the blades to the ends of the HUB


hub, taking care to align them at
the prescribed 5° angle of forward
sweep, as shown at right.
17 FORWARD SWEEP OF BLADES

160 
18 Cut the stabilizer bar from 1/32-inch wire,
bend it to shape, and glue it to the bottom
HUB
STABILIZER BAR of the rotor hub with epoxy. The stabilizer
weights are pieces of 3/32-inch-diameter solder
drilled with a 1/32-inch hole and glued to the
ends of the bar.
18

SOLDER

19 After balancing the completed


rotor, make the hinge pin from 19 WASHERS
HINGE PIN
1/32-inch music wire and sharpen
its downturned ends with a file.
Slip two 1/8-inch-by-1/8-inch
aluminum washers onto the
hinge pin, pass it through the
diamond-shaped eye at the top
of the main-rotor shaft, and
then press its ends down into
the balsa rotor blade and hub. INSTALLED
Secure the pin and washers to
the hub with epoxy.

BLADE ANGLE
15°
20

20 Prefabricate the tail-rotor components:


two blades, cut from 1/32-inch balsa
and sanded to an airfoil shape; and the
TAIL ROTOR cylindrical hub, carved from a piece
HUB of 3/16-inch-square balsa. (If 3/16-inch
15°
balsa is not available, laminate two
layers of 3/32-inch sheet.) Glue the
blades to the shaft at a 15° angle of
TAIL ROTOR attack, as shown below.
BLADE

Project 8 » Rubber-Powered Helicopter 161


PULLEY AND SHAFT
21 Install the tail rotor by passing its shaft through a washer,
then through the tail boom, then another washer, and
finally the rotor hub. Make a 90° bend in the end of the WASHER
shaft and secure it to the hub with epoxy. 21 TAIL
ROTOR
22 The tail-rotor belt drive is a piece of sewing thread that
is tied snugly around the two pulleys. Be sure to coat the
knot with glue and then cut off the loose ends so that they
will not foul the pulleys. INSTALLED

90°
BEND
BELT DRIVE

22

23 The rubber motor consists of 4 to 6 strands


of 1/8-inch rubber assembled and lubricated
using the same techniques we employed in
our model airplane project. Use 4 strands for
testing and indoor flying; try 6 strands for RUBBER
longer flights outdoors. MOTOR

23

The rubber-powered helicopter is now complete and ready for testing!

162 
TESTING PROCEDURE

Before the helicopter can be flight tested, it must be balanced so that the center of gravity is aligned with
the main-rotor shaft. If not, add weight to the nose or tail until the proper balance is achieved.

Next, wind the rubber motor about 20 turns and then release the rotor while holding the helicopter.
The two objectives of this test are to ensure that the main rotor is tracking properly and the belt drive is
spinning the tail rotor without slipping. If the main rotor is not tracking properly, one blade will fly higher
than the other. (To distinguish one blade from the other, it is helpful to color the tip of one blade with a
marker.) If one blade does fly higher than the other, bend up the stabilizer bar that precedes the high blade.
If this condition persists on the next test, bend down the stabilizer bar that follows the high blade.

Once the rotor is tracking satisfactorily, wind the rubber motor 20 to 30 turns for a low-power flight test.
When launching the aircraft, allow the main rotor to spin for a second or two and then release the fuselage.
With such a small number of turns, the aircraft should only fly a short distance; nonetheless, this test
is essential to ensure that the tail-rotor pitch is approximately correct. Specifically, observe whether the
helicopter maintains a more-or-less constant heading in flight. If so, the tail-rotor pitch is approximately
correct. If the helicopter pivots sharply to the left, the tail rotor is providing too much thrust. Correct the
problem by trimming a bit of balsa from the tips of the tail-rotor blades. If the aircraft pivots to the right,
the tail rotor is not providing enough thrust. Correct the problem by gently twisting the tail-rotor blades to
increase their pitch.

For full-power testing, wind the rubber motor until it forms two full rows of knots. To initiate a vertical
climb, hold both the fuselage and main rotor level when launching the aircraft. To initiate forward flight,
tilt the helicopter forward about 10°.

Project 8 » Rubber-Powered Helicopter 163


PROJECT 9
CAMERA-
EQUIPPED ROCKET
DESIGN, BUILD, AND TEST A MODEL
ROCKET CAPABLE OF RECORDING
A VIDEO OF ITS OWN FLIGHT.
LESSONS IN WHICH THIS PROJECT IS COVERED:
13 » This Is Rocket Science
14 » Build a Rocket
16 » Let’s Do Launch!

PROBLEM DEFINITION
Requirements
}} Must achieve a velocity of at least 13 meters per second by the time it reaches the top of the launch rod.
}} Must be capable of carrying an operating video camera safely to an altitude of at least 500 feet (150 meters).
}} Must return itself and its payload safely to Earth.

Constraints:
}} Use a readily available mini digital camcorder.
}} Use standard model rocket components and engines.

Because all modern research and development in the space sciences is performed and communicated in the
International System of Units (the metric system), we will use metric units exclusively for this project.

164 
REFERENCES
Hibbeler, Engineering Mechanics.
Westerfield, Make.

Rocket Propulsion
How does a rocket work? Combustion of propellant within the engine directs a high-velocity stream of exhaust
gases rearward. These gases are further accelerated through a nozzle at the back of the engine. If exhaust gases are
being accelerated rearward, it means that the rocket must be exerting a rearward force on these gases. Therefore, in
accordance with Newton’s third law, the exhaust gases must also exert an equal and opposite force on the rocket. This
force is the thrust that propels the rocket forward.

Full-scale rockets use liquid-fuel engines, solid-fuel engines, or both. For example, NASA’s space shuttle used liquid-fuel
main engines augmented by solid-fuel booster rockets to provide additional thrust during the first phase of each launch.

Most model rockets use disposable solid-fuel engines, which are factory-made and extensively tested for safety and reliability.

A cutaway view of a typical model rocket engine is shown below. When ignited, the black powder propellant will burn
intensely for a short period of time, typically on the order of one second. This combustion causes exhaust gases to be
expelled at high velocity through the ceramic nozzle at the rear, producing the thrust required to launch the rocket.
Once the propellant has been completely consumed, the delay charge is ignited. It burns much more slowly and
produces no thrust, thus allowing the rocket to coast upward to its highest altitude—a point called the apogee. Once
the delay charge has burned completely, it ignites the ejection charge, an explosive material that creates a surge of
pressure to pop the nose cone off the rocket and deploy the parachute or other recovery device.

end cap

ejection charge
delay charge

propellant

ceramic nozzle

Rocket engines like this one are available in many sizes, with widely varying performance characteristics. Selecting the
right engine to meet the requirements of this project is our most important design challenge.

Project 9 » Camera-Equipped Rocket 165


DESIGN CONCEPT
NOSE CONE

PAYLOAD BAY
Key features:
(BT-80)
TRANSITION SECTION
}} To fully enclose the camera, the payload bay
will use a standard BT-80-size cardboard body CAMERA LAUNCH LUG
tube, which is 66 millimeters in diameter, and
a plastic nose cone of the same size. BOOSTER BODY
}} The camera will be angled downward such (BT-60)
that the launch site will be within the field
of view throughout the rocket’s ascent. This
can only be achieved if the body tube of
the booster is smaller in diameter than the ELLIPTICAL FIN
payload bay, with the camera mounted within
the transition section that connects the two
tubes. Thus, the design will employ a standard
BT60-size tube for the booster.
}} The camera is removable, so its memory card can be installed and then removed for downloading the
video file after each flight.
}} The fins are elliptical—the theoretically optimal shape for minimizing aerodynamic drag.
}} Two launch lugs, mounted on the booster, will engage the 3/16-inch-diameter steel launch rod
mounted on the launchpad. The purpose of this system is to guide the rocket until it is moving fast
enough for the fins to provide aerodynamic stability.

DETAILED DESIGN
Overall Approach

To satisfy our design requirements, we will develop mathematical models that can predict the trajectory of a
model rocket powered by a particular standard-sized engine. This trajectory consists of three phases:

1 the engine burn phase, which starts at ignition and ends when the propellant is completely expended;

2 the coasting phase, during which the engine provides zero thrust, but the vehicle continues its ascent to
the apogee by virtue of the initial velocity it acquired during the engine burn; and

3 the recovery phase, in which the vehicle descends back to the Earth suspended beneath a parachute.

166 
Analysis of Engine Burn Phase

To analyze the engine burn phase, we will apply the impulse-momentum principle, a mathematical
adaptation of Newton’s second law. The impulse-momentum principle states that the total impulse
imparted to a body over a finite period of time is equal to the change in the momentum of that body
during the same period. Defined mathematically:

where:

Thus, the impulse-momentum principle can be written as:

where:

Assuming our rocket will launch vertically, the relevant forces acting on it are:

W = weight of the rocket, which acts downward


T = thrust, which acts upward
D = aerodynamic drag, which opposes the rocket’s motion and thus acts downward

Including aerodynamic drag will greatly complicate this analysis; thus, for now, we will ignore it.

The impulse-momentum equation for our rocket can now be written as:

(T − W) is the total upward force applied to the rocket—the net thrust, which causes the vehicle to
accelerate upward. Δt is the duration of the engine burn.

Project 9 » Camera-Equipped Rocket 167


We can estimate the weight of the rocket by weighing all known components (i.e., the body tubes, the nose
cone, the transition section, a typical engine, engine-mounting hardware, balsa wood for fins, and the video
camera) and then adding 10% (i.e., multiplying by 1.1) to account for the additional weight of glue, paint,
and various fittings. The result is:

measured mass = 188 gr


adjusted mass = 1.1(188 gr) = 207 gr

Converting from mass to weight:

Model Rocket Engine Classification Codes


All model rocket engines are classified according to a standardized code consisting of a letter and two numbers—for
example, C11-5. In this example:

†† The letter C indicates the impulse provided by the engine. The chart below shows the range of impulse values
associated with each of the seven common classes of model rocket engines, ranging from ¼A to E. Thus, a class
C engine delivers an impulse between 5 and 10 newton-seconds.
†† The number 11 represents the engine’s average thrust measured in newtons. Thus, a C11 engine will deliver an
average thrust of 11 newtons.
†† The number 5 is the duration of the delay charge in seconds. Thus, for a C11-5 engine, the ejection charge will
fire 5 seconds after the engine burn finishes.

If we were only interested in getting a rocket of a given weight MODEL ROCKET ENGINE CLASSES
into the air, this classification code would be entirely adequate as
a basis for engine selection. But because we are designing a rocket
Class Impulse (N-sec)
to achieve specific performance criteria, the code is not quite
adequate. We need to know the engine’s actual impulse, not just
¼A 0 - 0.625
a range of impulses, and we need to know how the thrust varies
with time, rather than just the average thrust. ½A 0.625 - 1.25
A 1.25 - 2.5
This information can be obtained from experimentally developed
graphs of thrust versus time—also called thrust curves—which B 2.5 - 5
are available from model rocket engine manufacturers and from
C 5 - 10
websites such as http://www.thrustcurve.org/. This website is
particularly useful, because it provides the relevant experimental D 10 - 20
data in a downloadable format that can be imported directly into E 20 - 40
a spreadsheet.

168 
For our first analysis, we will use a C11 engine, with the thrust curve shown below. The area underneath
this curve is equal to the impulse delivered by this engine.

25
THRUST CURVE
C11 Engine

20
Thrust (N)

15

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Time (sec)

To calculate the area of this complex shape, we will use a computational technique called the trapezoidal
rule. To apply this rule, divide the large, irregular area into a series of vertical slices, defined by the data
points used to plot the curve. Because all of these slices are either triangles or trapezoids, their individual
areas can be calculated as shown below. Each trapezoidal area is an increment of impulse. (Note that the
units work out correctly: impulse = force × time.) By adding these areas together, we can determine the
engine’s total impulse.

TRAPEZOIDAL RULE
area = 1
2 (b + B ) h area = 1
2 (T i + T j )(t j − t i )

5
Tj
Thrust (N)

B
b
h
Ti
0
ti tj 0.1
0
Time (sec)

Project 9 » Camera-Equipped Rocket 169


These computations can be performed quite efficiently with a computer spreadsheet (a copy of which can
be obtained from http://stephenjressler.com/diy-engineering/). This spreadsheet was used to generate the
graph of velocity versus time, as shown below.

VELOCITY VS. TIME


40 C11 Engine
34.8 m/sec
35

30
Velocity (m/sec)

25

20

15

10

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Time (sec)

This graph provides much useful information about the engine burn phase of our rocket’s trajectory. Note
that the rocket does not even start moving until about 1/20 second after ignition, when the engine has
built up enough thrust to overcome the rocket’s weight. Next, we see a sharp acceleration—an increase in
velocity, represented by the slope of the velocity-versus-time curve—followed by a lower level of acceleration
for the remainder of the engine burn. By the time the engine burn is complete, the rocket has reached a
velocity of 34.8 meters per second (about 78 miles per hour).

This result does not fully satisfy our needs because our design must be based on the rocket’s altitude at the
apogee of its trajectory. To determine altitude from the velocity-versus-time curve, recognize that:

distance = velocity × time

Thus, just as we were able to calculate impulse (force × time) as the area under the thrust-versus-time curve,
we can calculate altitude (velocity × time) as the area under the velocity-versus-time curve. As with impulse,
this computation can be performed by applying the trapezoidal rule with a computer spreadsheet. The result is
shown on the next page, plotted on the same horizontal axis as our previous graph of velocity versus time.

This graph tells us that, at the conclusion of the engine burn, our rocket will have traveled 14.6 meters
upward. More importantly, we can use both graphs to check one of our critical design requirements: that
the vehicle must attain a velocity of at least 13 meters per second by the time it reaches the top of the launch
rod. For a rocket of this size and weight, it is typical to use a launch rod that is 4 feet, or 1.2 meters, long.
Thus, we need to know the rocket’s velocity when its altitude is 1.2 meters.

170 
40
VELOCITY AND ALTITUDE VS. TIME 16
C11 Engine 14.6 m
Velocity (meters/second) 35 14
velocity
30 12

Altitude (meters)
altitude
25 10
20 8
15 14 m/sec 6
10 4
5 2
1.2 m
0 0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
Time (seconds)

The graph indicates that, with a C11 engine, the rocket will reach 1.2 meters at 0.3 seconds after ignition,
and at this same instant, the velocity will be 14 meters per second—comfortably exceeding the 13-meters-
per-second standard.

Analysis of Coasting Phase

To determine the rocket’s maximum altitude, we will apply the principle of conservation of energy to the
coasting phase of the rocket’s trajectory, from state 1—the point at which engine burnout occurs—to state
2—the apogee. Between these two points, the total energy of the system must remain constant. Therefore,
assuming that the rocket’s motion is perfectly vertical and ignoring the effect of aerodynamic drag:

This result falls far short of our design requirement, 150 meters. Thus, the C11 engine will not work.

Project 9 » Camera-Equipped Rocket 171


Second Analysis Iteration

We will now repeat the analysis with a D12 engine, which is in the next-higher impulse category and has
a slightly higher average thrust than the C11. This second iteration is greatly facilitated by the use of the
spreadsheet developed previously. The result is a maximum altitude of 280 meters, which exceeds our
design requirement by a comfortable margin. Before drawing any final conclusions, however, we must look
critically at four simplifying assumptions that are embedded in the analysis above.

First, in both our impulse-momentum and conservation of energy formulations, we assumed that the
rocket’s trajectory would be perfectly vertical. But in practice, a truly vertical trajectory is quite impossible.
If there is any wind at all, a stable rocket will tend to incline its trajectory into the wind, a phenomenon
called weather cocking. Furthermore, if the wind is relatively strong, we will deliberately incline the
launch rod slightly into the wind so that the downwind drift of the deployed parachute will bring the
vehicle back to Earth relatively close to the launch site. Any inclination of the rocket’s trajectory will reduce
its maximum altitude. During an angled ascent, only the vertical component of the vehicle’s velocity will
actually contribute to altitude gain. We cannot eliminate this source of inaccuracy in our analysis, but we
can minimize it by launching when the wind speed is relatively low.

Second, our analysis ignores the effect of aerodynamic drag. We might expect that this effect is minimal,
because the rocket is relatively small and highly streamlined. But, in fact, the adverse effect is quite
substantial, primarily because drag increases as a function of velocity squared. It is possible to account for
aerodynamic drag in an analysis of rocket performance; however, the process is computationally complex
and thus is beyond the scope of this course. The two key results of this analysis are as follows:

}} Aerodynamic drag has only a small effect on our prediction of the rocket’s velocity as it clears the
launch rod.
}} Aerodynamic drag has a substantial effect on our prediction of maximum altitude—dropping it
about 35%, from 280 meters to 180 meters. Nonetheless, because of the more-than-adequate power
provided by the D12 engine, our design still meets the 150-meter altitude requirement.

This more sophisticated analysis to determine the effects of aerodynamic drag has a very useful by-product: an
estimate of the elapsed time between engine burnout and apogee. This time interval works out to be just short
of 5 seconds, indicating that we should use an engine with a 5-second delay charge (i.e., a D12-5) so that the
parachute will deploy very soon after the rocket has passed through the apogee and started its descent.

The third issue with our spreadsheet-based analysis is that it implicitly assumes that the rocket’s weight is
constant. But, in practice, all rockets decrease in weight as their engines burn fuel. For a full-scale rocket,
the effect of fuel consumption is huge, because the weight of fuel constitutes such a large proportion of the
vehicle’s total weight. For our model, the effect is much less significant, because the weight of our engine’s
propellant is only about 10% of the total vehicle weight. And unlike aerodynamic drag, the effect of
propellant consumption is favorable because as the rocket gets lighter, it will accelerate more rapidly.

Like drag, the effect of propellant consumption can also be incorporated into our spreadsheet-based
analysis—though again at the cost of considerable additional complexity. The result is a 5% increase in
maximum altitude.

172 
The fourth potential issue with our analysis is that it is based on our initial estimate of the rocket’s mass: 207
grams. To ensure the validity of our analysis, we must verify this estimate after the rocket has been built.

The actual measured mass of our rocket prototype is 203 grams, a very favorable result. Because the actual
mass is just a bit less than the estimate, the actual performance of our rocket should be slightly better than
the analytical results, so there is no compelling need to redo the analysis with the corrected mass.

Parachute Design

Parachute Physics
The main components of a parachute are the canopy,
suspension lines, and payload, as shown here. For canopy
simplicity, a model rocket parachute typically uses
6 suspension lines attached to the payload with a
metal fitting called a snap swivel, which facilitates
easy removal and allows the parachute to rotate freely
suspension
without twisting the suspension lines. When using 6
suspension lines, the shape of the canopy is generally lines
a hexagon. To design the parachute for our rocket, snap swivel
we will need to determine an appropriate size for this
hexagonal canopy.

The purpose of the parachute is to control the rocket’s


payload
rate of descent during the recovery phase so that the
vehicle returns to the Earth without doing damage to
itself or anything below.

A parachute is best understood as a device that is


designed to develop as much aerodynamic drag as
possible. The general equation for aerodynamic drag is:

where:

Continued on next page…

Project 9 » Camera-Equipped Rocket 173


Now consider the mechanics of a parachute deployment. When the canopy first opens, the parachute
and payload are typically in free fall, moving at relatively high velocity. But the drag force developed
by the newly opened chute is quite high, resulting in rapid deceleration. As the parachute slows down,
it also develops less drag. Very quickly, the system reaches an equilibrium state—called terminal
velocity—in which the drag force is exactly equal to the weight of the rocket. Beyond this point, the
velocity remains constant for the remainder of the descent.

This condition can be represented mathematically by setting the drag force equal to the weight of the rocket.

The required area of the canopy can now be determined by solving for A:

To design the parachute, we must first determine the weight of the rocket during the recovery phase by
converting mass to weight and by accounting for the loss of propellant mass due to combustion:

The rocket’s measured mass is 203 grams, and the propellant mass is 25 grams, as determined from the
engine manufacturer’s specifications. Thus:

174 
By setting aerodynamic drag equal to the rocket’s weight (see “Parachute Physics” sidebar), we can now
calculate A, the required area of the canopy, as follows:

Because the canopy will be hexagonal, we can use the formula for the area of a hexagon to solve for the
canopy width, d, as follows:

The calculated width, d = 20.9 inches, corresponds to a radius of 12 inches, measured from the center of the
hexagon to one vertex.

12"
d

Project 9 » Camera-Equipped Rocket 175


FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/. Full-size cutting patterns and templates are provided in the PDF version of this guidebook
and are formatted to print on 8½-by-11-inch sheets.

Drawing 9.1 » Rocket,


perspective view

R
D G
A B H F
C G
N-Q H F
K J
K
Drawing 9.2 » Rocket, F L M I
exploded view
F

Drawing 9.3 » Camera


mount, exploded view O
Q

176 
MATERIALS LIST
COMPONENT DESCRIPTION PRODUCT*
A Plastic nose cone (BT-80 size) Apogee #20080
B Payload bay body tube, BT-80, 3″ long Apogee #10198
C Clear plastic BT-60 to BT-80 transition section Apogee #17077
D Coupler, BT-60 to BT-80 transition section
E Booster body tube, BT-60, 18″ long Apogee #10141
F Fins, 1/8″ balsa sheet (4) SIG #SIGB154
G Launch lug for 3/16″ launch rod (2) Apogee #13059
H Launch lug extender, 1/8″ balsa (2) SIG #SIGB154
I Engine mount tube
J Engine mount hook
K Centering ring, BT-60/BT-80 (2) Apogee #12009
L Large retaining ring
M Small retaining ring
N Camera mount, 1/8″ balsa SIG #SIGB154
O Camera mount, 1/4″ balsa SIG #SIGB139
P Camera mount, 1/8″ balsa SIG #SIGB154
Q Camera mount, 1/4″ balsa SIG #SIGB139
R BrightTea Mini DVR 808 #16 V3 Lens D Car Key Chain Micro **
Camera HD 720P Pocket Camcorder
S Small wood screw
Parachute canopy, plastic
Parachute suspension lines, nylon thread
Snap swivel Apogee #14510
Shock cord, 3/16″ rubber Apogee #30330
Recovery wadding Apogee #05750
Estes D12-5 engine Apogee #05782
Estes igniter Apogee #03077

*Components with an Apogee product number were obtained from Apogee Components (https://www.apogeerockets.com/).
Components with a SIG product code were obtained from SIG Manufacturing Co. (http://www.sigmfg.com).

**Purchased from Amazon.com. Many similar products are available; however, a different camera will require modifications
to the camera mount design.

The other supplies needed for this project are:

}} wood glue
}} cyanoacrylate glue
}} sanding sealer
}} spray enamel primer and paint

Project 9 » Camera-Equipped Rocket 177


}} medium- and fine-grit sandpaper
}} talcum powder

The only tools required for this project are:

}} hobby knife or single-edge razor blade


}} metal straightedge
}} small screwdriver (for camera mount)

FABRICATION AND ASSEMBLY PROCEDURE

airfoil shape 1
1 Print the fin cutting patterns provided in this
guidebook. Use spray adhesive to attach the patterns F
to a sheet of ½-inch-by-4-inch balsa, ensuring that
the wood grain is oriented as shown on the patterns.
Cut out the four fins (component F) and sand them
to a streamlined airfoil shape.

cutting pattern

2 The BT-60 body tube (component E) is sold in the correct 18-inch length; however, the larger BT-80
body tube (component B) must be cut to a length of 3 inches. After marking the tube, use a razor blade
or hobby knife to make the cut, pressing lightly on the blade and making several passes around the
circumference of the tube, rather than trying to cut all the way through the cardboard in a single pass.
After completing the cut, square it up
with a sheet of fine sandpaper glued to a
flat board, as demonstrated in lesson 14. G

3 The launch lug extenders (component G


H) are 1-inch-by-1/2-inch-by-1/8-inch
H
balsa with the grain oriented in the short
direction. These, too, are sanded to an airfoil shape
airfoil shape and then each is glued to a
1-inch launch lug (component G).
H
3

178 
4 Using cyanoacrylate glue, attach the plastic nose cone (component A) to the BT-80 body tube
(component B) and attach the clear plastic transition section (component C) to its cardboard coupler
(component D). Loop the Kevlar cord (which is included with component C) through two holes
punched into the rear of the transition section, tie it in place, and add a drop of cyanoacrylate to the
knot to prevent it from slipping. Do not glue component B to component D.

A
B 4
D
C

kevlar cord

2 1/4"
5 With a hobby knife, cut a short slot 2¼
inches from the aft end of the engine mount slot
tube (component I).

K
6 J

6 Insert one end of the metal engine hook


(component J) into this slot and then apply
glue and mount one of the two centering
rings (component K) ¼ inch from the aft end
1/4" of the tube.

7 To further strengthen the connection


between the engine hook and the tube, 7 K
glue the larger of the two cardboard L J
K
retaining rings (component L) about
halfway along the length of the steel
hook; then, mount the second centering
ring (component K) at the front end of
the engine mount tube.

Project 9 » Camera-Equipped Rocket 179


K
8 L
K 8 Use a piece of scrap balsa to apply glue
around the inside of the tube and then
M quickly insert the smaller retaining ring
(component M) from front to rear, all the
position of M way back to the metal engine hook, as
shown below.

9
9 Using the fin marking template (which is included with
the cutting patterns), mark the four fin center lines and fin center line
the launch lug center line at the rear of the booster launch lug
body tube (component E). Then, use a sharp pencil center line
and a block of wood to extend the fin center lines
forward about 4 inches, as demonstrated in lesson
14. Using the same procedure, extend the launch
lug center line along the full length of the tube.

fin marking
template

10

F 10 Glue the first fin (component F) in position,


E checking to ensure that it is perfectly aligned
with the center line mark and perfectly
fin
perpendicular perpendicular to the tube.
to tube

align launch lugs

11 Repeat this procedure for the remaining three


fins and then for the launch lugs, which are
positioned along the marked center line—one
at the rear and one at the front of the booster.
Ensure that both lugs are aligned.
11

180 
12 After all glue joints have dried fully,
install the engine mount in the rear of
the booster. First, smear glue generously
12
around the inside of the body tube at the
locations of the two centering rings; then,
insert the engine mount until the rear
centering ring is approximately 1/8 inch
inside the body tube.

13 Add a generous fillet of glue around the rear ring.


13 The full 30-newton thrust of the D12 engine will
be transmitted across these glue joints, so they need
to be strong. The fins also experience substantial
loads—both during powered flight and recovery—
fillet
so they should also be reinforced with a glue fillet
along both sides of each fin. The fin fillets should
be formed with the body tube oriented horizontally.
Create only two fillets at a time—on the two
upward-facing fin surfaces—so that the glue will
not run while it is drying.
fillet

14 Assemble the camera mount from


components N, O, P, and Q. Ensure that
14
the mount fits snugly inside the plastic C
transition section (component C). (If you N
use a different camera or choose to mount Q
O
the camera at a different angle, some
P
redesign of this mount will be required.)

15 Fasten the camera to the mount with a


rubber band or a piece of tape. Fasten the
mount to the transition section with a
small wood screw (component S).

15

Project 9 » Camera-Equipped Rocket 181


16 To attach the rubber shock cord to the 16
booster, cut out a piece of paper that is paper shock cord mount
approximately 3/4 inch by 2 inches and shock cord
coat one side with glue. Place one end of
the shock cord onto the glued surface,
as shown below, and then fold the paper
twice to lock the cord securely inside this
glue and fold
mount. Finally, glue the mount to the
inside of the booster body tube about
1 inch from the top. The other end of
the cord is tied to the bottom of the
transition section.
glue and fold
17 With construction complete, seal the
balsa and cardboard parts with three
coats of sanding sealer, sanding between
coats with fine sandpaper.

18 Apply a coat of primer and then paint the


rocket with spray enamel.
19 canopy outline

19 Lay out the hexagonal canopy by placing


template
the paper template underneath a sheet of
lightweight plastic and then measuring 12
"
outward from the center point a distance
of 12 inches along each of the six radial g ea
ch
radi
al
lin
e
center of parachute

alon

lines. After locating and marking the


ard
tw
" ou
12
re
su
mea

six vertices, use a straightedge and razor DIY Project #9 - Model Rocket
Parachute Template

blade to cut along the entire canopy


outline from vertex to vertex.

20 Cut three suspension lines, each 38


inches long, from strong nylon thread.
Form each end of these lines into a
20
loop and then fasten to the canopy’s six tape
canopy
vertices with small squares of duct tape. suspension
The loops help prevent the thread from line
slipping under load. To prevent tangles,
fasten the suspension lines as shown here.

21 Extend the suspension lines so that all six


legs are equal in length; then, tie them
to one eye of the snap swivel. Attach the
snap swivel to the rear of the payload bay.

182 
ADDITIONAL EQUIPMENT REQUIRED FOR TESTING

Before the rocket can be flight tested, the following equipment must be made or purchased:

}} electric launch controller


}} launchpad
}} device for measuring altitude

Electric Launch Controller

An electric launch controller is a device for firing a rocket engine igniter safely. We will design and build
this device in project 10.

Launchpad

Drawings for the DIY launchpad used in lesson 16 are provided below.

Drawing 9.4 » Launchpad, perspective view

A
D
B C

Drawing 9.5 » Launchpad,


exploded view
H E
D

Project 9 » Camera-Equipped Rocket 183


8 7/8" 1' 2 15/16"
3/4" 3/4" 45°

1/4" dia. 45°


1 1/2" 1 1/2"
45° 45°

11 1/8" 1' 2 15/16"


Component E - Adjustable Leg Component D - Fixed Leg (2 required)
5 1/4"

3/4" 3/4" 45° 4 1/2"

3/4"

3/16" dia.
4 1/2" 4 1/2" 1/4" dia.
1/8" dia.
45°

3 5/16"
4 1/2" 3 1/4"
Component A - Base Component B - Fixed Leg Mount Component C - Pivot

Drawing 9.6 » Launchpad, dimensions


of major components

Key features of this design are as follows:

}} The device is a simple wooden tripod with two fixed legs (component D) and one adjustable one
(component E). The adjustable leg can rotate, as a means of inclining the launcher. Rotation of this
leg is constrained such that the launch rod cannot be angled at more than 30° from vertical.
}} All wooden components are made of poplar and are fastened together with wood glue and screws.
}} The pivot for the adjustable leg is a 1/4-inch-by-2-inch carriage bolt with a washer and a wing nut
(component H).
}} On top of the tripod is a wooden platform (component A) with a 1/8-inch-diameter hole and a
3/16-inch-diameter hole drilled into it. These holes will accommodate the two most common launch
rod sizes used in the United States: 1/8 inch for smaller rockets and 3/16 inch for larger rockets.
}} To ensure a sufficiently high launch velocity, the 3/16-inch-diameter launch rod (component G) is 48
inches long. Rods longer than 36 inches are typically not sold in hardware stores but can be obtained
from McMaster-Carr. (See https://www.mcmaster.com/, product 9120K31.)
}} The launcher incorporates a blast deflector (component F), the purpose of which is to redirect the
stream of hot exhaust gases propelled downward by the rocket engine during the launch. Not only
does the blast deflector protect the wooden launchpad, but it also prevents any flammable materials
in the vicinity of the launcher from catching fire. This component is a 6-inch-by-6-inch-by-1/16-inch
piece of sheet steel.

184 
}} A tennis ball will be used as a safety cap, which must be placed on the tip of the launch rod any time
a launch is not underway. The purpose of the safety cap is to prevent accidental eye injuries and any
other bodily harm that might result from a close encounter with the tip of a steel rod. The launch
controller’s arming key should also be attached to the safety cap with a lanyard.

Measuring Altitude

Although not strictly required for launching, an altitude-measuring device will allow us to determine
definitively whether the design requirements for this project have been satisfied.

The preferred method for measuring altitude—called dual-axis tracking—is illustrated below.

altitude
h

a d
tracking b c tracking
station station
#1 #2
L

In this system, each tracking station is responsible for measuring two angles: one vertical (a or d) and one
horizontal (b or c). Both tracking stations must make their sightings simultaneously. The device used to
make simultaneous measurements of horizontal and vertical angles is called a theodolite.

Drawings and instructions for building a theodolite are provided in Mike Westerfield’s excellent book Make:
Rockets, cited in the References section above. Two theodolites are required to implement dual-axis tracking.

Project 9 » Camera-Equipped Rocket 185


As a rule of thumb, the distance between the two tracking stations should be between 50% and 200%
of the expected altitude. When launching in the Northern Hemisphere, the tracking stations should be
positioned generally to the south of the launcher so that the trackers will not be looking into the Sun when
they make their sightings. Each theodolite should be carefully leveled and then aimed directly at the other
theodolite, with 0° indicated on its horizontal protractor.

To perform the tracking operation effectively, the two trackers should be able to communicate with each
other during the launch. If they are too far apart to communicate orally, cell phones or walkie-talkies can
be used.

Prior to the launch, both tracking stations should rotate their instruments toward the launcher. Then, at
ignition, they will track the rocket as it ascends. When either tracker believes that the rocket has reached its
apogee, he or she will call out an appropriate prearranged command. At this instant, both tracking stations
will aim their instruments as precisely as possible at the rocket’s position in the sky and then record the
measured horizontal and vertical angles.

Once the angles a, b, c, and d are known, the altitude, h, is calculated with the following two equations:

These equations are mathematically equivalent; thus, in theory, we do not need both of them. But in
practice, some human error is inevitable in measuring angles a, b, c, and d. If we take sightings from
two tracking stations and then use the measured angles to solve these two equations, the two calculated
altitudes will inevitably be somewhat different. Thus, to minimize error, we will do both calculations and
then average the results.

TESTING PROCEDURE
Check Stability

Before the completed rocket is launched for the first time, it must be checked for stability—the vehicle’s
capacity to resist wind-induced deviations from its original attitude, through the aerodynamic action of
its fins.

Our rocket will be stable if its center of gravity (CG) is ahead of its center of pressure (CP) by a distance at
least equal to the diameter of the main body tube.

186 
The CG is the rocket’s balance point. To locate the CG, first ensure that the engine, parachute, and camera
are installed in the rocket; then, tie a loop of string around the booster body and adjust its position until the
rocket hangs level.

The CP is the point at which the lateral force can be assumed to act when the rocket’s attitude is not
aligned with its direction of flight—a condition that might occur when the vehicle is hit by a wind gust.

The location of the CP can be determined very accurately, using a complex mathematical approach
described in Westerfield’s book Make: Rockets, cited above. But the CP can also be located approximately
and conservatively, using a simpler experimental procedure.

To implement this procedure, create a cardboard cutout that accurately depicts the profile of the rocket.
The balance point of this cutout provides a conservative estimate of the CP location. If this point is at least
one body tube diameter behind the CG, the rocket is guaranteed to be stable.

If the rocket is found to be unstable, this condition can be corrected by (1) adding weight to the nose, (2)
increasing the length of the body tube, or (3) making the fins larger. For a rocket that has already been
built, only option (1) is a practical solution.

Prepare the Rocket for Flight

To prepare the rocket for flight:

}} Inspect the vehicle carefully to ensure that it has no broken components, that the rubber shock cord
has no nicks or cracks, and that both the shock cord and parachute are firmly attached to the rear of
the payload bay.
}} To install the engine, bend the metal engine hook slightly outward and slide the engine into the engine
mount tube with the nozzle oriented toward the rear until the hook snaps back to its original position.
}} Inspect the igniter to ensure that the two wire leads do not touch each other at any point along their
length. If they do, it will create a short circuit, and the igniter will not fire. Some manufacturers add
a paper strip across the leads to prevent short-circuiting, but it is still possible for the wires to touch
above or below the paper. If so, spread them gently apart.
}} Insert the igniter into the engine nozzle until it will not go any farther, indicating that the pyrogen
at its tip is in contact with the black powder propellant inside. Many engines are packaged with
disposable plastic plugs, which are designed to hold the igniter in place. If you have a plug, insert it
now. If not, secure the igniter to the engine with a small piece of tape.
}} Crumple a few sheets of recovery wadding into a ball and insert it into the booster from the top
until it comes into contact with the engine. Recovery wadding is made of a lightweight paper treated
with fire retardant. Its purpose is to prevent the plastic parachute from being melted by the hot gases
produced by the ejection charge.
}} Pack the parachute, using the procedure demonstrated in lesson 16.
}} Install the camera in the camera mount, but do not install the mount in the rocket at this time.

Project 9 » Camera-Equipped Rocket 187


Safety and Site Selection

In designing and building a model rocket, preparing for a launch, selecting the launch site, and conducting
the launch, it is essential to comply with the Model Rocket Safety Code, promulgated by the National
Association of Rocketry. This code includes the following provisions:

}} Use only lightweight, nonmetal parts for the nose, body, and fins of the rocket.
}} Use only certified, commercially made model rocket engines.
}} Use a recovery system such as a streamer or parachute, and use only flame-resistant or fireproof
recovery wadding.
}} Launch the rocket with an electric launch system and electric igniters; use a launch system with a
safety interlock and a launch switch that returns to the “off” position when released.
}} Launch from a launch rod that is pointed to within 30° of the vertical, use a blast deflector, and cap
the end of the launch rod when it is not in use.
}} For a class D engine, use a launch site with at least 500 feet of open space in all directions.
}} Wind speed at the time of launch must be less than 20 miles per hour.
}} Ensure that the launch area is clear of any flammable materials that might pose a fire hazard.
}} Use a countdown before launch.
}} Ensure that everyone is paying attention and is at least 15 feet away from the launcher.
}} If uncertain about the safety or stability of an untested rocket, check stability before flight.
}} Fly only after warning spectators and clearing them to a safe distance.
}} If a rocket engine misfires, remove the launcher’s safety interlock and wait 60 seconds after the last
launch attempt before allowing anyone to approach the rocket.
}} Do not launch a rocket at targets, into clouds, or near airplanes.
}} Do not attempt to recover a rocket from power lines, tall trees, or other dangerous places.

Launch Procedures

On the launch site:

}} Designate a launch control officer, who assumes overall responsibility for the safety of the launch site.
Measure the wind speed and direction. Do not launch if the wind speed exceeds 20 miles per hour.
}} Place the launchpad at a location that is clear of flammable materials. Orient the tripod with its
adjustable leg pointing into the wind. Incline the launch rod into the wind at an angle consistent with
the wind speed, but never greater than 30°. Once the launchpad is set, place the safety cap on the tip
of the rod. Ensure that the launch controller’s arming key is attached to the safety cap.
}} Place the launch controller 30 feet away from the launchpad and unroll the firing wires from the
controller to the pad.
}} Site the two tracking stations and measure the distance between them. Level and orient the theodolites.
}} To prevent the rocket from resting directly on the blast deflector plate, wrap a piece of tape around
the launch rod a few inches above the plate.
}} Place the rocket on the pad—first removing the safety cap, then guiding both launch lugs over the
rod, and then replacing the cap once more.

188 
}} Switch on the camera. Then, insert the camera mount into the transition section and secure it with a
screw. Slip the nose cone and upper segment of the payload bay onto the transition section, and secure
it with a wrap of tape to prevent it from detaching when the ejection charge fires.
}} Connect the two firing wires to the wire leads of the igniter. While making these connections,
double-check to ensure that there are no short circuits—that neither the leads nor the connectors are
touching each other, nor are they touching the steel blast deflector.
}} The rocket is now prepared for launch. Remove the safety cap and arming key from the launch rod
and clear the launchpad area.
}} Insert the arming key into the launch controller and check that the arming light is illuminated.
}} Check that the tracking stations are prepared for the launch and are in communication with each other.
}} The launch control officer takes one last look around the site to ensure that there is no one within 15
feet of the pad and that there are no aircraft flying overhead.
}} Count down from five to one and press the firing button. Immediately remove the arming key from
the launch controller and return the safety cap to the launch rod.
Trackers note when the rocket reaches its apogee and then take their readings.
}} Recover the rocket. Disassemble the payload bay, remove the camera, and shut it off. Check the
rocket for damage and prepare it for its next flight.

After the launch, download the video from the camera and perform the calculations to determine the
measured altitude.

Project 9 » Camera-Equipped Rocket 189


PROJECT 10
ELECTRIC LAUNCH
CONTROLLER
DESIGN, BUILD, AND TEST AN ELECTRIC LAUNCH
CONTROLLER TO FIRE A MODEL ROCKET IGNITER.
LESSON IN WHICH THIS PROJECT IS COVERED:
15 » Make an Electric Launch Controller

PROBLEM DEFINITION
Requirements
}} Incorporate all relevant safety features specified in the National Association of Rocketry (NAR)
Model Rocket Safety Code:
§§ Safety interlock in series with the launch switch.
§§ Launch switch that returns to the “off” position when released.
}} Provide at least 30 feet of standoff from the launchpad.
}} Produce a current of at least 2 amps to fire a standard igniter.
}} Produce a current of less than 0.05 amps to check continuity.

Constraints:
}} Powered by household AA batteries.
}} Use only readily available materials and components.

190 
The design requirements above are specific to the type of igniter sold by the Estes-Cox Corporation
(http://www.estesrockets.com/). For this particular product, the current must be at least 2 amps to ensure
that the igniter fires reliably, but when the circuit is tested for continuity, the current must be less than
0.05 amps to ensure that the igniter does not fire unintentionally. Recognize, however, that igniters
are manufactured and sold by several different companies, and their products all have slightly different
specifications. Thus, if you build a launch controller that will be using a different brand of igniter, some
adjustments to the forthcoming design will be required.

REFERENCES
Kemp, The Makerspace Workbench.
Westerfield, Make.

Electrical Circuits
In the broadest sense, electricity is defined as the flow of electrical charge. But in the more specific form
that we will be using in this project, electricity is the movement of electrons through a conductor due to
a difference in voltage. Voltage is a measure of the difference in electrical potential energy between two
points, such as the two terminals of a battery. Voltage is measured in volts (V).

To make use of this electrical potential, we need a circuit—a continuous conducting pathway
connecting the battery’s positive and negative terminals. The movement of electrons through a circuit is
called a current. It is typically measured in units of amperes, or amps (A).

The current produced by a battery is called direct current (DC) because it is steady and unvarying. The
associated circuit is called a DC circuit.

The relationship between voltage and current is defined mathematically by Ohm’s law:

where:

I = current through a conductor in amps


V = voltage between any two points along the conductor in volts
R = resistance of the conductor between the same two points

Resistance is a measure of the extent to which a conductor impedes the movement of electrons. The
standard unit of resistance is the ohm, represented by the Greek letter omega (Ω).

Continued on next page…

PRoject 10 » Electric Launch Controller 191


Although resistance impedes the flow of current, there are many circumstances in which resistance is
beneficial. Indeed, in this project, resistance will be deliberately added to our launch controller circuit to
enhance its function. In such circumstances, an electrical component called a resistor is used.

It is worth noting that incandescent lightbulbs work because of resistance—more specifically, a


phenomenon called resistive heating. Any time current flows through a conductor, it generates heat.
Higher resistance generates more heat. The filament of a lightbulb is a very fine piece of tungsten wire
with resistance so high that a sufficiently strong current causes it to glow.

In addition to incandescent bulbs, resistive heating is used in all sorts of other useful devices: toasters,
space heaters, your car’s rear window defroster, the soldering iron we will be using in this project, and
the igniter that will fire our rocket engine.

DESIGN CONCEPT
FIRING WIRE IGNITER

Key features:

}} The power source consists of one or more RESISTOR LED


AA batteries connected to an igniter by a FIRING
pair of firing wires. WIRE
}} At the heart of the circuit are two parallel
branches: a firing branch (which includes ARMING
the firing button) and a continuity check KEY
branch (which includes a resistor and a
light-emitting diode (LED)).
FIRING
BATTERY BUTTON
}} The firing button is spring-loaded, so it
will only close the circuit when it is being
pressed. The circuit will be broken as soon as the button is released.
}} Just ahead of the firing button is the arming key, which fulfills the NAR code requirement for a
“safety interlock.” When this key is engaged, the circuit is complete; when it is removed, the circuit is
broken, and an accidental press of the firing button will not fire the igniter.
}} In the continuity check branch, the LED serves as an arming light, which warns the user that the
launch controller is prepared for operation.

Note that, with arming key inserted, current can flow through the circuit even when the firing button is
not pressed. This current will light the LED, telling us that the circuit is intact—that the igniter is properly
connected and the firing wires have no breaks in continuity. Thus, when the LED is lit, the controller is
considered to be “armed.” This current will not fire the igniter because the large resistor in the continuity
check branch will cause the associated current to be less than 0.05 amps.

192 
When the firing button is pressed, current can flow through both parallel branches of the circuit; however,
nearly all of this current will actually flow through the firing circuit, because of its significantly lower
resistance. We will observe this phenomenon when we test the controller. When the firing button is pressed,
the LED will go dark, indicating that the current in the continuity check branch has dropped to nearly zero.

DETAILED DESIGN
Circuit Design: Firing Branch

The objective of this design task is to determine the number of 1.5-volt AA batteries required to power the
firing circuit. Applying Ohm’s law:

where:

Solving for n:

Therefore, five AA batteries are required to fire an igniter reliably.

PRoject 10 » Electric Launch Controller 193


In practice, however, we will be using a plastic battery holder to combine all of our batteries in series, and
six-battery holders are somewhat more common than five-battery holders. Thus, we will use six batteries.
There is no problem with providing some additional voltage; however, we must take this higher voltage into
account in the design of the continuity check circuit below.

Circuit Design: Continuity Check Branch

The objective of this design task is to determine the required resistance of the resistor in the continuity
check branch. Applying Ohm’s law:

where:

In the circuit analysis above, the term VF (the forward voltage drop of the LED) must be subtracted from
the battery voltage because the LED is classified as a non-ohmic device—an electrical component that does
not obey Ohm’s law. When current passes through an LED, there is a significant voltage drop; however, the
device has essentially zero resistance. Thus, for an LED, I does not equal V over R.

The LED manufacturer’s specifications indicate that VF is between 1.8 and 2.2 volts. We will use the
average—2 volts—for our analysis.

In addition, the forward current is specified as 20 milliamps, or 0.02 amps. This is the current the LED
can handle continuously without damage. A lower current will light the bulb, but not as brightly. The peak
forward current—specified as 30 milliamps—is the maximum current the device can handle for short
periods. Any higher and the LED will self-destruct.

We will need to design the continuity check branch of the launch controller circuit such that its current
does not exceed the LED’s forward current specification: 20 milliamps = 0.02 amps.

194 
The design requirement for this circuit is that the current must be less than 50 milliamps to guarantee
that the igniter will not fire during the continuity check. Note that this requirement is less stringent than
the LED’s maximum specified forward current; thus, if we design to meet the needs of the LED, we are
guaranteed to satisfy the design requirement as well.

Solving the equation above for R R , the resistance of the resistor in the continuity check branch:

Rounding up to the next-larger-available resistor size, we will use a 680Ω resistor.

Why round up, rather than down? Ohm’s law says that higher resistance will result in lower current. If the
actual current is somewhat lower than the 20 milliamps on which this calculation is based, we will have a
larger margin of safety with respect to both LED burnout and accidental firing of an igniter. If we rounded
down, we would be erring in the unsafe direction.

FINAL DESIGN DRAWINGS

arming key (1/2” dia.)


arming light (7/32” dia.)
firing button (1/2” dia.)

slot for firing wires


screw hook

1 1/2"

Drawing 10.1 » Launch


8" controller, wooden case

6 1/2"

PRoject 10 » Electric Launch Controller 195


push
button Drawing 10.2 » Launch
controller circuit

stereo
jack 640 Ω LED
resistor
firing
wire alligator clips
+ -

6 AA batteries
= soldered connection

MATERIALS LIST
Electrical Components

COMPONENT PRODUCT DESCRIPTION


Battery holder 6 × 1.5V AA battery holder with wire leads (Uxcell 036185)
Firing wire 18-gauge, 2-conductor, stranded-copper speaker wire
Firing button SPST push-button switch (RadioShack 2750618)
Arming key 1/8-inch phone plug (RadioShack 2740287)
Arming jack 1/8-inch stereo panel-mount audio jack (RadioShack 2740249)
Alligator clips Fully insulated mini 11/4″ alligator clips (RadioShack 2700378)
680Ω resistor Resistor kit, 1/4W (SparkFun Electronics COM-10969)
LED LED, assorted – 20 pack (SparkFun Electronics COM-12062)

Other Components

COMPONENT PRODUCT DESCRIPTION


Case sides (2) 11/2″ × 3/4″ × 8″ poplar
Case ends (2) 11/2″ × 3/4″ × 5″ poplar
Case bottom 61/2″ × 8″ × 1/4″ plywood
Case lid 61/2″ × 8″ × 1/4″ plywood
Battery hold-down (2) 1/2″ screw hooks
Lid retainer (4) 1/2″ panhead screws

196 
Additional tools and supplies needed for this project are:

}} soldering iron with stand


}} helping hands
}} 1-millimeter rosin-core 60/40 solder
}} soldering flux (if rosin-core solder is not available)
}} steel wool or sandpaper
}} heat-shrink tubing
}} wire stripper
}} pliers
}} woodworking tools (for wooden case)
}} wood glue
}} cyanoacrylate glue
}} digital multimeter (recommended but not required)

FABRICATION AND ASSEMBLY PROCEDURE

1 Build the wooden case as shown in drawing 10.1 above. Connect the sides, ends, and bottom with wood
glue. Drill holes in the lid for the arming key, LED, and firing button, as indicated in the drawing.

2 Insert the stereo jack and push button into their designated holes in the case lid and fasten them in
position with the nuts provided. Fasten the LED in its hole with a drop of cyanoacrylate glue.

3 Using drawing 10.2 as a guide and following the procedure demonstrated in lesson 15, make the
following soldered connections:

}} Negative battery lead to one firing wire.


}} Positive battery lead to both lower connectors on the stereo jack.
}} One end of the 640Ω resistor to the upper connector on the stereo jack.
}} Other end of the 640Ω resistor to the positive (long) lead of the LED.
}} One end of a short length of wire to the upper connector on the stereo jack; other end of this wire to
the first connector on the push button.
}} One end of another short length of wire to the negative (short) lead of the LED; other end of this wire
to the second connector on the push button.
}} Second firing wire to the second connector on the push button.
}} One alligator clip to the opposite end of each firing wire.

4 Install six AA batteries in the battery holder and secure the battery holder inside the case with a rubber band.

PRoject 10 » Electric Launch Controller 197


5 Carefully insert all wiring into the case. Feed the firing wires through the slot at the bottom of the case.

6 Fasten the lid to the case with four panhead screws.

The electric launch controller is now complete and ready for testing!

TESTING PROCEDURE

1 Connect the leads of a digital multimeter to the alligator clips at the ends of the firing wires.

2 Set the multimeter to measure current.

3 Insert the arming key into its socket. The LED should illuminate, and the multimeter should read less
than 20 milliamps (0.02 amps).

4 Press the firing button. The LED should go dark, and the multimeter should read more than 2 amps.

5 Remove the arming key. The launch controller is now ready for operation.

198 
PROJECT 11
BALLISTA,
ONAGER, AND
TREBUCHET
DESIGN, BUILD, AND TEST WORKING MODELS OF
THREE HISTORICALLY IMPORTANT CATAPULTS: THE
BALLISTA, THE ONAGER, AND THE TREBUCHET.
LESSONS IN WHICH THIS PROJECT IS COVERED:
17 » A Tale of Three Catapults
18 » Build a Ballista, Onager, and Trebuchet

THE BALLISTA

PROBLEM DEFINITION
Requirements:
}} Must be capable of flinging a golf ball at least 200 feet.
}} Use historically documented proportions, with a module size of 1 inch.

Constraints:
}} Use nylon rope for the torsion springs.
}} Use commonly available materials for all other components.

Project 11 » Ballista, Onager, and Trebuchet 199


REFERENCES
Campbell, Besieged.
———, Greek and Roman Artillery 399 BC–363 AD.
Landels, Engineering in the Ancient World.
Marsden, Greek and Roman Artillery: Historical Development.
———, Greek and Roman Artillery: Technical Treatises.
Oleson, ed., The Oxford Handbook of Engineering and Technology in the Classical World.

The Ballista
The ballista was developed by Hellenistic-era Greeks in the 3rd century B.C. and subsequently perfected
by Roman military engineers. It looks something like a giant crossbow; however, its source of power
is not a bending bow but a pair of torsion springs. The term torsion refers to twisting; indeed, the
ballista’s torsion springs are bundles of twisted rope, which drive a pair of throwing arms connected to a
sling. In antiquity, these torsion springs were usually fabricated from animal tendons woven into rope.

One of the most fascinating aspects of this machine is that its technological characteristics have been so
thoroughly documented in the ancient sources—most notably, a technical treatise written by a Greek
named Philo of Byzantium in the 3rd century B.C. Philo describes the mechanical system in great detail,
with all dimensions expressed in terms of a single module—defined as the diameter of one torsion
spring. For example, Philo specifies that each throwing arm should be six modules long and one-half
module wide, that the stock should be 19 modules long and one module high, and so on for every major
dimension of the machine.

In effect, what Philo has described is a mathematical model of the ballista—one that can be used to
design a catapult of practically any size, from the small desktop model we will be building to the 40-foot
monsters used by the Roman legions.

Thus, our second design requirement specifies that we will be using Philo’s mathematical model, with a
module size of 1 inch, resulting in overall dimensions of approximately 20 inches long by 15 inches wide
by 10 inches high.

DESIGN CONCEPT

Because the overall design concept for the ballista has been largely defined by the ancient sources, we can
develop it at a much higher level of detail than has been the case in previous projects, using a 3-D computer
model rather than a simple sketch.

200 
steel lever
bronze washer retainer pin
steel washer
wooden washer throwing arm
hole-carrier
slider
trigger
diagonal strut

ratchet

torsion
spring
stanchion

stock

throwing
arm
pedestal
windlass

Key features:

}} Each torsion spring is fabricated from nylon rope wrapped around a pair of steel levers. The nominal
diameter of these springs is 1 inch, which actually means that they will be composed of as many
strands of nylon rope as can be fit into the 1-inch-diameter holes in the hole-carriers.1
}} Each steel lever is supported on three washers—one of hardwood, one of steel, and one of bronze—
designed to allow the lever to rotate for the purpose of pretwisting the spring.
}} Each torsion spring holds a throwing arm.
}} A sling extends between the ends of the two arms.
}} The spring frame is a wooden framework that supports the torsion springs and is composed of two
hole-carriers and four vertical stanchions, which are configured to allow for maximum rotation of the
throwing arms.
}} Diagonal struts prevent the upper hole-carrier from shifting rearward when the ballista is cocked.
}} The stock is mounted on a pedestal, which allows the weapon to pivot both vertically and
horizontally for aiming.
}} On top of the stock is a movable component, called the slider, which incorporates a trigger
mechanism and a trough that will guide the projectile during its launch.
}} At the rear is a windlass, a hand-operated winch that is used to cock the catapult.

1
The term “hole-carrier” is translated from the Greek περίτρητον, as discussed by
Marsden in Greek and Roman Artillery: Technical Treatises, cited above.

Project 11 » Ballista, Onager, and Trebuchet 201


The mechanical system operates as follows:

}} To prepare for a fling, the slider is moved forward to engage the trigger with a metal loop on the rear
of the sling; then, the projectile is loaded.
}} As the slider is winched rearward by the windlass, the sling pulls the two throwing arms rearward
and inward, twisting the torsion springs through an angle of approximately 45°, thus storing the
elastic energy that will ultimately launch the projectile.
}} During the draw, the ratchet mechanism ensures that the slider will not fly forward if the windlass is
released. It also allows for range adjustment by controlling the length of the draw.

throwing arm
sling

slider
ratchet
trigger

windlass

trigger claw
trigger lever }} The projectile is launched
trigger pivot by pulling the trigger, which
trigger base plate
consists of a claw, pivot, and
lever, mounted on a base
plate. When this mechanism
is engaged, the lever prevents
the claw from rotating, thus
holding it firmly in contact with
the base plate. To release the
trigger, the lever provides the
mechanical advantage necessary
ratchet pawl windlass to rotate it out from underneath
ratchet the claw with very little force,
allowing the sling to push the
claw up and out of the way
during a projectile launch.

202 
DETAILED DESIGN
Analysis of Torsion Springs

The objective of this design task is to ensure that two 1-inch-diameter torsion springs made of nylon rope
will be capable of propelling a golf ball at least 200 feet. To accomplish this task, we will apply the principle
of conservation of energy and the projectile range equation, as we did in project 1.

Our model of the mechanical system consists of throwing arm


the two arms, sling, and projectile, with applied
torques at the base of each arm representing the a
torsion springs. The rotational position of both
arms is indicated by the angle a. torsion spring sling

Two discrete states of this system are defined T


below. At state 1, the catapult is cocked and
projectile
ready to shoot. At state 2, the arms and sling
have just come to rest in their forward position
and the projectile is moving at velocity v2,
having just cleared the sling. T

State 1 State 2
a2 a

a1

T1 T2

v2
T1 T2

a1

a2

The two principal forms of energy relevant to this problem are kinetic energy, the energy associated with
a mass in motion, and elastic energy, the energy stored though the elastic deformation of a material. The
term elastic refers to a material that, when stretched and then released, returns to its original shape. Nylon
rope is not perfectly elastic, but its behavior is sufficiently elastic for our purposes.

Project 11 » Ballista, Onager, and Trebuchet 203


Applying conservation of energy and recognizing that the kinetic energy at state 1 is zero:

This formulation incorporates a number of


simplifying assumptions. First, it ignores energy Mechanical Efficiency
losses such as friction and aerodynamic drag that
inevitably occur in real-world mechanical systems. and the Ballista
It also ignores the small change in gravitational
potential energy that will occur between states 1 Mechanical efficiency is a quantitative
and 2 as the projectile moves up the slider. measure of performance that is typically
defined as the ratio of useful output to input,
expressed as a percentage. For a ballista,
These effects are relatively small and can be the input is the elastic energy imparted to,
ignored without introducing excessively large and stored in, the torsion springs when the
errors into our analysis. But what cannot be machine is cocked. The useful output is the
ignored is the implicit assumption about what kinetic energy of the projectile.
actually occurs at state 2, the point at which
the arms and sling have come to rest and the The ballista is relatively unique in that—
subject to a few reasonable simplifying
projectile has just cleared the sling. The key issue assumptions—it has a maximum theoretical
is what actually stops the forward motion of the mechanical efficiency of 100%. Most of
arms. There are two possibilities: the machines we encounter in our everyday
lives—for example, automobile engines,
1 The arms might be stopped by colliding refrigerators, pumps, and turbines—have
with the vertical stanchions. If this occurs, inherent limitations in their mechanisms for
energy conversion that limit their maximum
our conservation of energy formulation is theoretical efficiency to substantially less
wildly inaccurate, because it fails to account than 100%.
for the substantial amount of energy that
would be lost in this collision. Any energy Of course, in the real world, actually
lost to the collision will not be transmitted achieving 100% efficiency from any device
into the projectile, and the result will be a is impossible. As a point of comparison, the
bow and arrow can also have a theoretical
significantly lower projectile velocity than mechanical efficiency of 100%, but
our analysis predicts. extensive experimental studies by archers
and bow manufacturers suggest that a well-
2 The sling will snap tight just before the designed and properly strung bow typically
arms make contact with the stanchions, thus has an actual measured mechanical
preventing the collision from occurring. efficiency of between 80% and 90%.

Based on this observation, we will assume


75% efficiency for our somewhat-less-
sophisticated ballista.

204 
In this case, the sling stops the arms; thus, nearly all of the arms’ kinetic energy is transmitted into
stretching the sling. And as the sling snaps taut, much of this energy is transmitted to the projectile as
additional kinetic energy.

It is much better to stop the forward motion of the arms with the sling than with the stanchions—because
the former is more consistent with our conservation of energy formulation and, more importantly, because
it will significantly improve the range of the ballista. We can achieve this end simply by adjusting the
length of the sling so that it prevents the arms from striking the stanchions.

Returning to the conservation of energy equation, we can now account for the less-than-perfect mechanical
efficiency (see “Mechanical Efficiency and the Ballista” sidebar) of our model by multiplying the change in
elastic energy by a factor of 0.75, indicating that only about 75% of this energy will actually find its way
into the projectile. The result is:

Because the behavior of nylon rope is highly nonlinear, we must determine the change in elastic energy of
our torsion springs experimentally. The procedure is as follows:

}} Fabricate a torsion spring that is identical to the ones we will be using in the ballista model.
}} For this spring, determine the relationship between torque and angle of rotation by determining the
torque required to twist the specimen through measured angles of 20° through 360° at 20° increments.
}} Plot the results of this test on a graph of torque versus rotation (below).
}} The energy stored in the spring is equal to the area under this graph, calculated between the
appropriate minimum and maximum angles of rotation.

This experimental procedure is demonstrated in lesson 17. The maximum angle used in this test was 360°,
because the inward pull occurring at higher angles of rotation is sufficiently large to damage the testing
machine (and therefore would damage the ballista as well).

TORQUE VS. ROTATION


180
160
torque (inch-pounds)

140
120
100
80 stored
energy
60
40
20
0
0 50 100 150 200 250 300 350
angle of rotation (degrees)

Project 11 » Ballista, Onager, and Trebuchet 205


Based on this graph, it is clear that elastic energy can be maximized by pretwisting each spring to an angle
of 315° such that the additional 45° rotation of the throwing arm will store energy corresponding to the
shaded area under the curve—the area between 315° and 360° of rotation.

This requirement for pretwisting explains the design of the steel levers and washers supporting the torsion
springs on our model (illustrated previously). The levers can be rotated to pretwist the springs. The heavy
steel washers serve as bearing plates to carry the huge compressive forces that will be applied to the wooden
structure of the spring frame as the springs twist, and the thinner bronze washers provide reduced friction
at the interface with the steel levers, allowing them to be rotated more easily. The small holes drilled into
the outer edge of the wooden washer allow for the insertion of steel retainer pins that will prevent the levers
from slipping after the springs have been pretwisted.

Having defined the required pretwist and the angle through which the throwing arms will rotate, we
can calculate the corresponding area under the torque-rotation curve using the trapezoidal rule and a
spreadsheet, just as we did for the impulse calculation in our rocket design. In doing this computation, we
must convert the angles from degrees to radians, or the units will not work out correctly. (A radian is a
dimensionless measure of an angle, with 2π radians equaling 360°.)

The computed result is 113 inch-pounds, or 9.4 foot-pounds, which we can now substitute into our
conservation of energy equation to solve for velocity.

Note that the spring energy is multiplied by two, because there are two torsion springs.

We can now substitute this velocity into the projectile range equation we first used in project 1. For the
optimum launch angle of 45°:

This is the theoretical maximum range, which we are unlikely to achieve in practice, primarily because the
projectile range equation does not account for aerodynamic drag. Nonetheless, it is reasonable to expect
that our machine will be capable of achieving the 200-foot range specified in the design requirements.

206 
Design of Sling and Cocking Mechanism

Designs for the sling and cocking


mechanism of the ballista are based on System Model
an equilibrium analysis of the sling under
its maximum loading condition with the
65°
ballista fully cocked. An idealized model of
the system in this worst-case configuration is 65° 49°
s
shown at right. The dimensions and angles 49°
are taken directly from the 3-D computer 175.5 in-lb
model shown above. Each torsion spring is P
subjected to a torque of 176 inch-pounds,
the maximum measured torque from our 175.5 in-lb
torsion spring experiments. The unknown
force P is the maximum rearward pull s
applied by the cocking system.
7.2"
Now we will isolate one throwing arm,
4.75"
replacing the sling with its tension force,
T, and resolving this force into its two
equivalent components: Tx parallel to
the arm and Ty perpendicular to it. As in
previous projects, we can use trigonometry
to define each component in terms of the
total force, T.

The relevant equilibrium equation is the sum


Free body diagram - one arm
of moments about the spring, s—a procedure
we first used to analyze our lever-based
175.5 in-lb
Ty = T cos 24° structural testing machine in project 2.

T
s 24°

4.75"
Tx = T sin 24°
y The sling will be fabricated from braided
fishing line, with a working tensile strength
of 30 pounds. Given the predicted sling
x tension of 40 pounds, at least two strands
of this material will be needed for the sling.
More is better.

Project 11 » Ballista, Onager, and Trebuchet 207


Now isolate the rear segment of the sling, where the Free body diagram - rear of sling
pulling force, P, is counterbalanced by the tension T
force, T, in both legs of the sling. Ty
49°

Consistent with this new set of coordinate axes, the


relevant equilibrium equation is the sum of forces in 49°
the x-direction: Tx = T sin 49° P

Tx = T sin 49°

y
49°
Ty
T x

This is a very useful number: It tells us how much


load the draw rope, windlass, and trigger mechanism must be capable of carrying, and it will also serve as
the basis for determining how much mechanical advantage our cocking system must provide.

The mechanical advantage of a windlass is the ratio of its outer radius to its shaft radius. Our model will
use a 1/4-inch-diameter steel shaft, and the largest outer radius for which the handles will not hit the
ground is 2 inches. Thus, the windlass will provide:

With a mechanical advantage of 16:1, every pound applied to one windlass handle will apply 16 pounds
of rearward pull to the slider. And because we will typically operate the windlass by applying force to two
handles simultaneously, we will be able to provide the full 61-pound cocking force with a pull of less than
two pounds on each handle.

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/. A full-size cutting pattern is provided in the PDF version of this guidebook. This pattern
measures 24 inches by 18 inches and must be printed at exactly 100% size. This can be accomplished by a
commercial printing establishment for just a few dollars.

208 
Drawing 11.1 »
Ballista, plan view

0 1" 2" 3" 4"


scale

left diagonal strut omitted for clarity

Drawing 11.2 » Ballista,


side elevation

0 1" 2" 3" 4"


scale

Drawing 11.3 » Ballista, front elevation

0 1" 2" 3" 4"


scale

Project 11 » Ballista, Onager, and Trebuchet 209


H
H
G
F
E Drawing 11.4 » Ballista spring frame, exploded view
D
I
I
I I
A

B
J
C
B C
K

K
J A U
S Q
T P
R
U U
Z R Q
X
O
S T
X L
M U
N X
L

N
X

V
W Drawing 11.5 »
Ballista stock,
W Y exploded view
trigger assembly Y
(see Drawing 11.7)

c
c
e
h

f
d

b
g

e
Drawing 11.6 » Ballista
slider, exploded view

210 
h

Drawing 11.7 » Ballista trigger, exploded view

u
l
v t
t
s
v
k n w

m
j
q
r r
r

i r
r p

p
r

Drawing 11.8 » Ballista pedestal, exploded view o

1/2" 1/4"

Elevation View
1 13/16"

Drawing 11.9 » 11/16"

Ballista throwing 5/16"

arm, elevation
and plan views 1/2" 1/4"

1 7/16"
6 1/4"

Plan View

Project 11 » Ballista, Onager, and Trebuchet 211


10 1/2"
1" 1/2"

J5 Drawing 11.10 » Sling


J2 fabrication jig
J4
1/2"
J5 J3
J3
J6

6 3/8" J1

MATERIALS LIST

COMPONENT DESCRIPTION
A Hole-carrier (2) 3/4″ × 3″ × 81/2″ poplar*
B Outer stanchion (2) 3/4″ × 2″ × 6″ poplar*
C Inner stanchion (2) 3/4″ × 11/2″ × 6″ poplar*
D Wooden washer (4) 3/4″ × 21/2″ diameter maple*
E Steel washer (4) 2″ outside diameter steel washer (McMaster-Carr product #98029A038)
F Bronze washer (4) 11/2″ outside diameter oil-embedded bronze thrust bearing (McMaster-Carr
product #5906K523)
G Steel lever (4) 1/4″ × 1/2″ × 3″ rectangular steel bar*
H Retainer pin (8) 1/8″ diameter × 21/2″ steel rod
I Dowel (8) 3/8″ diameter × 11/2″ hardwood dowel
J Retainer pin (2) 3/8″ diameter × 11/2″ hardwood dowel
K Diagonal strut (2) 1/4″ × 1/2″ × 14″ poplar*
L Ratchet (2) 1/4″ × 13/4″ × 19″ poplar*
M Stock base 3/4″ × 21/8″ × 9″ poplar
N Stock side (2) 1/4″ × 3/4″ × 151/2″ poplar
O Stock cross member 1/4″ × 1″ × 21/8″ poplar*
P Stock rear 3/4″ × 13/4″ × 25/8″ poplar
Q Windlass mount (2) 1/4″ × 13/4″ × 43/4″ poplar*
R Doubler (2) 1/4″ × 3/4″ × 13/4″ poplar
S Spacer (2) 1/4″ × 3/4″ diameter poplar*
T Windlass hub (2) 3/4″ × 11/2″ diameter poplar*
U Windlass handle (8) 1/4″ diameter × 13/4″ hardwood dowel

212 
V Stock bottom 1/4″ × 25/8″ × 9″ poplar
W Spring frame retainer 3/4″ × 1/2″ × 25/8″ poplar
(2)
X Dowel (4) 1/4″ diameter × 1″ hardwood dowel
Y Dowel (4) 1/4″ diameter × 11/4″ hardwood dowel
Z Windlass shaft 1/4″ diameter × 51/2″ steel rod
a Slider base 1/4″ × 23/32″ × 12″ poplar
b Slider 3/4″ × 119/32″ × 12″ poplar
c Trough sides (2) 1/4″ × 1/4″ × 101/4″ poplar
d Trigger block 3/4″ × 21/8″ × 13/4″ poplar
e Ratchet pawl (2) 1/4″ × 3/8″ × 2″ maple*
f Ratchet pawl axle 1/8″ diameter × 23/4″ steel rod
g Draw plate 1/16″ × 13/32″ × 13/4″ 6061 aluminum plate (McMaster-Carr product #8975K199)*
h Trigger mounting bolt (5) 4-40 × 2″ steel bolt (McMaster-Carr product #90272A121)
i Nut (5) 4-40 he× nut (McMaster-Carr product #90480A005)
j Trigger plate 1/16″ × 11/2″ × 15/8″ 6061 aluminum plate (McMaster-Carr product #8975K199)*
k Trigger bracket 1/16″ × 1/2″ × 1/2″ 6061 aluminum angle (McMaster-Carr product #8982K54)*
l Trigger claw 3/32″ × 1/2″ × 2″ 6061 aluminum bar (McMaster-Carr product #8975K297)*
m Trigger pivot 1/8″ diameter × 1/2″ steel rod
n Trigger lever 1/8″ × 1/2″ × 2″ 6061 aluminum bar (McMaster-Carr product #8975K577)*
o Pedestal transverse base 3/4″ × 3/4″ × 12″ poplar*
p Pedestal longitudinal 3/4″ × 3/4″ × 10″ poplar*
base (2)
q Pedestal column 3/4″ × 31/8″ × 2 11/16″ poplar
r Pedestal diagonal (6) 3/4″ × 3/4″ × 5″ poplar*
s Bracket base 3/4″ × 1″ × 25/8″ poplar*
t Bracket side (2) 1/2″ × 1″ × 23/4″ poplar*
u Horizontal pivot 1/4″ diameter × 4″ hardwood dowel
v Dowel (4) 3/16″ diameter × 1″ hardwood dowel
w Vertical pivot 3/8″ × 11/2″ hardwood dowel
x Dowel 3/8″ × 11/2″ hardwood dowel
- Torsion spring (2) 3/16″ nylon rope, 16 feet per spring
- Throwing arm (2) 1/2″ × 1/2″ × 61/4″ maple (see drawing 11.9)
- Sling braided fishing line (30-pound test), 15 feet
J1 Sling jig base 3/4″ × 8″ × 12″ MDF or plywood
J2 Sling jig back 3/4″ × 13/4″ × 12″ MDF or plywood
J3 Sling jig former (2) 1/16″ × 13/4″ × 51/2″ plywood*
J4 Sling jig spacer 3/4″ × 1″ × 4″ wood
J5 Arm loop peg (2) 1/4″ diameter × 2″ hardwood dowel (notched)
J6 Trigger loop peg 3/8″ × 13/4″ hardwood dowel (top 3/8″ filed flat)
*Cutting pattern provided

Project 11 » Ballista, Onager, and Trebuchet 213


In addition to common hand tools, this project requires extensive woodworking equipment, including:

}} table saw (for rip-cutting)


}} miter saw or radial arm saw (for crosscutting)
}} scroll saw or bandsaw (for curved cuts)
}} power-sanding station (for shaping and tapering)
}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} hand drill
}} files
}} clamps

Other supplies needed for this project include nylon cord, soft metal wire, wood glue, epoxy, cyanoacrylate
glue, and spray adhesive.

FABRICATION AND ASSEMBLY PROCEDURE

1 Prefabricate all of the wooden components specified in the materials list above. For components
that have cutting patterns, use spray adhesive to affix the patterns to wood of the appropriate type
and thickness. Then, cut to size. For components that do not have cutting patterns, simply use the
dimensions provided in the materials list. Refine the shape with a sander, as needed. Drill holes as
indicated on the cutting patterns.

The following wooden components require special attention during the prefabrication process:

}} Note in drawing 11.4 that the upper hole-carrier (component A) has two 1/4-inch holes drilled into its
rear edge to accommodate the diagonal struts (component K) while the lower hole-carrier does not.
The two hole-carriers are otherwise identical.
}} The throwing arms should be made from carefully selected straight-grained maple. After cutting two
1/2-inch-by-1/2-inch-by-61/4-inch arms, use a power sander to taper each one as shown in drawing 11.9;
then, add hardwood cleats, as shown, to prevent the arms from slipping out of the torsion springs.
}} Prefabricate the four wooden washers using the procedure demonstrated in lesson 18. Start by using
a 2-inch Forstner bit in the drill press to cut a 1/8-inch-deep recess into the face of a 3/4-inch maple
board. This recess will hold the 2-inch steel washer in position. Next, drill a 1-inch hole all the way
through the board at the center of the recess. Then, drill 8 equally spaced 1/8-inch holes around
the perimeter, as indicated on the cutting pattern. Finally, cut the outer 21/2-inch-diameter circular
outline with a scroll saw or bandsaw and sand it smooth.

214 
holes for handles
}} To prefabricate the windlass
1
hubs (component T), first cut
out two 11/2-inch-by-11/2-inch
squares of 3/4-inch poplar; then,
drill the 1/4-inch holes for the
hole
windlass shaft and four handles. for shaft
After all five holes have been
drilled, cut and sand the hub to
its final circular shape.

sand to final
circular shape

2 Prefabricate all of the metal components specified in the materials list above. After affixing the
appropriate cutting patterns to metal bars or sheets of the appropriate type and size, cut out each
component with a hack saw or bandsaw; then, use a file to refine the shape and smooth the edges of
each piece. When drilling holes in metal, be sure to use a clamp or vise to hold the workpiece.

3 Assemble the spring frame using wood glue and clamps to attach the two hole-carriers (component A)
to the four stanchions (components B and C), as shown in drawing 11.4. Once the glue has dried, drill
a 3/8-inch hole through each joint at the locations shown on the cutting pattern; then, tap a glue-soaked
3/8-inch wooden dowel (component I) into each hole, trim the dowel flush with the surface of the wood,
and sand it smooth.

4 Assemble each diagonal strut (component K) by gluing a 3/4-inch length of 1/4-inch hardwood dowel
into the hole in its lower end. The upper protrusion—which is an integral part of the wooden
component—must be rounded with a file or sandpaper so that it will fit into the 1/4-inch hole in the
upper hole-carrier.

5 Glue the wooden washers (component D) onto the hole-carriers as shown in drawing 11.4, ensuring
that the 1-inch hole in each washer is aligned with the corresponding 1-inch hole in the hole-carrier.

6 Using epoxy, glue the four steel washers (component E) into the recesses in the wooden washers
(component D).

7 Assemble the stock as shown in drawing 11.5. Begin by gluing and clamping components L, N, Q, R,
and S together to create the two side assemblies. Similarly, glue and clamp components M, V, and W
together to create the base. Ensure that the two spring frame retainers (component W) are spaced such
that the lower hole-carrier (component A) fits snugly between them. Drill 1/4-inch holes through the
base and install dowels (component Y) as indicated in drawing 11.5. Then, glue both sides to the base
and to components O and P. Reinforce these joints with 1/4-inch dowels (component X).

Project 11 » Ballista, Onager, and Trebuchet 215


8 Glue four handles to each windlass hub 8
(component T). File a flat face on both
ends of the steel windlass shaft (component
Z); then, insert the shaft into the windlass
mount (component Q) on either side of
the stock and epoxy one windlass hub
flat face
(component T) to each end of the shaft.

9 Assemble the slider as shown in drawing 11.6. Glue


and clamp components a, b, and c together. Use a chisel to cut
a 1/16-inch-deep recess into the bottom surface of the trigger block
(component d) to accommodate the aluminum draw plate (component g) while
also allowing for wood-to-wood contact between components b and d. After drilling
the transverse 1/8-inch hole in component d, glue components d and g in position on the slider. After
all glue has dried, use a drill press to drill the 5 vertical 1/8-inch holes for the trigger mounting bolts
through components d, g, b, and a, using the trigger plate (component j) as a template for the locations
of the holes. Use a 3/8-inch Forstner bit to countersink the bottoms of these holes so that the nuts
(component i) will not interfere with the forward movement of the slider within the stock.

10 Insert the ratchet pawl axle (component f) into the transverse hole in the trigger block (component d).
Epoxy a ratchet pawl (component e) to each end of the axle, taking care to ensure that the axle can still
rotate freely in the trigger block.

11 Insert the slider into the stock. Check that it slides smoothly, and check that the ratchet pawls on the
slider engage properly with the ratchets on the stock. Make adjustments if necessary.

12 Assemble the trigger as shown in drawing 11.7. Four mounting


bolts pass through the trigger brackets (component k), the hammer
trigger plate (component j), and the slider. The fifth mounting
bolt passes through the trigger lever (component n), the trigger 12
plate (component j), and the slider. The trigger claw (component
m) is fastened to the trigger brackets (component k) with a
short length of 1/8-inch-diameter steel rod—the trigger pivot
(component m). To retain the pivot within its supporting
brackets, use a hammer to flatten its ends, as shown below.
trigger pivot
13 Assemble the pedestal as shown in drawing 11.8. Use 3/16- (Component m)
inch hardwood dowels (component v) to reinforce the bracket
(components s and t) and use a 3/8-inch hardwood dowel
(component x) to reinforce the base (components o and q).
The horizontal and vertical pivots (components u and w) are
also made of hardwood dowels. Because of variability in the
metal surface
diameters of commercially available dowels, some sanding might
be required to ensure that these pivots are able to rotate freely in
their respective mounts.

216 
14 Fabricate the two torsion springs from 3/16-inch nylon rope, as demonstrated in lesson 18.

15 To pretwist the torsion springs, make a wrench from a 3-foot board and a pair of aluminum or steel
angles, as pictured below. As demonstrated in lesson 18, clamp the spring frame to the workbench
and insert a throwing arm into each torsion spring. Then, use the wrench to rotate each steel lever
(component G) through approximately 7/8 of a turn—315°—in the direction that pulls the arm into
contact with the outer stanchion. To prevent the spring from untwisting, insert two retainer pins
(component H) into the rim of each wooden washer.

1/4" x 1/4" notch


15
steel or aluminum angle

36" wooden handle

A
K
J K

16 To assemble the ballista,


slide the stock into C
the spring frame
between the two
inner stanchions
(component C)
J
and insert the
wooden retainer pins
u
(component J) through
the stanchions and into the
corresponding holes in the stock. Then,
16
attach the diagonal struts (component K),
first inserting the upper end of each strut
into the upper hole-carrier (component A) and then inserting the
lower end into the stock. Connect the pedestal to the stock with the
horizontal pivot, component u.

Project 11 » Ballista, Onager, and Trebuchet 217


17 Assemble the sling fabrication jig as shown in drawing 11.10. Fabricate the sling from one continuous
length of high-strength braided fishing line formed into 10 strands on the jig, as shown below. Tie the
two free ends of the line together tightly and secure with a drop of cyanoacrylate glue.

17

10 strands of
braided fishing line

18 Near each end of the sling, bind binding


the strands together to form an
eye. These two eyes will be used
to fasten the sling to the
throwing arms. 18

19 At the center of the sling, use thicker nylon


cord to weave a pouch for the projectile.
The curved shape of this pouch will
prevent the sling from slipping over the
top of, or underneath, the projectile
during a fling. Apply a few
drops of cyanoacrylate
glue to the pouch
to preserve its
curved shape.
woven pouch

19

218 
20 Fabricate the trigger ring by wrapping soft
metal wire around a 1/2-inch dowel. The
required number of wraps will depend
on the type and thickness of the wire. To
ensure that the trigger ring can carry the
61-pound cocking force with an appropriate
margin of safety, use enough wraps of wire
to provide a total tensile strength of at least trigger ring
120 pounds.

21 Connect the trigger ring to the sling with


10 loops (20 strands) of the same braided 20
fishing line used to fabricate the sling, as
shown below.

21 22 Remove the completed sling from the jig. Install it


on the ballista by placing the eyes of the sling into
the notches in the ends of the throwing arms. If
the sling is the correct length, the throwing arms
will need to be rotated inward about 5° before the
sling can be installed. If the torsion springs have
been properly pretwisted, this task will require
two people—one to pull the arms inward
and one to install the sling.

braided fishing line

notch

22 sling

23 For the cocking rope, use a 24-inch


length of nylon cord with a working
tensile strength of at least 75 pounds.
Tie one end of this cord to the shaft of notch
the windlass (component Z) and the
other end to the draw plate of the slider
(component g). Rotate the windlass until
all of the slack in the cord is taken up.

The ballista is now complete and ready for testing!

Project 11 » Ballista, Onager, and Trebuchet 219


TESTING PROCEDURE

Please employ your ballista safely. When the machine is fully cocked, it will launch a golf ball at approximately
65 miles per hour—not as fast as a typical golf drive, but fast enough to cause injury or property damage.
Never use the ballista to launch a projectile other than the golf ball for which it was designed.

To test the machine:

1 Find a location with at least 250 feet of unobstructed space.

2 Place the ballista on level ground, with the stock angled upward at 45°.

3 Engage the trigger claw with the trigger ring on the rear of the sling. Place a golf ball into the trough of
the slider, just ahead of the sling.

4 Ensuring that there are no people or property in the line of fire, rotate the windlass until the slider has
moved rearward four “ratchet clicks.” Engage both ratchet pawls with their respective ratchets.

5 Carefully inspect the throwing arms, sling, trigger ring, trigger, and ratchet pawls to ensure that there
are no indications of distress.

6 Aim the ballista. Double-check that there are no people or property in or near the line of fire. Pull
the trigger.

7 Check the throwing arms for damage. If the arms show evidence of having collided with the stanchions
during the fling, the sling will need to be shortened.

8 Retrieve the golf ball and repeat


the test, increasing the number
of “ratchet clicks” by one for each
test, until the maximum of seven
is achieved.

9 If the fully cocked ballista fails


to achieve a range of 200 feet,
disassemble the spring frame from
the stock and add another 45° of
pretwist to each torsion spring.

220 
THE ONAGER

The onager was developed several centuries after the ballista, during the Roman imperial era, and it was
used also used in the Middle Ages. This machine has a single throwing arm powered by one horizontally
oriented torsion spring. Because of this configuration, the design also incorporates a heavy framework to
stop the forward motion of the throwing arm.

The onager model demonstrated in lesson 18 was designed such that its single large torsion spring is
able to store approximately the same elastic energy as the two smaller torsion springs used in the ballista
model. Thus, both machines have the same theoretical maximum range. In practice, however, the range
of the onager is much less than that of the ballista, because of the onager’s substantially lower mechanical
efficiency. Much of the elastic energy stored in the onager’s torsion spring is wasted in stopping the forward
motion of the throwing arm at the conclusion of each fling.

Drawings and construction notes for the onager model are provided below.

projectile holder
Drawing 11.11 » Onager,
throwing arm front perspective view
eye bolt

framework for trigger


arresting arm

spring frame

axis of
torsion spring

lever
washers
framework for
arresting arm

throwing arm
lever
washers projectile
holder

trigger

eye bolt

Drawing 11.12 » Onager, spring frame


rear perspective view

Project 11 » Ballista, Onager, and Trebuchet 221


Drawing 11.13 » Onager,
8 1/2"
1 1/4"
side elevation

0 2" 4" 6"

scale

3 3/4"
2 1/4"

7" 10 1/2"

6 3/4"
4 1/4"

Drawing 11.14 » Onager, rear elevation

8 3/4"
0 2" 4" 6"

scale

axis of Drawing 11.15 »


torsion spring 0 2" 4" 6" Onager, plan view
scale

222 
1' 1/8"

1" 3/4"

Plan View
Drawing 11.16 »
Onager, throwing
arm detail

1" 1/2"

1' 1"
Side Elevation

Notes:

1 The torsion spring of the onager is composed of 14 loops of 3/16-inch nylon rope, with an overall
diameter of 11/4 inches.

2 The steel washers used for the torsion spring are 21/2″-diameter grade 8 washers, available from
McMaster-Carr (product #98023A041).

3 The bronze washers used for the torsion spring are 2-inch oil-embedded thrust bearings, available from
McMaster-Carr (product #5906K525).

4 The onager’s throwing arm is made of maple. All other components are poplar.

5 The projectile holder is cut from a used plastic pill bottle.

6 The trigger is fabricated from 1/8-


inch 6061 aluminum sheet.

7 The frame that stops the motion


of the throwing arm must be
padded with foam rubber to
absorb the shock of the arm’s
impact.

8 Although the drawings do not


show a windlass, one can be added
to facilitate the process of cocking
the machine.

Project 11 » Ballista, Onager, and Trebuchet 223


THE TREBUCHET

The trebuchet was the principal heavy artillery weapon of the medieval era. It was a single-armed machine
that was powered by a falling counterweight, with its projectile velocity significantly increased through the
use of a sling. Medieval sources show two substantially different configurations: one with its counterweight
rigidly fixed to the end of the throwing arm and one with its counterweight suspended like a pendulum
from the arm. Modern mathematical analyses have demonstrated that the pendulum counterweight caused
a substantial improvement in mechanical efficiency.

The trebuchet model demonstrated in lesson 18 uses a pendulum counterweight and was designed so that
the gravitational potential energy stored in its raised counterweight is approximately equal to the elastic
energy stored in the torsion springs of the ballista and onager models. Thus, the trebuchet model has the
same theoretical maximum range as the two torsion catapults. The actual performance of the trebuchet is
comparable to that of the ballista, primarily because of the beneficial effect of the trebuchet’s pendulum
counterweight. Nonetheless, the trebuchet is—and must be—a much larger machine to achieve this
comparable range.

Drawings and construction notes for the trebuchet model are provided below.

throwing arm

frame

pendulum sling
counterweight

Drawing 11.17 » Trebuchet,


perspective view

projectile

224 
throwing arm

trigger fixed end


of sling
Drawing 11.18 » Trebuchet,
trigger and sling details

free end
prong of sling
trigger pivot

15 3/4" 48"

16 3/4"

Drawing 11.19 » Trebuchet,


34"
side elevation

19 1/2" 34 3/4" 0 2" 4" 6" 8"


scale

Project 11 » Ballista, Onager, and Trebuchet 225


4"

0 2" 4" 6" 8"


scale

Drawing 11.20 » Trebuchet, front elevation

34"

Drawing 11.21 »
Trebuchet, plan view
29 3/4"

0 2" 4" 6" 8"


scale

2"

4"
Drawing 11.22 » Trebuchet, pouch (cotton cloth)
sling detail
fold tabs over
nylon cord cord and glue
ring
pouch

fixed end free end

42" 42"
sling (not to scale)

226 
Notes:

1 The trebuchet frame, throwing arm, and counterweight are constructed entirely of poplar.

2 The counterweight is an open-topped wooden box filled with concrete. Its total weight should be
approximately 10 pounds.

3 The sling is fabricated from nylon string and cotton cloth, as shown in drawing 11.22. The fixed end
of the sling is tied to a screw eye driven into the bottom of the throwing arm near its tip. The free end
of the sling incorporates a ring, made from steel wire, which slips over a steel prong projecting from the
tip of the throwing arm, as shown in drawing 11.18. The prong is fabricated from a 1/8-inch-diameter
steel rod.

4 The trigger is fabricated from 1/8-inch 6061 aluminum sheet.

5 The pivot of the main throwing arm is an axle made from a 3/8-inch-diameter steel rod, turning in a
pair of bronze bushings (McMaster-Carr product #6391K176). The axle is fixed to the throwing arm
with epoxy glue; the bushings are mounted in the supporting wooden frames.

6 The pivot of the pendulum counterweight is an axle made from a 1/4-inch-diameter steel rod, turning in
a pair of bronze bushings (McMaster-Carr product #6391K142). The axle is fixed to the throwing arm
with epoxy glue; the bushings are mounted in the vertical arms of the pendulum counterweight.

7 Although not shown in the drawings, the trebuchet frame must be rigidly attached to a flat base with
screws or glue.

8 During testing, if the launch angle is too high (i.e., if the projectile releases at an angle significantly
greater than 45°), bend the steel prong at the tip of the throwing arm upward or lengthen the sling.

If the launch angle is too low (i.e., if the projectile releases at an angle significantly less than 45°), bend the
steel prong downward or shorten the sling.

Project 11 » Ballista, Onager, and Trebuchet 227


PROJECT 12
HYDRAULIC ARM
DESIGN, BUILD, AND TEST A HYDRAULICALLY
POWERED MECHANICAL ARM.
LESSON IN WHICH THIS PROJECT IS COVERED:
19 » Design a Hydraulic Arm

PROBLEM DEFINITION
Requirements
}} Using hydraulic power alone, the mechanical arm must be able to:
§§ Grasp and lift a 1/2-pound concrete block measuring 1 inch by 2 inches by 3 inches.
§§ Lift the block from position X to position Y (in the graphic below) and then place it anywhere
between positions X and Z.
§§ Rotate at least 90° about its own vertical axis.
§§ Operate with no more than 5 pounds applied manually to any of the hydraulic actuators.

Constraints: Y
}} Use off-the-shelf plastic syringes for hydraulic
cylinders and actuators.
}} Use readily available materials for all other
15"
components.
}} Use water as the hydraulic fluid.
Z X

2"
20"

228 
REFERENCES
Hibbeler, Engineering Mechanics.
White, Fluid Mechanics.

Exploring Hydraulic Machines


In general, a machine is a device consisting of a source of power and a means of controlling its application. In science
and engineering, power is defined as the rate at which work is done. And work is the transfer of energy associated with
a force moving through a distance.

We can say, then, that machines are used for the controlled application of force through controlled distances. And the
mechanism by which many such devices move and apply force is the subject of this project: hydraulics.

A hydraulic machine uses fluid pressure to transmit, amplify, and control the application of forces.

Fluid pressure, p, in a hydraulic device can be calculated as:

F
p=
A

where:

F = an applied force
A = area over which the force is distributed

It should be noted that fluid pressure in a hydraulic system also increases with the depth of the fluid, as a result of the
fluid’s own weight. However, for our project, this increased pressure is insignificant in comparison with the pressure
caused by applied forces; thus, it can be safely ignored.

A typical hydraulic machine is composed of one or more of each of the following components:

†† actuator, which applies pressure to fluid contained within the system


†† hydraulic cylinder, which uses system pressure to support a load or move this load through a controlled distance
†† hydraulic ram, the movable elements of a hydraulic cylinder
†† hydraulic line, which connects the actuator and hydraulic cylinder so that the fluid pressure is equal in both
†† hydraulic fluid, which serves as the medium for transmission of force through the system via fluid pressure.

Our hydraulic machine will use standard off-the-shelf syringes for both the actuators and hydraulic cylinders, plastic
tubing for the hydraulic lines, and water as the hydraulic fluid. For a manually operated system with this configuration,
the hydraulic rams can be extended and retracted by applying pressure and suction, respectively, to the actuators.

Real-world hydraulic machinery works in much the same way as our model, except that system pressure is usually
supplied by a pump rather than a manually operated actuator. The pump only provides pressure—not suction.
Consequently, a reversing valve must be incorporated into the system so that the hydraulic rams can be both extended
and retracted.

Project 12 » Hydraulic Arm 229


DESIGN CONCEPT
GRIPPER
OUTER ARM

INNER ARM

ACTUATORS

COLUMN

BASE

Key features:

}} The base supports the column at its top and bottom and incorporates bearings that facilitate the
column’s rotation.
}} The inner arm connects to the column at a hinged joint called the shoulder.
}} The outer arm connects to the inner arm at a joint called the elbow.
}} The gripper is capable of grasping, holding, and releasing the concrete block on command.
}} The system uses four separate hydraulic cylinders: one to rotate the column about its own axis, one to
rotate the inner arm about the shoulder joint, one to rotate the outer arm about the elbow, and one to
operate the gripper.
}} The four associated actuators will be mounted on a standalone control station and connected to the
cylinders with plastic tubing.

DETAILED DESIGN
Selection of Hydraulic Cylinders

The following three off-the-shelf plastic syringes are available for this project:

SIZE PLUNGER AREA RANGE OF MOTION


12 mL 0.33 in2
2.3 in
20 mL 0.45 in2
3 in
35 mL 0.66 in2 3.4 in

230 
Hydraulic systems can be used to provide mechanical advantage by pairing an actuator that has a small
plunger area with a hydraulic cylinder that has a larger area. The mechanical advantage is the ratio of the
two plunger areas. Thus, using the 20-milliliter syringe as an actuator and a 35-milliliter syringe as the
associated hydraulic cylinder would yield a mechanical advantage of:

Similarly, using the 12-milliliter and 35-milliliter syringes would result in:

But there is a major pitfall associated with both of these options. Mechanical advantage always entails
a trade-off between force and distance. Thus, if we couple a 12-milliliter actuator with a 35-milliliter
hydraulic cylinder, we will double the applied force but also halve the distance through which the hydraulic
ram moves. The total range of motion of a 12-milliliter syringe is only 2.3 inches. Thus, for this pairing
of syringes, we would only be able to move the hydraulic ram 1.15 inches, even though this 35-milliliter
syringe is actually capable of moving three times that distance.

This is important because the design requirements for our mechanical arm specify a horizontal reach of 20
inches, and this reach must be achieved entirely through the action of just two hydraulic cylinders: one at
the shoulder and one at the elbow. Achieving a 20-inch reach with two cylinders that can move only 1 inch
each would be quite difficult, and the resulting design would have excessively high forces in the hydraulic
cylinders—even after accounting for the mechanical advantage of the hydraulic system.

Thus, for this project, we should not select syringes that provide mechanical advantage, because the cost
in reduced range of motion would exceed the benefit of multiplying the applied force. Rather, given the
demanding design requirement for a total reach of 20 inches, we should use pairings of syringes that give
us the largest-possible range of motion. On this basis, we will use the 35-milliliter syringe for all of our
actuators and hydraulic cylinders.

This selection has two important implications for our design:

1 We can design for the full 3.4-inch range of motion from each hydraulic ram.

2 Because the design requirements specify a maximum force of 5 pounds applied to the actuators (and
because the hydraulic system will not be providing any mechanical advantage), we will need to design
the arm such that the force developed in all of our hydraulic cylinders never exceeds 5 pounds.

Project 12 » Hydraulic Arm 231


Mounting the Hydraulic Cylinders

We must now devise a system for mounting these syringes in the mechanical arm in a manner that allows
for transmission of a 5-pound force—either pulling or pushing—and free rotation at both ends. We will
accomplish this task by designing end caps for the syringe and then fabricating these fittings through the
use of 3-D printing (see “3-D Printing” sidebar).

Pictured below is our SketchUp-based solution to the challenge of mounting off-the-shelf syringes in our
mechanical arm: a pair of plastic end caps, one that fits over the syringe’s tip and one that fits over the
end of the plunger. Each cap incorporates a cylindrical pivot—called a trunnion—that will serve as the
connector to the mechanical arm. On the plunger cap, the trunnion is placed on the center line of the
syringe, but on the other cap, it must be offset to provide clearance for the syringe tip, which protrudes
through the cap.

syringe The original plan for securing these caps to


trunnion
the syringe was to glue the tip cap in place
and to secure the plunger cap with a small
screw. (The small hole in the plunger cap
was intended to accommodate this screw.)
In practice, however, both caps fit so well
that no adhesive or mechanical connection
trunnion was needed to hold them in place.

With both caps installed on a syringe, the


center-to-center distance between the two
trunnions is 5 inches with the plunger 3-D Printing
retracted and 8.4 inches with the plunger
fully extended. These dimensions are Also known as additive manufacturing, 3-D
important for the design of the printing is a process by which a 3-D computer model
mechanical arm. is translated into a physical three-dimensional object
by building up layers of material—typically plastic
or metal—under precise computer control.
With the SketchUp computer model
complete, the two end caps are each The SketchUp software we have been using throughout
exported to a separate file in the DAE this course can be used to create models that are fully
format, which is compatible with most compatible with the 3-D printing process.
3-D printers. These files are then uploaded
to the Shapeways website (https:// If you don’t happen to have a 3-D printer, there
are several online 3-D printing services that will
www.shapeways.com/), and the desired
happily do the job, while also providing quality
material—called “strong and flexible control checks to ensure that the design satisfies the
plastic”—is specified. Shapeways provides stringent modeling requirements for 3-D printing.
detailed instructions and excellent tutorials
for developing models that are properly The 3-D printing service used for this project is
configured for 3-D printing (see https:// Shapeways (https://www.shapeways.com/).
www.shapeways.com/tutorials). These

232 
stringent modeling requirements are checked by automated tools when the design is uploaded. Once the
design has passed all required checks, it is sent to the printer for fabrication. The printing process typically
takes about two weeks.

Design of the Mechanical Arm

In designing the mechanical arm, we must simultaneously satisfy three closely related criteria:

}} The arm must have an operating envelope—an overall range of motion—that satisfies the design
requirements.
}} The hydraulic cylinders—operating between their minimum and maximum lengths of 5 inches and
8.4 inches—must provide the required range of rotation at the shoulder and elbow joints.
}} The forces developed in the hydraulic cylinders cannot exceed 5 pounds while lifting the 1/2-pound
load at any position within the operating envelope.

Satisfying all of these criteria simultaneously will require a series of trial-and-error analyses that are best
accomplished with a computer spreadsheet.

To formulate a spreadsheet-based analysis of our mechanical arm, we begin with a mathematical model—
an idealized graphical representation with all key joints designated as A through H and all dimensions and
angles defined as variables.

y LFG
LEF LGH
LDE F
LCD θF dG
θ’F
dE
dD G H
C E Cartesian coordinates
θC D
joint x-coordinate y-coordinate
A 0 0
LBC
B dB LAB
C 0 LAB + LBC
dB D LCD sinθ C − d D cos θ C y C − LCD cos θ C − d D sinθ C
B y C − (LCD + LDE ) cos θ C − d E sinθ C
E (LCD + LDE ) sinθ C − d E cosθ C
LAB F (LCD + LDE + LEF ) sinθ C y C − (LCD + LDE + LEF ) cos θ C
G x F + LFG sinθ F′ − d G cosθ F′ y F − LFG cos θ F′ − d G sinθ F′
A
x H x F + (LFG + LGH ) sinθ F′ y F − (LFG + LGH ) cos θ F′
(θ F′ = θ C + θ F − 180° )

Note that this model allows for the hydraulic cylinders—designated BD and EG—to be offset from the
center line of the arm, with the offset distances defined as dB , dD, dE , and dG .

Project 12 » Hydraulic Arm 233


The Cartesian coordinates shown on this graphic are paired x- and y-values defining the position of each
joint in two-dimensional space with respect to an origin (x = 0, y = 0). For this model, the origin is defined
at joint A, the base of the column. Thus, the coordinates of all other joints are defined with respect to A.

For example, joint B is located at a distance dB to the right of joint A and at a distance L AB above A; thus,
its coordinates are x = dB and y = L AB . Joint C is located at the same horizontal position as joint A and at a
distance L AB + LBC above A. Thus, its coordinates are x = 0 and y = L AB + LBC.

Beyond joint C, the coordinates get significantly more


complicated, because the positions of these joints depend y
not just on the arm lengths and offsets, but also on the
angles at the shoulder and elbow, designated as θC and LCDx
θF , respectively. Thus, all of these joint positions must be C
defined with trigonometric functions. θC
LCDy
To illustrate this process, the formulation of Cartesian LCD θC
coordinates for joint D is shown here. This diagram shows
only segment CD, the innermost segment of the inner arm. D’
The shoulder angle θC is shown as an acute angle (less than dD
90°) so that the relevant trigonometric relationships will θC
dDy
be clearer. Note also that line segments CD’ and DD’ can D
each be visualized as the hypotenuse of a right triangle, the dDx
dimensions of which are annotated on the diagram.
x
Based on these dimensions, the x-coordinate of joint D can
be determined as follows:

Similarly, the y-coordinate is:

234 
In developing these Cartesian coordinates, we have created a set of mathematical expressions that will
describe the precise position of the arm for any dimensions and angles we choose. These expressions are
very easily incorporated into a spreadsheet, which must also perform two additional calculations that are
critical to the design:

}} By applying the Pythagorean theorem, it must calculate the length of each hydraulic cylinder for any
arm position.
}} By applying the principle of equilibrium, it must calculate the required force in each cylinder for any
arm position.

This spreadsheet is available for download from http://stephenjressler.com/diy-engineering/. This


computational tool is used to facilitate the design of the mechanical arm as follows:

}} Choose a trial set of dimensions for the arm.


}} At both the shoulder (joint C) and the elbow (joint F), determine the allowable range of angles, θC
and θF , corresponding to cylinder lengths ranging from 5 inches to 8.4 inches.
}} Check whether the gripper (joint H) can be moved to positions X, Y, and Z (as defined in the design
requirements) within the allowable range of angles for both θC and θF .
}} Check that the force in both hydraulic cylinders never exceeds 5 pounds for any pair of angles, θC
and θF , within the allowable ranges.
}} If the arm cannot reach positions x, y, and z, or if the force in either cylinder exceeds 5 pounds, adjust
the dimensions accordingly and repeat these checks.

In making these adjustments, it is helpful to recognize that:

}} Maximum cylinder forces tend to occur at the extremes of the arm’s range of motion.
}} The force in a cylinder can generally be reduced by moving that cylinder farther away from its
associated joint. For example, the force in cylinder BD can be reduced by increasing LCD or dD, both
of which will move the cylinder farther away from joint C through most of its range of motion.

Through this systematic trial-and-error process, the 2" 9.5"


geometric design solution below was achieved. (It 6.75"
is worth noting, however, that many other fully 2.25"
50° y
successful design configurations are possible.) 4.0"

150°
115°
5.75" 65°

50°

5.75"
150°
z x

Project 12 » Hydraulic Arm 235


If we plot every possible position of the tip of the outer arm (joint H), the result is a graphical representation
of the arm’s operating envelope, as shown below. Given that points X, Y, and Z from the design
requirements all fall within this envelope, we can be confident that this design will work.

z x

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer models for the complete mechanical arm and for the syringe end caps can be
downloaded from http://stephenjressler.com/diy-engineering/. A set of full-size cutting patterns is provided in
the PDF version of this guidebook. The layout drawing measures 24 inches by 18 inches and must be printed
at exactly this size. This can be accomplished by a commercial printing establishment for just a few dollars.

Drawing 12.1 » Mechanical


arm, perspective view

236 
a
j b b
b
i c c
a R R J
I gripper
C H (see Drawing 12.3)
G K
F z
b z
G
H L
A
z D J
a
c c b b
a E
D b
M
c j
k
P
c b i
O B Drawing 12.2 » Mechanical
N arm, exploded view
z O
Q
c c

d d e
U e

Drawing 12.3 » Gripper, U


exploded view W
T
g
S
V
g g
V W
T d
d g
h
W

nut
washer
outer brass tube

inner brass tube


Drawing 12.4 » Typical
connection, exploded view washer

screw

Project 12 » Hydraulic Arm 237


Y

Drawing 12.5 » Actuator station, exploded view

MATERIALS LIST
COMPONENT DESCRIPTION**
A Post 3/4″ poplar*
B Base 3/4″ poplar*
C Top pivot mount 3/4″ poplar*
D Column side (2) 1/8″ plywood* (SIG #SIGPW015)
E Column bottom spacer 3/4″ poplar*
F Column top spacer 3/4″ poplar*
G Inner arm spacer (2) 3/4″-diameter hardwood dowel, 13/8″ long, with 3/16″ hole drilled
through the center
H Inner arm side (2) 1/8″ plywood* (SIG #SIGPW015)
I Outer arm spacer 1/4″ poplar*
J Outer arm side (2) 1/8″ plywood* (SIG #SIGPW015)
K Bell crank mount 1/4″ poplar*
L Outer arm spacer 3/4″ × 1″ × 15/8″ poplar, rounded
M Column lower crank 1/8″ plywood* (SIG #SIGPW015)
N Column upper crank 1/8″ plywood* (SIG #SIGPW015)
O Crank mount spacer (2) 1/4″ poplar*
P Upper cylinder mount 1/8″ plywood* (SIG #SIGPW015)
Q Lower cylinder mount 1/8″ plywood* (SIG #SIGPW015)
R Cylinder spacer (6) 1/8″ plywood* (SIG #SIGPW015)
S Slider top 1/4″ poplar*
T Slider side (2) 1/8″ plywood* (SIG #SIGPW015)
U Pushrod (2) 1/8″ plywood* (SIG #SIGPW015)
V Bell crank (2) 1/4″ plywood* (SIG #SIGPW021)
W O-ring mount (4) 1/2″-diameter hardwood dowel, 3/8″ long, with 1/8″ hole drilled
through center

238 
X Actuator station base 3/4″ × 41/4″ × 12″ MDF, plywood, or poplar
Y Actuator station back 1/4″ plywood* (SIG #SIGPW021)
Z Actuator station front 1/4″ plywood* (SIG #SIGPW021)
a Pivot, arm connection (2) See Schedule of Connections
b Pivot, hydraulic cylinder (4) See Schedule of Connections
c Pivot, hydraulic cylinder (4) See Schedule of Connections
d Pivot, pushrod and bell crank (6) See Schedule of Connections
e O-ring retainer (2) See Schedule of Connections
g O-ring Rubber O-ring, 7/16″ inside diameter (McMaster-Carr #2418T136)
h Slider pin (2) 3/16″-diameter hardwood dowel, 1/2″ long
i Sleeve bearing Bronze sleeve bearing for 1/4″ shaft (McMaster-Carr #6391K142)
j Column pivot 1/4″-diameter steel rod
k Thrust bearing Bronze thrust bearing (McMaster-Carr #5906K531)
z Hydraulic cylinder Syringe with end caps (Pitsco Education #52526)
- Hydraulic line Flexible plastic tubing (Pitsco Education #38980)

*Cutting pattern provided in PDF version of this guidebook.


**For hard-to-find items, product numbers are provided for appropriate items at the following online vendors:
SIG Manufacturing (http://www.sigmfg.com/)
McMaster-Carr (https://www.mcmaster.com/)
Pitsco Education (http://www.pitsco.com/)

SCHEDULE OF CONNECTIONS

COMPONENT # REQ’D SCREW SIZE INNER BRASS TUBE OUTER TUBE


a 2 4-40 × 21/2″ 5/32″ outside diameter, 3/16″ outside
17/8″ long diameter,
11/2″ long
b 4 4-40 × 21/2″ 5/32″ outside diameter, none
17/8″ long
c 4 4-40 × 2″ 5/32″ outside diameter, none
15/8″ long
d 6 4-40 × 3/4″ 5/32″ outside diameter, 3/16″ outside
3/8″ long diameter,
1/4″ long
e 2 4-40 × 11/2″ none none

Note: All required screws, nuts, washers, and brass tubing can be obtained from McMaster-Carr (https://www.mcmaster.com/).

Project 12 » Hydraulic Arm 239


In addition to common hand tools, this project requires the following equipment:

}} miter saw or radial arm saw (for crosscutting)


}} scroll saw or bandsaw (for curved cuts)
}} power-sanding station (for shaping and tapering)
}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} files
}} clamps

Other supplies needed for this project include wood glue, epoxy, cyanoacrylate glue, and spray adhesive.

FABRICATION AND ASSEMBLY PROCEDURE

1 Fabricate plastic end caps for the four hydraulic cylinders using the 3-D printing process as described
previously. The SketchUp model for the end caps is provided at http://stephenjressler.com/diy-engineering/.

2 Prefabricate all of the wooden components specified in the materials list above. For components that
have cutting patterns, use spray adhesive to affix the patterns to wood of the appropriate type and
thickness and then cut to size. For components that do not have cutting patterns, simply use the
dimensions provided in the materials list. Shape with a sander as needed. Drill holes as precisely as
possible, as indicated on the cutting patterns.

3 Prefabricate all brass tubes for the hinged connections, as specified in the schedule of connections.

4 Cut the two column pivots (component j) from a 1/4-inch-diameter steel rod.

5 Fabricate the four hydraulic cylinders by installing 3-D-printed plastic end caps on four 35-milliliter
syringes.

6 Using wood glue and clamps, assemble


G
the column, inner arm, and outer arm, R R
I K
as shown at right. Note that spacers
L
(component R) are glued to the inside F G H J
faces of the column and outer arm at
the holes where the hydraulic cylinders R
will be fastened. These spacers will D
ensure that the cylinders remain
6
E
centered within the arm.
M

240 
7 Connect the arms and column together at the shoulder and elbow joints, as shown below, using two
pivots (component a), as specified in the schedule of connections and as illustrated in drawing 12.4. For
each connection, ensure that the tightened screw and
nut bear only on the inner brass tube so that the
outer tube can rotate freely.

a 7

8 Assemble the gripper as shown in drawing 12.3. Glue the slider sides (component T) to the slider top
(component S), but do not install the slider pins (component h) at this time. Install the two pushrods
(component U) on the slider top (component S) using two pivots (component d). Install the inner ends
of the two bell cranks (component V) to the bell crank mount (component K) on the outer arm using
two pivots (component d). For both of these connections, the screws should be inserted upward, with
the nut installed on top, to provide adequate clearance for the hydraulic cylinder below. Fasten the four
rubber O-rings (component g) to the front ends of the bell cranks using components W and e.

9 Slide the gripper into position within the outer arm. Line up the holes in the slider sides (component
T) with the corresponding slots in component J. Then, glue the two slider pins (component h) into
the rear holes in the slider sides such that they project through the slots, as shown below. Connect the
gripper pushrods (component U) to the elbows of the bell cranks (component V) using two pivots
(component d).

d V 10 Install the hydraulic cylinder for the


U
gripper using two pivots (component b).
d The forward pivot must pass through
the forward slots in the outer arm sides
U V (component J), the forward holes in
J b the slider sides (component T), and the
forward trunnion of the hydraulic cylinder.

b slot
h 11 Install the hydraulic cylinders at the
9 shoulder and elbow joints using pivots
(components b and c), as indicated in
drawing 12.2.

Project 12 » Hydraulic Arm 241


12 Build the base. First, glue and clamp the cylinder mount, consisting of components P, Q, and O, to
the post (component A), as shown in drawing 12.2. Then, glue and clamp components A, B, and C
together, ensuring that the post (component A) is perfectly vertical and the pivot holes in components
B and C are aligned. Reinforce these joints with wood screws.

13 Use epoxy to glue bronze sleeve bearings (component i) into the large holes in components B and C.

14 Glue the upper crank and its spacer (components N and O) to the column base.

15 Glue the lower column pivot (component j) into bottom of the column (components E and M) with
epoxy, allowing 3/4 inch of the pivot to project below the bottom surface of component M.

16 Slide the bronze thrust bearing (component k) onto the lower column pivot (component j); then, insert
the pivot into the lower bronze sleeve bearing (component i) in the base.

17 Insert the upper column pivot (component j) through the upper bronze sleeve bearing (component i)
and into the top of the column (component F). Secure it to the column with a drop of cyanoacrylate
glue applied from below.

18 Install the final hydraulic cylinder in the base, connecting its lower trunnion to the cylinder mounts
(components P and Q) and its upper trunnion to the upper and lower cranks (components M and N)
on the base of the column using two pivots (component c).

19 Construct the actuator station, as shown in drawing 12.5, and install the four actuators.

20 Connect the four actuators to the four hydraulic cylinders with plastic tubing.

The hydraulic arm is now complete and ready for testing!

TESTING PROCEDURE

In preparing the hydraulic arm for operation, the greatest challenge is filling the hydraulic system with
water without also introducing large air bubbles into the system. Air is significantly more compressible than
water, so air bubbles in the hydraulic system will significantly hamper its performance.

242 
To fill the hydraulic system with water:

}} Disconnect the plastic tubes from all four hydraulic cylinders.


}} Manually manipulate the mechanical arm and gripper such that all four hydraulic cylinders are fully
retracted.
}} Manually retract all four actuator plungers.
}} For each actuator, dip the end of the connected plastic tube into a basin of water. Withdraw the
actuator plunger slowly until the tube and syringe are filled with water.
}} There will probably be a large air bubble inside the syringe as well, because the tube was initially
filled with air. To remove this bubble, raise the end of the plastic tube above the elevation of the
actuator and slowly retract the plunger, expelling water while also allowing the bubble to float up to
the top of the tube. Once the air bubble has been expelled, insert the end of the tube back into the
basin of water and withdraw the actuator plunger to refill the syringe.
}} Repeat this process until the actuator and tube are completely filled with water, with the minimum-
possible amount of air. Now attach the end of the tube to its associated hydraulic cylinder.

Once the hydraulic system is filled, screw or clamp the bases of the mechanical arm and actuator station to
a firm, level surface. Then, operate the machine by extending and retracting the appropriate actuators.

Project 12 » Hydraulic Arm 243


PROJECT 13
WATER TURBINE
DESIGN, BUILD, AND TEST A WATER TURBINE.
LESSON IN WHICH THIS PROJECT IS COVERED:
20 » Make a Water Turbine

PROBLEM DEFINITION
Requirements
}} Must be capable of lifting a 2.2-pound weight through a vertical distance of 2 feet in the minimum-
possible time.
}} Power output must be measurable.

The Turbine
Constraints:
}} Use an elevated water reservoir as the source of energy. Of all the mechanical devices that
}} Use readily available materials for all components. power our technology-intensive
world, none is more important
than the turbine—a rotary
machine that uses a moving fluid
to do work. There are many types
of turbines, their most important
REFERENCES distinguishing characteristic
being the type of fluid that is used
Hibbeler, Engineering Mechanics. to power the device. Turbines
powered by steam, natural gas,
Roberts, Making Things Move. water, and wind generate most of
White, Fluid Mechanics. the world’s electric power. And gas
turbines are also used extensively
as engines for aircraft, ships,
and locomotives.

244 
DESIGN CONCEPT
RESERVOIR

Key features: RUNNER

}} This configuration is an impulse


turbine, which uses a vertically oriented
wheel, called a runner, with radial blades
extending outward from its rim. For
simplicity, the blades will be flat plates. PENSTOCK
}} The runner will spin on a 3/8-inch- BASIN
diameter steel axle.
}} The runner will be powered by a jet of LOAD
water directed from an outlet pipe called
a penstock.
}} The penstock will be fabricated from 3/4-inch-diameter polyvinylchloride (PVC) pipe with a 1/2-inch-
diameter nozzle at its end to concentrate the flow.
}} The water source will be an elevated reservoir made from a 30-gallon plastic trash can.
}} A large catch basin will be placed below the wheel to capture the water after it passes through the turbine.
}} The 2.2-pound load will be lifted by a length of nylon cord wound around the turbine shaft.

DETAILED DESIGN

The detailed design of this machine will address three questions:

}} How large and how high must the reservoir be to provide the water jet with enough energy to power
the turbine?
}} What diameter should the turbine runner be for optimum performance?
}} How much time will it take for the optimally designed turbine to lift a 2.2-pound weight through a
distance of 2 feet?

Reservoir Size and Height

The mathematical relationship between velocity, pressure, and elevation at two points in a fluid flow is
described by Bernoulli’s equation, as follows:

Project 13 » Water Turbine 245


1
where:

h1

To apply this equation, we first designate two key points in


the flow: point 1 on the surface of the reservoir and point v2 2
2 just outside the nozzle from which the water jet is h2=0
flowing, as shown at right.

Both points 1 and 2 are at the same pressure—atmospheric pressure. Thus, the pressure terms on both sides of
the equation cancel out. Furthermore, if the reservoir is very large, then the drop in its surface elevation due to
the outflow of water will be so slow that the velocity v1 is effectively zero. And if we designate the elevation of
the nozzle as the base level, then h2 also equals zero. Consequently, Bernoulli’s equation reduces to:

Solving for v2, the flow velocity at the nozzle:

This solution has two interesting implications:

}} The size of the penstock has no effect on the velocity of the water flowing from it.
}} The only characteristic of the reservoir that matters is the elevation of its surface.

The latter observation suggests that, in theory, v2 is independent of the reservoir size. Thus, we could
simplify our project by using a reservoir much smaller than the 30-gallon trash can shown in our concept
sketch. In practice, however, using a small reservoir would be problematic in two ways:

1 Recall that the solution to Bernoulli’s equation relied on the assumption that v1, the velocity of the
reservoir surface, is essentially zero. This is a very reasonable assumption for a large reservoir but quite
incorrect for a small one.

2 Using a small reservoir would require a long penstock extending all the way down to the nozzle. Water
flowing through a pipe always experiences friction, and the energy losses due to friction increase with the
length of the pipe. Thus, using a small reservoir would significantly reduce the velocity of the water jet.

246 
For these reasons, we will use the 30-gallon trash can.

Now that we have made this decision, it is important to recognize that even this large reservoir will be
subject to some real-world energy losses that are not accounted for in the theoretical solution to Bernoulli’s
equation. As water flows from reservoir to nozzle, it will still experience some friction—as well as some
turbulence at both the nozzle and the connection between the penstock and the reservoir. The resulting
energy losses are typically accounted for by appending an empirical coefficient to the velocity calculation.
We will use a value of 0.8 to account for the rather crude construction of our reservoir and penstock.

For h1, the reservoir elevation, we will use 2.5 feet. In general, greater height results in more potential
energy, which translates into higher velocity for the water jet. However, from a practical perspective, a
reservoir higher than 2.5 feet would be quite difficult to fill because the turbine apparatus will be mounted
on a tabletop.

Substituting this value into Bernoulli’s equation:

Power Output and Runner Diameter

By definition, power is the rate at which work is done. Work is a form of energy transfer that is associated
with a force moving through a distance. Thus:

Substituting the definition of velocity:

turbine runner

The idealized model we will use to estimate the


power output of our turbine is shown here. The water jet
model represents a single turbine blade at the vB
FB vJ
instant when it is oriented vertically at the bottom
area=AJ
of the runner. The water jet, with cross-sectional
area AJ, is moving at velocity vJ and is impacting blade
the blade, which is also moving at velocity vB .
After striking the blade, the water jet is deflected
in all directions and flows off the blade laterally.

Project 13 » Water Turbine 247


By applying the principle of conservation of momentum, the force FB applied by the water jet to the blade
can be calculated as:

where:

The power produced by this force is: r2

The weight that can be lifted by this force is called the


load and is represented by the variable W in the free body r1
W
diagram to the right.

As shown in this diagram, the force FB and the force W each


apply a torque to the runner shaft. FB

Applying the principle of equilibrium, the magnitudes of these two torques must be equal. Thus:

Using the equations above, we can now plot graphs of load versus runner speed and power versus runner
speed for a runner of any diameter. Runner speed is defined in revolutions per minute (rpm). For a given
runner speed, in rpm, the associated blade velocity, vB , is:

248 
In this equation, r is the radius of the turbine runner measured from its axis to the point at which the center
of the water jet impacts the blade. For example, for a 1-foot-diameter runner spinning at a rotational speed
of 10 rpm:

Four graphs of load and power versus rotor speed are presented below. Each of these graphs was generated
as follows:

}} Define a range of rotational speeds from a minimum of zero to a maximum corresponding to vB =


10.1 feet per second—the velocity of the water jet.
}} For each rotational speed, calculate the blade velocity, vB.
}} For each calculated value of vB , calculate power and load using the equations developed above.
}} Plot power versus rotational speed and load versus rotational speed on the same graph.

0.5 6-inch runner diameter 10 0.5 8-inch runner diameter 10

0.41 ft-lb/sec 0.41 ft-lb/sec


0.4 8 0.4 8
power power
Power (ft-lb/sec)
Power (ft-lb/sec)

0.3 6 0.3 6

Load (lb)
Load (lb)

load
0.2 4 0.2 4
load
optimum load = 2.5 lb
0.1 optimum load = 1.9 lb 2 0.1 2

0 0 0 0
0 100 200 300 0 100 200 300
Rotational Speed (rpm) Rotational Speed (rpm)

0.5 10-inch runner diameter 10 0.5 12-inch runner diameter 10

0.41 ft-lb/sec 0.41 ft-lb/sec


0.4 8 0.4 8
power power
Power (ft-lb/sec)
Power (ft-lb/sec)

0.3 6 0.3
Load (lb)

6
Load (lb)

optimum load = 4.2 lb


0.2 4 0.2 4
optimum load = 3.1 lb

load load
0.1 2 0.1 2

0 0 0 0
0 100 200 300 0 100 200 300
Rotational Speed (rpm) Rotational speed (rpm)

Project 13 » Water Turbine 249


These graphs clearly illustrate the following key points about turbine performance:

}} The speed and power of a turbine are determined not just by the water jet applying force to its blades,
but also by the load placed on its shaft.
}} There is a critical load above which the turbine stalls and produces no power.
}} As the load is reduced below this critical level, power production increases, because the reduced load
allows the runner to rotate faster.
}} As load is reduced, this increase in power only continues to a point, represented by the peak of the
power curve. Beyond this point, power production drops off as the beneficial effect of higher turbine
speed is increasingly offset by the adverse effect of reduced force on the turbine blades.
}} When the load is removed entirely, the blades move at the same velocity as the water jet. Even though
this is the highest speed the runner is capable of achieving, it produces no power, because the force on
the blades is zero.
}} The peak power output (0.409 foot-pounds per second = 0.00075 horsepower) is identical for all four
runner diameters. Evidently, peak power depends only on the characteristics of the water jet, not on
the turbine size.
}} Lifting capacity increases significantly with runner size, because a larger runner produces more torque
for a given blade force.
}} Most importantly, every turbine has an optimal load at which power production is maximized.

These optimum loads are as follows:

RUNNER DIAMETER OPTIMUM LOAD


6 inches 1.9 pounds
8 inches 2.5 pounds
10 inches 3.1 pounds
12 inches 4.2 pounds

Clearly, the theoretically optimal runner diameter for lifting our 2.2-pound weight is just under 8 inches.
We will use 8 inches for simplicity.

Required Lift Time

Solving the basic definition of power in terms of time:

250 
Substituting the required load (2.2 pounds), the required lift distance (2 feet), and the turbine’s peak power
output (0.409 foot-pounds per second), the theoretical lift time is:

Number of Runner Blades

The three scale drawings below illustrate 8-inch-diameter runners with 8, 10, and 12 blades being impacted
by a 1/2-inch water jet. In all three cases, optimum transmission of force from the jet to the runner occurs
when a blade—indicated as the “target blade”—is oriented vertically at the bottom of the runner. Note that
for the 12-blade runner, the blade following the target blade impinges on the water jet and interrupts its
flow, reducing the force on the target. And for the 8-blade runner, there is a sizable gap between the water
jet and the following blade, resulting in the target blade having to rotate well beyond its optimum vertical
orientation before the following blade intercepts the jet.

8 blades 10 blades 12 blades

target blade water jet

Given these observations, we will use a 10-blade runner for our turbine design.

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/
diy-engineering/. Full-size cutting patterns for wooden components are provided in the PDF version of this
guidebook. These drawings are sized for printing on standard 8 1/2-inch-by-14-inch (legal size) paper.

Project 13 » Water Turbine 251


Drawing 13.1 » Water
turbine, perspective view

Q
Drawing 13.2 » Water P
turbine, exploded view K
H
J
G
W C R
I
O M
D A
V
V
F
B E V
U C
T D
S W

e
d
e
b
c
b a
Drawing 13.3 » Jig for
fabrication of turbine runner

252 
MATERIALS LIST
COMPONENT DESCRIPTION
A Runner disk 1/2″ hardwood plywood*
B Runner blade (10) 3/32″ × 3/4″ × 11/2″ poplar
C Runner mount (2) 3/4″ poplar*
D Spacer (2) 1/4″ poplar*
E Penstock mount, front 3/4″ poplar*
F Penstock mount, rear 3/4″ poplar*
G Penstock mount, top 3/4″ × 2″ × 53/4″ poplar
H Runner mount, rear 3/4″ × 2″ × 23/4″ poplar
I Platform, front 3/4″ × 121/4″ × 143/4″ plywood with notch for component M and 11/4″
hole for penstock
J Platform, center 3/4″ × 121/4″ × 16″ plywood
K Platform, rear 3/4″ × 121/4″ × 143/4″ plywood with notch for component M
L Platform, top 3/4″ × 18″ × 18″ plywood with notch for penstock
M Base, longitudinal rail 3/4″ × 11/2″ × 24″ poplar
N Water reservoir 30-gallon plastic trash can
O Lower penstock 3/4″ PVC pipe (McMaster-Carr #49035K24), 10″ long
P Upper penstock 3/4″ PVC pipe (McMaster-Carr #49035K24), 71/4″ long
Q Penstock-reservoir PVC straight adapter with hex body, 3/4″ socket female, NPT male
connection (McMaster-Carr #4880K62)
R Elbow connection 90° elbow for 3/4″ PVC pipe (Mcmaster-Carr #4880K22)
S Nozzle 3-D-printed from SketchUp model provided at http://stephenjressler.
com/diy-engineering/
T Stopper Tapered rubber plug, 1/2″ outside diameter (McMaster-Carr
#6448K93)
U Runner shaft 3/8″ diameter steel rod, 8″ long
V Shaft collar (3) Set screw shaft collar, 3/8″ diameter (McMaster-Carr #9414T8)
W Shaft bearing (2) Flanged ball bearing, open, for 3/8″ shaft diameter (McMaster-Carr
#6383K232)
- Lifting cord Nylon string
2.2-lb load Slotted mass set (PASCO ME-7589)**
a Jig core 1/2″ hardwood plywood*
b Jig end (2) 3/4″ × 4″ × 81/2″ plywood or MDF
c Jig base 3/4″ × 41/2″ × 81/2″ plywood or MDF
d Jig side 3/4″ × 4″ × 61/2″ plywood or MDF
e Connector (2) 1/4″ × 3″ machine bolt with washers and wing nut

*Cutting pattern provided in PDF version of this guidebook.


**Available from PASCO Scientific (https://www.pasco.com/).

Project 13 » Water Turbine 253


In addition to common hand tools, this project requires the following equipment:

}} table saw or circular saw (for rip-cutting)


}} miter saw or radial arm saw (for cross-cutting)
}} scroll saw or bandsaw (for curved cuts)
}} power-sanding station (for shaping)
}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} files
}} Allen wrenches
}} clamps

Other supplies needed for this project include wood glue, epoxy, PVC pipe glue, silicone sealant, and
spray adhesive.

FABRICATION AND ASSEMBLY PROCEDURE

1 Fabricate the penstock nozzle (component S) using the 3-D printing process described in project 12.
The SketchUp model for the nozzle is provided at http://stephenjressler.com/diy-engineering/.

2 Prefabricate all of the wooden components specified in the materials list above. For components
that have cutting patterns, use spray adhesive to affix the patterns to wood of the appropriate type
and thickness; then, cut to size. For components that do not have cutting patterns, simply use the
dimensions provided in the materials list. Shape with a sander as needed. It is especially important to
achieve a high degree of precision in fabricating the runner disk (component A) and drilling its 3/8-inch
shaft hole. Do not cut the blade slots in the runner disk at this time, and do not remove the paper
cutting pattern from this component.

3 Fabricate the runner shaft (component U) by cutting an 8-inch length of 3/8-inch-diameter steel rod
and filing its ends smooth.

4 Use a hole saw or spade bit to cut a 1-inch-diameter hole in the bottom of a 30-gallon plastic trash can
(component N). Determine the hole location carefully. It should be located on a relatively flat portion
of the can’s bottom surface so that the pipe fitting (component Q) can be properly sealed. And it
should be located approximately 5 inches radially outward from the center of the can’s circular bottom
so that it will line up with the penstock when the trash can is centered on its supporting platform
(component L).

5 Fabricate the penstock tubes (components O and P) from 3/4-inch PVC pipe as indicated in the
materials list above. The lengths of these components might need to be adjusted somewhat, depending
on the location of the hole in the trash can.

254 
6 Build the jig illustrated in drawing 13.3. After the jig is fully assembled, drill two 1/4-inch holes for the
bolts that will be used to clamp the runner disk inside the jig. The central slot in the jig is cut on the
table saw during the jig’s first use.

7 Following the procedure demonstrated in lesson 20, clamp the runner disk (component A) into the jig
and cut the 10 equally spaced radial slots for the turbine blades. Ensure that the table saw’s blade height
is set such that the slot cut into the runner disk is 1/2-inch deep.

8 Using a waterproof adhesive (epoxy or polyurethane glue), glue the 10 blades (component B) into the
slots in the runner disk, as demonstrated in lesson 20.

9 Glue one 3/8-inch shaft collar (component V) onto either side of the runner disk’s central hole, using
a piece of 3/8-inch steel rod to align the collars with the hole. This procedure is also demonstrated in
lesson 20.

10 Paint the completed runner disk with primer and several coats of enamel paint to enhance its water
resistance.

11 Build the turbine’s wooden framework by assembling components C, D, E, F, G, H, I, J, K, L, and M as


shown in drawings 13.1 and 13.2. Glue and clamp all connections; then, reinforce with wood screws.

12 Build the penstock by assembling components O, P, Q, and R with PVC glue. Insert the lower
penstock tube (component O) through the holes in components I, F, and E; then, glue the nozzle
(component S) to the outlet end of the penstock.

13 Attach the trash can to the penstock by inserting the pipe fitting (component Q) into the hole in the
bottom of the can, applying silicone sealant to the joint from the inside of the trash can, and then
tightening the nut on the fitting with a wrench.

14 To mount the runner, insert the two flanged ball bearings (component W) into the runner mounts
(component C). Then, insert the steel runner shaft (component U) through one ball bearing and then
through a 3/8-inch washer, the runner, another 3/8-inch washer, and the opposite ball bearing. Fasten the
runner to the shaft by using an Allen wrench to tighten the set screws in the two shaft collars mounted
on the runner.

15 Install another shaft collar (component V) on the end of the runner shaft to serve as a retainer for the
lifting cord.

16 Tie the upper end of the lifting cord tightly to the runner shaft. Tie the lower end to the 2.2-pound load.

The water turbine is now complete and ready for testing!

Project 13 » Water Turbine 255


TESTING PROCEDURE

When the turbine is operated, the water jet’s


impact with the runner blades will spray a
large quantity of water for a considerable
distance around the machine. A catch basin
placed below the runner will not capture
all of this water, unless it is quite large. For
this reason, the turbine should be operated
outdoors, in an area where this overspray
will not cause damage.

It is impossible to operate the turbine


without getting wet. Dress accordingly.

To set up the test, place the turbine on a


table or bench, as shown here, with the
runner shaft extending beyond the edge
of the table to provide a clear path for the
lifting cord. Ensure that the load is attached
to the lower end of the lifting cord. Spin
the runner until the lifting cord is taut.
Insert the stopper (component T) into the
end of the penstock nozzle and then fill the
reservoir with water.

To operate the turbine, simply pull the


stopper from the penstock nozzle. When
the load has reached the runner shaft, insert
the stopper back into the nozzle. Refill the
reservoir prior to each test.

256 
PROJECT 14
GEAR TRAIN
DESIGN, BUILD, AND TEST A GEAR TRAIN THAT
WILL ENABLE AN EXISTING WATER TURBINE TO
LIFT A SIGNIFICANTLY LARGER LOAD OPTIMALLY.
LESSON IN WHICH THIS PROJECT IS COVERED:
21 » Design a Gear Train

PROBLEM DEFINITION
Requirement
}} The turbine must be capable of lifting 6 pounds at optimum power.

Constraints:
}} Use the same reservoir and penstock from project 13.
}} Use the same 8-inch runner from project 13.
}} Use readily available materials for all components.

REFERENCES
Roberts, Making Things Move.
Learn Engineering, “Spur Gear Design.”

Project 14 » Gear Train 257


Gears and Gear Trains
Gears are toothed wheels that engage with each other for the purpose of transmitting torque from one shaft to another.
Although gears are not the only means of performing this function, they perform it particularly well because meshed
gear teeth are durable and cannot slip.

When two or more gears are mounted with their teeth meshed, the resulting system is called a gear train. In any gear
train, there is always a driving gear, which receives power from the machine, and a driven gear, which transmits this
power onward.

In general, gear trains can be used for three purposes:


1 Increasing torque, with a smaller gear driving a larger one
2 Increasing speed, with a larger gear driving a smaller one
3 Changing the direction of power transmission

It is important to recognize that, in all three of these cases, gear trains do not affect power output. Except for minor
losses due to friction, the shaft of the driven gear transmits the same power as the shaft of the driving gear.

There are many different types of gears, but for this project, we will work exclusively
with the familiar spur gear configuration, shown here.

The key geometric characteristics of a spur gear are as follows:

†† The pitch circle is an imaginary circle that is centered on the gear, with its arc
passing approximately through mid-height of the teeth. When two gears are
properly meshed, their respective pitch circles are tangent to each other.

†† The diameter of a gear’s pitch circle is called the pitch diameter, represented by
D. The associated pitch radius, designated r, is one-half of the pitch diameter.

†† The number of teeth in a gear is designated N.

†† The diametrical pitch is designated P and calculated as:

This quantity does not correspond to an identifiable physical characteristic of the gear; nonetheless, it is an
important design parameter because paired gears must have the same diametrical pitch to mesh properly.

†† An alternative form of the diametrical pitch is the circular pitch, CP, calculated as:


This parameter does correspond to an observable physical characteristic: the arc length measured from the
center of one tooth to the center of an adjacent tooth.

258 
†† The pressure angle is the orientation of the line of force transmission between the teeth on two meshed
gears, measured with respect to the tangent to the pitch circle. Most manufactured gears use standard pressure
angles—14.5°, 20°, or 25°. Like the diametrical pitch, the pressure angles of both gears in a pair must be
identical or the teeth will not mesh properly.

Improperly meshing gears will result in excessive wear, noise, rough running, and possibly jamming; thus, matching
gears properly is quite important.

The mechanical advantage provided by a pair of meshed gears is also called the gear ratio and is given by:

r2 N 2
mechanical advantage = =
r1 N 1

where:

DESIGN CONCEPT

Given that the purpose of this design is to increase torque—and, therefore, to gain mechanical advantage—
we will use a smaller gear driving a larger one. (See “Gears and Gear Trains” sidebar.) The smaller driving
gear will be mounted directly to our turbine’s 3/8-inch runner shaft, and the larger driven gear will rotate on
a parallel 1/4-inch shaft. On this second shaft, we will mount a 3/4-inch-diameter wooden drum. The nylon
lifting cord will be wound around this drum such that rotation of the driven gear will lift the 6-pound
load. Both gears will be fabricated from medium-density fiberboard (MDF).

DETAILED DESIGN
Required Gear Ratio

For a pair of meshed gears, the relationship between the input torque, T1, and the output torque, T2, is given by:

Project 14 » Gear Train 259


In this equation, N1/N2 is the gear ratio. Solving for this ratio:

The input torque, obtained from our analysis of the 8-inch turbine runner in project 13, is T1 = 0.0391 foot-
pounds. The output torque, T2, is equal to the load, W, multiplied by the radius of the 3/4-inch-diameter
drum on which the lifting cord is wound:

Therefore, the required gear ratio is:

For simplicity, we will round up to 5.

Number of Teeth

In choosing the number of teeth in the driving gear, N1, we must keep three principal considerations
in mind:

1 The driving gear should be reasonably small, such that the driven gear (which must be five times
larger) will not be excessively large.

2 The individual gear teeth must be wide enough that they will not fail in flexure or shear under
load. This is particularly true for gears made of MDF, a relatively weak material. From a structural
perspective, a gear tooth loaded by its contact with another gear is really just a small beam. As such,
one tooth can be analyzed for structural strength using the same method we employed for our wooden
beam bridge in project 3. Based on this analysis, a gear tooth that is approximately 3/16 inch wide at its
base will provide more than adequate strength.

3 In determining N1, we must take care to avoid a problem called interference, which occurs when the
tips of one gear’s teeth cut into the bases of the other’s teeth. This phenomenon happens—even with a
properly designed tooth profile—if the number of teeth on one gear is too small.

260 
The mathematical test for interference is as follows:

where:

Given that N2 is not yet known, this expression must be solved iteratively. First, assume a value of N1 and
calculate the associated value of N2 based on the required gear ratio; then, determine the minimum N1
required to avoid interference using the inequality above. If this minimum N1 is greater than the initial
assumed value, try a larger value for N1 and repeat the calculation. Continue this process until the assumed
value of N1 is greater than or equal to the calculated value.

For more information on the application of this analytical method, see “Spur Gear Design” at
http://www.learnengineering.org/2013/02/spur-gear-design.html.

For the parameters of this project, the result is teeth.

Thus, for the required gear ratio of 5:1, our driven gear must have:

Designing and Fabricating Gears

To achieve an appropriate degree of precision, we will use laser cutting to fabricate our gears. This
technology uses a computer-controlled laser to cut or engrave any shape—simple or complex—from a
variety of different sheet materials, such as cardboard, wood, plastic, and even metal. The gears for this
project were fabricated by Ponoko (https://www.ponoko.com/), an online laser-cutting service.

Project 14 » Gear Train 261


Before any component can be made by laser cutting, it must be drawn with an appropriate computer
graphics or modeling software package. SketchUp software works quite well for this task. For this
particular project, however, it is necessary to augment SketchUp with a software plug-in for drawing gears.
One of several available gear-drawing plug-ins is provided at http://suplugins.com/podium/free-plugins.php.
(See the “Involute Gears” drop-down on this page.)

To install this plug-in:

}} Download the file Cadalog_Gear_v3.1.rbz from the website above and save it in a folder.
}} Open SketchUp. Click the “Preferences” menu item, select “Extensions,” and click the “Install
Extension” button.
}} Navigate to the folder where Cadalog_Gear_v3.1.rbz is saved; select this file and click the “Open” button.

When the plug-in has been installed, a new menu item—“Key Involute Gear”—will appear under the
“SketchUp “Draw” menu. When this menu item is clicked, SketchUp will display a dialog box requiring
the following pieces of input data:

}} number of teeth
}} pressure angle
}} pitch radius
}} shaft radius
}} keyway width
}} keyway depth

Based on our design, the number of teeth will be 18 for the driving gear and 90 for the driven gear. We
will use the standard pressure angle of 20° for both gears. As noted in the design concept above, we will be
fastening the driving gear directly to the existing 3/8-inch runner shaft; thus, its shaft radius is 3/16 inch. And
we will use a 1/4-inch steel rod for the shaft of the driven gear, so its radius is 1/8 inch. Our gears will not use
a keyway (a mechanical connection between the gear and shaft); thus, the keyway width and depth will be
set to zero.

The only pieces of input data still needed are the pitch radii for the two gears.

To determine these radii, recall our decision to set the width of each gear tooth at approximately 3/16 inch to
ensure that the individual teeth will have adequate structural strength. Recall, also, that the circular pitch,
CP, is defined as the arc length measured from the center of one tooth to the center of an adjacent tooth.
If the width of one tooth is 3/16 inch, then the distance from one tooth to the next should be approximately
twice that width: 3/8 inch. Thus, CP = 3/8 inch.

Given the mathematical definition of circular pitch, we can solve for the diametrical pitch, P, as follows:

262 
Given the definition of diametrical pitch, we can determine the pitch diameter, D1, for the driving gear
as follows:

The pitch radius, r1, is half of the pitch diameter: 1.074 inches.

The pitch radius, r 2, of the driven gear is:

Therefore, our input to the SketchUp gear-drawing plug-in for the driving gear is as follows:

}} number of teeth = 18
}} pressure angle = 20°
}} pitch radius = 1.074”
}} shaft radius = 0.1875”
}} keyway width = 0
}} keyway depth = 0

Input for the driven gear is:

}} number of teeth = 90
}} pressure angle = 20°
}} pitch radius = 5.37”
}} shaft radius = 0.125”
}} keyway width = 0
}} keyway depth = 0

To facilitate laser cutting, the two resulting gear drawings must be exported from SketchUp as separate
files in the EPS format. These files are imported into Adobe Illustrator—a graphics software package—
and placed on a template, as shown on the next page. This template, provided by the Ponoko laser-cutting
service, is the same size and shape as the piece of material from which the gears will be cut. The Ponoko
website (https://www.ponoko.com/starter-kits/adobe-illustrator) provides detailed instructions for this
process, as well as templates of various sizes.

Project 14 » Gear Train 263


As an alternative to Adobe Illustrator, the AutoCAD, CorelDRAW, or Inkscape graphics program can also
be used. See the Ponoko website for details and templates.

Once the graphics file has been prepared, it is uploaded to the Ponoko website, and a material is selected.
The laser-cut components are typically shipped approximately two weeks after completion of an order.

Distance between Gears

The required distance between the two gears in our gear train is:

Because laser-cut gears made of MDF have slightly rough edges, we will increase this distance by 1/32 inch
to prevent the gears from binding. The resulting distance is 6.475 inches.

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/diy-
engineering/. The gears in this model can be exported for laser cutting as described on the previous page.

264 
Drawing 14.1 » Gear train
mounted on water turbine

Drawing 14.2 » Gear


train, exploded view
H
G
G
H E
B C D

A
F

MATERIALS LIST
COMPONENT DESCRIPTION
A Driving gear Laser cut from 1/4″ MDF*
B Driven gear Laser cut from 1/4″ MDF*
C Drum 3/4″-diameter hardwood dowel, 2″ long
D Retainer 1″-diameter disk, cut from 1/16″ plywood
E Shaft 1/4″-diameter steel rod, 7″ long
F Shaft collar Set screw shaft collar for 3/8″ diameter (McMaster-Carr #9414T8)
G Shaft collar (2) Set screw shaft collar for 1/4″ diameter (McMaster-Carr #9414T6)
H Bearing Ball bearing, open-flanged, 1/4″ shaft diameter (McMaster-Carr
#6383K214)
- Lifting cord Nylon string
*The prototype for this project used the Ponoko online laser-cutting service (https://www.ponoko.com/), as outlined above.

Project 14 » Gear Train 265


In addition to common hand tools, this project requires:

}} scroll saw or bandsaw (for curved cuts)


}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} files
}} a set of Allen wrenches

Wood glue and epoxy glue are also used.

FABRICATION AND ASSEMBLY PROCEDURE

1 Fabricate the two gears (components A and B) using the laser-cutting process described above.

2 Drill 11/8-inch-diameter holes in the water turbine’s runner mounts (component C from project 13).
The locations of these holes are shown on the full-size cutting pattern for project 13.

3 Fabricate the wooden drum and retainer (components C and D). Glue them together with wood glue;
then, use a drill press to drill a 1/4-inch-diameter hole through the center of the drum, as shown below.

C
D

1/4” diameter
longitudinal
hole

4 Cut a 7-inch length of 1/4-inch steel rod for the shaft of the driven gear (component E) and file its
ends smooth.

5 After roughening the end of this shaft (component E) with a file or sandpaper, use epoxy to glue the
drum assembly (components C and D) onto the shaft as shown in drawings 14.1 and 14.2 above.

6 Slide the driven gear (component B) onto the shaft and glue it to the inner end of the drum
(component C) with wood glue. Ensure that the gear is perfectly perpendicular to the shaft.

7 Use epoxy to glue the 3/8-inch shaft collar (component F) to the driving gear (component A). Use a
3/8-inch-diameter steel rod to ensure that the holes in the collar and gear are aligned.

266 
8 Insert two ball bearings (component H) into the 7/8-inch-diameter holes in the runner mounts
(component C from project 13).

9 Slide the driving gear (components A and F) onto the water turbine’s runner shaft.

10 Slide the driven gear, shaft, and drum assembly (components B, C, D, and E) through a 1/4-inch shaft
collar (component G), then through the two ball bearings (component H), and then through another
1/4-inch shaft collar, as shown in drawing 14.2.

11 After aligning the two gears, tighten the set screws in all three shaft collars with Allen wrenches of the
appropriate size.

12 Tie the upper end of the lifting cord tightly to the drum (component C). Tie the lower end to the
6-pound load.

The gear train is now complete and ready for testing!

TESTING PROCEDURE

Set up the water turbine as described in project 13.

With the 6-pound load attached to the lower end


of the lifting cord, spin the runner until the lifting
cord is taut. Insert the rubber stopper into the end
of the penstock nozzle and then fill the reservoir
with water.

To operate the turbine, pull the stopper from the


penstock nozzle. When the load has reached the
drum, insert the stopper back into the nozzle.
Refill the reservoir prior to each test.

Project 14 » Gear Train 267


PROJECT 15
WOODEN
PENDULUM CLOCK
DESIGN, BUILD, AND TEST A
WOODEN PENDULUM CLOCK.
LESSON IN WHICH THIS PROJECT IS COVERED:
22 » Make a Mechanical Clock

PROBLEM DEFINITION
Requirement
}} Must display both hours and minutes.
}} Must be adjustable, to maintain accurate timekeeping.

Constraints:
}} Use laser cutting for all principal wooden components.
}} Use readily available materials for all other components.

REFERENCES
Roberts, Making Things Move.
Law, Brian Law’s Wooden Clocks.

268 
The Pendulum Clock
The pendulum clock was first conceived by Galileo—the great mathematician, scientist, and engineer—
who did pioneering investigations of pendulums in the early 17th century. Galileo was not able to
bring his ideas to fruition before he died, but these ideas did inspire the work of Christiaan Huygens,
the Dutch scientist who is credited with inventing the pendulum clock in 1656 and patenting it the
following year.

Huygens’s invention was a revolutionary development that improved the accuracy of timekeeping by
nearly two orders of magnitude. Most clocks built before the pendulum clock had no minute hands,
because these devices’ poor accuracy made minutes largely irrelevant. In this sense, Huygens reinvented
time itself.

Today, having been replaced by more accurate quartz and electric clocks, pendulum clocks live on mostly
as historical curiosities and decorative furnishings, such as grandfather clocks and cuckoo clocks. To DIY
engineers, however, they also represent a wonderful opportunity to learn about the science of harmonic
motion and to apply our knowledge of gear trains at a level far beyond the simple two-gear apparatus we
used to augment the torque of our water turbine in project 14.

DESIGN CONCEPT

Our clock will consist of five major subassemblies, which we will design separately and sequentially:

}} the pendulum and its associated mechanism


}} a gear train to drive the minute hand
}} a separate gear train to drive the hour hand
}} a mechanism to supply power to the clock
}} a case to hold it all together.

DETAILED DESIGN
Pendulum

The period, T, of a pendulum is given by:

Project 15 » Wooden Pendulum Clock 269


where:

Note that, as expected, the period does not depend on either the mass of the pendulum or the amplitude of
its motion.

Our clock will use a pendulum with a period of exactly 2 seconds. To determine the length of this
pendulum, solve the above expression for L and substitute T = 2 seconds:

Pendulum Motion
The oscillatory motion of a pendulum is characterized by three measurements:

†† The amplitude is the pendulum’s maximum displacement, measured in terms of the angle it
swings away from vertical.

†† The period is the time it takes for one full oscillation.

†† The frequency is the number of oscillations per time.

The mathematical relationship between period and frequency is as follows:

The most important characteristic of a pendulum is that for small amplitudes, its period is independent
of both its mass and its amplitude. This latter characteristic—independence of period from amplitude—
is called isochronism. This discovery, made by Galileo, ultimately led to Huygens’s successful invention
of the pendulum clock. Isochronism is essential for accurate timekeeping, because the minor amplitude
variations that inevitably occur in a real-world pendulum will not significantly change the time required
for each oscillation.

In practice, the period of a pendulum is not perfectly independent of its amplitude, though its deviation
from perfect isochronism is extremely small, as long as the amplitude is kept relatively small. This is why
clock pendulums are typically designed for amplitudes of 3° or less. At this angle, the deviation from
perfect isochronism is only about 0.0002, or 0.02%, but at 20°, this error grows to nearly 1%.

270 
dowel pivot

The configuration of our pendulum is shown here. The weight—called a


pendulum bob—will be a hardwood disk suspended from a thin metal rod,
which is threaded for adjustability. The upper end of the rod is anchored to a wooden ring
wooden ring in which a chisel-shaped dowel is mounted. The point of the chisel
will serve as the pendulum’s pivot point.

Escapement
metal rod
The most important component of a mechanical clock is the escapement—a
device that sustains the pendulum’s motion while also translating the oscillatory
motion of the pendulum into the controlled unidirectional rotation of a shaft
that will drive the clock hands.

The escapement has three main components: the timing wheel, anchor, and pendulum bob
yoke—as shown below.

The timing wheel is a gear with 30 uniquely shaped teeth. It


is fixed to a shaft that is subjected to a constant torque by the
clock’s power source: a hanging weight, which will be
discussed below.
yoke

The anchor is a pivoting arm, which is mechanically linked to


anchor the pendulum by the yoke. When the pendulum rotates, the
anchor rotates along with it. At the ends of the anchor are two
prongs, called pallets, which engage with the timing wheel in a
very ingenious way. Because the timing wheel is subjected to a
constant torque, the wheel would spin continuously if it were not
prevented from doing so by the anchor. As the anchor rocks back
and forth with the pendulum, the pallets engage with the wheel
timing wheel alternately, allowing it to advance by exactly one tooth for each
pendulum cycle. Furthermore, the pallets are configured such
that as each one engages with a tooth of the timing wheel, it gives
the pendulum a little push. This repetitive nudge reinforces the
swing of the pendulum, offsetting friction losses and keeping the
system in motion indefinitely.

Why does the timing wheel have 30 teeth? Given that it advances one tooth for each full oscillation of the
pendulum and given that the period of the pendulum is 2 seconds, the 30-tooth wheel will rotate through
exactly one turn every 60 seconds. Thus, the motion of the entire mechanical system will be effectively
calibrated by the rotation of the timing wheel at exactly 1 revolution per minute (rpm).

Project 15 » Wooden Pendulum Clock 271


Gear Train for the Minute Hand

The minute hand must rotate at 1 revolution per 60 minutes, or 1/60 rpm—60 times more slowly than the
timing wheel. Theoretically, we could accomplish this with two gears: a small gear driving a larger one, to
decrease the speed of rotation. But this solution is impractical
(particularly for a device that uses wooden gears) because it
would result in an excessively large driven gear. 1

As an alternative, we will use a series of compound gearsets,


each of which consists of two gears fixed to the same shaft, as
shown right. A
B
Gear A has 16 teeth. Because it is fixed to the same shaft (2) as 2
the timing wheel, it turns at 1 rpm. Gear A engages with gear
C
B—a 60-tooth gear on a parallel shaft (3)—achieving a speed
reduction of 60:16, or 3.75:1. Also on shaft 3 is a 15-tooth gear F
D
(C), which engages with another 60-tooth gear (D) on another
parallel shaft (4) for an additional speed reduction of 60:15, or
5 3
4:1. This arrangement is repeated once more, with gear E on
E
shaft 4 engaging gear F on shaft 5 for another 4:1 reduction.
4
The most important characteristic of compound gearsets is that
their individual gear ratios are multiplicative. Thus, the total
speed reduction of this gear train, with its six gears and four shafts, is 3.75 × 4 × 4 = 60:1. Thus, shaft 5—
the main clock shaft—will turn the minute hand at 1/60 of a revolution per minute, or 1 revolution per
hour, as required.

Gear Train for the Hour Hand

The hour hand must rotate 12 times more slowly than the
minute hand. This speed reduction will be accomplished with
another compound gearset. minute hand

As shown here, a 10-tooth gear (G) is fixed to the main clock hour
shaft (5). It engages with a 30-tooth gear (H) on a short parallel hand
J
shaft (6), providing a 3:1 speed reduction. Also on shaft 6 is an
G
8-tooth gear (I) that engages with a 32-tooth gear (J) on the
5
main clock shaft (5) for a further 4:1 reduction. Overall, the
3:1 and 4:1 ratios provide the additional 12:1 speed reduction
required for the hour hand. 6
I
H

272 
On a standard clock dial, the hour hand must rotate on the same axis as the minute hand, but independent
of it. Thus, gear J is actually mounted on a short length of brass tubing that fits over the main clock shaft
(5), and the hour hand is fixed directly to this gear.

Power Source

The design of the escapement assumes that a constant torque will be applied to the shaft of the timing wheel.
This torque is supplied by a weight, which hangs from a wooden drum fixed to the main clock shaft. This
torque is transmitted through the entire gear train to the timing wheel. As the clock runs, the weight will
slowly fall, and eventually it will have to be raised manually. (In other words, we will need to wind the clock.)

There is one potential problem with this arrangement. If the drum were
permanently fixed to the main clock shaft, then raising the weight would
also entail dragging the entire gear train through many hundreds of
rotations—a process that would require considerable effort and
would subject the gears to considerable wear and tear.

To prevent this problem, we will allow the drum to rotate


freely on the main clock shaft and attach a ratchet
mechanism to the drum. The mechanism consists of a
ratchet toothed wheel and two pawls, which can be engaged
F to prevent the wheel from turning clockwise but
pawl allow it to rotate counterclockwise freely. The pawls
are mounted on gear F, so engaging the ratchet has
the effect of locking the drum to the gear.
drum
F Gear F is also not fixed to the main clock shaft,
though it must be a tight fit. Through this
arrangement, the clock hands can be rotated
manually to set the time without having to move the
entire gear train.

Case, Clock Face, and Hands

With the design for the mechanical system essentially complete, the wooden case, clock face, and hands
are added, as illustrated in the drawings on the next page. The principal purpose of the case is to hold the
six shafts of the two gear trains in their proper positions. It is designed to be easily disassembled so that
the mechanism can be adjusted and cleaned. The case configuration allows the clock to be either hung
on a wall or placed on a shelf; however, in the latter mode, the shelf must have a suitable cutout for the
pendulum, as shown in lesson 22.

Project 15 » Wooden Pendulum Clock 273


FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project and full-size layout files for all laser-cut components
can be downloaded from http://stephenjressler.com/diy-engineering/.

Drawing 15.1 »
Pendulum clock,
front perspective
view (without
pendulum)

Drawing 15.2 » Pendulum Drawing 15.3 »


clock, rear perspective view Pendulum clock, side
perspective view

U i
Drawing 15.4 » m
Pendulum clock, d N
8 1 M 7
exploded view 9
7 K L
j
A B 2'
V g
a
F C
b g
2 3'
O Q
7
c J 3 e
G
R f
5' 7 P
5
6 S
4' h
6' 4 E
I H m
9 d D m
g g h
8
8 9 d

274 
W k

11

Drawing 15.5 » Pendulum,


exploded view

10
n
12

MATERIALS LIST
COMPONENT DESCRIPTION
A 16-tooth gear Laser cut from 1/4″ hardwood-faced MDF
B 60-tooth gear
C 15-tooth gear
D 60-tooth gear
E 15-tooth gear
F 60-tooth gear
G 10-tooth gear
H 30-tooth gear
I 8-tooth gear
J 32-tooth gear
K Timing wheel
L Anchor
M Spacer
N Yoke
O Drum end
P Drum body (3)
Q Ratchet
R Pawl (2)
S Shaft 6 mount
T Clock case back
U Clock case front
V Clock face
W Pendulum ring (2)

Project 15 » Wooden Pendulum Clock 275


a Clock face arc (12) Laser cut from 1/8″ bamboo
b Minute hand
c Hour hand
d Front face retainer (3)
e Pawl retainer (2)
f Pawl retainer dowel (2) 1/4″-diameter hardwood dowel, 9/16″ long
g Face mount dowel (4) 3/8″-diameter hardwood dowel, 11/4″ long
h Front-to-back frame dowel (2) 3/4″-diameter hardwood dowel, 51/4″ long
i Frame and pendulum support 3/4″-diameter hardwood dowel, 51/4″ long, notched for pendulum
dowel pivot
j Yoke dowel (2) 1/4″-diameter hardwood dowel, 1″ long
k Pendulum pivot 1/4″-diameter hardwood dowel, 1″ long, sanded to a chisel point
m Front-to-back frame piece (3) 11/2″ × 11/2″ × 3″ hardwood
n Pendulum bob 5″-diameter disk cut from 1″-thick hardwood
- Winding cord Nylon string, approximately 5 feet long
1 Shaft for M and N 1/4″-diameter brass rod, 4″ long*
2 Shaft for K and A 1/4″-diameter brass rod, 31/2″ long*
2’ Sleeve for shaft 2 9/32″-diameter brass tube, 3″ long**
3 Shaft for B and C 1/4″-diameter brass rod, 31/2″ long*
3’ Sleeve for shaft 3 9/32″-diameter brass tube, 3″ long**
4 Shaft for D and E 1/4″-diameter brass rod, 31/2″ long*
4’ Sleeve for shaft 4 9/32″-diameter brass tube, 3″ long**
5 Main clock shaft 1/4″-diameter brass rod, 43/4″ long*
5’ Sleeve for shaft 5 9/32″-diameter brass tube, 1/2″ long**
6 Shaft for H and I 1/4″-diameter brass rod, 11/4″ long*
6’ Sleeve for shaft 6 9/32″-diameter brass tube, 5/8″ long**
7 Bushing (4) 9/32″-diameter brass tube, 1/4″ long**
8 Brass screw Brass decorative round-head screw, #12, 1″ long (McMaster-Carr
#92407A296)
9 Brass washer Brass washer for 1/4″ screw size (McMaster-Carr #92916A365)
10 Pendulum rod Steel threaded rod, 4-40 size, 36″ long (McMaster-Carr
#98841A005)
11 Decorative pendulum sheath 5/32″ diameter brass tube, 35″ long (McMaster-Carr #8859K235)
12 Threaded coupler Steel threaded round standoff, 1/4″ diameter, 4-40 thread (McMaster-
Carr #91125A435)
* McMaster-Carr product #8859K351
**McMaster-Carr product #8859K25

In addition to common hand tools, this project requires the following equipment:

}} miter saw or radial arm saw (for cross-cutting)


}} scroll saw or bandsaw (for curved cuts)

276 
}} power-sanding station (for shaping)
}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} chisels
}} files
}} clamps

Other supplies needed for this project include wood glue, cyanoacrylate glue, epoxy glue, medium and fine
sandpaper, and a wood finish (e.g., varnish or tung oil) of your choice.

FABRICATION AND ASSEMBLY PROCEDURE


1 Fabricate all laser-cut components using the procedure described in project 14 and the materials
specified in the materials list above. Full-size layout files for these components are provided at http://
stephenjressler.com/diy-engineering/. These files use templates provided by the Ponoko laser-cutting
service (https://www.ponoko.com/) and can only be used with this service.

2 Build each of the three front-to-back frame pieces (component m) from three 1-inch-by-11/2-inch-by-1
1/2-inch hardwood blocks and a 3/4-inch hardwood dowel (component h), as shown below. (If 1-inch-
thick hardwood is not available, four 3/4-inch-thick laminations can be used.) Each hardwood block has
a 3/4-inch hole drilled exactly through its center to accommodate the dowel. The dowel should project
1/4 inch from the front face of the completed piece.

3/4" hole h

hardwood block

h
1/4"

2
m

Project 15 » Wooden Pendulum Clock 277


3 Using a sharp chisel, cut a V-groove into the 3/4-inch
3
dowel (component i) on the top front-to-back frame
piece, as shown here. This groove will serve as the
support for the pendulum pivot. If the clock will be hung m
on a wall, drill a 1/4-inch mounting hole into the rear of V-groove
the dowels on all three frame pieces as well.

12 i

1/4" hole 4 1/4" hole

4 The pendulum bob (component n) is a 5-inch-diameter


disk cut from 1-inch-thick hardwood. Once the disk
n has been cut out and sanded smooth, drill a 1/4-inch-
diameter hole into its edge, as shown here, and glue the
steel threaded coupler (component 12) into this hole
with epoxy.

5 Fabricate all metal shafts, sleeves, and bushings from 1/4-inch brass rod and 9/32-inch brass tubing, as
specified in the materials list above. After each cut, file the ends smooth. Ensure that all shafts can
rotate freely when inserted into their respective sleeves.

6 Construct the drum-ratchet assembly by gluing and clamping component O, three laminations of
component P, and component Q together as shown below. During this process, temporarily insert a
1/4-inch metal rod through all four components to ensure that their shaft holes are aligned. The ratchet
teeth on component Q must be oriented as shown here.
(If you install the ratchet backward, your clock
will need to run counterclockwise!) Q 1/4" rod
P
O P P
Once this assembly has dried, drill
through the shaft hole with a 1/4-
inch drill bit to smooth the inside
of the hole so that the drum-ratchet
assembly will rotate freely on
6
its shaft.

7 Glue the two yoke dowels (component j) into the holes at the bottom of the yoke (component N).

278 
8 Glue the spacer (component M) to the rear of the anchor (component L), temporarily inserting a 1/4-
inch metal rod into the shaft holes to ensure that they are aligned.

9 Assemble the pendulum support ring by gluing and clamping the two laminations of component W
together. Once the glue has dried, drill a 1/8-inch hole and a 1/4-inch hole into the edge of the ring, as
shown below. Fabricate the pendulum pivot (component k) by sanding the tip of a 1/4-inch hardwood
dowel to a chisel point. Then, glue the pivot into the 1/4-inch hole in the pendulum support ring.
Finally, use epoxy to glue the upper end of the pendulum rod (component 10) into the 1/8-inch hole in
the pendulum support ring.

1/4" hole k

W
1/8" hole
9
10

10 Cut the decorative pendulum sheath (component 11) to a length of 35 inches and then slide it over the
pendulum rod (component 10). Screw the lower end of the pendulum rod into the threaded coupler
(component 12) in the pendulum bob (component n).

11 Glue the three front-to-back frame pieces into the 3/4-inch holes in the back of the clock case
(component T). To ensure that the frame pieces are properly aligned, slip the front of the clock case
(component U) over the front ends of the frame pieces. Then, clamp the entire assembly together. (Do
not glue component U to the frame pieces!) Ensure that the entire assembly is square before setting it
aside to dry thoroughly.

12 Sand all wooden components with fine sandpaper; then, give all of them a light coat of finishing oil
or varnish.

13 Glue the anchor (components L and M) to its shaft (component 1), with the rear face of the spacer
(component M) exactly 9/16 inches from the rear end of the shaft. Position the anchor carefully. Check
to ensure that it is exactly perpendicular to the shaft; then, glue it in position with cyanoacrylate.

14 Glue the timing wheel (component K) and gears A, B, C, D, and E to their respective shaft sleeves
(components 2’, 3’, and 4’), as shown below, using the procedure demonstrated in lesson 22. It is
critically important that each gear is precisely positioned on its shaft sleeve and that each gear is

Project 15 » Wooden Pendulum Clock 279


perfectly perpendicular to its shaft. To achieve these outcomes, use the jigs demonstrated in lesson 22.
These jigs are constructed from the same veneered MDF of which the gears are made. To facilitate their
use, the gear positions (shown below) are defined in terms of t, the thickness of the MDF material.

K B D

sleeve 2’ sleeve 3’ sleeve 4’


A C E

14

t 3t 5t
3t 5t 7t
t = thickness of MDF material of which the gears are made

15 Use cyanoacrylate to glue the four brass shaft bushings (component 7) into their respective holes in the
front and back of the clock case (components U and T).

16 Slide gear F and the drum-ratchet assembly (components O, P, and Q) onto the main clock shaft
(component 5). Do not glue them to the shaft. Fasten the two ratchet pawls (component r) to gear
F as shown in drawing 15.4, using a 1/4-inch hardwood dowel (component f) and a laser-cut retainer
(component e) to connect each pawl. Glue the ends of each dowel to the pawl and the retainer, but not
to the gear, so that the pawl is free to rotate. Ensure that the pawls engage properly with the ratchet.

17 To assemble the clock case, insert the anchor shaft (component 1) and the main clock shaft (component
5) into their respective bushings in the back of the clock case. Then, position the front of the clock
case, inserting these same two shafts and the three front-to-back frame pieces into their respective
holes. Fasten the front to the frame pieces with decorative brass screws (component 8), brass washers
(component 9), and laser-cut wooden retainers (component d), as shown in drawing 15.4 above.

18 Install the three compound gearsets inside the clock case, as demonstrated in lesson 22. Hold each
sleeve in position and then insert the corresponding shaft through the front of the case, the sleeve, and
the rear of the case. Each gearset should rotate freely, and its gears should spin without wobbling. Each

280 
gear should engage its partner without binding or jamming at any point in its 360° rotation. If any
binding occurs, remove the gearset and smooth the edges of the offending gear teeth with sandpaper or
a fine file.

19 Mount shaft 6 in the front of the clock case, inserting it through its designated hole in the case and
into the laser-cut mount (component S), which is glued to both the shaft and the case.

20 Slide gear G onto the front of the main clock shaft (component 5) and glue it in position with
cyanoacrylate. Leave a gap of approximately 1/16 inch between the rear face of the gear and the front
face of the clock case.

21 Glue gears H and I onto their shaft sleeve (component 6’) with cyanoacrylate, leaving a 1/16-inch gap
between the two gears and ensuring that both gears are perpendicular to the shaft. Slip this assembly
onto shaft 6, ensuring that gear H engages with gear G on the main clock shaft.

22 Glue the remaining shaft sleeve (component 5’) into the shaft hole of gear J. The rear end of the sleeve
should be flush with the rear face of the gear. Glue the hour hand to the front end of this sleeve; then,
slide this entire assembly onto the end of the main clock shaft (component 5), engaging gear J with gear I.

23 Slip the minute hand onto the end of the main clock shaft. Rotate both hands to the 12 o’clock
position; then, secure the minute hand to the shaft with cyanoacrylate.

24 Slide the yoke (component N) onto the rear end of the anchor shaft (component 1). Do not glue it in
place at this time; we will do this after adjusting the angle of the yoke (see below). If the fit is not snug,
use a piece of tape to prevent the yoke from rotating on the shaft.

25 Fabricate the winding cord from a 5-foot length of nylon string. Wrap the cord four or five turns
around the drum (component P). Tie one end to a weight of approximately 8 pounds and the other end
to a much smaller weight (e.g., a lead fishing sinker), which will keep the cord taut. If the cord slips on
the drum, increase the number of wraps.

26 Install the pendulum by inserting the pivot (component k) into the V-groove in the top of component i.
Ensure that the pendulum rod falls between the two dowels (component j) projecting from the rear of
the yoke.

The wooden pendulum clock is now complete and ready for testing!

Project 15 » Wooden Pendulum Clock 281


TESTING PROCEDURE
1 Mount the clock securely on a wall, or place it on a shelf that has a suitable cutout to accommodate
the pendulum.

2 Lubricate the gears with a dry film lubricant.

3 Wind the clock by raising the 8-pound weight. Ensure that both ratchet pawls are engaged before
releasing the weight. Start the clock by giving the pendulum a gentle push.

4 As the clock runs, observe the operation of the escapement. As the anchor rotates, both pallets should
engage equally with the teeth of the timing wheel. If not, rotate the yoke slightly on its shaft and
restart the clock. Once both pallets are engaging the timing wheel equally, glue the yoke to its shaft
with cyanoacrylate.

5 To prevent long-term damage to the gear train, it is advisable to determine the minimum amount of
weight necessary to keep the clock operating. Starting at 8 pounds, reduce the weight in small increments
until the clock no longer runs consistently. Then, increase the weight to the previous increment.

282 
PROJECT 16
ELECTRIC-
POWERED CRANE
DESIGN, BUILD, AND TEST AN ELECTRIC
MOTOR–DRIVEN CRANE.
LESSON IN WHICH THIS PROJECT IS COVERED:
23 » Design a Motor-Powered Crane

PROBLEM DEFINITION
Requirement
}} Must be capable of lifting 100 pounds.
}} Must operate at or near its optimum level of power output.

Constraints:
}} Use an inexpensive off-the-shelf electric hobby motor as the source of power.
}} Use readily available materials for all other components.

REFERENCES
Roberts, Making Things Move.

Project 16 » Electric-Powered Crane 283


How an Electric Motor Works
At the heart of any direct current (DC) motor are two major components, called the rotor and stator.
The rotor consists of an electric coil that is fixed to the motor shaft such that when the coil spins, the
shaft spins along with it. The stator, which is stationary, consists of two permanent magnets: one with its
north pole facing inward and the other with its south pole facing inward.

Whenever electric current flows through a conductor, it creates a magnetic field around the conductor.
In a motor, current flows through the coil of the rotor, causing a magnetic field. And the interaction
between this magnetic field and the one caused by the two fixed magnets of the stator produces the
torque that spins the motor shaft.

However, the shaft would only rotate a half-turn and then stop if it were not for a device called the
commutator, which reverses the direction of the current flow through the coil with every half-turn. The
commutator allows the forces of magnetic attraction and repulsion to propel the coil through its full
360° rotation.

Torque, Power, and Speed


The graph below shows the relationship between torque and rotational speed for a typical DC hobby motor. Note that
the very low level of torque produced by these small motors is typically expressed in units of inch-ounces, rather than
inch-pounds or foot-pounds.

0.6
stall torque
0.5 0.479 in-oz @ 1000 rpm
Torque (in-oz)

0.4

0.3
no-load
0.2
speed
0.1

0
0 1000 2000 3000 4000 5000 6000 7000
Rotational speed (rpm)

On this graph, the no-load speed is the rotational speed at which the torque applied to the shaft is zero. This is the
highest speed the motor is capable of achieving. The stall torque is the level of applied torque that prevents the shaft
from spinning. Thus, it is the highest level of torque the motor is capable of supplying. In between these two extremes,
the torque-speed relationship for most electric motors is linear, as shown. As torque decreases, speed increases.

284 
The power delivered by a rotating shaft can be calculated as torque times rotational speed, where the rotational speed
must be expressed in radians per second. For example, a rotational speed of 1000 revolutions per minute (rpm) is
expressed in radians per second as follows:

According to the graph, the motor will produce a torque of 0.479 inch-ounces at this speed. The resulting power
output is:

If we do these same calculations for a series of representative points along the torque curve and then plot power versus
speed on the same graph, the result is as shown below. The power curve is parabolic, and it peaks at exactly half of the
motor’s no-load speed: 3300 rpm. For this small motor, the peak power output is approximately 98 inch-ounces per
second—almost 25% more than the peak power produced by the water turbine we designed in project 13.

0.6 120

0.5 torque peak power 100

Power (in-oz/sec)
power
Torque (in-oz)

0.4 80

0.3 60

0.2 40

0.1 20

0 0
0 1000 2000 3000 4000 5000 6000 7000
3300 rpm
Rotational speed (rpm)
Would it be possible to use this little hobby motor to fulfill our project requirement for a power source capable of
lifting 100 pounds? Theoretically, the answer is yes; however, to do it successfully, we would need to design and build
a gear train capable of providing a mechanical advantage of roughly 1000:1. This is certainly possible, but not terribly
practical—particularly for gears made of wood.

Fortunately, we can solve this problem by using a gearhead motor, a standard hobby motor with its own built-in gear
train. Thanks to this feature, a gearhead motor provides substantially more torque—though at lower speed—than a
comparable hobby motor.

Project 16 » Electric-Powered Crane 285


DESIGN CONCEPT

The basic configuration of our crane is shown in the SketchUp model below.

backstay
boom

mast

pulley

lifting
rope

motor

motor main frame


mount
ball bearing
shaft

Key features:

}} The main frame will incorporate a mount for a gearhead motor (see “Torque, Power, and Speed”
sidebar), with the motor’s shaft rigidly connected to a steel shaft that spans across the wooden frame
and rotates in a pair of ball bearings.
}} To ensure that the shaft is not bent by the 100-pound pulling force in the lifting rope, we will use a
3/8-inch-diameter steel rod for the shaft.
}} The lifting rope will be 1/4-inch nylon rope with a working load of 120 pounds. This rope will be
wound around the shaft and will extend out to the end of an angled boom, around a pulley, and
down to the load.
}} The boom will be supported by backstays that pass over the top of a mast and then connect to the
rear of the frame.

286 
DETAILED DESIGN
P
Selection of Motor

To lift a 100-pound load from the floor up to desk lifting rope


height, the crane will wind about 3 feet of rope onto the
shaft, resulting in the configuration shown here. With two
shaft
layers of 1/4-inch rope wound around the shaft, the pulling force, P,
will be applied at an effective radius of approximately
T
0.6 inches.
re = 0.6"
Thus, the torque, T, applied by the motor to the shaft is:

where:

Before using this theoretical equation, however, we must first account for the inevitable adverse effects of
friction, which will occur in both the shaft bearings and the pulley.

The purpose of a pulley is to change the direction of the tension force in a rope. If a pulley had no friction,
then the magnitude of the tension force would be exactly equal on either side of the pulley. However,
because a pulley wheel always experiences some friction as it rotates, the pulling force applied to a lifting
rope that passes through a pulley must always be slightly higher than the weight being lifted. For a typical
single-wheel pulley, the difference is about 10%, meaning that a 110-pound pull would be required to lift a
100-pound load.

Similarly, if the ball bearings on which our shaft will rotate were perfectly frictionless, then 100% of the
torque produced by the motor would be translated into tension in the lifting rope. But, in practice, a small
portion of the output torque is lost in overcoming friction. Again, we will assume that this loss is 10%.

In the required-torque equation, we will account for these losses by incorporating two empirical friction
factors as follows:

Project 16 » Electric-Powered Crane 287


where:

Substituting W = 100 pounds and re = 0.6 inches, the required torque is:

SparkFun Electronics (https://www.sparkfun.com/) offers 14 different gearhead motors with a wide range
of performance characteristics. The manufacturer’s specifications for these motors are summarized below.
Based on these data, it appears that only motor 14 will meet our need, because it is the only available
product that will supply torque (at peak power) greater than the 1173 inch-ounce requirement.

STALL TORQUE AT
GEAR NO-LOAD SPEED
MOTOR TORQUE PEAK POWER
RATIO (RPM)
(IN-OZ) (IN-OZ)
1 10 303 17 8

2 18 168 27 14

3 30 101 46 23

4 50 81 57 29

5 56 51 77 39

6 40 40 103 52

7 100 30 137 69

8 150 20 185 93

9 300 10 368 184

10 500 6 613 307

11 900 4 992 496

288 
12 1000 3 1102 551

13 1500 2 1497 749

14 3000 1 2995 1498

Design of the Lifting Apparatus

In practice, a motor with a lower torque rating could be used in conjunction with a pulley system
configured to provide mechanical advantage. Most real-world cranes use pulley systems for this purpose.

A pulley system achieves mechanical advantage when its configuration results in the load, W, being
supported by multiple legs of the lifting rope, each of which carries the applied pulling force, P. Consistent
with this principle, various pulley systems and their associated mechanical advantages are illustrated below.

P P P P

W=2P W=3P W=4P W=5P

2:1 3:1 4:1 5:1

We will consider alternative crane designs that incorporate 2:1 and 4:1 pulley systems in the lifting apparatus.

The required-torque calculation can be modified to incorporate mechanical advantage, M, as follows:

For the 2:1 pulley system, assuming an increase in friction losses from 10% to 15% due to the addition of a
second pulley wheel:

Project 16 » Electric-Powered Crane 289


For the 4:1 pulley system, assuming 20% friction losses:

These calculations show that the pulley systems result in substantial reductions in the required torque.
Comparing these results against the manufacturer’s specifications above, we can draw the following
conclusions:

}} With a single pulley, there is no mechanical advantage, and motor 14 is required to lift the load.
}} With a 2:1 pulley system, motor 13 will work.
}} With a 4:1 pulley system, motor 11 will work.

Why might we want to pursue either of the pulley-assisted alternatives? The cost of all three motors is the
same, so using a lower-torque motor will not provide any cost advantage. But using a pulley system does have
the distinct advantage of lowering the required pulling force. With a 2:1 system, the required pull drops from
110 pounds to 58 pounds, and with the 4:1 system, it is only 32 pounds. As a result, the lifting rope can be
lighter, and the structural demands on the shaft, bearings, and supports will be substantially reduced.

Based on this advantage—and on the fact that pulley systems are a lot of fun to build—we will use motor
11 and a 4:1 pulley system.

FINAL DESIGN DRAWINGS

The SketchUp 3-D computer model for this project can be downloaded from http://stephenjressler.com/diy-
engineering/. A full-size cutting pattern for several of the key components is provided in the PDF version of this
guidebook. This pattern is designed to print on a single 81/2-by-11-inch sheet with the printing scale set to 100%.

Drawing 16.1 » Electric


motor-powered crane,
perspective view

290 
Drawing 16.2 » Electric
motor-powered crane,
perspective view

H
K J I

F f

L e

L G
J
K d N
c e
B F
A

B
c
E b Drawing 16.3 » Electric
M motor-powered crane,
a exploded view
D
D
C

3 1/2" 5"

2"
Plan View

1 1/8" dia. 4 1/2" 1 1/2"

3/4" dia. 3/8" dia.

3 1/2"
2 1/4" 2"

12 1/2" 27 1/2" 4 1/4" 3 3/4"

Side Elevation

Drawing 16.4 » Main frame, plan and elevation views

Project 16 » Electric-Powered Crane 291


Drawing 16.5 » Boom, plan and elevation views

1/4" dia.

3 1/2" 2"

12 1/4" 15" 1"

Plan View

3/8" dia.

2 1/2"

15 1/4" 12" 4 3/4"

34 1/2"

Side Elevation

6 1/2" 5"

1 3/4" 7" 13 3/4" 5"


Plan View
3/8" dia.

1 1/2"

8 3/4" 7 1/2" 11 1/4"


28 1/4"
Side Elevation

Drawing 16.6 » Mast, plan and elevation views

292 
MATERIALS LIST
COMPONENT DESCRIPTION

A Main frame base 3/4″ × 3.5″ × 48″ poplar


B Main frame side (2) 3/4″ × 3.5″ × 48″ poplar
C Motor mount base 3/4″ × 3″ × 5″ poplar*
D Motor mount side (2) 3/4″ × 3″ × 2″ poplar
E Backstay anchorage 3/4″-diameter hardwood dowel, 7″ long
F Boom side (2) 3/4″ × 21/2″ × 281/2″ poplar
G Boom base 3/4″ × 2″ × 21″ poplar
H Upper backstay mount 3/4″ × 3/4″ × 13/4″ poplar*
I Cleat 3/4″ × 1/2″ × 11/2″ poplar
J Mast column (2) 3/4″ × 11/2″ × 281/4″ poplar
K Mast transverse member (2) 3/4″ × 11/2″ × 5″ poplar
L Mast diagonal brace (2) 3/4″ × 3/4″ × 10″ poplar
M Motor mount plate 1/16″ × 3″ × 3″ aluminum plate* (McMaster-Carr #8975K295)
N Shaft 3/8″-diameter steel rod, 61/2″ long
a Gearhead motor Standard gearmotor, 3 rpm (SparkFun #ROB-12262)
b Shaft coupler 0.375″ to 6 mm set screw shaft coupler (ServoCity #625202)
c Ball bearing (2) Flanged ball bearing, 3/8″ shaft diameter (McMaster-Carr #6383K232)
d Boom and tower pivot 3/8″ × 8″ he× bolt with nut and washer
e Pulley Double pulley (McMaster-Carr #3099T23)
f Pulley attachment Steel U-bolt, 1/4″-20 thread, 3/4″ inside diameter (McMaster-Carr
#3201T31)
Motor mount screws (4) Metric steel screw, M3 × 0.5 mm × 5 mm (McMaster-Carr
#92005A114)
- Backstay 1/4″ nylon rope (working load at least 100 pounds)
- Lifting rope 1/8″ nylon cord (working load at least 50 pounds)
*Full-size cutting pattern provided in the PDF version of this guidebook.

The vendors referenced in the materials list above are:

}} McMaster-Carr (https://www.mcmaster.com/)
}} ServoCity (https://www.servocity.com/)
}} SparkFun Electronics (https://www.sparkfun.com/)

All other materials are available from most building supply stores.

Project 16 » Electric-Powered Crane 293


In addition to common hand tools, this project requires extensive woodworking equipment, including:

}} table saw (for rip-cutting)


}} miter saw or radial arm saw (for cross-cutting)
}} scroll saw or bandsaw (for curved cuts and tapers)
}} power-sanding station (for shaping)
}} drill press (for high-precision drilling)
}} hacksaw or bandsaw (for metal cutting)
}} files
}} clamps

Other supplies needed for this project include wood glue, spray adhesive, 11/4-inch wood screws, 11/2-inch
wood screws, and 1/2-inch panhead screws.

FABRICATION AND ASSEMBLY PROCEDURE

1 Prefabricate the wooden and metal components specified as A–N in the materials list and drawing
16.3 above. For components that have cutting patterns, use spray adhesive to affix the patterns to
the appropriate material and then cut to size. Refine the shape with a sander or file as needed. For
components that do not have cutting patterns, simply use the dimensions provided in the materials list
and in drawings 16.4–16.6. Drill holes as indicated on the drawings. When drilling holes in metal, be
sure to use a clamp or vise to hold the workpiece.

2 Glue and clamp the two motor mount sides (component D) to one of the main frame sides (component
B) as shown in drawing 16.4; then, reinforce these joints with 11/4-inch wood screws.

3 Attach the gearhead motor (component a) to the motor mount plate (component M) with four M3 ×
0.5 mm × 5 mm screws. The four holes in the front face of the motor are threaded to accept this metric
screw size.

4 Attach the motor mount plate (component M) to the motor mount base (component C) with four 1/2-
inch panhead screws, one at each corner.

5 Attach the motor mount base (component C) to the two motor mount sides (component D) with four 11/2-
inch wood screws. To facilitate disassembly of the motor mount, do not glue components C and D together.

6 Assemble the main frame (components A, B, and E) with glue and clamps as shown in drawing 16.4;
then, reinforce all joints with 11/4-inch wood screws.

7 Assemble the boom (components F, G, H, and I) with glue and clamps as shown in drawing 16.5; then,
reinforce the joints between components F, G, and I with 11/4-inch wood screws.

294 
8 Attach the upper pulley (component e) to the boom with a U-bolt (component f) and two 1/4-inch nuts.

9 Assemble the mast (components J, K, and L) with glue and clamps as shown in drawing 16.6; then,
reinforce all joints with 11/4-inch wood screws.

10 Insert the two ball bearings (component c) into the 11/8-inch-diameter holes in the main frame sides.
Then, insert the steel shaft (component N) through these bearings.

11 Connect the shaft (component N) to the gearhead motor (component a) with the shaft coupler
(component b). Ensure that the set screws are tight.

12 Install the boom and mast on the main frame with a 3/8-inch bolt, as shown in drawings 16.1 and 16.2.

13 Rig the boom by running a length of 1/4-inch nylon rope from the backstay anchorage (component E)
up over the top of one mast column, through the upper backstay mount (component H) at the top of
the boom, back over the top of the other mast column, and down to the opposite side of the backstay
anchorage (component E). Tie both ends of the rope securely to the anchorage. With the backstay
rigged, the boom should be angled at approximately 45°.

14 Rig the lifting rope. Tie one end of the 1/8-inch nylon cord tightly to the shaft (component N). Wrap
several turns around the shaft so that the rope will not slip. Then, run the cord up through one wheel
of the upper pulley, down through one wheel of the lower pulley, up through the second wheel of the
upper pulley, down through the second wheel of the lower pulley, and up through the hole at the top of
the boom. Secure the free end with several wraps around the cleat (component I).

The electric-powered crane is now complete and ready for testing!

TESTING PROCEDURE
1 Clamp the rear of the crane’s main frame to the top of a sturdy bench or table.

2 Connect a DC benchtop power supply to the motor and set it to 12 volts. (As an alternative, four 1.5-
volt household batteries connected in series can be used.)

3 With the motor running, check the operation of the system. Ensure that the direction of the motor’s
rotation is winding—rather than unwinding—the lifting rope. If necessary, the motor’s rotation can
be reversed by reversing the electrical connections. Ensure that all four pulley wheels are rotating freely.
Then, disconnect the power source.

Project 16 » Electric-Powered Crane 295


4 Use 1/4-inch nylon rope to rig a lifting sling to the 100-pound load, as shown in lesson 23. Fasten the
sling to the crane’s lower pulley.

5 Reconnect the power source to lift the load. Be patient! Operating at its rated speed, the motor will lift
the load at less than 1 inch per minute. This is the cost of mechanical advantage!

296 
PROJECT 17
TRIBUTE TO RUBE
GOLDBERG
DESIGN, BUILD, AND TEST A MACHINE IN THE
SPIRIT OF RUBE GOLDBERG’S CARTOONS.
LESSON IN WHICH THIS PROJECT IS COVERED:
24 » Creative Design: A Tribute to Rube Goldberg

PROBLEM DEFINITION
Requirement
}} The purpose of the machine is to click and release a mouse button.
}} The machine must incorporate all of the following elements:
§§ linear spring
§§ braced frame
§§ lever
§§ beam
§§ suspension bridge
§§ sail
§§ lighter-than-air object
§§ aerodynamic lift
§§ aerodynamic stability
§§ propeller
§§ electric circuit
§§ torsion spring
§§ counterweight
§§ hydraulic system
§§ bell crank
§§ hydropower
§§ gear train

Project 17 » Tribute to Rube Goldberg 297


§§ pendulum
§§ electric motor
§§ pulley
}} All elements must contribute to the machine’s purpose.
}} Operation will be initiated by only one human action.

Constraints:
}} Must be buildable.
}} Must operate safely indoors.
}} Must use commonly available materials.

A Tribute to Rube Goldberg


Throughout this course, we have seen that the modern engineering enterprise is characterized by the design and
construction of technologies that perform important functions while operating as efficiently as possible. Yet one of
America’s most famous engineers earned great notoriety by doing exactly the opposite—designing excessively complex,
inefficient devices that accomplished laughably simple tasks. His name was Rube Goldberg.

Reuben Lucius Goldberg lived from 1883 to 1970. He earned an engineering degree from the University of California,
Berkeley, in 1904 and practiced engineering briefly before changing course and pursuing what would become a highly
successful career as a cartoonist and author.

The cartoons that have earned Rube Goldberg lasting fame depict farcical inventions that perform simple tasks by
fantastically complicated means. A typical example, Goldberg’s “Self-Operating Napkin” (See the cartoon in lesson 24).

How does it work? As the diner raises the spoon (A) to his mouth, a string (B) pulls downward on the handle of a
ladle (C), throwing a cracker (D) past a hungry parrot (E), which flies from its perch to retrieve the meal, thereby
causing the perch (F) to tilt, tipping a bowl of seeds (G) into the adjacent bucket (H). The bucket’s added weight pulls
a cord (I), which activates a cigarette lighter (J), igniting the fuse of a skyrocket (K). When the rocket launches, the
attached sickle (L) cuts the string (M), which releases the clock pendulum, to which a napkin has been attached. As
the pendulum oscillates back and forth, the napkin wipes the diner’s mouth. This wonderfully ridiculous machine is
configured as a hat, which enhances its portability while also making a rather dramatic fashion statement.

This is Rube Goldberg at his finest.

Goldberg created these “inventions” to entertain us—and certainly not to communicate the ethic of engineering
efficiency. Yet they do have potential learning value in at least three respects:

†† Goldberg’s inventions reflect the integrative nature of modern engineering, demonstrating how a wide variety
of mechanical (and biomechanical) subsystems can be combined to accomplish a single function.

†† They illustrate the creative dimension of engineering, demonstrating quite vividly that there is always more
than one valid approach to solving a technological problem.

298 
†† Most importantly, from the perspective of this course, Goldberg’s inventions provide us with a
wonderful opportunity to review key concepts from earlier projects while also testing our newly
acquired DIY engineering know-how.

Note that the self-operating napkin strongly resembles a graphical syllabus for this course, addressing
such topics as structural frames, parabolic cables, catapults, projectile motion, aeronautical systems,
levers, torque, application of load with a bucket full of stuff, potential and kinetic energy, pulleys, friction,
rocketry, impulse-momentum, pendulum motion, and—we may safely infer—gear trains inside the clock.
If Rube Goldberg was able to summarize this course so effectively without even trying, how well might we
be able to do if we really put our minds to it?

DESIGN

One possible solution to this design problem is presented in lesson 24. There are infinitely many more just
waiting to be conceived, sketched, developed, built, and tested. The greatest joy in this sort of project lies in
the exercise of unconstrained creativity. It is now time for you to experience this joy by developing a design
of your own.

Project 17 » Tribute to Rube Goldberg 299


GLOSSARY
3-D printing: A process by which a computer model is translated into a physical three-dimensional object
by building up layers of material—typically plastic or metal—under precise computer control. Also called
additive manufacturing.

acceleration: The rate of change of velocity, expressed in units of distance per time squared.

actuator: A hydraulic machine component used to apply pressure to the hydraulic system.

additive manufacturing: A process by which a computer model is translated into a physical three-
dimensional object by building up layers of material—typically plastic or metal—under precise computer
control. Also called 3-D printing.

advancing blade: In helicopter aeronautics, the rotor blade moving in the same direction as the aircraft at a
given instant.

adverse pressure gradient: A condition characterized by the movement of a fluid from a region of
relatively low pressure to a region of higher pressure. In aerodynamics, an adverse pressure gradient is the
fundamental cause of stalling.

aerodynamic stability: The capacity of an aircraft to maintain its attitude despite disturbances caused by
wind gusts, turbulence, etc.

airfoil: The shape of a wing, blade, or sail as viewed in cross section. An airfoil moving through a fluid
develops an aerodynamic force.

airframe: The structure of an aircraft.

ampere: A unit of measure for electric current. The term is often shortened to amp and abbreviated A.

amplitude: In simple harmonic motion, the maximum displacement of an oscillating object, measured
from its static equilibrium position.

anchor: A component of an escapement in a pendulum clock. This pivoting arm engages with the timing
wheel to sustain the motion of the pendulum while also translating the oscillatory motion of the pendulum
into the controlled unidirectional rotation of a shaft that drives the clock hands. At each end of the anchor
is a prong called a pallet.

anchorage: A structure at which one end of one main cable of a suspension bridge is anchored.

angle of attack: The angle of a wing or airfoil, measured with respect to the flow of undisturbed air.

300 
anti-torque rotor: On a helicopter, a small vertically oriented propeller that counterbalances the torque
effect of the main rotor. Also called a tail rotor.

apogee: The highest altitude achieved in the trajectory of a rocket.

apparent wind: The relative wind velocity and direction felt from the perspective of a moving reference
frame (typically a vehicle), as distinct from the true wind.

arming key: In model rocketry, a safety switch that must be activated by the insertion of a key before the
launch controller can fire an igniter.

aspect ratio: In aerodynamics and aerospace engineering, the ratio of the span length to the chord of a
wing or fin.

attitude: The orientation of an aircraft flying in three-dimensional space, defined in terms of the pitch
axis, yaw axis, and roll axis of the aircraft.

backstay: A cable running from the top of a tower or mast to a fixed support below and to the rear.

balance: In airplane design, adjusting the weight distribution of an aircraft to achieve a suitable center of
gravity location.

ballast: A heavy material added to improve the stability or balance of a vehicle or vessel.

ballista: A two-armed stone-throwing torsion catapult developed by Hellenistic-era Greeks in the 3rd
century B.C. and subsequently perfected by Roman military engineers. Also called a palintone.

beam: A structural member that carries load primarily in flexure, or bending.

bearing pad: A structural component used to transmit a concentrated reaction force between a member
and its supporting foundation.

Bernoulli’s equation: A law of physics that states the mathematical relationship between velocity, pressure,
and elevation at two points in a steady fluid flow.

Bernoulli’s principle: A law of physics stating that when the speed of a moving fluid increases, its pressure
decreases, and when the speed decreases, the pressure increases.

blast deflector: In model rocketry, a metal plate that redirects the stream of hot exhaust gases propelled
downward by the rocket engine during a launch. The blast deflector is a component of the launchpad.

booster: A rocket component that provides the thrust and stability necessary to lift a payload to the
desired altitude.

braced frame: A structural frame that uses diagonal braces as its principal means of carrying lateral loads.

Glossary 301
buckling strength: The largest compression force a member can carry before buckling.

buckling: A failure mode in which a member subjected to compressive loading suddenly and
catastrophically displaces laterally and collapses.

buoyant force: An upward force caused by immersing a solid body in a fluid. According to the principle
of buoyancy, the magnitude of the buoyant force is equal to the weight of the displaced fluid.

camber: In aerodynamics, the shape of an imaginary line drawn from the leading edge to the trailing
edge of an airfoil at exactly mid-height between the top and bottom of the airfoil.

canopy: The upper portion of a parachute. The canopy provides the aerodynamic drag that slows the
descent of the parachute’s payload.

center of effort (CE): On a sailing vessel, the point of application of the aerodynamic force caused by wind.

center of gravity (CG): The balance point of an object. The weight of the object is considered to act at its
center of gravity.

center of lateral resistance (CLR): On a sailing vessel, the point of application of the lateral resisting force
caused by the hull moving through the water.

center of pressure: When the attitude of a rocket is not aligned with its direction of flight, the point at
which the resulting lateral force acts.

chord: The width of a wing.

circuit: A continuous conducting pathway along which electric current can flow.

circular pitch: A geometric property of a gear, calculated as π divided by the diametrical pitch.

cleat: A stationary fitting used to secure the end of a rope.

close-hauled: Sailing as close to the wind direction as possible.

coil spring: A spring made from metal wire formed into a coil.

collective pitch: A helicopter cockpit control that varies the pitch of all rotor blades collectively, causing
the aircraft to ascend or descend.

column: A structural member—usually oriented vertically—that carries load primarily in compression.

commutator: In an electric motor, a device that reverses the direction of the current flow through the coil
of the rotor with every half-turn.

302 
component of a force: The portion of a force that can be considered to act in a given direction, determined
in accordance with the rules of vector mathematics.

compound gearset: Two or more gears fixed to a single shaft. When multiple compound gearsets are
combined, large gear ratios can be obtained with gears of relatively small size.

compression spring: A coil spring that is designed to be compressed or shortened.

compression: An internal force that causes shortening in a structural member or spring.

compressive deformation: Physical shortening of a structural member subjected to internal compression.

compressive strength: The largest compression force a member can carry before failing.

concrete: A rocklike structural material formed by mixing Portland cement, aggregate (typically gravel
and sand), and water.

conductor: A material that conducts electricity with minimal resistance.

conservation of energy: A law of physics that states that the total energy of an isolated system remains
constant. An implication of this principle is that energy can neither be created nor destroyed but can be
transformed from one form to another.

cosine (abbreviated cos): In trigonometry, the adjacent side of a right triangle divided by the hypotenuse.

cross section: The shape defined by passing a plane through a structural member perpendicular to the
longitudinal axis of the member.

cross-sectional area: The area of a cross-sectional shape.

current: A measure of the movement of electrons through a circuit due to a difference in voltage. Current
is measured in units of amperes, or amps (A).

cyclic pitch: A helicopter cockpit control that varies the pitch of all rotor blades cyclically, causing the
aircraft to move forward, rearward, or to one side.

deck: A structural element that forms the floor of a bridge—the horizontal surface on which vehicles and
pedestrians are supported.

deflection: The vertical distance a beam bends in response to its applied loads.

deformation: A change in the physical shape or dimensions of a structural element or spring.

Glossary 303
delay charge: In model rocketry, a block of flammable material contained within a solid-fuel rocket engine
between the propellant and the ejection charge to provide a time delay between propellant burnout and
the ignition of the ejection charge. The delay charge provides no thrust.

design basis: A mathematical relationship used as the basis for a structural design. A design basis can take
many forms, but it always derives from the requirement that an actual condition must be less than or equal
to an allowable condition.

diagonal brace: A diagonally oriented structural member that carries lateral loads in a braced frame.

diametrical pitch: A geometric property of a gear, calculated as the number of teeth divided by the pitch
diameter.

diaphragm: A planar structural component that resists lateral deformation and shearing in the plane of
the diaphragm.

dihedral angle: In aircraft design, the upward tilt of the two halves of a wing, measured with respect to
the horizontal.

direct current: Steady, unvarying electric current, as is produced by a battery.

dissymmetry of lift: In helicopter aeronautics, the development of unequal lift by individual rotor blades
due to the forward movement of the aircraft.

distributed load: A load that is applied continuously along a structural element and is typically expressed
in units of force per length or force per area.

drag: The resisting force exerted on a body as it moves through a fluid.

driven gear: In a gear train, the gear that receives power from the driving gear and transmits it onward.

driving gear: In a gear train, the gear that receives power from the machine and transmits it onward to the
driven gear.

dual-axis tracking: A method of measuring altitude using two ground-based tracking stations.

ejection charge: In model rocketry, an explosive material contained in a solid-fuel rocket engine for the
purpose of blowing the nose cone off the rocket and deploying the parachute or other recovery device.

elastic energy: The energy stored though the elastic deformation of a material.

elastic section modulus: A measure of the resistance of a cross-sectional shape to bending.

elastic: Material behavior characterized by the material returning to its original shape after being unloaded.

304 
elbow: On a robotic arm, a pivoting joint near the center of the arm.

electricity: The flow of electrical charge.

escapement: A mechanism that sustains the motion of the pendulum in a pendulum clock while also
translating the oscillatory motion of the pendulum into the controlled unidirectional rotation of a shaft
that drives the clock hands.

factor of safety: A number, always greater than 1, that provides a margin of error to account for sources
of uncertainty inherent in structural design (e.g., unanticipated loads, natural variability in material
properties, fabrication errors).

feathering: Rotation of a helicopter rotor blade about its own longitudinal axis.

fillet: In model rocketry, a line of glue or putty that is used to round and strengthen interior corners (e.g.,
at the joints between the fins and body tube).

fin: A projecting planar surface (also called a vertical stabilizer) added to an aircraft or rocket for the
purpose of enhancing stability.

flange: One of two horizontal elements of an I-shaped cross section.

flexural stress: Tensile or compressive stress developed in a bending beam.

flexure formula: An equation used to calculate flexural stress in a beam.

flexure: Structural behavior characterized by bending.

fluid: A liquid or gas.

foot: The bottom edge of a sail.

footing: A structural foundation element that transmits the compressive force in a column or wall
downward to the soil below.

force: A push or a pull applied to a body.

form drag: Resistance caused by variations of pressure associated with the flow of air around a body. A
streamlined shape has low form drag while an angular shape typically has higher form drag.

former: A transverse bulkhead used to form the shape of a curved three-dimensional object.

forward current: The electric current a light-emitting diode (LED) can handle continuously without
damage.

Glossary 305
forward voltage drop: The voltage drop associated with current flowing through a diode.

frame: A structural system typically composed primarily of beams and columns.

free body diagram: A graphical representation of a body isolated from its surroundings and showing all
forces and moments acting on the body. The free body diagram is a tool used to solve problems involving
the principle of equilibrium.

freeboard: The vertical distance between the gunwale of a boat or ship and the surface of the water.

freewheeling: Free rotation of a propeller when it is not being driven by its power source.

frequency: In simple harmonic motion, the number of oscillations per unit of time.

fulcrum: The support on which a lever pivots.

fuselage: The body of an aircraft.

gear ratio: The mechanical advantage provided by a pair of meshed gears, calculated as the number of
teeth on the driven gear divided by the number of teeth on the driving gear.

gear train: A system consisting of two or more gears mounted with their teeth meshed.

gear: A toothed wheel that engages with one or more other gears for the purpose of transmitting torque
from one shaft to another.

gearhead motor: A small electric motor with its own built-in gear train to increase output torque.

gondola: The crew cab of a blimp or airship.

gravitational potential energy: The energy associated with an elevated mass subjected to the acceleration
of gravity.

gravity load: A downward force associated with anything that has weight (e.g., structural self-weight,
occupants, furniture, vehicles).

gunwale: The uppermost edge of the hull of a boat or ship.

gusset: A plate or bracket used to strengthen a structural joint.

gusset plate: A structural component used to connect two or more members together at a joint.

heading: The direction of travel.

heel angle: The angle at which a sailboat is tipped laterally due to the force of the wind acting on its sails.

306 
heeling: The lateral tipping of a sailboat due to the force of the wind acting on its sails.

heeling moment: The moment caused by the aerodynamic force of wind acting on a sail. The heeling
moment causes the sailboat to tip laterally and is counteracted by the righting moment.

hole-carrier: A ballista component on which one or more torsion springs are mounted. The term “hole-
carrier” is translated from the Greek περίτρητον.

hydraulic cylinder: A hydraulic machine component that uses system pressure to support a load or move
this load through a controlled distance.

hydraulic fluid: A liquid that serves as the medium for transmission of force through a hydraulic
machine via fluid pressure.

hydraulic line: A hydraulic machine component that connects the actuator and hydraulic cylinder such
that the fluid pressure is equal in both.

hydraulic machine: A machine that uses fluid pressure to transmit, amplify, and control the application
of forces.

hydraulic ram: The movable element of a hydraulic cylinder.

hydrostatics: The study of equilibrium and stability of objects immersed in a fluid that is not moving.

impulse: The integral of force with respect to time over a finite time period. Impulse is expressed in units
of force times time.

impulse-momentum principle: A law of physics stating that the total impulse imparted to a body over a
finite period of time is equal to the change in the momentum of that body during the same period.

impulse turbine: A turbine that uses a high-velocity jet of water striking the blades of a vertically oriented
runner to produce power.

incidence angle: In aircraft design, the angle at which the wing is mounted on the fuselage.

induced drag: Resistance occurring when high-pressure air below a wing spills over the wingtips toward
the low-pressure region above the wing, causing vortex that trails behind the airplane and resists its forward
movement. Also called vortex drag.

interference: A phenomenon that occurs in a gear train when the tips of one gear’s teeth cut into the teeth
of the adjacent gear. This phenomenon occurs when the number of teeth on one gear is too small.

internal force: A force developed within a structural member in response to applied loads.

Glossary 307
internal moment: The tendency of external loads to cause bending of a structural member. Flexural
stress is proportional to internal moment.

isochronism: The independence of period from amplitude in an oscillating system.

joint: A node at which two or more structural members are connected together.

keel: On a sailboat, a vertical fin extending downward from the bottom of the hull. The keel prevents the
boat from being blown sideways.

kinetic energy: The energy associated with a mass in motion.

lateral load: A predominantly horizontal force associated with wind or earthquake effects.

lateral-torsional buckling: A flexural failure that occurs when the compression force in the top half of a
beam causes it to buckle sideways and twist.

launch lug: In model rocketry, a small paper tube that engages with a metal launch rod to guide the
rocket upward from the launchpad until its velocity is high enough for the fins to provide stability. One or
more launch lugs are permanently attached to the body tube of the rocket booster.

launch rod: In model rocketry, a metal rod that engages with the rocket’s launch lugs to guide the rocket
off the launchpad until its velocity is high enough for the fins to provide stability. The launch rod is a
component of the launchpad.

leach: The aft edge of a sail.

lead: On a sailing vessel, the longitudinal distance between the center of effort and center of lateral
resistance.

leading edge: The forward edge of a wing or propeller.

leeward: Nautical term for the downwind direction—the direction toward which the wind is blowing.

leeward drift: In sailing, movement through the water in the leeward direction due to a component of the
aerodynamic force that is perpendicular to the vessel’s forward motion. Also called leeway.

leeway: In sailing, movement through the water in the leeward direction due to a component of the
aerodynamic force that is perpendicular to the vessel’s forward motion. Also called leeward drift.

left-handed propeller: A propeller that produces forward thrust when rotating counterclockwise viewed
from the perspective of the pilot looking forward.

lever: A rod or bar that pivots on a fulcrum and is used to gain mechanical advantage.

308 
lift: The upward force that counterbalances the weight of an aircraft.

light-emitting diode (LED): An electronic component that emits light when subjected to a suitable voltage.

line loading: A load that is distributed continuously and uniformly along a specified length of a structural
member and is expressed in units of force per length.

load: In structural engineering, an external force applied to a structure. In mechanical engineering, the
demand (in force or torque) placed on the shaft of a turbine or motor.

luff: The forward edge of a sail.

machine: A device consisting of a source of power and a means of controlling its application.

main rotor: On a helicopter, a large horizontally oriented propeller that provides lift.

mechanical advantage: The multiplication of force caused by a lever, pulley system, gear train,
hydraulic machine, etc.

mechanical efficiency: A quantitative measure of machine performance typically defined as the ratio of
useful output to input, expressed as a percentage.

method of joints: A method by which the principle of equilibrium is used to calculate internal forces in
the members of a structure.

model: A graphical representation of the essential components of a physical system defined in a way that
facilitates the application of scientific principles to predict the system’s performance.

modulus of elasticity: The stiffness of a material, expressed in units of force per area.

modulus of rupture: The stress at which a given type of wood fails in flexure.

moment: The tendency of a force to cause rotation, expressed in units of force times distance.

momentum: The mass of an object times its velocity.

neutral surface: In flexure, a horizontal plane along which there is no deformation.

neutrally stable: At the borderline between stable and unstable.

no-go zone: In sailing, a range of headings at or near the windward direction within which it is
impossible for a sailboat to make forward headway.

no-load speed: The rotational speed of a motor or turbine when no load (or resisting torque) is applied to
the shaft.

Glossary 309
ohm: The standard unit of measure for electric resistance, represented by the Greek letter omega (Ω).

Ohm’s law: A law of physics that states that electric current is equal to voltage divided by resistance.

onager: A one-armed stone-throwing torsion catapult developed in the late Roman imperial era and also
used in the medieval era.

operating envelope: The overall range of motion of the outer end of a robotic arm.

palintone: A two-armed stone-throwing torsion catapult developed by Hellenistic-era Greeks in the 3rd
century B.C. and subsequently perfected by Roman military engineers. Also called a ballista.

pallet: A component of an escapement in a pendulum clock. This prong at each end of the anchor
engages with the timing wheel such that a small force is transmitted from the timing wheel to the
pendulum with each oscillation. This force sustains the motion of the pendulum.

payload: The useful cargo carried aloft by a rocket. Also, the load suspended beneath a parachute canopy.

payload bay: A rocket component that encloses and protects the payload.

peak forward current: The maximum electric current a light-emitting diode (LED) can handle for
short periods without damage.

pendulum: An object that oscillates with a particular period and frequency by rotating about its
upper end.

pendulum bob: The weight mounted at the lower end of the pendulum on a pendulum clock.

penstock: In a hydropower system, a pipe that delivers fluid from the reservoir to the turbine.

period: In simple harmonic motion, the time required for one full oscillation.

pitch: The forward distance through which a propeller would advance during one full rotation if it did
not slip with respect to the fluid through which it is moving.

pitch axis: One of three axes defining the attitude of an aircraft. The axis about which an aircraft
rotates when it dives or climbs.

pitch circle: An imaginary circle centered on a gear with its arc passing approximately through mid-
height of the gear teeth. When two gears are properly meshed, their respective pitch circles are tangent
to each other.

pitch diameter: The diameter of the pitch circle of a gear.

pitch linkage: A rod that controls the feathering rotation of a helicopter rotor blade.

310 
pitch radius: The radius of the pitch circle of a gear.

point of sail: The direction of a sailboat, defined with respect to the wind direction.

Portland cement: A fine gray powder manufactured from limestone, clay, and other ingredients and
used as one of the principal constituents of concrete.

power: The rate at which work is done.

pressure angle: The orientation of the line of force transmission between the teeth on two meshed
gears, measured with respect to the tangent to the pitch circle. Most manufactured gears use standard
pressure angles of 14.5°, 20°, or 25°.

principle of buoyancy: A law of physics that states that immersing an object in a fluid causes an
upward force—called the buoyant force—with a magnitude equal to the weight of the displaced fluid.

principle of conservation of energy: See conservation of energy.

principle of equilibrium: A law of physics that states that for a body that is at rest or moving at
constant velocity, the vector sum of all forces and moments acting on the body is zero.

propeller: A device that converts rotational motion into thrust.

pulley: A wheel on an axle designed to change the direction of a taut rope or cable without changing its
tension force.

pulley system: Two or more pulleys interconnected in a way that provides mechanical advantage.

radian: A non-dimensional measure of an angle. An angle of 2π radians is equal to 360°.

reaching: Sailing in a direction approximately perpendicular to the wind direction. Also called sailing
on a beam reach.

reaction: A force or moment developed at a structural support that keeps the structure in equilibrium
with its applied loads.

relative wind: In helicopter aerodynamics, the total airflow felt by a main rotor blade that is moving
forward (as a result of blade rotation) and vertically (as a result of teetering).

resistance: A measure of the extent to which a conductor or electrical component impedes the
movement of electrons. The standard unit of resistance is the ohm, represented by the Greek letter
omega (Ω).

resisting force: An internal force developed in a spring or structural member in response to


deformation, either in tension or compression.

Glossary 311
resistive heating: Heat generated by the flow of current through a conductor that has resistance.

resistor: An electrical component used to add a specific resistance to a circuit.

retreating blade: In helicopter aeronautics, the rotor blade moving in the opposite direction as the
aircraft at a given instant.

right-handed propeller: A propeller that produces forward thrust when rotating clockwise viewed
from the perspective of the pilot looking forward.

righting moment: The moment that counterbalances the heeling moment in a sailboat. The righting
moment is caused by the hull’s buoyant force offset laterally from its center of gravity.

roll axis: One of three axes defining the attitude of an aircraft. The axis about which an aircraft rotates
when it banks to the left or right.

rotary-wing aircraft: A helicopter.

rotor: The rotating element of an electric motor. The rotor usually incorporates an electric coil,
which creates a magnetic field that interacts with another magnetic field produced by the stator, thus
providing the torque that propels the motor.

rotor disk: An imaginary circular disk defined by the sweep of a helicopter’s spinning main rotor.

rubber peg: On a rubber-powered model airplane, the dowel that anchors the rear end of the rubber
motor.

rudder post: The vertical pivot, or post, on which the rudder rotates.

runner: The rotating element of a turbine consisting of a wheel with radial blades, or vanes, extending
outward from its rim.

running: Sailing at or very near the direction toward which the wind is blowing.

rupture: A structural failure mode in which a member physically breaks under tensile loading.

saddle: A saddle-shaped structural component located at the top of a suspension bridge tower and used to
transmit force from the main cable downward into the tower.

sag: The vertical distance between the lowest point on a draped cable and the cable supports.

sailing on a beam reach: Sailing in a direction approximately perpendicular to the wind direction. Also
called reaching.

312 
segment of a circle: A geometric shape defined by an arc of a circle and a chord connecting the two ends of
the arc.

semirigid rotor: A type of helicopter rotor using two blades that can rotate about both the feathering and
teetering axes. Also called a teetering rotor.

sheet: On a sailing vessel, a rope used to control the position of the sail relative to the direction of the wind.

shoulder: On a robotic arm, a pivoting joint at the inner end of the arm.

sine (abbreviated sin): In trigonometry, the opposite side of a right triangle divided by the hypotenuse.

skin friction drag: Resistance caused by the friction associated with a fluid moving across a surface. If the
surface is smooth, skin friction drag is low; if the surface is rough or bumpy, skin friction drag is higher.

slider: A movable ballista component that incorporates a trigger mechanism and a trough that guides the
projectile launch.

snap swivel: A metal fitting used to connect two objects together while allowing them to rotate freely
and independently.

span rating: A rating system used by the Engineered Wood Association for plywood loaded in flexure.
The rating consists of two numbers separated by a slash. The left number is the maximum allowable
spacing of supports in inches when the plywood is used for roof sheathing in residential construction; the
right number is the maximum spacing of supports in inches for flooring.

spring: A device that stores energy through the elastic deformation of a material.

spring energy: The elastic energy stored in a spring that has undergone a change in length or a rotational
deformation.

spur gear: A conventional planar gear with radially oriented teeth.

stabilizer bar: A helicopter main rotor component consisting of a horizontal metal bar mounted
perpendicular to the rotor blades with a weight at each end. The purpose of the stabilizer bar is to hold the
rotor in its current plane of rotation, resisting perturbations caused by wind gusts or other disturbances.

stall: In aerodynamics, the sharp loss of lift and increase in drag caused by the separation of airflow on the
top of a wing.

stall torque: The lowest level of resisting torque that prevents the shaft of a motor or turbine from spinning.

stator: A stationary element of an electric motor. The stator usually incorporates one or more magnets,
which create a magnetic field that interacts with another magnetic field produced by the rotor, thus
providing the torque that propels the motor.

Glossary 313
stiffness: The resisting force developed in a spring or structural member per unit change of length.

streamline: A line depicting the steady flow of a fluid.

strength: The maximum internal force or moment a structural member can resist before failure.

stress: A measure of the intensity of internal force in a member, expressed in units of force per area.

structural analysis: The systematic application of engineering mechanics principles to determine the
internal forces and moments in one or more members of a structural system.

structural model: A graphical representation of a structural system defined in a way that facilitates a
structural analysis.

suspender: A vertical wire used to suspend the deck structure of a suspension bridge from one of the
main cables.

suspension line: A string or cord that connects a parachute’s canopy to its payload.

swash plate: A component of a helicopter’s main rotor hub that translates the pilot’s collective and cyclic
pitch control inputs into corresponding changes in the pitch of the individual rotor blades.

tail rotor: On a helicopter, a small vertically oriented propeller that counterbalances the torque effect of
the main rotor. Also called an anti-torque rotor.

tangent (abbreviated tan): In trigonometry, the opposite side of a right triangle divided by the adjacent side.

teetering: Tipping (seesaw) rotation of a helicopter’s main rotor about a horizontal axis perpendicular to
the blades’ feathering axis.

teetering rotor: A semirigid rotor of a helicopter.

tensile deformation: Physical elongation of a structural member subjected to internal tension.

tensile strength: The largest tension force a member can carry before failing.

tension: An internal force that causes elongation in a structural member or spring.

tension spring: A coil spring that is designed to be stretched or elongated.

terminal velocity: The velocity at which the drag developed by a falling body is exactly equal to the
weight of the body. When a falling body reaches terminal velocity, its acceleration stops.

theodolite: An instrument used to measure both horizontal and vertical angles for determining altitudes by
the dual-axis tracking method.

314 
thickness form: A mathematical function defining the thickness of an airfoil at every point along the
camber line. The thickness form is always defined in terms of pairs of equal offsets, one up and one down.

thrust: A force provided by a propulsion system of a vehicle.

thrust bearing: A rotary bearing that reduces friction at a rotating joint that is also subjected to axial
compression.

thrust curve: A graph of thrust versus time for a model rocket engine. The area underneath the thrust
curve corresponds to the impulse delivered by the engine.

thrust line: In aircraft design, an imaginary line corresponding to the axis of the propeller.

tiller: On a boat, a lever attached to the top of the rudder post and used for steering.

timing wheel: A component of an escapement in a pendulum clock. This gear engages with the anchor to
sustain the motion of the pendulum while also translating the oscillatory motion of the pendulum into the
controlled unidirectional rotation of a shaft that drives the clock hands.

torque: The tendency of a force to cause rotation about a shaft or axis. Torque is a type of moment.

torque effect: The tendency of an aircraft to roll in the opposite direction of its propeller’s rotation.

torsion: The twisting behavior of a structural member, shaft, or spring.

torsion spring: A spring that stores elastic energy by twisting.

trailing edge: The rear edge of a wing or propeller.

transom: The flat surface at the stern of a boat hull.

trapezoidal rule: A numerical technique for calculating the area under a curve by subdividing this area
into trapezoidal shapes and summing up their individual areas.

trebuchet: A medieval one-armed stone-throwing catapult that used a falling weight as the source of energy
for launching projectiles.

trimming: The process of adjusting the balance and controls of an aircraft to achieve optimum flight
characteristics.

true wind: The actual wind velocity and direction, as distinct from the apparent wind.

trunnion: A cylindrical mount for a pivoting bearing on a machine.

turbine: A rotary machine that uses a moving fluid to do work.

Glossary 315
vector: A mathematical entity that has both magnitude and direction.

vertical stabilizer: A projecting planar surface (also called a fin) added to an aircraft or rocket for the
purpose of enhancing stability.

voltage: A measure of the difference in electrical potential energy between two points. Voltage is measured
in volts (V ).

vortex: A spiraling wake of air, which is the cause of vortex drag.

vortex drag: Resistance occurring when high-pressure air below a wing spills over the wingtips toward the
low-pressure region above the wing, causing a vortex that trails behind the airplane and resists its forward
movement. Also called induced drag.

weather cocking: The tendency of a rocket to incline its trajectory into the wind.

weather helm: The tendency of a sailboat to turn windward with no change in the rudder position.

web: The vertical element of an I-shaped cross section.

wind tunnel: A device used in aerodynamic research to study the effects of air moving around solid objects.

windlass: A hand-operated winch that is used to pull a heavy load by winding rope onto its rotating drum.

windward: Nautical term for the direction from which the wind is blowing.

wing area: The area of a wing, calculated by multiplying the span length by the chord.

work: The transfer of energy associated with a force moving through a distance.

yaw axis: One of three axes defining the attitude of an aircraft. The axis about which an aircraft rotates
when it turns to the left or right in a horizontal plane.

yoke: A component of an escapement in a pendulum clock. This pivoting arm mechanically links the
pendulum to the anchor.

316 
BIBLIOGRAPHY
Blattenberger, Kirt. “Penni Helicopter.” Airplanes and Rockets. http://www.airplanesandrockets.com/
helicopters/penni-helicopter-jan-1970-aam.htm.

Brewer, Ted. Understanding Boat Design. Camden, ME: International Marine, 1994.

Breyer, Donald E., Kenneth Fridley, Kelly Cobeen, and David Pollock. Design of Wood Structures—ASD.
New York: McGraw-Hill, 2003.

Campbell, Duncan. Besieged: Siege Warfare in the Ancient World. New York: Osprey, 2006.

———. Greek and Roman Artillery 399 BC–363 AD. New York: Osprey, 2003.

Hibbeler, R. C. Engineering Mechanics: Statics & Dynamics. Upper Saddle River, NJ: Pearson Prentice
Hall, 2007.

———. Mechanics of Materials. Upper Saddle River, NJ: Pearson, 2016.

Kemp, Adam. The Makerspace Workbench: Tools, Technologies, and Techniques for Making. Sebastopol, CA:
Maker Media, 2013.

Kosky, Philip, Robert Balmer, William Keat, and George Wise. Exploring Engineering: An Introduction to
Engineering and Design. New York: Elsevier, 2013.

Landels, J. G. Engineering in the Ancient World. New York: Barnes & Noble, 1978.

Law, Brian. Brian Law’s Wooden Clocks. http://www.woodenclocks.co.uk/.

Learn Engineering. “Spur Gear Design.” http://www.learnengineering.org/2013/02/spur-gear-design.html.

Marsden, E. W. Greek and Roman Artillery: Historical Development. Oxford: Oxford University Press, 1969.

———. Greek and Roman Artillery: Technical Treatises. Oxford: Oxford University Press, 1971.

Mindess, Sidney, J. Francis Young, and David Darwin. Concrete. Upper Saddle River, NJ: Pearson Prentice
Hall, 2003.

Oleson, John Peter, ed. The Oxford Handbook of Engineering and Technology in the Classical World. New
York: Oxford University Press, 2008.

Bibliography 317
Roberts, Dustyn. Making Things Move: DIY Mechanisms for Inventors, Hobbyists, and Artists. New York:
McGraw-Hill, 2011.

Ross, Don. Rubber Powered Model Airplanes. Hummelstown, PA: Aeronautical Publishers, 2016.

Simons, Martin. Model Aircraft Aerodynamics. Kent, UK: Nexus, 1999.

Westerfield, Mike. Make: Rockets. Sebastopol, CA: Maker Media, 2014.

White, Frank. Fluid Mechanics. New York: McGraw-Hill, 2011.

318 
APPENDIX A
STRUCTURAL
TESTING MACHINE
LEVER-BASED TENSION-
COMPRESSION TESTING MACHINE
USED IN DIY PROJECT #2

Drawing A.1 » Tension-compression


testing machine, perspective view

APPENDIX A » Structural Testing Machine 319


G

L
A E
F E
J
I
K
I
B J
H C
B

D
Drawing A.2 » Tension-
compression testing
machine, exploded view
H

MATERIALS LIST
Wooden Components
NAME DESCRIPTION
A Lever arm 603/4″ × 21/2″ × 3/4″ poplar
B Fulcrum (2) 201/4″ × 21/2″ × 3/4″ poplar
C Lower specimen mount 45″ × 21/2″ × 3/4″ poplar
D Base 45″ × 71/2″ × 3/4″ poplar or plywood
E Lateral arm support (2) 201/4″ × 11/2″ × 3/4″ poplar
F Spacer 11/2″ × 21/2″ × 3/4″ poplar
G Spacer 11/2″ × 11/2″ × 3/4″ poplar
H Brace (2) 8″ × 21/2″ × 3/4″ poplar
I Brace (2) 5″ × 21/2″ × 3/4″ poplar
J Bearing plate (2) 21/2″ × 21/2″ × 3/4″ poplar

320 
Hardware
NAME DESCRIPTION
K Pivot 1/2″ hex bolt
L Safety pin 1/4″ hex bolt

The use of this machine to test cardboard specimens in both tension and compression is demonstrated in
lesson 3.

APPENDIX A » Structural Testing Machine 321


APPENDIX B
CUTTING
PATTERNS D
AND LAYOUTS
J I
K
PROJECT #1 This symbol, when shown,
indicates the preferred
grain direction.

F
G

B C

0 1" 2" 3" 4"


PROJECT #2

0 1" 2" 3" 4"

DIY Project #2 - Cardboard Tower


Note: 32 required Connecting Plate Cutting Pattern
0 1" 2" 3" 4"

Note: 12 required DIY Project #2 - Cardboard Tower


Gusset Plate A Template
0 1" 2" 3" 4"

Note: 12 required DIY Project #2 - Cardboard Tower


Gusset Plate B Cutting Pattern
0 1" 2" 3" 4"

Note: 36 required
DIY Project #2 - Cardboard Tower
Gusset Plate C Cutting Pattern
A C C C B

A C C C B

A C C C B

A C C C B

0 1" 2" 3" 4"


DIY Project #2 - Cardboard Tower
Layout Drawing
PROJECT #5

0 1" 2" 3" 4"

DIY Project #5 - Concrete Sailboat


Cutting Pattern 1 of 6
L

0 1" 2" 3" 4"

DIY Project #5 - Concrete Sailboat


Cutting Pattern 2 of 6
H
12
⁄ " diameter hole
Bevel edge to this
line on front side

0 1" 2" 3" 4"

DIY Project #5 - Concrete Sailboat


Cutting Pattern 3 of 6
0
1"
2"
3"

E
4"

Bevel top edge to match formers C and D

14
⁄ " diameter
hole
F
R
Q

DIY Project #5 - Concrete Sailboat


Cutting Pattern 4 of 6
O
1 32
⁄ " gap
P
rudder

S
14
⁄ " diameter R
hole Q
A
deck

0 1" 2" 3" 4"

DIY Project #5 - Concrete Sailboat


Cutting Pattern 5 of 6
N

0 1" 2" 3" 4"


DIY Project #5 - Concrete Sailboat Cutting Pattern 6 of 6
PROJECT #6

K
J
J
B C C L
D D D D

A' I G G
H
A

I I I I

A A'

Left Side Elevation Right Side Elevation

0 1" 2" 3" 4" DIY Project #6 - Radio-controlled Blimp


Layout Drawing & Cutting Patterns
PROJECT #7

0 1" 2" 3" 4"

S5

F11 F11 W1
W1
S4
S4 W2
F8
F10 F10
W2 F9 F8
F4 F9

F5 S3
F7 F5 S3
S2
grain
S1 direction
F4
3 32
⁄ " balsa sheet
R3 W3
S6
R3
grain
1 16
⁄ " balsa sheet aluminum sheet
direction

F6
L1 L1
F1
F6

W5
W4 1 32
⁄ " plywood

R1 template N5
N3 N2
R1 template

N4 N1
18
⁄ " balsa sheet

1 16
⁄ " plywood
DIY Project #7 - Model Airplane Cutting Pattern
⁄ " square basswood leading edge
3 32

W2
⁄ " balsa
3 32

gusset (typ.) W4
R1 R1 R1 R1 R1 R1 R1 R2 W3
W4
⁄ " × 5⁄16"
1 32 W3
balsa spar
Cover top of center panel
W1 reinforcement W5
with 1⁄32" balsa sheet
(typ.)
after wing is assembled.
WING CENTER PANEL DOUBLERS
W5
3 32⁄ " × 1⁄2" balsa trailing edge
F4
LEFT WING PLAN VIEW WING CENTER PANEL PLAN VIEW
⁄ " square basswood leading edge
3 32

W2
R1

R2 R1 R1 R1 R1 R1 R1 R1 R2
⁄ " music wire
18

R3

W1 WING RIBS
LANDING GEAR

⁄ " music wire


3 64

⁄ " × 1⁄2" balsa trailing edge


3 32

PROPELLER SHAFT
RIGHT WING PLAN VIEW
⁄ " square bass
3 32

Fuselage framework is 3⁄32" square balsa,


except as indicated. F8 F9
N4
N3 Top ⁄ " square bass
3 32
N4 N2
N5 N1 N5
⁄ " square bass
3 32

F5
F10 ⁄ " balsa
3 32
F4
gusset (typ.)

Rear Front Side F1


F2 F3 ⁄ " square bass
3 32

NOSE BLOCK FUSELAGE SIDE ELEVATION

S5 ⁄ " dia. dowel


1 16 ⁄ " dia. dowel
18 ⁄ " dia. dowel
18

⁄ " square balsa


14 ⁄ " aluminum tubing
14

F6 rubber peg

F11 F7
S6

FUSELAGE PLAN VIEW ⁄ " × 1⁄4" balsa (typ.)


3 32
⁄ " × 3⁄16" balsa (typ.)
3 32

⁄ " square balsa (typ.)


3 32

⁄ " square balsa (typ.)


3 32 S1

S3 S3
L1
F2 F3
S2
⁄ " dia. dowel
14 S4 S4
wheel hub F1
(with 1⁄16" hole)

HORIZONTAL STABILIZER PLAN VIEW


VERTICAL STABILIZER PLAN VIEW FUSELAGE FRAMES
0 1" 2" 3" 4"

DIY Project #7 - Rubber-Powered Model Airplane


Layout Drawing
PROJECT #8

0 1" 2" 3" 4"

copper wire
⁄ " square brass washer
5 16 Plan View

main rotor shaft


(3⁄64" music wire)
18" × 3⁄16" balsa (tapered)

front
⁄ " music wire nose
1 32

rear

landing-gear legs
stabilizer bar (1⁄32" music wire)
(1⁄32" music wire)
18" × 3⁄16" balsa

tail rotor hub


(3⁄16" balsa)

gussets
1⁄32" balsa)

tail-rotor blades
(1⁄32" balsa) ⁄ " music wire
1 32
⁄ " square bamboo skids
1 16
hook

Elevation View
main rotor hub
main-rotor blades ( ⁄ " balsa)
1 32
(1⁄8" balsa) pulleys (3⁄32" balsa) Project 8 - Rubber-Powered Helicopter
Cutting Patterns Full-Size Layout Drawing and Cutting Patterns
PROJECT #9

grain direction grain direction

F F
F F
grain direction grain direction

O
N

Q
0 1" 2" 3" 4"

Fin marking template DIY Project #9 - Model Rocket


Cutting Patterns & Fin Marking Template
0 1" 2" 3" 4"

center of parachute

e
l lin
a dia
chr
ea
long
arda
tw
" ou
12
ure
as
me

DIY Project #9 - Model Rocket


Parachute Template
PROJECT #11

K R
g j
K
R

p l
D D n
O k k G
p
G
o

D D e
B
e J3

s
r
B J3
r

A r
C
r

r
C S
A r T

Q Q T

L t

t
L
DIY Engineering Project #11 - Ballista
0 1" 2" 3" 4"

Cutting Patterns
PROJECT #12

O O

3/16” dia. 3/16” dia.

1/2” dia.

I
B

Z
S
3/16” dia.
H H D D

1/8”
dia.
Y J J
1/8”
dia.

1/2” dia. 3/16” dia.


V
C
N
3/16” dia.
V

M
P
3/16” dia. T
1/4”
1/4” dia. dia.

Q 3/16” dia. 3/16” dia.

3/16” dia.
T E

3/16” dia. 0 1" 2" 3" 4"

U 1/4” dia.

K R
DIY Engineerign Project #12 - Hydraulic Arm
U F
3/16” dia. NOTE: All holes are 5/32” diameter unless indicated otherwise. Cutting Patterns
PROJECT #13

0 1" 2" 3" 4"

1 1/8” dia.
7/8” dia.
C

NOTE: This hole is only required for Project #14

NOTE: This hole is only required for Project #14 1 1/8” dia.

C
7/8” dia.

DIY Engineering Project #13 - Water Turbine


Cutting Patterns (1 of 3)
1 1/4” dia.

0 1" 2" 3" 4" DIY Engineering Project #13 - Water Turbine
Cutting Patterns (2 of 3)
a
1” dia.

3/8” dia.

3/8” dia.

0 1" 2" 3" 4" DIY Engineering Project #13 - Water Turbine
Cutting Patterns (3 of 3)
PROJECT #16

0 1" 2" 3" 4"

1 1/2" dia.
M
1/2" dia.

1/8" dia.(typ.)
C

1/4" dia.

DIY Project #16 - Electric Motor-Driven Crane

Cutting Patterns

Вам также может понравиться