13.11.2015 Views

Dynamics cheat sheet

my dynamics notes - 12000.org

my dynamics notes - 12000.org

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Dynamics</strong> <strong>cheat</strong> <strong>sheet</strong><br />

Nasser M. Abbasi<br />

September 27, 2015<br />

page compiled on September 27, 2015 at 5:28pm<br />

Contents<br />

1 Modal analysis for two degrees of freedom system 3<br />

1.1 Step one, finding the equations of motion in normal coordinates space . . . . . . . . . . . . . . . 3<br />

1.2 Step 2, solving the eigenvalue problem, finding the natural frequencies . . . . . . . . . . . . . . . 4<br />

1.3 Step 3, finding the non-normalized eigenvectors . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5<br />

1.4 Step 4. Mass normalization of the shape vectors (or the eigenvectors) . . . . . . . . . . . . . . . . 6<br />

1.5 Step 5, obtain the modal transformation matrix Φ . . . . . . . . . . . . . . . . . . . . . . . . . . 7<br />

1.6 Step 6, applying modal transformation to decouple the original equations of motion . . . . . . . . 8<br />

1.7 Step 7. Converting modal solution to normal coordinates solution . . . . . . . . . . . . . . . . . . 9<br />

1.8 Numerical solution using modal analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9<br />

2 Fourier series representation of a periodic function 14<br />

3 Transfer functions for different vibration systems 16<br />

3.1 Force transmissibility . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

3.2 vibration isolation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16<br />

3.3 accelerometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

3.4 seismometer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18<br />

3.5 Summary of vibration transfer functions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

4 Steady state solutions to single degree freedom system 20<br />

4.1 Summary: . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

4.2 Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20<br />

4.3 Undamped case . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

4.4 damped cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21<br />

5 Summary table of solutions 22<br />

5.1 Harmonic loading mu ′′ + cu ′ + ku = F sin ϖt . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22<br />

5.2 constant loading mu ′′ + cu ′ + ku = F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24<br />

5.3 no loading (free) mu ′′ + cu ′ + ku = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25<br />

5.4 impulse F 0 δ (t) loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26<br />

5.5 Impulse sin function . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27<br />

6 Tree view look at the different cases 28<br />

7 Some common formulas 31<br />

8 Derivation of the solutions of the equations of motion 33<br />

8.1 Free vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

8.1.1 Undamped free vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

1


8.1.2 under-damped free vibration c < c r , ξ < 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

8.1.3 critically damped free vibration ξ = c<br />

c r<br />

= 1 . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

8.1.4 over-damped free vibration ξ = c<br />

c r<br />

> 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

8.2 Constant input F . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

8.2.1 Undamped case ζ = 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34<br />

8.2.2 under-damped ζ < 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

8.2.3 Critical damping ζ = 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

8.2.4 Over-damped ζ > 0 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35<br />

8.3 Harmonic forced input F sin (ϖt) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

8.3.1 Undamped forced vibration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

8.3.2 Under-damped forced vibration c < c r , ξ < 1 . . . . . . . . . . . . . . . . . . . . . . . . . 38<br />

8.3.3 critically damping forced vibration ξ = c<br />

c r<br />

= 1 . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

8.3.4 over-damped forced vibration ξ = c<br />

c r<br />

> 1 . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40<br />

8.4 Response to impulsive loading . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

8.4.1 impulse input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41<br />

8.4.2 sin impulse input . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

8.5 references . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49<br />

9 Derivation of rotation formula 49<br />

10 Miscellaneous hints 50<br />

11 Formulas 51<br />

12 Velocity and acceleration diagrams 53<br />

12.1 Spring pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53<br />

12.2 pendulum with blob moving in slot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54<br />

12.3 spring pendulum with block moving in slot . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55<br />

12.4 double pendulum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

13 Velocity and acceleration of rigid body 2D 57<br />

14 Velocity and acceleration of rigid body 3D 58<br />

14.1 Using Vehicle dynamics notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

14.2 3D Not Using vehicle dynamics notations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

14.2.1 Derivation for F = ma in 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

14.2.2 Derivation for τ = Iω in 3D . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61<br />

14.2.3 Derivation for τ = Iω in 3D using principle axes . . . . . . . . . . . . . . . . . . . . . . . 62<br />

15 Wheel spinning precession 63<br />

16 References 63<br />

17 Misc. items 64<br />

18 Astrodynamics 65<br />

18.1 Ellipse main parameters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

18.2 Table of common equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65<br />

18.3 Flight path angle for ellipse γ . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

18.4 Parabolic trajectory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

18.5 Hyperbolic trajectory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70<br />

18.6 showing that energy is constant . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

18.7 Earth satellite Transfer orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

18.7.1 Hohmann transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72<br />

2


18.7.2 Bi-Elliptic transfer orbit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74<br />

18.7.3 semi-tangential elliptical transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75<br />

18.8 Rocket engines, Hohmann transfer, plane change at equator . . . . . . . . . . . . . . . . . . . . . 75<br />

18.8.1 Rocket equation with plane change not at equator . . . . . . . . . . . . . . . . . . . . . . 77<br />

18.9 Spherical coordinates . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78<br />

18.10interplanetary transfer orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

18.10.1 interplanetary hohmann transfer orbit, case one . . . . . . . . . . . . . . . . . . . . . . . . 79<br />

18.11rendezvous orbits . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81<br />

18.11.1 Two satellite, walking rendezvous using Hohmann transfer . . . . . . . . . . . . . . . . . . 81<br />

18.11.2 Two satellite, separate orbits, rendezvous using Hohmann transfer, coplaner . . . . . . . . 83<br />

18.11.3 Two satellite, separate orbits, rendezvous using bi-elliptic transfer, coplaner . . . . . . . . 85<br />

18.12Semi-tangential transfers, elliptical, parabolic and hyperbolic . . . . . . . . . . . . . . . . . . . . 88<br />

18.13Lagrange points . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90<br />

18.14Orbit changing by low contiuous thrust . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

18.15References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91<br />

19 Dynamic of flights 92<br />

19.1 Wing geometry . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 92<br />

19.2 Summary of main equations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94<br />

19.2.1 Writing the equations in linear form . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96<br />

19.3 definitions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98<br />

19.4 images and plots collected . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100<br />

19.5 Some strange shaped airplanes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

19.6 links . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

19.7 references . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 119<br />

1 Modal analysis for two degrees of freedom system<br />

Detailed steps to perform modal analysis are given below for a standard undamped two degrees of freedom<br />

system. The main advantage of solving a multidegree system using modal analysis is that it decouples the<br />

equations of motion (assuming they are coupled) making solving them much simpler.<br />

In addition it shows the fundamental shapes that the system can vibrate in, which gives more insight into<br />

the system. Starting with standard 2 degrees of freedom system<br />

x 1<br />

x 2<br />

k 1 k 2<br />

m 1 m 2<br />

f 1 t<br />

f 2 t<br />

In the above the generalized coordinates are x 1 and x 2 . Hence the system requires two equations of motion<br />

(EOM’s).<br />

1.1 Step one, finding the equations of motion in normal coordinates space<br />

The two EOM’s are found using any method such as Newton’s method or Lagrangian method. Using Newton’s<br />

method, free body diagram is made of each mass and then F = ma is written for each mass resulting in the<br />

equations of motion. In the following it is assumed that both masses are moving in the positive direction and<br />

that x 2 is larger than x 1 when these equations of equilibrium are written<br />

3


x 1<br />

x 2<br />

k 1 x 1<br />

k 2 x 2 x 1 <br />

m<br />

k 2 x 2 x 1 <br />

m 2<br />

1<br />

f 1 t<br />

f 2 t<br />

F m 1 x 1<br />

<br />

k 1 x 1 k 2 x 2 x 1 f 1 t m 1 x 1<br />

<br />

F m 2 x 2<br />

<br />

k 2 x 2 x 1 f 2 t m 2 x 2<br />

<br />

Hence, from the above the equations of motion are<br />

or<br />

In Matrix form<br />

⎡ ⎤ ⎧<br />

⎣ m 1 0 ⎨<br />

⎦<br />

0 m<br />

⎩<br />

2<br />

m 1 x ′′<br />

1 + k 1 x 1 − k 2 (x 2 − x 1 ) = f 1 (t)<br />

m 2 x ′′<br />

2 + k 2 (x 2 − x 1 ) = f 2 (t)<br />

m 1 x ′′<br />

1 + x 1 (k 1 + k 2 ) − k 2 x 2 = f 1 (t)<br />

m 2 x ′′<br />

2 + k 2 x 2 − k 2 x 1 = f 2 (t)<br />

x ′′<br />

1<br />

x ′′<br />

2<br />

⎫ ⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎬<br />

⎭ + ⎣ k 1 + k 2 −k 2<br />

⎨x ⎦ 1 ⎬ ⎨<br />

−k 2 k<br />

⎩<br />

2 x<br />

⎭ = f 1 (t) ⎬<br />

⎩<br />

2 f 2 (t)<br />

⎭<br />

The above two EOM are coupled in stiffness, but not mass coupled. Using short notations, the above is<br />

written as<br />

[M]{x ′′ } + [K]{x} = {f}<br />

Modal analysis now starts with the goal to decouple the EOM and obtain the fundamental shape functions that<br />

the system can vibrate in. To make these derivations more general, the mass matrix and the stiffness matrix are<br />

written in general notations as follows<br />

⎡ ⎤ ⎧<br />

⎣ m 11 m 12<br />

⎨<br />

⎦<br />

m 21 m<br />

⎩<br />

22<br />

x ′′<br />

1<br />

x ′′<br />

2<br />

⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎬<br />

⎭ + ⎣ k 11 k 12<br />

⎨x ⎦ 1 ⎬ ⎨<br />

k<br />

⎩<br />

22 x<br />

⎭ = f 1 (t) ⎬<br />

⎩<br />

2 f 2 (t)<br />

⎭<br />

k 21<br />

The mass matrix [M] and the stiffness matrix [K] must always come out to be symmetric. If they are not<br />

symmetric, then a mistake was made in obtaining them. As a general rule, the mass matrix [M] is PSD (positive<br />

definite matrix) and the [K] matrix is positive semi-definite matrix. The reason the [M] is PSD is that x T [M]{x}<br />

represents the kinetic energy of the system, which is typically positive and not zero. But reading some other<br />

references 1 it is possible that [M] can be positive semi-definite. It depends on the application being modeled.<br />

1.2 Step 2, solving the eigenvalue problem, finding the natural frequencies<br />

The first step in modal analysis is to solve the eigenvalue problem det ( [K] − ω 2 [M] ) = 0 in order to determine<br />

the natural frequencies of the system. This equations leads to a polynomial in ω and the roots of this polynomial<br />

1 http://en.wikipedia.org/wiki/Fundamental_equation_of_constrained_motion<br />

4


are the natural frequencies of the system. Since there are two degrees of freedom, there will be two natural<br />

frequencies ω 1 , ω 2 for the system.<br />

det ( [K] − ω 2 [M] ) = 0<br />

⎛⎡<br />

⎤ ⎡ ⎤⎞<br />

det ⎝⎣ k 11 k 12<br />

⎦ − ω 2 ⎣ m 11 m 12<br />

⎦⎠ = 0<br />

k 21 k 22 m 21 m 22<br />

⎡<br />

⎤<br />

det ⎣ k 11 − ω 2 m 11 k 12 − ω 2 m 12<br />

⎦ = 0<br />

k 21 − ω 2 m 21 k 22 − ω 2 m 22<br />

(<br />

k11 − ω 2 ) (<br />

m 11 k22 − ω 2 ) (<br />

m 22 − k12 − ω 2 ) (<br />

m 12 k21 − ω 2 )<br />

m 21 = 0<br />

ω 4 (m 11 m 22 − m 12 m 21 ) + ω 2 (−k 11 m 22 + k 12 m 21 + k 21 m 12 − k 22 m 11 ) + k 11 k 22 − k 12 k 21 = 0<br />

The above is a polynomial in ω 4 . Let ω 2 = λ it becomes<br />

λ 2 (m 11 m 22 − m 12 m 21 ) + λ (−k 11 m 22 + k 12 m 21 + k 21 m 12 − k 22 m 11 ) + k 11 k 22 − k 12 k 21 = 0<br />

This quadratic polynomial in λ which is now solved using the quadratic formula. Then the positive square<br />

root of each λ root to obtain ω 1 and ω 2 which are the roots of the original eigenvalue problem. Assuming from<br />

now that these roots are ω 1 and ω 2 the next step is to obtain the non-normalized shape vectors ϕ 1 , ϕ 2 also<br />

called the eigenvectors associated with ω 1 and ω 2<br />

1.3 Step 3, finding the non-normalized eigenvectors<br />

For each natural frequency ω 1 and ω 2 the corresponding shape function is found by solving the following two<br />

sets of equations for the vectors ϕ 1 , ϕ 2<br />

⎡<br />

⎡<br />

⎧<br />

⎫<br />

⎣ k 11 k 12<br />

k 21<br />

⎤<br />

⎦ − ω1<br />

2<br />

k 22<br />

⎣ m 11 m 12<br />

m 21<br />

⎤ ⎫ ⎧<br />

⎨ϕ ⎦ 11 ⎬ ⎨<br />

m<br />

⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

and<br />

⎡ ⎤<br />

⎣ k 11 k 12<br />

⎦ − ω2<br />

2<br />

k 21 k 22<br />

⎡<br />

⎣ m 11 m 12<br />

m 21<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎨ϕ ⎦ 12 ⎬ ⎨<br />

m<br />

⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

22 0<br />

⎭<br />

For ω 1 , let ϕ 11 = 1 and solve for<br />

⎡ ⎤ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ k 11 k 12<br />

⎦ − ω1<br />

2 ⎣ m 11 m 12<br />

⎨ 1 ⎬ ⎨<br />

⎦<br />

k 21 k 22 m 21 m<br />

⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ k 11 − ω1 2m 11 k 12 − ω1 2m ⎨<br />

12 1 ⎬ ⎨<br />

⎦<br />

k 21 − ω1 2m 21 k 22 − ω1 2m ⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

Which gives one equation now to solve for ϕ 21 (the first row equation is only used)<br />

(<br />

k11 − ω 2 1m 11<br />

)<br />

+ ϕ21<br />

(<br />

k12 − ω 2 1m 12<br />

)<br />

= 0<br />

Hence<br />

ϕ 21 = − ( k 11 − ω1 2m )<br />

11<br />

(<br />

k12 − ω1 2m 12)<br />

5


Therefore the first shape vector is<br />

⎧ ⎫ ⎧<br />

⎨ϕ 11 ⎬ ⎪⎨<br />

ϕ 1 =<br />

⎩<br />

ϕ<br />

⎭ = 1<br />

⎪<br />

21<br />

⎩<br />

− ( k 11 −ω1 2m )<br />

( 11<br />

k12 −ω1 2m 12)<br />

⎫<br />

⎪⎬<br />

⎪⎭<br />

Similarly the second shape function is obtained. For ω 2 , let ϕ 12 = 1 and solve for<br />

⎡ ⎤ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ k 11 k 12<br />

⎦ − ω2<br />

2 ⎣ m 11 m 12<br />

⎨ 1 ⎬ ⎨<br />

⎦<br />

k 21 k 22 m 21 m<br />

⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

22 0<br />

⎭<br />

⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ k 11 − ω2 2m 11 k 12 − ω2 2m ⎨<br />

12 1 ⎬ ⎨<br />

⎦<br />

k 21 − ω2 2m 21 k 22 − ω2 2m ⎩<br />

22 ϕ<br />

⎭ = 0⎬<br />

⎩<br />

22 0<br />

⎭<br />

Which gives one equation now to solve for ϕ 22 (the first row equation is only used)<br />

(<br />

k11 − ω 2 2m 11<br />

)<br />

+ ϕ22<br />

(<br />

k12 − ω 2 2m 12<br />

)<br />

= 0<br />

Hence<br />

ϕ 22 = − ( k 11 − ω2 2m )<br />

11<br />

(<br />

k12 − ω2 2m 12)<br />

Therefore the second shape vector is<br />

⎧ ⎫ ⎧<br />

⎨ϕ 12 ⎬ ⎪⎨<br />

ϕ 2 =<br />

⎩<br />

ϕ<br />

⎭ = 1<br />

⎪<br />

22<br />

⎩<br />

− ( k 11 −ω2 2m )<br />

( 11<br />

k12 −ω2 2m 12)<br />

Now that the two non-normalized shape vectors are found, the next step is to perform mass normalization<br />

1.4 Step 4. Mass normalization of the shape vectors (or the eigenvectors)<br />

Let<br />

µ 1 = ϕ T 1 [M] ϕ 1<br />

This results in a scalar value µ 1 , which is later used to normalize ϕ 1 . Similarly<br />

µ 2 = ϕ T 2 [M] ϕ 2<br />

⎫<br />

⎪⎬<br />

⎪⎭<br />

For example, to find µ 1<br />

⎧ ⎫T ⎡ ⎤ ⎧ ⎫<br />

⎨ϕ 11 ⎬<br />

µ 1 = ⎣ m 11 m 12<br />

⎨ϕ ⎦ 11 ⎬<br />

⎩<br />

ϕ<br />

⎭<br />

21 m 21 m<br />

⎩<br />

22 ϕ<br />

⎭<br />

21<br />

{ } ⎡ ⎤ ⎧ ⎫<br />

= ϕ 11 ϕ ⎣ m 11 m 12<br />

⎨ϕ ⎦ 11 ⎬<br />

21<br />

m<br />

⎩<br />

22 ϕ<br />

⎭<br />

21<br />

m 21<br />

}<br />

=<br />

{ϕ ⎧ ⎫<br />

⎨ϕ 11 ⎬<br />

11 m 11 + ϕ 21 m 21 ϕ 11 m 12 + ϕ 21 m 22<br />

⎩<br />

ϕ<br />

⎭<br />

21<br />

= ϕ 11 (ϕ 11 m 11 + ϕ 21 m 21 ) + ϕ 21 (ϕ 11 m 12 + ϕ 21 m 22 )<br />

6


Similarly, µ 2 is found<br />

⎧ ⎫T ⎡ ⎤ ⎧ ⎫<br />

⎨ϕ 12 ⎬<br />

µ 2 = ⎣ m 11 m 12<br />

⎨ϕ ⎦ 12 ⎬<br />

⎩<br />

ϕ<br />

⎭<br />

22 m 21 m<br />

⎩<br />

22 ϕ<br />

⎭<br />

22<br />

{ } ⎡ ⎤ ⎧ ⎫<br />

= ϕ 12 ϕ ⎣ m 11 m 12<br />

⎨ϕ ⎦ 12 ⎬<br />

22<br />

m<br />

⎩<br />

22 ϕ<br />

⎭<br />

22<br />

m 21<br />

}<br />

=<br />

{ϕ ⎧ ⎫<br />

⎨ϕ 12 ⎬<br />

12 m 11 + ϕ 22 m 21 ϕ 12 m 12 + ϕ 22 m 22<br />

⎩<br />

ϕ<br />

⎭<br />

22<br />

= ϕ 12 (ϕ 12 m 11 + ϕ 22 m 21 ) + ϕ 22 (ϕ 12 m 12 + ϕ 22 m 22 )<br />

Now that µ 1 , µ 2 are obtained, the mass normalized shape vectors are found. They are called Φ 1 , Φ 2<br />

⎧ ⎫<br />

⎨ϕ 11 ⎬<br />

⎧ ⎫<br />

Φ 1 = ϕ ⎩<br />

ϕ<br />

⎭<br />

1 21<br />

⎨ √µ1<br />

ϕ 11 ⎬<br />

√ = √ =<br />

µ1 µ1 ⎩ √µ1<br />

ϕ 21 ⎭<br />

Similarly<br />

Φ 2 = ϕ 2<br />

√<br />

µ2<br />

=<br />

⎧<br />

⎨<br />

⎩<br />

ϕ 12<br />

ϕ 22<br />

⎫<br />

⎬<br />

⎭<br />

√<br />

µ2<br />

=<br />

⎧<br />

⎨<br />

õ2<br />

ϕ 12<br />

⎫<br />

⎬<br />

⎩ √µ2<br />

ϕ 22 ⎭<br />

1.5 Step 5, obtain the modal transformation matrix Φ<br />

The modal transformation matrix is the 2 × 2 matrix made of of Φ 1 , Φ 2 in each of its columns<br />

[Φ] = [Φ 1 Φ 2 ]<br />

⎡<br />

= ⎣<br />

⎤<br />

õ1<br />

ϕ 11 √µ2<br />

ϕ 12<br />

⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

Now the [Φ] is found, the transformation from the normal coordinates {x} to modal coordinates, which is<br />

called {η} is found<br />

{x} = [Φ] {η}<br />

⎧ ⎫ ⎡<br />

⎨x 1 (t) ⎬ ϕ 11<br />

⎩<br />

x 2 (t)<br />

⎭ = ⎣<br />

⎤<br />

õ1 õ2<br />

ϕ 12<br />

⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

⎧ ⎫<br />

⎨η 1 (t) ⎬<br />

⎩<br />

η 2 (t)<br />

⎭<br />

The transformation from modal coordinates back to normal coordinates is<br />

{η} = [Φ] −1 {x}<br />

⎧ ⎫ ⎡<br />

⎨η 1 (t) ⎬ ϕ 11<br />

⎩<br />

η 2 (t)<br />

⎭ = ⎣<br />

⎤<br />

õ1 õ2<br />

ϕ 12<br />

⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

−1 ⎧ ⎫<br />

⎨ x 1 (t) ⎬<br />

⎩<br />

x 2 (t)<br />

⎭<br />

7


However, [Φ] −1 = [Φ] T [M] therefore<br />

{η} = [Φ] T [M] {x}<br />

⎧ ⎫ ⎡ ⎤<br />

⎨η 1 (t) ⎬ √µ1<br />

ϕ 11 √µ2<br />

ϕ 12<br />

⎩<br />

η 2 (t)<br />

⎭ = ⎣ ⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

⎤ ⎧ ⎫<br />

⎣ m 11 m 12<br />

⎨x ⎦ 1 (t) ⎬<br />

m 21 m<br />

⎩<br />

22 x 2 (t)<br />

⎭<br />

T ⎡<br />

The next step is to apply this transformation to the original equations of motion in order to decouple them<br />

1.6 Step 6, applying modal transformation to decouple the original equations of motion<br />

The EOM in normal coordinates is<br />

⎡<br />

⎣ m 11 m 12<br />

m 21<br />

⎤ ⎧<br />

⎨<br />

⎦<br />

m<br />

⎩<br />

22<br />

x ′′<br />

1<br />

x ′′<br />

2<br />

⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎬<br />

⎭ + ⎣ k 11 k 12<br />

⎨x ⎦ 1 ⎬ ⎨<br />

k<br />

⎩<br />

22 x<br />

⎭ = f 1 (t) ⎬<br />

⎩<br />

2 f 2 (t)<br />

⎭<br />

Applying the above modal transformation {x} = [Φ] {η} on the above results in<br />

⎡ ⎤ ⎧ ⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ m 11 m 12<br />

⎨η 1<br />

⎦ ′′ ⎬<br />

[Φ]<br />

m 21 m<br />

⎩<br />

22 η ′′ ⎭ + ⎣ k 11 k 12<br />

⎨η ⎦ 1 ⎬ ⎨<br />

[Φ]<br />

k 21 k<br />

⎩<br />

22 η<br />

⎭ = f 1 (t) ⎬<br />

⎩<br />

2 f 2 (t)<br />

⎭<br />

pre-multiplying by [Φ] T results in<br />

⎡ ⎤ ⎧<br />

[Φ] T ⎣ m 11 m 12<br />

⎨<br />

⎦ [Φ]<br />

m<br />

⎩<br />

22<br />

m 21<br />

η 1<br />

′′<br />

η 2<br />

′′<br />

2<br />

k 21<br />

⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎬<br />

⎭ + [Φ]T ⎣ k 11 k 12<br />

⎨η ⎦ 1 ⎬ ⎨<br />

[Φ]<br />

k<br />

⎩<br />

22 η<br />

⎭ = f 1 (t) ⎬<br />

[Φ]T ⎩<br />

2 f 2 (t)<br />

⎭<br />

⎡ ⎤<br />

⎡ ⎤<br />

The result of [Φ] T ⎣ m 11 m 12<br />

⎦ [Φ] will always be ⎣ 1 0 ⎦. This is because mass normalized shape vectors<br />

m 21 m 22 0 1<br />

are used. If the shape functions were not mass normalized, then the diagonal values will not be 1 as shown.<br />

⎡ ⎤<br />

⎡ ⎤<br />

The result of [Φ] T ⎣ k 11 k 12<br />

⎦ [Φ] will be ⎣ ω2 1 0<br />

⎦.<br />

k 21 k 22 0 ω2<br />

2<br />

⎧ ⎫ ⎧ ⎫<br />

⎨<br />

Let the result of [Φ] T f 1 (t) ⎬ ⎨ ˜f<br />

⎩<br />

f 2 (t)<br />

⎭ be 1 (t) ⎬<br />

,Therefore, in modal coordinates the original EOM becomes<br />

⎩ ˜f 2 (t)<br />

⎭<br />

⎡ ⎤ ⎧<br />

⎣ 1 0 ⎨<br />

⎦<br />

0 1<br />

⎩<br />

η 1<br />

′′<br />

η 2<br />

′′<br />

k 21<br />

⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎬<br />

⎭ + ⎣ ω2 1 0 ⎨η ⎦ 1 ⎬ ⎨ ˜f<br />

0 ω2<br />

2 ⎩<br />

η<br />

⎭ = 1 (t) ⎬<br />

⎩<br />

2<br />

˜f 2 (t)<br />

⎭<br />

The EOM are now decouples and each can be solved as follows<br />

η ′′<br />

1 (t) + ω 2 1η 1 (t) = ˜f 1 (t)<br />

η ′′<br />

2 (t) + ω 2 2η 2 (t) = ˜f 2 (t)<br />

To solve these EOM’s, the initial conditions in normal coordinates must be transformed to modal coordinates<br />

using the above transformation rules<br />

{η (0)} = [Φ] T [M] {x (0)}<br />

{<br />

η ′ (0) } = [Φ] T [M] { x ′ (0) }<br />

8


Or in full form<br />

and<br />

⎧ ⎫ ⎡<br />

⎨η 1 (0) ⎬<br />

⎩<br />

η 2 (0)<br />

⎭ = ⎣<br />

⎧ ⎫ ⎡<br />

⎨η 1 ′ (0) ⎬<br />

⎩<br />

η 2 ′ (0) ⎭ = ⎣<br />

⎤T ⎡<br />

õ1<br />

ϕ 11 √µ2<br />

ϕ 12<br />

⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

⎤T ⎡<br />

õ1<br />

ϕ 11 √µ2<br />

ϕ 12<br />

⎦<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22<br />

⎣ m 11 m 12<br />

m 21<br />

⎣ m 11 m 12<br />

m 21<br />

⎤ ⎧ ⎫<br />

⎨x ⎦ 1 (0) ⎬<br />

m<br />

⎩<br />

22 x 2 (0)<br />

⎭<br />

⎤ ⎧ ⎫<br />

⎨x ⎦<br />

′ 1 (0) ⎬<br />

m<br />

⎩<br />

22 x ′ 2 (0) ⎭<br />

Each of these EOM are solved using any of the standard methods. This will result is solutions η 1 (t) and<br />

η 2 (t)<br />

1.7 Step 7. Converting modal solution to normal coordinates solution<br />

The solutions found above are in modal coordinates η 1 (t) , η 2 (t). The solution needed is x 1 (t) , x 2 (t). Therefore,<br />

the transformation {x} = [Φ] {η} is now applied to convert the solution to normal coordinates<br />

⎧ ⎫ ⎡ ⎤ ⎧ ⎫<br />

⎨x 1 (t) ⎬ √µ1<br />

ϕ 11 √µ2<br />

ϕ 12 ⎨<br />

⎩<br />

x 2 (t)<br />

⎭ = η<br />

⎣ ⎦ 1 (t) ⎬<br />

õ1<br />

ϕ 21 √µ2<br />

ϕ 22 ⎩<br />

η 2 (t)<br />

⎭<br />

⎧<br />

⎫<br />

⎨ √µ1<br />

ϕ 11<br />

η 1 (t) + √ ϕ 12<br />

µ2<br />

η 2 (t) ⎬<br />

=<br />

⎩ √µ1<br />

ϕ 21<br />

η 1 (t) + √ ϕ 22<br />

µ2<br />

η 2 (t)<br />

⎭<br />

Hence<br />

and<br />

x 1 (t) = ϕ 11<br />

√<br />

µ1<br />

η 1 (t) + ϕ 12<br />

√<br />

µ2<br />

η 2 (t)<br />

x 2 (t) = ϕ 21<br />

√<br />

µ1<br />

η 1 (t) + ϕ 22<br />

√<br />

µ2<br />

η 2 (t)<br />

Notice that the solution in normal coordinates is a linear combination of the modal solutions. The terms<br />

ϕ<br />

√ ij<br />

µ<br />

are just scaling factors that represent the contribution of each modal solution to the final solution. This<br />

completes modal analysis<br />

1.8 Numerical solution using modal analysis<br />

This is a numerical example that implements the above steps using a numerical values for [K] and [M].<br />

Let k 1 = 1, k 2 = 2, m 1 = 1, m 2 = 3 and let f 1 (t) = 0 and f 2 (t) = sin (5t). Let initial conditions be<br />

x 1 (0) = 0, x ′ 1 (0) = 1, x 2 (0) = 1.5, x ′ 2 (0) = 3, hence<br />

⎧ ⎫ ⎫<br />

⎨x 1 (0) ⎬ 0⎬<br />

⎧<br />

⎨<br />

⎩<br />

x 2 (0)<br />

⎭ = ⎩<br />

1<br />

⎭<br />

and<br />

⎧ ⎫ ⎧ ⎫<br />

⎨x ′ 1 (0) ⎬ ⎨<br />

⎩<br />

x ′ 2 (0) ⎭ = 1.5⎬<br />

⎩<br />

3<br />

⎭<br />

In normal coordinates, the EOM are<br />

9


⎡ ⎤ ⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ m 1 0 ⎨x ⎦<br />

′′ ⎬<br />

1<br />

0 m<br />

⎩<br />

2 x ′′ ⎭ + ⎣ k 1 + k 2 −k 2<br />

⎨x ⎦ 1 ⎬ ⎨<br />

2 −k 2 k<br />

⎩<br />

2 x<br />

⎭ = f 1 (t) ⎬<br />

⎩<br />

2 f 2 (t)<br />

⎭<br />

⎡ ⎤ ⎧ ⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ 1 0 ⎨x ⎦<br />

′′ ⎬<br />

1<br />

0 3<br />

⎩<br />

x ′′ ⎭ + ⎣ 3 −2 ⎨x ⎦ 1 ⎬ ⎨<br />

−2 2<br />

⎩<br />

x<br />

⎭ = 0 ⎬<br />

⎩<br />

2 sin (5t)<br />

⎭<br />

2<br />

In this example m 11 = 1, m 12 = 0, m 21 = 0, m 22 = 3 and k 11 = 3, k 12 = −2, k 21 = −2, k 22 = 2 and f 1 (t) = 0<br />

and f 2 (t) = sin (5t)<br />

step 2 is now applied which solves the eigenvalue problem in order to find the two natural frequencies<br />

Let ω 2 = λ hence<br />

The solution is λ 1 = 3. 475 and λ 2 = 0.192, therefore<br />

And<br />

det ( [K] − ω 2 [M] ) = 0<br />

⎛⎡<br />

⎤ ⎡ ⎤⎞<br />

det ⎝⎣ 3 −2 ⎦ − ω 2 ⎣ 1 0 ⎦⎠ = 0<br />

−2 2 0 3<br />

⎡<br />

⎤<br />

det ⎣ 3 − ω2 −2<br />

⎦ = 0<br />

−2 2 − 3ω 2<br />

(<br />

3 − ω<br />

2 ) ( 2 − 3ω 2) − (−2) (−2) = 0<br />

3ω 4 − 11ω 2 + 2 = 0<br />

3λ 2 − 11λ + 2 = 0<br />

ω 1 = √ 3.475 = 1.864<br />

ω 2 = √ 0.192 = 0.438<br />

step 3 is now applied which finds the non-normalized eigenvectors. For each natural frequency ω 1 and ω 2 the<br />

corresponding shape function is found by solving the following two sets of equations for the eigen vectors ϕ 1 , ϕ 2<br />

⎛⎡<br />

⎤ ⎡ ⎤⎞<br />

⎧ ⎫ ⎧ ⎫<br />

⎝⎣ 3 −2 ⎦ − ω1<br />

2 ⎣ 1 0 ⎨ϕ ⎦⎠<br />

11 ⎬ ⎨<br />

−2 2 0 3<br />

⎩<br />

ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

For ω 1 = 1. 864<br />

⎛⎡<br />

⎤ ⎡ ⎤⎞<br />

⎧ ⎫ ⎧ ⎫<br />

⎝⎣ 3 −2 ⎦ − 1.864 2 ⎣ 1 0 ⎨ϕ ⎦⎠<br />

11 ⎬ ⎨<br />

−2 2<br />

0 3<br />

⎩<br />

ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ −0.475 −2 ⎨ 1 ⎬ ⎨<br />

⎦<br />

−2 −8.424<br />

⎩<br />

ϕ<br />

⎭ = 0⎬<br />

⎩<br />

21 0<br />

⎭<br />

This gives one equation to solve for ϕ 21 (the first row equation is only used)<br />

−0.475 − 2ϕ 21 = 0<br />

10


Hence<br />

The first eigen vector is<br />

Similarly for ω 2 = 0.438<br />

ϕ 21 = 0.475<br />

−2 = −0.237<br />

⎧ ⎫ ⎧ ⎫<br />

⎨ϕ 11 ⎬ ⎨<br />

ϕ 1 =<br />

⎩<br />

ϕ<br />

⎭ = 1 ⎬<br />

⎩<br />

21 −0.237<br />

⎭<br />

⎛⎡<br />

⎤ ⎡ ⎤⎞<br />

⎧ ⎫ ⎧ ⎫<br />

⎝⎣ 3 −2 ⎦ − 0.438 2 ⎣ 1 0 ⎨ϕ ⎦⎠<br />

12 ⎬ ⎨<br />

−2 2<br />

0 3<br />

⎩<br />

ϕ<br />

⎭ = 0⎬<br />

⎩<br />

22 0<br />

⎭<br />

⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ 2.808 −2 ⎨ 1 ⎬ ⎨<br />

⎦<br />

−2 1.425<br />

⎩<br />

ϕ<br />

⎭ = 0⎬<br />

⎩<br />

22 0<br />

⎭<br />

This gives one equation to solve for ϕ 22 (the first row equation is only used)<br />

2.808 − 2ϕ 22 = 0<br />

Hence<br />

The second eigen vector is<br />

ϕ 22 = −2.808<br />

−2<br />

= 1.404<br />

⎧ ⎫ ⎧ ⎫<br />

⎨ϕ 12 ⎬ ⎨<br />

ϕ 2 =<br />

⎩<br />

ϕ<br />

⎭ = 1 ⎬<br />

⎩<br />

22 1.404<br />

⎭<br />

Now step 4 is applied, which is mass normalization of the shape vectors (or the eigenvectors)<br />

µ 1 = ϕ T 1 [M] ϕ 1<br />

Hence<br />

Similarly, µ 2 is found<br />

Hence<br />

⎧ ⎫<br />

⎨ 1 ⎬<br />

µ 1 =<br />

⎩<br />

−0.237<br />

⎭<br />

= 1. 169<br />

⎧ ⎫<br />

⎨ 1 ⎬<br />

µ 2 =<br />

⎩<br />

1.404<br />

⎭<br />

= 6.914<br />

⎤ ⎧ ⎫<br />

⎣ 1 0 ⎨ 1 ⎬<br />

⎦<br />

0 3<br />

⎩<br />

−0.237<br />

⎭<br />

T ⎡<br />

µ 2 = ϕ T 2 [M] ϕ 2<br />

⎤ ⎧ ⎫<br />

⎣ 1 0 ⎨ 1 ⎬<br />

⎦<br />

0 3<br />

⎩<br />

1.404<br />

⎭<br />

T ⎡<br />

11


Now that µ 1 , µ 2 are found, the mass normalized eigen vectors are found. They are called Φ 1 , Φ 2<br />

⎧ ⎧ ⎫<br />

⎨ ⎨ 1 ⎬<br />

Φ 1 = ϕ 1<br />

√<br />

µ1<br />

=<br />

⎩<br />

ϕ 11<br />

⎧ ⎫<br />

⎨ 0.925 ⎬<br />

=<br />

⎩<br />

−0.219<br />

⎭<br />

ϕ 21<br />

⎫<br />

⎬<br />

⎭<br />

√<br />

µ1<br />

=<br />

⎩<br />

−0.237<br />

⎭<br />

√<br />

1.169<br />

Similarly<br />

Φ 2 = ϕ 2<br />

√<br />

µ2<br />

=<br />

⎧<br />

⎨<br />

⎩<br />

⎧ ⎫<br />

⎨0.380⎬<br />

=<br />

⎩<br />

0.534<br />

⎭<br />

ϕ 12<br />

ϕ 22<br />

⎫<br />

⎬<br />

⎭<br />

√<br />

µ2<br />

=<br />

⎧<br />

⎨<br />

1<br />

⎫<br />

⎬<br />

⎩<br />

1.404<br />

⎭<br />

√<br />

6.914<br />

Therefore, the modal transformation matrix is<br />

[Φ] = [Φ 1 Φ 2 ]<br />

⎡<br />

⎤<br />

0.925<br />

= ⎣<br />

0.380<br />

⎦<br />

−0.219 0.534<br />

This result can be verified using Matlab’s eig function as follows<br />

EDU>> K=[3 -2;-2 2]; M=[1 0;0 3];<br />

EDU>> [phi,lam]=eig(K,M)<br />

phi =<br />

-0.3803 -0.9249<br />

-0.5340 0.2196<br />

EDU>> diag(sqrt(lam))<br />

0.4380<br />

1.8641<br />

Matlab result agrees with the result obtained above. The sign difference is not important. Now step 5 is<br />

applied. Matlab generates mass normalized eigenvectors by default.<br />

Now that [Φ] is found, the transformation from the normal coordinates {x} to modal coordinates, called<br />

{η} , is obtained<br />

{x} = [Φ] {η}<br />

⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫<br />

⎨x 1 (t) ⎬<br />

⎩<br />

x 2 (t)<br />

⎭ = 0.925 0.380 ⎨η ⎣ ⎦ 1 (t) ⎬<br />

−0.219 0.534<br />

⎩<br />

η 2 (t)<br />

⎭<br />

12


The transformation from modal coordinates back to normal coordinates is<br />

However, [Φ] −1 = [Φ] T [M] therefore<br />

{η} = [Φ] −1 {x}<br />

⎧ ⎫ ⎡<br />

⎤−1 ⎧ ⎫<br />

⎨η 1 (t) ⎬<br />

⎩<br />

η 2 (t)<br />

⎭ = 0.925 0.380 ⎨ x<br />

⎣ ⎦ 1 (t) ⎬<br />

−0.219 0.534<br />

⎩<br />

x 2 (t)<br />

⎭<br />

{η} = [Φ] T [M] {x}<br />

⎧ ⎫ ⎡<br />

⎤T ⎡ ⎤ ⎧ ⎫<br />

⎨η 1 (t) ⎬<br />

⎩<br />

η 2 (t)<br />

⎭ = 0.925 0.380<br />

⎣ ⎦ ⎣ 1 0 ⎨x ⎦ 1 (t) ⎬<br />

−0.219 0.534 0 3<br />

⎩<br />

x 2 (t)<br />

⎭<br />

⎡<br />

⎤ ⎧ ⎫<br />

0.925 −0.657 ⎨x = ⎣ ⎦ 1 (t) ⎬<br />

0.38 1.6<br />

⎩<br />

x 2 (t)<br />

⎭<br />

The next step is to apply this transformation to the original equations of motion in order to decouple them.<br />

Applying step 6 results in<br />

⎡ ⎤ ⎧ ⎫ ⎡ ⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ 1 0 ⎨η 1<br />

⎦<br />

′′ ⎬<br />

0 1<br />

⎩<br />

η 2<br />

′′ ⎭ + ⎣ ω2 1 0 ⎨η ⎦ 1 ⎬ ⎨<br />

0 ω2<br />

2 ⎩<br />

η<br />

⎭ = 0 ⎬<br />

[Φ]T ⎩<br />

2 sin (5t)<br />

⎭<br />

⎡ ⎤ ⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫ ⎡<br />

⎤T ⎧ ⎫<br />

⎣ 1 0 ⎨η 1<br />

⎦<br />

′′ ⎬<br />

0 1<br />

⎩<br />

η 2<br />

′′ ⎭ + ⎣ 1.8642 0 ⎨η ⎦ 1 ⎬<br />

0 0.438 2 ⎩<br />

η<br />

⎭ = 0.925 0.380 ⎨ 0 ⎬<br />

⎣ ⎦<br />

2 −0.219 0.534<br />

⎩<br />

sin (5t)<br />

⎭<br />

⎡ ⎤ ⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫ ⎧ ⎫<br />

⎣ 1 0 ⎨η 1<br />

⎦<br />

′′ ⎬<br />

0 1<br />

⎩<br />

η ′′ ⎭ + 3. 47 0 ⎨η ⎣ ⎦ 1 ⎬ ⎨<br />

0 0.192<br />

⎩<br />

η<br />

⎭ = −0.219 sin (5t) ⎬<br />

⎩<br />

2 0.534 sin (5t)<br />

⎭<br />

2<br />

The EOM are now decoupled and each EOM can be solved easily as follows<br />

η ′′<br />

1 (t) + 3.47η 1 (t) = −0.219 sin (5t)<br />

η ′′<br />

2 (t) + 0.192η 2 (t) = 0.534 sin (5t)<br />

To solve these EOM’s, the initial conditions in normal coordinates must be transformed to modal coordinates<br />

using the above transformation rules<br />

⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫<br />

⎨η 1 (0) ⎬<br />

⎩<br />

η 2 (0)<br />

⎭ = 0.925 −0.657 ⎨x ⎣ ⎦ 1 (0) ⎬<br />

0.38 1.6<br />

⎩<br />

x 2 (0)<br />

⎭<br />

⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫<br />

⎨η 1 (0) ⎬<br />

⎩<br />

η 2 (0)<br />

⎭ = 0.925 −0.657 ⎨0⎬<br />

⎣ ⎦<br />

0.38 1.6<br />

⎩<br />

1<br />

⎭<br />

⎧ ⎫<br />

⎨−0.657⎬<br />

=<br />

⎩<br />

1.6<br />

⎭<br />

13


and<br />

⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫<br />

⎨η 1 ′ (0) ⎬<br />

⎩<br />

η 2 ′ (0) ⎭ = 0.925 −0.657 ⎨x ⎣ ⎦<br />

′ 1 (0) ⎬<br />

0.38 1.6<br />

⎩<br />

x ′ 2 (0) ⎭<br />

⎡<br />

⎤ ⎧ ⎫<br />

0.925 −0.657 ⎨1.5⎬<br />

= ⎣ ⎦<br />

0.38 1.6<br />

⎩<br />

3<br />

⎭<br />

⎧ ⎫<br />

⎨−0.584⎬<br />

=<br />

⎩<br />

5.37<br />

⎭<br />

Each of these EOM are solved using any of the standard methods. This results in solutions η 1 (t) and η 2 (t) .<br />

Hence the following EOM’s are solved<br />

and also<br />

η ′′<br />

1 (t) + 3.47η 1 (t) = −0.219 sin (5t)<br />

η 1 (0) = −0.657<br />

η ′ 1 (0) = −0.584<br />

η ′′<br />

2 (t) + 0.192η 2 (t) = 0.534 sin (5t)<br />

η 2 (0) = 1.6<br />

η ′ 2 (0) = 5.37<br />

The solutions η 1 (t) , η 2 (t) are found using basic methods shown in other parts of these notes. The last step<br />

is to transform back to normal coordinates by applying step 7<br />

⎧ ⎫ ⎡<br />

⎤ ⎧ ⎫<br />

⎨x 1 (t) ⎬<br />

⎩<br />

x 2 (t)<br />

⎭ = 0.925 0.380 ⎨η ⎣ ⎦ 1 (t) ⎬<br />

−0.219 0.534<br />

⎩<br />

η 2 (t)<br />

⎭<br />

⎧<br />

⎫<br />

⎨ 0.925 η 1 + 0.38η 2 ⎬<br />

=<br />

⎩<br />

0.534 η 2 − 0.219 η<br />

⎭<br />

1<br />

Hence<br />

and<br />

x 1 (t) = 0.925η 1 (t) + 0.38η 2 (t)<br />

x 2 (t) = 0.534η 1 (t) − 0.219η 2 (t)<br />

The above shows that the solution x 1 (t) and x 2 (t) has contributions from both nodal solutions.<br />

2 Fourier series representation of a periodic function<br />

Given a periodic function f (t) with period T then its Fourier series approximation ˜f (t) using N terms is<br />

(<br />

˜f (t) = 1 N<br />

)<br />

2 F ∑<br />

2π<br />

in<br />

0 + Re F n e T t<br />

= 1 2 F 0 + 1 2<br />

= 1 2<br />

N∑<br />

n=−N<br />

n=1<br />

N∑<br />

F n e<br />

n=1<br />

2π<br />

in<br />

F n e T t<br />

in<br />

2π<br />

T t + F ∗ ne<br />

2π<br />

−in T t<br />

14


Where<br />

F n = 2 T<br />

F 0 = 2 T<br />

∫ T<br />

0<br />

∫ T<br />

0<br />

2π<br />

−in<br />

f (t) e T t dt<br />

f (t) dt<br />

Another way to write the above is to use the classical representation using cos and sin. The same coefficients<br />

(i.e. the same series) will result.<br />

˜f (t) = a 0 +<br />

a 0 = 1 T<br />

n=1<br />

∫ T<br />

0<br />

a n = 1<br />

T/2<br />

b n = 1<br />

T/2<br />

N∑<br />

a n cos n 2π N T t + ∑<br />

b n sin n 2π<br />

T t<br />

f (t) dt<br />

∫ T<br />

0<br />

∫ T<br />

0<br />

n=1<br />

(<br />

f (t) cos n 2π )<br />

T t dt<br />

(<br />

f (t) sin n 2π )<br />

T t dt<br />

Just watch out in the above, that we divide by the full period when finding a 0 and divide by half the period<br />

for all the other coefficients. In the end, when we find ˜f (t) we can convert that to complex form. The complex<br />

form seems easier to use.<br />

15


3 Transfer functions for different vibration systems<br />

y<br />

c<br />

m<br />

k<br />

z<br />

u y z<br />

3.1 Force transmissibility<br />

Let steady state<br />

x ss = Re<br />

{ ˆF<br />

k D (r, ζ) eiϖt }<br />

Then<br />

f tr (t) = f spring + f damper<br />

= kx + cx ′<br />

{<br />

= Re k ˆF<br />

} {<br />

k D (r, ζ) eiϖt + Re ciϖ ˆF<br />

}<br />

k D (r, ζ) eiϖt<br />

{(<br />

= Re ˆF + ciϖ ˆF<br />

) }<br />

D (r, ζ) e iϖt<br />

k<br />

Hence<br />

|f tr (t)| max<br />

= ∣ ˆF<br />

√<br />

∣ √<br />

∣ |D| 1 + c 2 ϖ2 ∣∣<br />

k 2 = ˆF ∣ |D| 1 + (2ζr) 2<br />

So TR or force transmissibility is<br />

T R = |f √<br />

tr (t)| ∣ max<br />

∣ ˆF<br />

= |D| 1 + (2ζr) 2<br />

∣<br />

If r > √ 2 then we want small ζ to reduce force transmitted to base. For r < √ 2, it is the other way round.<br />

3.2 vibration isolation<br />

We need transfer function between y and z.<br />

Equation of motion<br />

my ′′ = −c ( y ′ − z ′) − k (y − z)<br />

my ′′ + cy ′ + ky = cz ′ + kz<br />

16


Let z = Re { Ze iωt} , z ′ = Re { iωZe iωt} and let y = Re { Y e iωt} , y ′ = Re { iωY e iωt} , y ′′ = Re { −ω 2 Y e iωt} ,<br />

hence the above becomes<br />

Hence |D (r, ζ)| =<br />

m Re { −ω 2 Y e iωt} + c Re { iωY e iωt} + k Re { Y e iωt} = c Re { iωZe iωt} + k Re { Ze iωt}<br />

√<br />

1+(2ζr) 2<br />

√<br />

(1−r 2 ) 2 +(2ζr) 2<br />

ciω + k<br />

Y =<br />

−ω 2 m + ciω + k Z<br />

i2ζω n mω + k<br />

=<br />

−ω 2 m + i2ζω n mω + k Z<br />

i2ζω n ω + ωn<br />

2 =<br />

−ω 2 + i2ζω n ω + ωn<br />

2 Z<br />

i2ζr + 1<br />

=<br />

(1 − r 2 ) + i2ζr Z<br />

( )<br />

and arg (D) = tan−1 (2ζr) − tan −1 2ζr<br />

1−r 2<br />

where r = ω ω n<br />

√<br />

Hence for good vibration isolation we need |Y | max<br />

|Z|<br />

to be small. i.e. |D| 1 + (2ζr) 2 to be small. This is the<br />

same TR as for force isolation above.<br />

For small |D|, we need small ζ and small k (the small k is to make r > √ 2) see plot<br />

In Matlab, the above can be plotted using<br />

17


close all;<br />

zeta=linspace(0.1, 0.7, 10);<br />

r=linspace(0, 3, 10);<br />

D=@(r,z) (sqrt(1+(2*z*r).^2)./sqrt((1-r.^2).^2+(2*z*r).^2));<br />

figure;<br />

hold on;<br />

for i=1:length(zeta)<br />

plot(r,D(r,zeta(i)));<br />

end<br />

grid on;<br />

legend<br />

3.3 accelerometer<br />

We need transfer function between u and z a where now z a is the amplitude of the ground acceleration.<br />

This device is used to measure base acceleration by relating it linearly to relative displacement of m to base.<br />

Equation of motion. We use relative distance now.<br />

m ( u ′′ + z ′′) + cu ′ + ku = 0<br />

mu ′′ + cu ′ + ku = −mz ′′<br />

Let z ′′ = Re { Z a e iωt} . Notice we here jumped right away to the z ′′ itself and wrote it as Re { Z a e iωt} and we<br />

did not go through the steps as above starting from base motion. This is because we want the transfer function<br />

between relative motion u and acceleration of base.<br />

Now, u = Re { Ue iωt} , u ′ = Re { iωUe iωt} , u ′′ = Re { −ω 2 Ue iωt} , hence the above becomes<br />

Hence |D (r, ζ)| =<br />

m Re { −ω 2 Ue iωt} + c Re { iωUe iωt} + k Re { Ue iωt} = −m Re { Z a e iωt}<br />

−m<br />

U =<br />

−ω 2 m + iωc + k Z a<br />

−1<br />

=<br />

−ω 2 + iω2ζω n + ωn<br />

2 Z a<br />

−1<br />

=<br />

(ωn 2 − ω 2 Z a<br />

) + iω2ζω n<br />

( )<br />

−1<br />

√(ω and arg (D) = −1800 − tan −1 2ωζωn<br />

n−ω 2 2 ) 2 +(2ωζω n) 2 ωn−ω 2 2<br />

When system is very stiff, which means ω n very large compared to ω , then D (r, ζ) ≈ −1<br />

ω 2 n<br />

Z a , hence by<br />

measuring u we estimate Z a the amplitude of the ground acceleration since ω 2 n is known. For accuracy, need<br />

ω n > 5ω at least.<br />

3.4 seismometer<br />

Now we need to measure the base motion (not base acceleration like above). But we still use the relative<br />

displacement. Now the transfer function is between u and z where now z is the base motion amplitude.<br />

Equation of motion. We use relative distance now.<br />

m ( u ′′ + z ′′) + cu ′ + ku = 0<br />

mu ′′ + cu ′ + ku = −mz ′′<br />

Let z = Re { Ze iωt} , z ′ = Re { iωZe iωt} ,z ′′ = Re { −ω 2 Ze iωt} ,and let u = Re { Ue iωt} , u ′ = Re { iωUe iωt} , u ′′ =<br />

Re { −ω 2 Ue iωt} , hence the above becomes<br />

18


Now, u = Re { Ue iωt} , u ′ = Re { iωUe iωt} , u ′′ = Re { −ω 2 Ue iωt} , hence the above becomes<br />

m Re { −ω 2 Ue iωt} + c Re { iωUe iωt} + k Re { Ue iωt} = −m Re { −ω 2 Ze iωt}<br />

U =<br />

mω 2<br />

−ω 2 m + iωc + k Z<br />

ω 2<br />

=<br />

−ω 2 + iω2ζω n + ωn<br />

2 Z<br />

r 2<br />

=<br />

(1 − r 2 ) + i2ζr Z<br />

( )<br />

r<br />

Hence |D (r, ζ)| = √ 2<br />

and arg (D) = − 2ζr<br />

(1−r 2 )+i2ζr tan−1 1−r 2 1<br />

Now if r is very large, which happens when ω n ≪ ω, then<br />

(1−r 2 )+i2ζr ⇒ 1 since r 2 is the dominant factor.<br />

−r 2<br />

r<br />

Therefore U = 2<br />

(1−r 2 )+i2ζr Z a now becomes U ≃ −Z a therefore measuring the relative displacement U gives<br />

linear estimate of the ground motion. However, this device requires that ω n be much smaller than ω, which<br />

means that m has to be massive. So this device is heavy compared to accelerometer.<br />

3.5 Summary of vibration transfer functions<br />

For good isolation of mass from ground motion, rule of thumb: Make damping low, and stiffness low (soft spring).<br />

system equation used to derive transfer function<br />

isolate base from force<br />

transmitted by machine<br />

isolate machine<br />

from motion of base<br />

accelerometer: Measure<br />

base acc. using relative<br />

displacement<br />

f tr (t) = f spring + f damper<br />

Use absolute mass position<br />

my ′′ + cy ′ + ky = cz ′ + kz<br />

|f tr(t)| ∣ max<br />

∣ ˆF<br />

∣<br />

|Y | max<br />

|Z|<br />

√<br />

= |D| 1 + (2ζr) 2<br />

√<br />

= |D| 1 + (2ζr) 2<br />

−1<br />

U =<br />

Use relative mass position (ωn 2 − ω 2 Z a ⇒ |D (r, ζ)|<br />

) + iω2ζω n<br />

−1<br />

mu ′′ + cu ′ + ku = −mz ′′ = √<br />

(ωn 2 − ω 2 ) 2 + (2ωζω n ) 2<br />

seismometer: Measure<br />

base motion using<br />

relative displacement<br />

Use relative mass position<br />

mu ′′ + cu ′ + ku = −mz ′′ U =<br />

r 2<br />

(1−r 2 )+i2ζr Z ⇒ |D (r, ζ)| = r 2 √<br />

(1−r 2 )+i2ζr<br />

19


4 Steady state solutions to single degree freedom system<br />

4.1 Summary:<br />

( )<br />

my ′′ + cy ′ + ky = Re ˆF e<br />

iϖt<br />

{ }<br />

x = Re ˆXe<br />

iϖt<br />

ˆX = ˆF D (r, ζ)<br />

k<br />

1<br />

D (r, ζ) =<br />

(1 − r 2 ) + 2iζr<br />

{ }<br />

ˆF<br />

x = Re<br />

k |D (r, ζ)| ei(ϖt−θ)<br />

θ = tan −1<br />

2ζr<br />

1 − r 2<br />

4.2 Details<br />

( )<br />

Let load be harmonic and represented in general as Re ˆF e<br />

iϖt<br />

where ˆF is the complex amplitude of the force.<br />

Hence system is represented by<br />

( )<br />

my ′′ + cy ′ + ky = Re ˆF e<br />

iϖt<br />

( ) ˆF<br />

y ′′ + 2ζω n y ′ + ωny 2 = Re<br />

m eiϖt<br />

)<br />

)<br />

)<br />

Let y = Re<br />

(Ŷ e<br />

iϖt<br />

Hence y ′ = Re<br />

(iϖŶ eiϖt , y ′′ = Re<br />

(−ϖ 2 Ŷ e iϖt , therefore the differential equation<br />

becomes<br />

(<br />

Re −ϖ 2 Ŷ e iϖt) (<br />

+ 2ζω n Re iϖŶ eiϖt) + ωn 2 Re<br />

(Ŷ ) ( )<br />

e<br />

iϖt ˆF<br />

= Re<br />

m eiϖt<br />

Dividing numerator and denominator ω 2 n gives<br />

Where r = ϖ ω n<br />

, hence the response is<br />

Therefore, the phase of the response is<br />

y = Re<br />

( )<br />

arg (y) = arg ˆF<br />

( ˆF<br />

−ϖ 2 + 2ζω n iϖ + ωn) 2 Ŷ =<br />

m<br />

ˆF<br />

m<br />

Ŷ =<br />

(−ϖ 2 + 2ζω n iϖ + ωn)<br />

2<br />

Ŷ = ˆF 1<br />

k (1 − r 2 ) + i2ζr<br />

( )<br />

ˆF 1<br />

k (1 − r 2 ) + i2ζr eiϖt<br />

( 2ζr<br />

− tan −1 (1 − r 2 )<br />

20<br />

)<br />

+ ϖt


Hence at t = 0 the phase of the response will be<br />

( ) ( ) 2ζr<br />

arg (y) = arg ˆF − tan −1 (1 − r 2 )<br />

So when ˆF<br />

( )<br />

is real, the phase of the response is simply − tan −1 2ζr<br />

(1−r 2 )<br />

4.3 Undamped case<br />

When ζ = 0 the above becomes<br />

For real force this becomes<br />

∣<br />

The magnitude ∣Ŷ ∣ = F k<br />

4.4 damped cases<br />

ζ > 0<br />

1<br />

(1−r 2 )<br />

( ˆF<br />

y = Re<br />

k<br />

∣ ˆF<br />

∣<br />

=<br />

k<br />

)<br />

1<br />

(1 − r 2 ) eiϖt<br />

1<br />

(<br />

(1 − r 2 ) cos<br />

y = F k<br />

and phase zero.<br />

∣<br />

∣Ŷ<br />

y = Re<br />

( ))<br />

ϖt + arg ˆF<br />

1<br />

(1 − r 2 cos (ϖt)<br />

)<br />

( )<br />

ˆF 1<br />

k (1 − r 2 ) + i2ζr eiϖt<br />

∣ ˆF<br />

∣ 1<br />

∣ = √<br />

k<br />

(1 − r 2 ) 2 + (2ζr) 2<br />

) (<br />

arg<br />

(Ŷ = φ = arg ˆF<br />

Hence for real force and at t = 0 the phase of displacement is<br />

( ) 2ζr<br />

− tan −1 1 − r 2<br />

)<br />

− tan −1 ( 2ζr<br />

1 − r 2 )<br />

+ ϖt<br />

lag behind the load.<br />

When r < 1 then φ goes from 0 to −90 0 Therefore phase of displacement is 0 to −90 0 behind force. The<br />

minus sign at the front was added since the complex number is in the denominator. Hence the response will<br />

always be lagging in phase relative for load.<br />

For r > 1<br />

Now 1 − r 2 is negative, hence the phase will be from −90 ◦ to −180 ◦<br />

When r = 1<br />

( ) ˆF 1<br />

y = Re<br />

k i2ζ eiϖt<br />

∣ ∣<br />

∣Ŷ<br />

)<br />

arg<br />

(Ŷ<br />

∣ ˆF<br />

∣<br />

∣ =<br />

k<br />

= −90 ◦<br />

1<br />

2ζ<br />

Now phase is −90 ◦ 21


Phase of response complex<br />

amplitude for underdamped<br />

and when r1<br />

Phase will be from -90 to -<br />

180 degrees<br />

Phase of response complex<br />

amplitude for<br />

underdamped and when<br />

r=1<br />

Phase will -90 degrees<br />

2r<br />

2r<br />

1 r 2<br />

1 r 2<br />

Examples. System has ζ =<br />

)<br />

0.1 and<br />

√<br />

m = 1, k = 1 subjected for force 3 cos (0.5t) find the steady state solution.<br />

Answer y (t) = Re<br />

(Ŷ e<br />

iϖt<br />

, ω n = = 1 rad/sec, hence r = 0.5 under the response is<br />

k<br />

m<br />

(∣ ∣∣ y (t) = Re Ŷ ∣ e iϖt)<br />

∣<br />

= ∣Ŷ ∣ cos (ϖt)<br />

= F k<br />

1<br />

√(1 − r 2 ) 2 + (2ζr) 2 cos<br />

(<br />

( ))<br />

2 (0.1) 0.5<br />

.5t − tan −1 1 − 0.5 2<br />

1<br />

= 3<br />

cos (.5t − 7.59 √(1 ◦ )<br />

− 0.5 2 ) 2 + (2 (0.1) 0.5) 2<br />

= 3. 964 9 cos (.5t − 7.59 ◦ )<br />

5 Summary table of solutions<br />

5.1 Harmonic loading mu ′′ + cu ′ + ku = F sin ϖt<br />

The equation of motion can also be written as u ′′ + 2ζωu ′ + ω 2 u = F m<br />

sin ϖt.<br />

22


The following table gives the solutions for initial conditions are u (0) and u ′ (0) under all damping conditions.<br />

The roots shown are the roots of the quadratic characteristic equation λ 2 + 2ζωλ + ω 2 λ = 0. Special handling is<br />

needed to obtain the solution of the differential equation for the case of ζ = 0 and ϖ = ω as described in the<br />

detailed section below.<br />

⎧<br />

⎨ −iω<br />

roots<br />

ζ = 0<br />

⎩<br />

⎧<br />

⎪⎨<br />

u (t)<br />

⎪⎩<br />

⎧<br />

⎨<br />

roots<br />

⎩<br />

+iω<br />

ϖ = ω → u (0) cos ϖt + u′ (0)<br />

ϖ<br />

sin ϖt − F K<br />

ϖ ≠ ω → u (0) cos ωt +<br />

−ξω + iω n<br />

√<br />

1 − ξ 2<br />

−ξω − iω n<br />

√<br />

1 − ξ 2<br />

(<br />

u ′ (0)<br />

ω<br />

− F K<br />

ϖt<br />

2<br />

cos (ϖt)<br />

)<br />

r<br />

sin ωt + F 1<br />

1−r 2 K<br />

sin ϖt<br />

1−r 2<br />

ζ < 1<br />

u (t) = e −ξωt (A cos ω d t + B sin ω d t) + F K<br />

A = u 0 + F K<br />

1 √(1−r 2 ) 2 +(2ξr) 2 sin θ<br />

1<br />

sin (ϖt − θ)<br />

√(1−r 2 ) 2 2<br />

+(2ξr)<br />

B = v 0<br />

ω d<br />

+ u 0ξω<br />

ω d<br />

⎧<br />

⎨ −ω<br />

roots<br />

⎩<br />

−ω<br />

+ F √ 1<br />

K<br />

ω d (1−r 2 ) 2 +(2ξr)<br />

2<br />

(ξω sin θ − ϖ cos θ)<br />

ζ = 1<br />

ζ > 1<br />

u (t) = (A + Bt) e −ωt + F K<br />

A = u 0 + F K<br />

1 √(1−r 2 ) 2 +(2r) 2 sin θ<br />

1<br />

sin (ϖt − θ)<br />

√(1−r 2 ) 2 2<br />

+(2r)<br />

F/k<br />

B = v 0 + u 0 ω +<br />

(ω sin θ − ϖ cos θ)<br />

√(1−r 2 ) 2 2<br />

+(2r)<br />

⎧<br />

√<br />

⎨ −ω n ξ + ω n ξ 2 − 1<br />

roots<br />

⎩<br />

√<br />

−ω n ξ − ω n ξ 2 − 1<br />

(<br />

−ξ+ √ (<br />

ξ<br />

u (t) = Ae<br />

2 −1<br />

)ω nt −ξ− √ )<br />

ξ + Be 2 −1 ω nt +<br />

F<br />

K<br />

β sin (ϖt − θ)<br />

A = v 0+u 0 ωξ+u 0 ω √ ((<br />

ξ 2 −1+ F K β ξ+ √ )<br />

)<br />

ξ 2 −1 ω sin θ−ϖ cos θ<br />

2ω √ ξ 2 −1<br />

B = − v 0+u 0 ωξ−u 0 ω √ ((<br />

ξ 2 −1+ F K β ξ− √ )<br />

)<br />

ξ 2 −1 ω sin θ−ϖ cos θ<br />

2ω √ ξ 2 −1<br />

23


5.2 constant loading mu ′′ + cu ′ + ku = F<br />

⎧<br />

⎨ −iω<br />

roots<br />

ζ = 0 ⎩<br />

+iω<br />

ζ < 1<br />

ζ = 1<br />

u (t) = ( u 0 − F )<br />

k cos ωt +<br />

v 0<br />

ω<br />

sin ωt + F k<br />

⎧<br />

√<br />

⎨ −ξω + iω n 1 − ξ 2<br />

roots<br />

⎩<br />

√<br />

−ξω − iω n 1 − ξ 2<br />

( (u0<br />

u (t) = e −ξωt − F )<br />

k cos ωd t +<br />

⎧<br />

⎨ −ω<br />

roots<br />

⎩<br />

−ω<br />

(<br />

) )<br />

v 0 +u 0 ξω− F k ξω<br />

ω d<br />

sin ω d t + F k<br />

u (t) = (( u 0 − F ) (<br />

k + v0 + u 0 ω − F k ω) t ) e −ωt + F k<br />

⎧<br />

√<br />

⎨ −ω n ξ + ω n ξ 2 − 1<br />

roots<br />

⎩<br />

√<br />

−ω n ξ − ω n ξ 2 − 1<br />

ζ > 1<br />

B = F k p 1−u 0 p 1 +v 0<br />

(p 2 −p 1 )<br />

A = u 0 − F k − B<br />

p 1 = − c<br />

2m + √ ( c<br />

2m<br />

p 2 = − c<br />

2m − √ ( c<br />

2m<br />

u (t) = Ae p 1t + Be p 2t + F k<br />

) 2 −<br />

k<br />

m = −ω √<br />

nξ + ω n ξ 2 − 1<br />

) 2 −<br />

k<br />

m = −ω √<br />

nξ − ω n ξ 2 − 1<br />

24


5.3 no loading (free) mu ′′ + cu ′ + ku = 0<br />

⎧<br />

⎨ −iω<br />

roots<br />

⎩<br />

+iω<br />

ζ = 0<br />

u ′′ + ω 2 u = 0<br />

ζ < 1<br />

ζ = 1<br />

u (t) = u(0) cos ωt + u′ (0)<br />

ω<br />

sin ωt<br />

⎧<br />

u (t) = A cos (ωt − φ)<br />

⎪⎨ √<br />

( )<br />

A = u(0) + u ′ 2<br />

(0)<br />

ω<br />

( )<br />

⎪⎩ φ = tan −1 u ′ (0)/ω<br />

u(0)<br />

⎧<br />

⎨ −ξω + iω √ 1 − ξ 2<br />

roots<br />

⎩<br />

−ξω − iω √ 1 − ξ 2<br />

(<br />

u (t) = e −ξωt u(0) cos ω d t + u′ (0)+u(0)ξω<br />

ω<br />

⎧<br />

d<br />

⎨ −ω<br />

roots<br />

⎩<br />

−ω<br />

)<br />

sin ω d t<br />

ζ > 1<br />

u (t) = (u(0)(1 + ωt) + u ′ (0)t) e −ωt<br />

⎧<br />

⎨ λ 1 = −ωξ + ω √ ξ 2 − 1<br />

roots<br />

⎩<br />

λ 2 = −ωξ − ω √ ξ 2 − 1<br />

⎧<br />

u (t) = Ae<br />

⎪⎨<br />

λ1ωt + Be λ 2ωt<br />

A = u′ (0)−u(0)λ 2<br />

2ω √ ξ 2 −1<br />

⎪⎩<br />

B = −u′ (0)+u(0)λ 1<br />

2ω √ ξ 2 −1<br />

25


5.4 impulse F 0 δ (t) loading<br />

⎧<br />

⎨ −iω<br />

roots<br />

ζ = 0<br />

⎩<br />

+iω<br />

u ′′ + ω 2 u = 0<br />

ζ < 1<br />

ζ = 1<br />

ζ > 1<br />

transient<br />

{ }} {<br />

u (t) = u(0) cos ωt + u′ (0)<br />

⎧<br />

ω<br />

⎨ −ξω + iω √ 1 − ξ 2<br />

roots<br />

⎩<br />

−ξω − iω √ 1 − ξ 2<br />

steady state<br />

{ }} {<br />

sin ωt + F 0<br />

sin ωt<br />

mω<br />

transient<br />

steady state<br />

{ (<br />

}} ){<br />

{ ( }} ){<br />

u (t) = e −ξωt u(0) cos ω d t + u′ (0) + u(0)ξω<br />

sin ω d t + e −ξωt F0<br />

sin ω d t<br />

⎧<br />

ω d<br />

mω d<br />

⎨ −ω<br />

roots<br />

⎩<br />

−ω<br />

u (t) = (u (0) (1 + ωt) + u ′ (0) t) e −ωt + F 0t<br />

⎧<br />

⎨ λ 1 = −ωξ + ω √ ξ 2 − 1<br />

roots<br />

⎩<br />

λ 2 = −ωξ − ω √ ξ 2 − 1<br />

⎧<br />

u (t) = Ae<br />

⎪⎨<br />

λ1ωt + Be λ2ωt +<br />

⎪⎩<br />

A = u′ (0)−u(0)λ 2<br />

2ω √ ξ 2 −1<br />

B = −u′ (0)+u(0)λ 1<br />

2ω √ ξ 2 −1<br />

m e−ωt<br />

F 0<br />

m<br />

2ω √ ξ 2 −1 eλ 1ωt −<br />

The impulse response can be implemented in Mathematica as<br />

parms = {m -> 10, c -> 1.2, k -> 4.3, a -> 1};<br />

tf = TransferFunctionModel[a/(m s^2 + c s + k) /. parms, s]<br />

sol = OutputResponse[tf, DiracDelta[t], t];<br />

F 0<br />

m<br />

2ω √ ξ 2 −1 eλ 2ωt<br />

Plot[sol, {t, 0, 60}, PlotRange -> All, Frame -> True,<br />

FrameLabel -> {{z[t], None}, {Row[{t, " (sec)"}], eq}},GridLines -> Automatic]<br />

26


5.5 Impulse sin function<br />

Input is given by F (t) = F 0 sin (ϖt) where ϖ = 2π<br />

2t 1<br />

= π t 1<br />

ζ = 0<br />

ζ < 1<br />

ζ = 1<br />

⎧<br />

⎨ −iω<br />

roots<br />

⎩<br />

+iω<br />

⎧<br />

⎧<br />

⎪⎨<br />

ϖ = ω →<br />

(<br />

⎪⎨<br />

⎪⎩ −u (0) +<br />

F 0 π<br />

k 2<br />

⎧<br />

(<br />

u (t)<br />

⎪⎨ u (0) cos ωt +<br />

ϖ ≠ ω →<br />

⎪⎩ ⎪⎩<br />

⎧<br />

√<br />

⎨ −ξω + iω n 1 − ξ 2<br />

roots<br />

⎩<br />

√<br />

−ξω − iω n 1 − ξ 2<br />

⎧<br />

⎪⎨<br />

u (t) =<br />

⎪⎩<br />

u (0) cos ωt + u′ (0)<br />

ω sin ωt − F 0<br />

k<br />

ωt<br />

)<br />

cos ωt +<br />

−u ′ (0)+ F 0<br />

k<br />

u ′ (0)<br />

ϖ<br />

( −<br />

(F π/t1<br />

0/k) ω<br />

π/t1<br />

1−(<br />

ω<br />

2 cos (ωt) 0 ≤ t ≤ t 1<br />

π<br />

2t 1<br />

))<br />

) 2<br />

ω<br />

sin ωt t > t 1<br />

sin ωt + F 0/k<br />

1−( π/t1<br />

ω<br />

( )<br />

) 2 sin πt<br />

t 1<br />

0 ≤ t ≤ t 1<br />

u (t 1 ) cos ωt + u′ (t 1 )<br />

ω<br />

sin ωt t > t 1<br />

time constant τ = 1<br />

ζω n<br />

e −ξωt (A cos ω d t + B sin ω d t) + F 0<br />

√ 1<br />

k<br />

(<br />

e −ξωt u (t 1 ) cos ω d t + u′ (t 1 )+u(t 1 )ξω<br />

ω d<br />

sin ω d t<br />

A = u (0) + F 0<br />

k<br />

1 √(1−r 2 ) 2 +(2ξr) 2 sin θ<br />

B = u′ (0)<br />

ω d<br />

⎧<br />

⎨<br />

roots<br />

+ u(0)ξω<br />

ω d<br />

+ F 0<br />

−ω<br />

√ 1<br />

k<br />

ω d (1−r 2 ) 2 +(2ξr)<br />

(<br />

sin<br />

(1−r 2 ) 2 +(2ξr)<br />

)<br />

2<br />

2<br />

(ξω sin θ − ϖ cos θ)<br />

)<br />

t 1<br />

− θ<br />

⎩<br />

−ω<br />

⎧<br />

( )<br />

⎪⎨ (A + Bt) e −ωt + F 0 √ 1<br />

sin k<br />

π t<br />

u (t) =<br />

(1−r 2 ) 2 +(2r) 2 t 1<br />

− θ 0 ≤ t ≤ t 1<br />

⎪⎩<br />

u (t) = (u (t 1 ) + (u ′ (t 1 ) + u (t 1 ) ω) t) e −ωt t > t 1<br />

π t<br />

0 ≤ t ≤ t 1<br />

t > t 1<br />

A = u (0) + F 0<br />

k<br />

1 √(1−r 2 ) 2 +(2r) 2 sin θ<br />

B = u ′ (0) + u (0) ω + F 0<br />

(<br />

√ 1<br />

k<br />

(1−r 2 ) 2 +(2r) 2<br />

)<br />

ω sin θ − π t 1<br />

cos θ<br />

27


ζ > 1<br />

⎧<br />

√<br />

⎨ −ω n ξ + ω n ξ 2 − 1<br />

roots<br />

⎩<br />

√<br />

−ω n ξ − ω n ξ 2 − 1<br />

⎧ (<br />

⎪⎨<br />

−ξ+ √ (<br />

ξ<br />

Ae<br />

2 −1<br />

)ωt + Be<br />

u (t) =<br />

(<br />

⎪⎩<br />

−ξ+ √ )<br />

ξ 2 −1<br />

ω<br />

u (t) = A 1 e<br />

nt + B1 e<br />

A = u′ (0)+u(0)ωξ+u(0)ω √ ((<br />

ξ 2 −1+ F k β ξ+ √ )<br />

ξ 2 −1<br />

2ω √ ξ 2 −1<br />

B = − u′ (0)+u(0)ωξ−u(0)ω √ ξ 2 −1+ F k β ((<br />

ξ− √ ξ 2 −1<br />

1<br />

β = √<br />

(1−r 2 ) 2 +(2ξr) 2<br />

(<br />

A 1 = − ˙u(t 1)+u(t 1 )ω n ξ− √ )<br />

ξ 2 −1<br />

√<br />

2ω n ( ξ 2 −1<br />

B 1 = ˙u(t 1)+u(t 1 )ξω n ξ+ √ )<br />

ξ 2 −1<br />

2ω n<br />

√<br />

ξ 2 −1<br />

2ω √ ξ 2 −1<br />

−ξ− √ )<br />

ξ 2 −1 ωt +<br />

F<br />

π t<br />

(<br />

−ξ− √ )<br />

ξ 2 −1 ω nt<br />

k β sin (<br />

)<br />

ω sin θ− π cos θ t 1<br />

)<br />

)<br />

ω sin θ− π cos θ t 1<br />

)<br />

t 1<br />

− θ<br />

0 ≤ t ≤ t 1<br />

t > t 1<br />

6 Tree view look at the different cases<br />

This tree illustrates the different cases that needs to be considered for the solution of single degree of freedom<br />

system with harmonic loading.<br />

There are 12 cases to consider. Resonance needs to be handled as special case when damping is absent due to<br />

the singularity in the standard solution when the forcing frequency is the same as the natural frequency. When<br />

damping is present, there is no resonance, however, there is what is called practical response which occur when<br />

the forcing frequency is almost the same as the natural frequency.<br />

28


Solution to single degree of freedom system, second order, linear, time invariant<br />

By Nasser M. Abbasi<br />

December 11, 2012<br />

Single_degree_system_tree.vsd<br />

undamped<br />

damped<br />

r <br />

<br />

c<br />

c cr<br />

free<br />

mü ku 0<br />

forced<br />

mü ku Fsint<br />

free<br />

forced<br />

mü cu ku 0<br />

mü cu ku Fsint<br />

1 1 1<br />

r 1 r 1<br />

r 1 r 1<br />

1<br />

1<br />

1 1 1<br />

1<br />

r 1 resonance<br />

1 underdamped, roots a ib with a real and negative<br />

1 critical damping roots a,a<br />

1 overdamped roots both real and negative a,b<br />

x<br />

x<br />

x<br />

xx<br />

The following is another diagram made sometime ago which contains more useful information and is kept<br />

here for reference.<br />

29


One degree freedom<br />

linear system unforced<br />

response<br />

y t c m y t k m yt 0<br />

y t 2 n y t n2 yt 0<br />

Some<br />

defintions<br />

k<br />

n m rad/sec<br />

damping ratio<br />

c c c<br />

r<br />

c<br />

2 km 2 n m<br />

Solution roots<br />

s 1,2 n n 2 1<br />

3 cases<br />

d n 1 2 30<br />

Defined only<br />

for under<br />

damped<br />

case(else zero)<br />

1 1 1<br />

Underdamped<br />

2 roots,<br />

complex<br />

conjugate<br />

Critical damped<br />

(one real root,<br />

double<br />

multiplicity)<br />

Over damped,<br />

2 real roots,<br />

distinct<br />

yt e nt Acos d t Bsin d t<br />

A y0, B y 0A n<br />

d<br />

yt A Bte nt<br />

A y0, B y 0 A n<br />

T d 2<br />

d<br />

1<br />

n<br />

Nasser M. Abbasi<br />

May 25, 2011<br />

One_DOF_system.vsd<br />

Damped period of oscillation<br />

Time constant<br />

yt Ae 2 1 n t Be 2 1 n t<br />

A y 0y0 n<br />

2 n<br />

B y 0y0 n<br />

2 n<br />

2 1<br />

2 1<br />

2 1<br />

2 1


7 Some common formulas<br />

These definitions are used throughout the derivations below.<br />

ξ = c c r<br />

=<br />

c<br />

2 √ km = c<br />

2ω n m<br />

u st = F static deflection<br />

√<br />

k<br />

k<br />

ω n =<br />

m<br />

ω D = ω n<br />

√<br />

1 − ξ 2 note: not defined for ξ > 1 since becomes complex<br />

r = ϖ ω n<br />

T d = 2π damped period of oscillation<br />

ω<br />

{ d<br />

−1<br />

τ = , −1 }<br />

time constants where λ i are roots of characteristic equation<br />

λ 1 λ2<br />

1<br />

β =<br />

magnification factor<br />

√(1 − r 2 ) 2 + (2rξ) 2<br />

β max when r = √ 1 − 2ξ 2<br />

1<br />

β max =<br />

2ξ √ 1 − ξ 2<br />

y n<br />

= e ζωn2π<br />

ω D<br />

y<br />

( n+1<br />

)<br />

yn<br />

ln = ζ2π<br />

y n+1<br />

( )<br />

1<br />

M ln yn<br />

= ζ2π<br />

y n+M<br />

small damping<br />

⇒<br />

√<br />

ζ2π<br />

1−ζ 2<br />

e<br />

⇒ e ζ2π<br />

This table shows many cycles it takes for the peak to decay by half its original value as a function of the<br />

damping ζ. For example, we see that when ζ = 2.7% then it takes 4 cycles for the peak (i.e. displacement) to<br />

reduce to half its value.<br />

data = Table[{i, (1/i Log[2]/(2*Pi)*100)}, {i, 1, 20}];<br />

TableForm[N@data,<br />

TableHeadings -> {None, {Column[{"number of cycles",<br />

"needed for peak", "to decay by half"}], "\[Zeta] (%)"}}]<br />

31


The roots of the characteristic equation for u ′′ + 2ξωu ′ + ω 2 u = 0 are given in this table<br />

ξ < 1<br />

roots<br />

{ √ √ }<br />

−ξω + jω n 1 − ξ 2 , −ξω − iω n 1 − ξ 2<br />

time constant τ<br />

1<br />

ξω<br />

1<br />

ξ = 1 {−ω, −ω}<br />

ω<br />

{ √ √ }<br />

ξ > 1 −ω n ξ + ω n ξ 2 − 1, −ω n ξ − ω n ξ 2 1<br />

− 1<br />

√ , 1√ (which to use? the bigger?)<br />

ω nξ−ω n ξ 2 −1 ω nξ+ω n ξ 2 −1<br />

32


8 Derivation of the solutions of the equations of motion<br />

8.1 Free vibration<br />

8.1.1 Undamped free vibration<br />

u<br />

K<br />

M<br />

mu ′′ + ku = 0<br />

u ′′ + ω 2 u = 0<br />

The solution is<br />

u (t) = u (0) cos ωt + u′ (0)<br />

ω<br />

sin ωt<br />

8.1.2 under-damped free vibration c < c r , ξ < 1<br />

u<br />

K<br />

M<br />

C<br />

The solution is<br />

mu ′′ + cu ′ + ku = 0<br />

u ′′ + 2ξωu ′ + ω 2 u = 0<br />

u = e −ξωt (A cos ω d t + B sin ω d t)<br />

Applying initial conditions gives A = u (0) and B = u′ (0)+u(0)ξω<br />

ω d<br />

. Therefore the solution becomes<br />

(<br />

)<br />

u (t) = e −ξωt u (0) cos ω d t + u′ (0) + u (0) ξω<br />

sin ω d t<br />

ω d<br />

33


8.1.3 critically damped free vibration ξ = c<br />

c r<br />

= 1<br />

The solution is<br />

u (t) = (A + Bt) e −( cr<br />

2m<br />

)<br />

t<br />

= (A + Bt) e −ωt<br />

where A, B are found from initial conditions A = u (0),B = u ′ (0) + u (0) ω, hence<br />

8.1.4 over-damped free vibration ξ = c<br />

c r<br />

> 1<br />

The solution is<br />

where A, B are found from initial conditions.<br />

u (t) = ( u (0) + ( u ′ (0) + u (0) ω ) t ) e −ωt<br />

u (t) = Ae λ 1t + Be λ 2t<br />

A = u′ (0) − u (0) λ 2<br />

2ω √ ξ 2 − 1<br />

B = −u′ (0) + u (0) λ 1<br />

2ω √ ξ 2 − 1<br />

where λ 1 and λ 2 are the roots of the characteristic equation<br />

8.2 Constant input F<br />

8.2.1 Undamped case ζ = 0<br />

λ 1 = − c<br />

√ ( c<br />

2m + 2m<br />

λ 2 = − c<br />

2m − √ ( c<br />

2m<br />

) 2 k −<br />

m = −ξω + ω√ ξ 2 − 1<br />

) 2 k −<br />

m = −ξω − ω√ ξ 2 − 1<br />

mu ′′ + ku = F<br />

u ′′ + ω 2 u = F<br />

u (t) = u h + u p<br />

Where u h = A cos ωt + B sin ωt and u p = F k<br />

, the solution is<br />

Applying initial conditions gives<br />

u (t) = A cos ωt + B sin ωt + F k<br />

And complete solution is<br />

A = u (0) − F k<br />

B = u′ (0)<br />

ω<br />

u (t) = F (<br />

k + u (0) − F )<br />

cos ωt + u′ (0)<br />

sin ωt<br />

k<br />

ω<br />

34


8.2.2 under-damped ζ < 1<br />

The general solution is<br />

u (t) = e −ξωt (A cos ω d t + B sin ω d t) + F k<br />

From initial conditions<br />

Hence the solution is<br />

u (t) = e −ξωt ( (<br />

u (0) − F k<br />

A = u (0) − F k<br />

B = u′ (0) + u (0) ξω − F k ξω<br />

ω d<br />

) (<br />

u ′ (0) + u (0) ξω − F k<br />

cos ω d t +<br />

ξω<br />

) )<br />

sin ω d t + F ω d k<br />

8.2.3 Critical damping ζ = 1<br />

The general solution is<br />

u(t) = (A + Bt)e −ωt + F k<br />

Where from initial conditions<br />

A = u(0) − F k<br />

B = u ′ (0) + u(0)ω − F k ω<br />

8.2.4 Over-damped ζ > 0<br />

The solution is<br />

u (t) = Ae p1t + Be p2t + F k<br />

Where now<br />

B =<br />

F<br />

k p 1 − u 0 p 1 + u ′ (0)<br />

(p 2 − p 1 )<br />

A = u (0) − F k − B<br />

Hence the solution is<br />

u (t) = Ae p 1t + Be p 2t + F k<br />

Where<br />

p 1 = − c<br />

√ ( c<br />

2m + 2m<br />

p 2 = − c<br />

2m − √ ( c<br />

2m<br />

) 2 k −<br />

m = −ωξ + ω √<br />

n ξ 2 − 1<br />

) 2 k −<br />

m = −ωξ − ω √<br />

n ξ 2 − 1<br />

35


8.3 Harmonic forced input F sin (ϖt)<br />

8.3.1 Undamped forced vibration<br />

u<br />

K<br />

M<br />

ReF e it <br />

mu ′′ + ku = F sin ϖt<br />

Since there is no damping in the system, then there is no steady state solution. In other words, the particular<br />

solution is not the same as the steady state solution in this case. We need to find the particular solution using<br />

method on undetermined coefficients.<br />

Let u = u h + u p . By guessing that u p = c 1 sin ϖt then we find the solution to be<br />

u = A cos ω n t + B sin ω n t + F 1<br />

sin ϖt<br />

k 1 − r2 Applying initial conditions is always done on the full solution. Applying initial conditions gives<br />

u (0) = A<br />

u ′ (t) = −Aω sin ω n t + Bω cos ω n t + ϖ F 1<br />

cos ϖt<br />

k 1 − r2 u ′ (0) = Bω n + ϖ F 1<br />

k 1 − r 2<br />

B = u′ (0)<br />

− F r<br />

ω n k 1 − r 2<br />

Where r = ϖ ω n<br />

The complete solution is<br />

( u ′ (0)<br />

u (t) = u (0) cos ω n t + − F k<br />

ω n<br />

r<br />

1 − r 2 )<br />

sin ω n t + F k<br />

1<br />

sin ϖt (1)<br />

1 − r2 Example: Given force f (t) = 3 sin (5t) then ϖ = 5 rad/sec, and ˆF = 3. Let m = 1, k = 1, then ω n = 1<br />

rad/sec. Hence r = 5, Let initial conditions be zero, then<br />

( )<br />

5<br />

1<br />

u = −3<br />

1 − 5 2 sin t + 3 sin 5t<br />

1 − 52 = 0.625 sin t − 0.125 sin 5.0t<br />

Resonance forced vibration When ϖ ≈ ω we obtain resonance since r → 1 in the solution given in Eq (1)<br />

above and as written the solution can not be used for analysis. To obtain a solution for resonance some calculus<br />

is needed. Eq (1) is written as<br />

( u ′ (0)<br />

u (t) = u (0) cos ωt +<br />

ω<br />

− F k<br />

ωϖ<br />

ω 2 − ϖ 2 )<br />

sin ωt + F k<br />

ω 2<br />

ω 2 sin ϖt<br />

− ϖ2 (1A)<br />

36


When ϖ ≈ ω but less than ω, letting<br />

where ∆ is very small positive quantity. And since ϖ ≈ ω let<br />

Multiplying Eq (2) and (3) gives<br />

Eq (1A) can now be written in terms of Eqs (2,3) as<br />

( u ′ (0)<br />

u (t) = u (0) cos ωt +<br />

ω<br />

(<br />

v0<br />

= u (0) cos ωt +<br />

ω − F k<br />

ω − ϖ = 2∆ (2)<br />

ω + ϖ ≈ 2ϖ (3)<br />

ω 2 − ϖ 2 = 4∆ϖ (4)<br />

− F ωϖ<br />

k 4∆ϖ<br />

ω<br />

4∆<br />

Since ϖ ≈ ω the above becomes<br />

( u ′ (0)<br />

u (t) = u (0) cos ωt +<br />

ω<br />

− F k<br />

= u (0) cos ωt + u′ (0)<br />

ω<br />

sin ωt + F k<br />

)<br />

sin ωt + F k<br />

)<br />

sin ωt + F k<br />

)<br />

ω<br />

sin ωt + F 4∆ k<br />

Using sin ϖt − sin ωt = 2 sin ( ϖ−ω<br />

2<br />

t ) cos ( ϖ+ω<br />

2<br />

t ) the above becomes<br />

u (t) = u (0) cos ωt + u′ (0)<br />

ω<br />

From Eqs (2,3) the above can be written as<br />

sin ωt + F k<br />

u (t) = u (0) cos ωt + u′ (0)<br />

ω<br />

ω<br />

2∆<br />

sin ωt + F k<br />

(<br />

sin<br />

ω 2<br />

sin ϖt<br />

4∆ϖ<br />

ω 2<br />

sin ϖt<br />

4∆ϖ<br />

ω<br />

sin ϖt<br />

4∆<br />

ω<br />

(sin ϖt − sin ωt)<br />

4∆<br />

( ϖ − ω<br />

2<br />

) ( )) ϖ + ω<br />

t cos t<br />

2<br />

ω<br />

(sin (−∆t) cos (ϖt))<br />

2∆<br />

Since lim ∆→0<br />

sin(∆t)<br />

∆<br />

= t the above becomes<br />

This is the solution to use for resonance.<br />

u (t) = u (0) cos ωt + u′ (0)<br />

ω<br />

sin ωt − F k<br />

ωt<br />

2<br />

cos (ωt)<br />

37


8.3.2 Under-damped forced vibration c < c r , ξ < 1<br />

u<br />

K<br />

C<br />

M<br />

Fsin t<br />

mu ′′ + cu ′ + ku = F sin ϖt<br />

u ′′ + 2ξωu ′ + ω 2 u = F sin ϖt<br />

m<br />

The solution is<br />

where<br />

and<br />

u (t) = u h + u p<br />

u h (t) = e −ξωt (A cos ω d t + B sin ω d t)<br />

u p (t) =<br />

F<br />

√(k − mϖ) 2 + (cϖ) 2 sin (ϖt − θ)<br />

where<br />

tan θ =<br />

cϖ<br />

k − mϖ 2 =<br />

2ξr<br />

1 − r 2<br />

Very important note here in the calculations of tan θ above, one should be careful on the sign of the<br />

denominator. When the forcing frequency ϖ > ω the denominator will become negative (the case of ϖ = ω is<br />

resonance and is handled separately). Therefore, one should use arctan that takes care of which quadrant the<br />

angle is. For example, in Mathematica use<br />

ArcTan[1 - r^2, 2 Zeta r]]<br />

and in Matlab use<br />

atan2(2 Zeta r,1 - r^2)<br />

Otherwise, wrong solution will result when ϖ > ω The full solution is<br />

u (t) = e −ξωt (A cos ω d t + B sin ω d t) + F k<br />

1<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin (ϖt − θ) (1)<br />

Applying initial conditions gives<br />

A = u (0) + F k<br />

B = u′ (0)<br />

ω d<br />

+<br />

1<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin θ<br />

u (0) ξω<br />

ω d<br />

+ F k<br />

Another form of these equations is given as follows<br />

1<br />

ω d<br />

√<br />

(1 − r 2 ) 2 + (2ξr) 2 (ξω sin θ − ϖ cos θ)<br />

38


Hence the full solution is<br />

u p = p 0 1 ((<br />

1 − r<br />

2 )<br />

k (1 − r 2 ) 2 + (2ζr) 2 sin ϖt − 2ζr cos ϖt )<br />

u (t) = e −ξωnt (A cos ω d t + B sin ω d t) + F k<br />

Applying initial conditions now gives<br />

1<br />

(1 − r 2 ) 2 + (2ζr) 2 ((<br />

1 − r<br />

2 ) sin ϖt − 2ζr cos ϖt ) (1)<br />

A = u (0) +<br />

B = u′ (0)<br />

ω d<br />

2F rξ<br />

k<br />

1<br />

(1 − r 2 ) 2 + (2ξr) 2<br />

+ u (0) ξω n<br />

− F ( 1 − r 2) ϖ 2F rζ<br />

ω d kω d (1 − r 2 ) 2 +<br />

2<br />

+ (2ζr) kω d<br />

ω n<br />

(1 − r 2 ) 2 + (2ζr) 2<br />

The above 2 sets of equations are equivalent. One uses the phase angle explicitly and the second ones do not.<br />

Also, the above assume the force is F sin ϖt and not F cos ϖt. If the force is F cos ϖt then in Eq 1 above, the<br />

term reverse places as in<br />

u (t) = e −ξωnt (A cos ω d t + B sin ω d t) + F k<br />

Applying initial conditions now gives<br />

(<br />

1 − r<br />

2 )<br />

A = u (0) + F k (1 − r 2 ) 2 + (2ξr) 2<br />

B = u′ (0)<br />

ω d<br />

+ u (0) ξω n<br />

ω d<br />

+<br />

2F rζ<br />

kω d<br />

1 ((<br />

1 − r<br />

2 )<br />

(1 − r 2 ) 2 + (2ζr) 2 cos ϖt − 2ζr sin ϖt )<br />

ϖ<br />

(1 − r 2 ) 2 + (2ζr) 2 − F ( 1 − r 2)<br />

kω d<br />

ω n<br />

(1 − r 2 ) 2 + (2ζr) 2<br />

When a system is damped, the problem with the divide by zero when r = 1 does not occur here as was the<br />

case with undamped system, since when when ϖ ≈ ω or r = 1, the solution in Eq (1) becomes<br />

u (t) = e −ξωt ((<br />

u (0) + F k<br />

) (<br />

1<br />

u ′<br />

2ξ sin θ (0) u (0) ξω<br />

cos ω d t + + + F ω d ω d k<br />

) )<br />

1<br />

2ω d ξ (ξω sin θ − ϖ cos θ) sin ω d t<br />

+ F k<br />

1<br />

sin (ϖt − θ)<br />

2<br />

and the problem with the denominator going to zero does not show up here. The amplitude when steady state<br />

response is maximum can be found as follows. The amplitude of steady state motion is F 1<br />

k<br />

√(1−r . This<br />

2 ) 2 +(2ξr)<br />

√<br />

2<br />

1<br />

is maximum when the magnification factor β = √<br />

is maximum or when (1 − r 2 ) 2 + (2ξr) 2 or<br />

(1−r 2 ) 2 +(2ξr)<br />

√ 2<br />

(<br />

1 − ( ) )<br />

ϖ 2 2 ( )<br />

ω + 2ξ<br />

ϖ 2<br />

ω is minimum. Taking derivative w.r.t. ϖ and equating the result to zero and solving<br />

for ϖ gives<br />

ϖ = ω √ 1 − 2ξ 2<br />

We are looking for positive ϖ, hence when ϖ = ω √ 1 − 2ξ 2 the under-damped response is maximum.<br />

39


8.3.3 critically damping forced vibration ξ = c<br />

c r<br />

= 1<br />

The solution is<br />

Where u h = (A + Bt) e −ωt and u p = F k<br />

arctan definition). Hence<br />

u (t) = u h + u p<br />

1<br />

2r<br />

sin (ϖt − θ) where tan θ = (making sure to use correct<br />

√(1−r 2 ) 2 2 1−r<br />

+(2r) 2<br />

u (t) = (A + Bt) e −ωt + F k<br />

1<br />

√(1 − r 2 ) 2 + (2r) 2 sin (ϖt − θ)<br />

where A, B are found from initial conditions<br />

A = u (0) + F k<br />

B = u ′ (0) + u (0) ω + F k<br />

1<br />

√(1 − r 2 ) 2 + (2r) 2 sin θ<br />

1<br />

√(1 − r 2 ) 2 + (2r) 2 (ω sin θ − ϖ cos θ)<br />

8.3.4 over-damped forced vibration ξ = c<br />

c r<br />

> 1<br />

The solution is<br />

where<br />

and<br />

u p (t) = F k<br />

u (t) = u h + u p<br />

u h (t) = Ae p 1t + Be p 2t<br />

1<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin (ϖt − θ)<br />

hence<br />

u = Ae p 1t + Be p 2t + F k<br />

1<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin (ϖt − θ)<br />

where tan θ = 2ξr<br />

1−r 2<br />

Hence the solution is<br />

and<br />

p 1 = − c<br />

√ ( c<br />

2m + 2m<br />

p 2 = − c<br />

2m − √ ( c<br />

2m<br />

) 2 k −<br />

m = −ωξ + ω √<br />

n ξ 2 − 1<br />

) 2 k −<br />

m = −ωξ − ω √<br />

n ξ 2 − 1<br />

(<br />

−ξ+ √ (<br />

ξ<br />

u (t) = Ae<br />

2 −1<br />

)ωt −ξ− √ )<br />

ξ + Be 2 −1 ωt F + β sin (ϖt − θ)<br />

k<br />

u ′ (0) + u (0) ωξ + u (0) ω √ ξ 2 − 1 + F k<br />

((ξ β + √ )<br />

ξ 2 − 1<br />

A =<br />

2ω √ ξ 2 − 1<br />

u ′ (0) + u (0) ωξ − u (0) ω √ ξ 2 − 1 + F k<br />

((ξ β − √ )<br />

ξ 2 − 1<br />

B = −<br />

2ω √ ξ 2 − 1<br />

)<br />

ω sin θ − ϖ cos θ<br />

)<br />

ω sin θ − ϖ cos θ<br />

40


8.4 Response to impulsive loading<br />

8.4.1 impulse input<br />

Undamped system with impulse<br />

mü + ku = F 0 δ(t)<br />

with initial conditions u (0) = 0 and u ′ (0) = 0.Assuming the impulse acts for a very short time period from<br />

0 to t 1 seconds, where t 1 is small amount. Integrating the above differential equation gives<br />

∫ t1<br />

0<br />

müdt +<br />

∫ t1<br />

0<br />

kudt =<br />

∫ t1<br />

0<br />

F 0 δ(t)<br />

Since t 1 is very small, it can be assumed that u changes is negligible, hence the above reduces to<br />

since we assumed u ′ (0) = 0 and since ∫ t 1<br />

0<br />

∫ t1<br />

0<br />

∫ t1<br />

müdt =<br />

0<br />

( ) d ˙u<br />

m dt =<br />

dt<br />

∫ ˙u(t1 )<br />

˙u(0)<br />

∫ t1<br />

0<br />

∫ t1<br />

0<br />

d ˙u = F 0<br />

m<br />

˙u (t 1 ) − ˙u (0) = F 0<br />

m<br />

˙u (t 1 ) = F 0<br />

m<br />

F 0 δ(t)<br />

F 0 δ(t)<br />

∫ t1<br />

0<br />

∫ t1<br />

0<br />

∫ t1<br />

0<br />

δ(t)<br />

δ(t)<br />

δ(t)<br />

δ(t) = 1 then the above reduces to<br />

˙u (t 1 ) = F 0<br />

m<br />

Therefore, the effect of the impulse is the same as if the system was a free system but with initial velocity given<br />

by F 0<br />

m<br />

and zero initial position. Hence the system is now solved as follows<br />

With u (0) = 0 and u ′ (0) = F 0<br />

m<br />

. The solution is<br />

mü + ku = 0<br />

u impulse (t) = F 0<br />

sin ωt<br />

mω<br />

If the initial conditions were not zero, then the solution for these are added to the above. From earlier, it was<br />

found that the solution is u (t) = u(0) cos ωt + u′ (0)<br />

ω<br />

sin ωt, therefore, the full solution is<br />

u (t) =<br />

under-damped with impulse c < c r , ξ < 1<br />

due to IC only<br />

due to impulse<br />

{ }} { { }} {<br />

u(0) cos ωt + u′ (0)<br />

ω<br />

sin ωt + F 0 sin ωt<br />

mω<br />

mü + c ˙u + ku = δ(t)<br />

ü + 2ξω ˙u + ω 2 u = δ(t)<br />

with initial conditions u (0) = 0 and u ′ (0) = 0.Integrating gives<br />

∫ t1<br />

0<br />

müdt +<br />

∫ t1<br />

0<br />

c ˙udt +<br />

41<br />

∫ t1<br />

0<br />

kudt =<br />

∫ t1<br />

0<br />

F 0 δ(t)


Since t 1 is very small, it can be assumed that u changes is negligible as well as the change in velocity, hence the<br />

above reduces to the same result as in the case of undamped. Therefore, the system is solved as free system, but<br />

with initial velocity u ′ (0) = F 0 /m and zero initial position.<br />

Initial conditions are u (0) = 0 and u ′ (0) = 0 then the solution is<br />

applying initial conditions gives A = 0 and B =<br />

u impulse = e −ξωt (A cos ω d t + B sin ω d t)<br />

( F0<br />

)<br />

m<br />

ω d<br />

, hence<br />

u impulse (t) = e −ξωt (<br />

F0<br />

mω d<br />

sin ω d t<br />

If the initial conditions were not zero, then ( the solution for these are added ) to the above. From earlier, it<br />

was found that the solution is u (t) = e −ξωt u(0) cos ω d t + u′ (0)+u(0)ξω<br />

ω d<br />

sin ω d t , therefore, the full solution is<br />

due to IC only<br />

{ (<br />

}} ){<br />

{ ( }} ){<br />

u (t) = e −ξωt u(0) cos ω d t + u′ (0) + u(0)ξω<br />

sin ω d t + e −ξωt F0<br />

sin ω d t<br />

ω d<br />

mω d<br />

)<br />

due to impulse<br />

critically damped with impulse input ξ = c<br />

c r<br />

= 1 with initial conditions u (0) = 0 and u ′ (0) = 0 then the<br />

solution is<br />

u (t) = (A + Bt) e −( cr<br />

2m<br />

)<br />

t<br />

= (A + Bt) e −ωt<br />

where A, B are found from initial conditions A = u (0) = 0 and B = u ′ (0) + u (0) ω = F 0<br />

m<br />

, hence the solution<br />

is<br />

u impulse (t) = F 0t<br />

m e−ωt<br />

If the initial conditions were not zero, then the solution for these are added to the above. From earlier, it<br />

was found that the solution is u (t) = (u 0 (1 + ωt) + u ′ (0) t) e −ωt , therefore, the full solution is<br />

u (t) =<br />

due to IC only<br />

{ }} {<br />

(<br />

u (0) (1 + ωt) + u ′ (0) t ) e −ωt +<br />

due to impulse<br />

{ }} {<br />

F 0 t<br />

m e−ωt<br />

over-damped with impulse input ξ = c<br />

c r<br />

> 1 With initial conditions are u (0) = 0 and u ′ (0) = 0 the<br />

solution is<br />

u impulse (t) = Ae λ 1ωt + Be λ 2ωt<br />

where A, B are found from initial conditions and<br />

λ 1 = −ωξ + ω √ ξ 2 − 1<br />

λ 2 = −ωξ − ω √ ξ 2 − 1<br />

A = u′ (0) − u (0) λ 2<br />

2ω √ ξ 2 − 1<br />

B = −u′ (0) + u (0) λ 1<br />

2ω √ ξ 2 − 1<br />

Hence the solution is<br />

(<br />

−ξ+ √ (<br />

ξ<br />

u impulse (t) = Ae<br />

2 −1<br />

)ωt −ξ− √ )<br />

ξ + Be 2 −1 ωt<br />

42


where<br />

A =<br />

B =<br />

F 0<br />

m<br />

2ω √ ξ 2 − 1<br />

− F 0<br />

m<br />

2ω √ ξ 2 − 1<br />

Hence<br />

u impulse (t) =<br />

F 0<br />

(<br />

m<br />

2ω √ ξ 2 − 1 e −ξ+ √ )<br />

ξ 2 F 0<br />

(<br />

−1 ωt −<br />

m<br />

2ω √ ξ 2 − 1 e −ξ− √ )<br />

ξ 2 −1 ωt<br />

If the initial conditions were not zero, then the solution for these are added to the above. From earlier, it<br />

was found that the solution is u (t) = Ae p 1t + Be p 2t , therefore, the full solution is<br />

u (t) = Ae λ 1ωt + Be λ 2ωt +<br />

F 0<br />

F 0<br />

m<br />

2ω √ ξ 2 − 1 eλ 1ωt m<br />

−<br />

2ω √ ξ 2 − 1 eλ 2ωt<br />

8.4.2 sin impulse input<br />

Now assume the input is as follows<br />

A = u′ (0) − u (0) λ 2<br />

2ω √ ξ 2 − 1<br />

B = −u′ (0) + u (0) λ 1<br />

2ω √ ξ 2 − 1<br />

given by F (t) = F 0 sin (ϖt) where ϖ = 2π<br />

2t 1<br />

= π t 1<br />

undamped system with sin impulse<br />

⎧<br />

⎨ F 0 sin (ϖt) 0 ≤ t ≤ t 1<br />

mü + ku =<br />

⎩<br />

0 t > t 1<br />

with u (0) = u 0 and ˙u (0) = v 0 . For 0 ≤ t ≤ t 1 the solution is<br />

( )<br />

( )<br />

v0<br />

u (t) = u 0 cos ωt +<br />

ω − u r<br />

1 π<br />

st<br />

1 − r 2 sin ωt + u st<br />

1 − r 2 sin t<br />

t 1<br />

43


where r = ϖ ω = π/t 1<br />

ω<br />

= T<br />

2t 1<br />

where T is the natural period of the system. u st = F 0<br />

k<br />

, hence the above becomes<br />

⎛<br />

⎜<br />

u (t) = u 0 cos ωt + ⎝ v 0<br />

ω − F 0<br />

k<br />

When u 0 = 0 and v 0 = 0 then<br />

u (t) = − F 0<br />

k<br />

u (t) = F 0<br />

k<br />

( )<br />

π/t1<br />

ω<br />

1 −<br />

(<br />

π/t1<br />

ω<br />

1<br />

1 −<br />

(<br />

π/t1<br />

ω<br />

( )<br />

π/t1<br />

ω<br />

1 −<br />

(<br />

π/t1<br />

ω<br />

) 2<br />

sin ωt + F 0<br />

k<br />

) 2<br />

(sin<br />

) 2<br />

⎞<br />

⎟<br />

⎠ sin ωt + F 0<br />

k<br />

1<br />

1 −<br />

(<br />

π/t1<br />

ω<br />

1<br />

1 −<br />

(<br />

π/t1<br />

ω<br />

(<br />

π t<br />

t 1<br />

)<br />

− π/t 1<br />

ω sin ωt )<br />

(<br />

) 2<br />

sin π t )<br />

t 1<br />

The above Eq (1) gives solution during the time 0 ≤ t ≤ t 1<br />

Now after t = t 1 the force will disappear, the differential equation becomes<br />

mü + ku = 0 t > t 1<br />

(<br />

) 2<br />

sin π t )<br />

t 1<br />

but with the initial conditions evaluate at t = t 1 . From (1)<br />

( )<br />

v0<br />

u (t 1 ) = u 0 cos ωt 1 +<br />

ω − u r<br />

1<br />

st<br />

1 − r 2 sin ωt 1 + u st<br />

1 − r 2 sin ϖt 1<br />

( )<br />

v0<br />

= u 0 cos ωt 1 +<br />

ω − u r r<br />

st<br />

1 − r 2 u st<br />

1 − r 2 sin ωt 1 (2)<br />

since sin ϖt 1 = 0. taking derivative of Eq (1)<br />

˙u (t) = −ωu 0 sin ωt + ω<br />

and at t = t 1 the above becomes<br />

˙u (t 1 ) = −ωu 0 sin ωt 1 + ω<br />

= −ωu 0 sin ωt 1 + ω<br />

(<br />

v0<br />

)<br />

ω − u r<br />

st<br />

1 − r 2 cos ωt + ϖ 1 cos ϖt<br />

1 − r2 (<br />

v0<br />

)<br />

ω − u r<br />

st<br />

1 − r 2 )<br />

ω − u r<br />

st<br />

1 − r 2<br />

(<br />

v0<br />

cos ωt 1 + ϖ 1<br />

1 − r 2 cos ϖt 1<br />

(1)<br />

cos ωt 1 − ϖ 1<br />

1 − r 2 (3)<br />

since cos ϖt 1 = −1. Now (2) and (3) are used as initial conditions to solve mü + ku = 0 . The solution for<br />

t > t 1 is<br />

u (t) = u (t 1 ) cos ωt + ˙u (t 1)<br />

ω<br />

sin ωt<br />

Resonance with undamped sin impulse When ϖ ≈ ω and t ≤ t 1 we obtain resonance since r → 1 in<br />

the solution shown up and as written the solution can’t be used for analysis in this case. To obtain a solution<br />

for resonance some calculus is needed. Eq (1) is written as<br />

(<br />

)<br />

ϖ<br />

v 0<br />

u (t) = u 0 cos ωt +<br />

ω − u ω<br />

st<br />

1 − ( 1<br />

)<br />

ϖ 2<br />

sin ωt + u st<br />

ω<br />

1 − ( )<br />

ϖ 2<br />

sin ϖt<br />

ω<br />

( )<br />

v0<br />

= u 0 cos ωt +<br />

ω − u ωϖ<br />

ω 2<br />

st<br />

ω 2 − ϖ 2 sin ωt + u st<br />

ω 2 sin ϖt<br />

− ϖ2 (1A)<br />

Now looking at case when ϖ ≈ ω but less than ω, hence let<br />

ω − ϖ = 2∆ (2)<br />

44


where ∆ is very small positive quantity. and we also have<br />

Multiplying Eq (2) and (3) with each others gives<br />

Going back to Eq (1A) and rewriting it as<br />

ω + ϖ ≈ 2ϖ (3)<br />

ω 2 − ϖ 2 = 4∆ϖ (4)<br />

( v0<br />

u (t) = u 0 cos ωt +<br />

ω − u ωϖ<br />

)<br />

ω 2<br />

st sin ωt + u st sin ϖt<br />

4∆ϖ<br />

4∆ϖ<br />

( v0<br />

= u 0 cos ωt +<br />

ω − u ω<br />

)<br />

ω 2<br />

st sin ωt + u st sin ϖt<br />

4∆<br />

4∆ϖ<br />

Since ϖ ≈ ω the above becomes<br />

( v0<br />

u (t) = u 0 cos ωt +<br />

ω − u ω<br />

)<br />

ω<br />

st sin ωt + u st sin ϖt<br />

4∆<br />

4∆<br />

= u 0 cos ωt + v 0<br />

ω sin ωt + u ω<br />

st (sin ϖt − sin ωt)<br />

4∆<br />

now using sin ϖt − sin ωt = 2 sin ( ϖ−ω<br />

2<br />

t ) cos ( ϖ+ω<br />

2<br />

t ) the above becomes<br />

u (t) = u 0 cos ωt + v ( ( )<br />

0<br />

ω sin ωt + u ω ϖ − ω<br />

st sin t cos<br />

2∆ 2<br />

From Eq(2) ϖ − ω = −2∆ and ω + ϖ ≈ 2ϖ hence the above becomes<br />

or since ϖ ≈ ω<br />

Now lim ∆→0<br />

sin(∆t)<br />

∆<br />

This can also be written as<br />

( ϖ + ω<br />

u (t) = u 0 cos ωt + v 0<br />

ω sin ωt + u ω<br />

st (sin (−∆t) cos (ϖt))<br />

2∆<br />

u (t) = u 0 cos ωt + v 0<br />

ω sin ωt − u ω<br />

st (sin (∆t) cos (ωt))<br />

2∆<br />

= t hence the above becomes<br />

u (t) = u 0 cos ωt + v 0<br />

ω sin ωt − u ωt<br />

st cos (ωt)<br />

2<br />

2<br />

))<br />

t<br />

u (t) = u 0 cos ϖt + v 0<br />

ϖ sin ϖt − u ϖt<br />

st cos (ϖt) (1)<br />

( )<br />

2<br />

π<br />

= u 0 cos t + v ( ) ( ) ( )<br />

0 π π π<br />

t 1 ϖ sin t − u st t cos t<br />

t 1 2t 1 t 1<br />

since ϖ ≈ ω in this case. This is the solution to use for resonance and for t ≤ t 1<br />

Hence for t > t 1 , the above equations is used to determine initial conditions at t = t 1<br />

u (t 1 ) = u 0 cos ϖt 1 + v 0<br />

ϖ sin ϖt 1 − u st<br />

ϖt 1<br />

2 cos (ϖt 1)<br />

but cos ϖt 1 = cos π t 1<br />

t 1 = −1 and sin ϖt 1 = 0 and ϖt 1<br />

2<br />

= π 2<br />

, hence the above becomes<br />

Taking derivative of Eq (1) gives<br />

u (t 1 ) = −u 0 + u st<br />

π<br />

2<br />

ϖ 2 t<br />

˙u (t) = −ϖu 0 sin ϖt + v 0 cos ϖt + u st<br />

2 sin (ϖt) − u ϖ<br />

st cos (ϖt)<br />

2<br />

45


and at t = t 1<br />

ϖ 2 t 1<br />

ϖ<br />

˙u (t 1 ) = −ϖu 0 sin ϖt 1 + v 0 cos ϖt 1 + u st sin (ϖt 1 ) − u st<br />

2<br />

2 cos (ϖt 1)<br />

ϖ<br />

= −v 0 + u st<br />

2<br />

Now the solution for t > t 1 is<br />

u (t) = u (t 1 ) cos ωt + ˙u (t 1)<br />

ω<br />

(<br />

π<br />

= −u (0) + u st<br />

2<br />

under-damped with sin impulse c < c r , ξ < 1<br />

sin ωt<br />

)<br />

cos ωt + −u′ (0) + u st<br />

π<br />

2t 1<br />

ω<br />

⎧<br />

⎨ F 0 sin (ϖ) 0 ≤ t ≤ t 1<br />

mü + c ˙u + ku =<br />

⎩<br />

0 t > t 1<br />

sin ωt<br />

or<br />

⎧<br />

⎨<br />

ü + 2ξω ˙u + ω 2 F 0 sin (ϖ) 0 ≤ t ≤ t 1<br />

u =<br />

⎩<br />

0 t > t 1<br />

mü + c ˙u + ku = F sin ϖt<br />

ü + 2ξω ˙u + ω 2 u = F sin ϖt<br />

m<br />

For t ≤ t 1 Initial conditions are u (0) = u 0 and ˙u (0) = v 0 and u st = F k<br />

u (t) = e −ξωt (A cos ω d t + B sin ω d t) +<br />

then the solution from above is<br />

u st<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin (ϖt − θ) (1)<br />

Applying initial conditions gives<br />

For t > t 1 . From (1)<br />

A = u 0 +<br />

B = v 0<br />

ω d<br />

+ u 0ξω<br />

ω d<br />

+<br />

u st<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin θ<br />

u st<br />

u (t 1 ) = e −ξωt 1<br />

(A cos ω d t 1 + B sin ω d t 1 ) +<br />

ω d<br />

√<br />

(1 − r 2 ) 2 + (2ξr) 2 (ξω sin θ − ϖ cos θ)<br />

u st<br />

√(1 − r 2 ) 2 + (2ξr) 2 sin (ϖt 1 − θ) (2)<br />

Taking derivative of (1) gives<br />

˙u (t) = −ξωe −ξωt (A cos ω d t + B sin ω d t) + e −ξωt (−Aω d sin ω d t + ω d B cos ω d t)<br />

u st<br />

+ ϖ<br />

cos (ϖt − θ)<br />

√(1 − r 2 ) 2 + (2ξr) 2<br />

46


at t = t 1<br />

˙u (t 1 ) = −ξωe −ξωt 1<br />

(A cos ω d t 1 + B sin ω d t 1 ) + e −ξωt 1<br />

(−Aω d sin ω d t 1 + ω d B cos ω d t 1 )<br />

u st<br />

+ ϖ<br />

cos (ϖt 1 − θ) (3)<br />

√(1 − r 2 ) 2 + (2ξr) 2<br />

Now for t > t 1 the equation becomes<br />

which has the solution<br />

where A = u (t 1 ) and B = ˙u(t 1)+u(t 1 )ξω<br />

ω d<br />

mü + c ˙u + ku = 0<br />

u = e −ξωt (A cos ω d t + B sin ω d t)<br />

critically damped with sin impulse ξ = c<br />

c r<br />

= 1 For t ≤ t 1 Initial conditions are u (0) = u 0 and ˙u (0) = v 0<br />

then the solution is from above<br />

u (t) = (A + Bt) e −ωt u st<br />

+<br />

sin (ϖt − θ) (1)<br />

√(1 − r 2 ) 2 + (2r) 2<br />

Where tan θ =<br />

cϖ<br />

k−mϖ 2<br />

= 2ξr<br />

1−r 2 . A, B are found from initial conditions<br />

A = u 0 +<br />

u st<br />

√(1 − r 2 ) 2 + (2r) 2 sin θ<br />

B = v 0 + u 0 ω +<br />

u st<br />

√(1 − r 2 ) 2 + (2r) 2 (ω sin θ − ϖ cos θ)<br />

For t > t 1 the solution is<br />

To find u (t 1 ) , from Eq(1)<br />

u (t) = (u (t 1 ) + ( ˙u (t 1 ) + u (t 1 ) ω) t) e −ωt (2)<br />

u (t 1 ) = (A + Bt) e −ωt 1<br />

+<br />

u st<br />

√(1 − r 2 ) 2 + (2r) 2 sin (ϖt 1 − θ)<br />

taking derivative of (1) gives<br />

˙u (t) = −ω (A + Bt) e −ωt + Be −ωt + ϖ<br />

sin (ϖt − θ) (3)<br />

√(1 − r 2 ) 2 + (2r) 2<br />

u st<br />

at t = t 1<br />

˙u (t 1 ) = −ω (A + Bt 1 ) e −ωt 1<br />

+ Be −ωt 1<br />

+ ϖ<br />

sin (ϖt 1 − θ) (4)<br />

√(1 − r 2 ) 2 + (2r) 2<br />

u st<br />

Hence Eq (2) can now be evaluated using Eq(3,4)<br />

over-damped with sin impulse ξ = c<br />

c r<br />

> 1 For t ≤ t 1 Initial conditions are u (0) = u 0 and ˙u (0) = v 0 then<br />

the solution is<br />

u = Ae p1t + Be p2t u st<br />

+<br />

sin (ϖt − θ)<br />

√(1 − r 2 ) 2 + (2ξr) 2<br />

where tan θ = 2ξr<br />

1−r 2 (make sure you use correct quadrant, see not above on arctan) and<br />

p 1 = − c<br />

√ ( c<br />

) 2<br />

2m + k −<br />

2m m<br />

= −ωξ + ω √ ξ 2 − 1<br />

47


and<br />

leading to the solution where tan θ = 2ξr<br />

1−r 2<br />

p 2 = − c<br />

√ ( c<br />

) 2<br />

2m − k −<br />

2m m<br />

= −ωξ − ω √ ξ 2 − 1<br />

and<br />

is<br />

p 1 = − c<br />

√ ( c<br />

2m + 2m<br />

p 2 = − c<br />

2m − √ ( c<br />

2m<br />

) 2 k −<br />

m = −ωξ + ω √<br />

n ξ 2 − 1<br />

) 2 k −<br />

m = −ωξ − ω √<br />

n ξ 2 − 1<br />

(<br />

−ξ+ √ (<br />

ξ<br />

u (t) = Ae<br />

2 −1<br />

)ωt −ξ− √ )<br />

ξ + Be 2 −1 ωt F + β sin (ϖt − θ)<br />

k<br />

u ′ (0) + u (0) ωξ + u (0) ω √ ξ 2 − 1 + F k<br />

((ξ β + √ )<br />

ξ 2 − 1<br />

A =<br />

2ω √ ξ 2 − 1<br />

u ′ (0) + u (0) ωξ − u (0) ω √ ξ 2 − 1 + F k<br />

((ξ β − √ )<br />

ξ 2 − 1<br />

B = −<br />

2ω √ ξ 2 − 1<br />

1<br />

β = √<br />

(1 − r 2 ) 2 + (2ξr) 2<br />

)<br />

ω sin θ − ϖ cos θ<br />

)<br />

ω sin θ − ϖ cos θ<br />

For t > t 1 . From Eq(1) and at t = t 1<br />

Taking derivative of Eq (1)<br />

(<br />

−ξ+ √ )<br />

(<br />

ξ<br />

u (t 1 ) = Ae<br />

2 −1 ωt 1 −ξ− √ )<br />

ξ<br />

+ Be<br />

2 −1 ωt 1<br />

+ D sin (ϖt 1 − θ) (2)<br />

(<br />

−ξ+ √ (<br />

ξ<br />

˙u (t) = ωAe<br />

2 −1<br />

)ωt −ξ− √ )<br />

ξ + ωBe 2 −1 ωt + ϖD cos (ϖt − θ)<br />

At t = t 1<br />

−ξ+<br />

˙u (t 1 ) = ωAe( √ )<br />

(<br />

ξ 2 −1 ωt 1 −ξ− √ )<br />

ξ<br />

+ ωBe<br />

2 −1 ωt 1<br />

+ ϖD cos (ϖt 1 − θ) (3)<br />

Equation of motion now is<br />

which has solution for over-damped given by<br />

ü + 2ξω ˙u + ω 2 u = 0<br />

(<br />

−ξ+ √ (<br />

ξ<br />

u (t) = Ae<br />

2 −1<br />

)ω nt −ξ− √ )<br />

ξ + Be 2 −1 ω nt<br />

where<br />

˙u (t 1 ) + u (t 1 ) ω n<br />

(ξ − √ )<br />

ξ 2 − 1<br />

A = −<br />

√<br />

2ω n ξ 2 − 1<br />

˙u (t 1 ) + u (t 1 ) ξω n<br />

(ξ + √ )<br />

ξ 2 − 1<br />

B =<br />

√<br />

2ω n ξ 2 − 1<br />

48


8.5 references<br />

1. Vibration analysis by Robert K. Vierck<br />

2. Structural dynamics theory and computation, 5th edition by Mario Paz, William Leigh<br />

3. Dynamic of structures, Ray W. Clough and Joseph Penzien<br />

4. Theory of vibration,volume 1, by A.A.Shabana<br />

5. Notes on Diffy Qs, Differential equations for engineers, by Jiri Lebl, online PDF book, chapter 2.6, oct<br />

1,2012 http://www.jirka.org/diffyqs/<br />

9 Derivation of rotation formula<br />

This formula is very important. Will show its derivation now in details. It is how to express vectors in rotating<br />

frames.<br />

Consider this diagram<br />

P<br />

r<br />

Y<br />

X<br />

r p<br />

y<br />

r o<br />

o<br />

<br />

x<br />

Moving frame of<br />

reference, attached<br />

to body of interest<br />

Absolute (or inertial frame of reference)<br />

In the above, the small axis x, y is a frame attached to some body which rotate around this axis with angular<br />

velocity ω (measured by the inertial frame of course). All laws derived below are based on the following one rule<br />

d<br />

dt r ∣<br />

∣∣∣absolute<br />

= d dt r ∣<br />

∣∣∣relative<br />

+ ω × r (1)<br />

Lets us see how to apply this rule. Let us express the position vector of the particle r p . We can see by<br />

normal vector additions that the position vector of particle is<br />

r p = r o + r (2)<br />

Notice that nothing special is needed here, since we have not yet looked at rate of change with time. The<br />

complexity (i.e. using rule (1)) appears only when we want to look at velocities and accelerations. This is when<br />

we need to use the above rule (1). Let us now find the velocity of the particle. From above<br />

ṙ p = ṙ o + ṙ<br />

49


Every time we take derivatives, we stop and look. For any vector that originates from the moving frame,<br />

we must apply rule (1) to it. That is all. In the above, only r needs rule (1) applied to it, since that is the<br />

only vector measure from the moving frame. Replacing ṙ p by V p and ṙ o by V o , meaning the velocity of P and o,<br />

Hence the above becomes<br />

V p = V o + ṙ<br />

and now we apply rule (1) to expand ṙ<br />

V p = V o + (V rel + ω × r) (3)<br />

where V rel is just d dt r∣ ∣<br />

relative<br />

The above is the final expression for the velocity of the particle V p using its velocity as measured by the<br />

moving frame in order to complete the expression.<br />

So the above says that the absolute velocity of the particle is equal to the absolute velocity of the base of the<br />

moving frame + something else and this something else was (V rel + ω × r)<br />

Now we will find the absolute acceleration of P . Taking time derivatives of (3) gives<br />

˙V p = ˙V<br />

( )<br />

o + ˙Vrel + ˙ω × r + ω × ṙ<br />

(4)<br />

As we said above, each time we take time derivatives, we stop and look for vectors which are based on the<br />

moving frame, and apply rule (1) to them. In the above, ˙Vrel and ṙ qualify. Apply rule (1) to ˙V rel gives<br />

˙V rel = a rel + ω × V rel (5)<br />

where a rel just means the acceleration relative to moving frame. And applying rule (1) to ṙ gives<br />

Replacing (5) and (6) into (4) gives<br />

ṙ = V rel + ω × r (6)<br />

a p = a o + (a rel + ω × V rel + ˙ω × r + ω × (V rel + ω × r))<br />

= a o + a rel + (ω × V rel ) + ( ˙ω × r) + (ω × V rel ) + (ω × (ω × r))<br />

= a o + a rel + 2 (ω × V rel ) + ( ˙ω × r) + (ω × (ω × r)) (7)<br />

Eq (7) says that the absolute acceleration a p of P is the sum of the acceleration of the base a o of the<br />

moving frame plus the relative acceleration a rel of the particle to the moving frame plus 2 (ω × V rel ) + ( ˙ω × r) +<br />

(ω × (ω × r))<br />

Hence, using Eq(3) and Eq(7) gives us the expressions we wanted for velocity and acceleration.<br />

10 Miscellaneous hints<br />

1. When finding the generalized force for the user with the Lagrangian method (the hardest step), using the<br />

virtual work method, if the force (or virtual work by the force) ADDS energy to the system, then make<br />

the sign of the force positive otherwise the sign is negative.<br />

2. For damping force, the sign is always negative.<br />

3. External forces such as linear forces applied, torque applied, in general, are positive.<br />

4. Friction force is negative (in general) as friction takes energy from the system like damping.<br />

50


11 Formulas<br />

sin 2 x 1cos2x<br />

2<br />

cos 2 x 1cos2x<br />

2<br />

sin 2x 2 sin x cosx<br />

cos2x cos 2 x sin 2 x<br />

1 2 sin 2 x<br />

V p V o r V rel<br />

Velocity of p<br />

relative to o<br />

a p a 0 r r 2 V rel a rel<br />

p<br />

IF rigid body is rotating AND translation at the same time THEN<br />

Take moments around its cg only<br />

ELSE<br />

can also take moments about any other points on it (but<br />

change I)<br />

END IF<br />

o<br />

r<br />

<br />

Y<br />

R Y<br />

X<br />

R y<br />

R X<br />

R<br />

x<br />

y<br />

<br />

R x<br />

BODY FIXED<br />

COORDINATES<br />

This gives the relation<br />

between the rate of<br />

change with respect to<br />

time of a vector expressed<br />

in a frame of reference<br />

which is body fixed (y,x)<br />

here, to the rate of change<br />

with respect to time of the<br />

same vector expressed<br />

using an inertial frame<br />

coordinates XY.<br />

The omega here is the<br />

angular velocity of the<br />

rigid body rotation, or the<br />

body fixed coordinates,<br />

with respect to the inertial<br />

frame.<br />

d<br />

dt R X,Y d dt R x,y R X,Y<br />

<br />

R<br />

r<br />

<br />

R r<br />

r<br />

The small disk rotates<br />

at angular speed of <br />

The condition of no slip can<br />

be seen to be<br />

R r r<br />

For the sign of generalized force:<br />

If work done by force takes away<br />

energy from system, then the<br />

sign is negative, else positive. So<br />

Friction will always have negative<br />

sign, so will damping force.<br />

d<br />

dt<br />

cosxt sinxtx t<br />

d<br />

dt<br />

df<br />

dxt<br />

<br />

d<br />

dxt<br />

d<br />

cosxt <br />

dt<br />

df<br />

dxt<br />

d<br />

d<br />

cos xt<br />

dxt<br />

d<br />

xt dt<br />

xt dt<br />

51


F ma<br />

linear momentum p mv<br />

F d p dt<br />

particle kinetic energy T 1 2 mv2<br />

I<br />

angular momentum H I cg <br />

d<br />

dt H<br />

rigid body T 1 Mv 2 cg<br />

2 1 I 2<br />

2<br />

cg<br />

my 2 n y n 2 y fy,t<br />

y cy k m y fy,t<br />

n <br />

For small<br />

epsilon<br />

1<br />

1<br />

1<br />

1<br />

1 <br />

1 <br />

k<br />

m ,c 2 n<br />

m<br />

conservation of angular momentum d dt<br />

H constant<br />

x n2 x 0<br />

If n 0 then sinusodial solution, ok<br />

if n 0 then solution blows up, exponential<br />

52


12 Velocity and acceleration diagrams<br />

12.1 Spring pendulum<br />

<br />

Length of pendulum fixed<br />

Length of pendulum changes with time<br />

2L <br />

L<br />

L<br />

L<br />

Velocity diagram<br />

acceleration<br />

diagram<br />

L 2 L Lt L 2<br />

<br />

L<br />

L<br />

Velocity diagram<br />

acceleration<br />

diagram<br />

y<br />

Length of pendulum changes with time and base of pendulum moves<br />

ÿ<br />

x<br />

x<br />

y<br />

L<br />

x<br />

L 2<br />

ÿ<br />

L<br />

x<br />

2L <br />

Velocity diagram<br />

<br />

L<br />

acceleration<br />

diagram<br />

r<br />

Nasser M. Abbasi<br />

Oct 10, 2013 (drawing2.vsd)<br />

53


12.2 pendulum with blob moving in slot<br />

<br />

r<br />

<br />

y<br />

y<br />

r<br />

y<br />

System. Pendulum with<br />

blob which moves in slot<br />

perpendicular to axis<br />

Velocity diagram<br />

2y <br />

r 2<br />

<br />

y<br />

ÿ<br />

r<br />

acceleration diagram<br />

y 2<br />

54


12.3 spring pendulum with block moving in slot<br />

<br />

r<br />

<br />

y<br />

y<br />

r<br />

System. Pendulum with blob which<br />

moves in slot perpendicular to axis. In<br />

addition, length of the axis of pendulum<br />

itself changes in time. y and r are<br />

measured from static equilibrium of<br />

springs<br />

y<br />

Velocity diagram<br />

r<br />

2y <br />

r 2<br />

2r<br />

<br />

y<br />

ÿ<br />

r<br />

acceleration diagram<br />

y 2<br />

r<br />

55


12.4 double pendulum<br />

<br />

<br />

Double<br />

pendulum<br />

L 1<br />

L 2<br />

<br />

<br />

<br />

m 1<br />

m 1<br />

L 1<br />

• 2<br />

m 2<br />

m 2<br />

L 1<br />

<br />

<br />

L 1<br />

L 2<br />

m 2<br />

••<br />

acceleration diagram<br />

L 1<br />

<br />

<br />

m 1<br />

L 2<br />

L 2<br />

• 2<br />

L 1<br />

• 2<br />

<br />

L 2<br />

••<br />

<br />

L 1<br />

•<br />

L 1<br />

••<br />

Velocity diagram<br />

L 2<br />

•<br />

<br />

<br />

L 1<br />

•<br />

56


13 Velocity and acceleration of rigid body 2D<br />

<br />

V<br />

y<br />

x<br />

U<br />

V U<br />

y<br />

x<br />

U V<br />

Y<br />

X<br />

Linear<br />

Velocity<br />

diagram<br />

Rigid body,<br />

rotating at<br />

angular <br />

speed<br />

Linear<br />

acceleration<br />

diagram<br />

Notice the<br />

sign<br />

difference<br />

Nasser M. Abbasi<br />

Drawing_rigid_body_rotati<br />

on_1.vsd<br />

May 23, 2011<br />

a <br />

a x<br />

a y<br />

<br />

U V<br />

V U<br />

Finding linear acceleration of center of mass of a rigid body under pure rotation using fixed body coordinates.<br />

In the above U is the speed of the center of mass in the direction of the x axis, where this axis is fixed on<br />

the body itself. Similarly, V is the speed of the center of mass in the direction of the y axis, where the y axis is<br />

attached to the body itself.<br />

Just remember that all these speeds (i.e. U,V ) and accelerations (a x , a y ) are still being measured by an<br />

observer in the inertial frame. It is only that the directions of the velocity components of the center of mass<br />

is along an axis fixed on the body. Only the direction. But actual speed measurements are still done by a<br />

stationary observer. Since clearly if the observer was sitting on the body itself, then they will measure the speeds<br />

to be zero in that case.<br />

57


14 Velocity and acceleration of rigid body 3D<br />

14.1 Using Vehicle dynamics notations<br />

F x<br />

Nasser M. Abbasi<br />

3d_1.vsd<br />

May 26, 2011<br />

L<br />

I <br />

N<br />

F z<br />

W<br />

z<br />

r<br />

Linear force<br />

torque<br />

Linear velocity<br />

angular velocity<br />

v <br />

y<br />

<br />

p q V<br />

x<br />

M<br />

F y<br />

U<br />

F <br />

Forces, Torques, linear<br />

velocities and angular velocities<br />

for a general 3D rigid body<br />

I xx I xy I xz<br />

<br />

I yx I yy I yz<br />

linear momentum<br />

U qW rV<br />

F d dt p d dt mvm d dt v m a<br />

V rU pW<br />

W pV qU<br />

U<br />

V<br />

W<br />

p<br />

q<br />

r<br />

F x<br />

F y<br />

F z<br />

L<br />

M<br />

N<br />

d dt H d dt I<br />

Angular momentum<br />

Derivation of this is much more complicated than with<br />

the case of linear motion (F=ma), since m is scalar<br />

there, but for rotation, I is matrix. See next page for the<br />

derivation<br />

58


14.2 3D Not Using vehicle dynamics notations<br />

Nasser M. Abbasi<br />

3d_1.vsd<br />

May 26, 2011<br />

z<br />

F z<br />

Linear force<br />

torque<br />

Linear velocity<br />

v <br />

v x<br />

v y<br />

v z<br />

angular velocity<br />

v z<br />

F x<br />

x<br />

I <br />

v x<br />

x<br />

x<br />

z<br />

z<br />

y<br />

y<br />

F y<br />

<br />

F <br />

Forces, Torques, linear<br />

velocities and angular<br />

velocities for a general 3D<br />

rigid body<br />

I xx I xy I xz<br />

<br />

I yx I yy I yz<br />

F d p dt d mvm d v m dt dt<br />

a y z v x x v z<br />

a<br />

a x y v z z v y<br />

v y<br />

y<br />

x<br />

y<br />

z<br />

F x<br />

F y<br />

F z<br />

x<br />

y<br />

z<br />

linear momentum<br />

a z x v y y v z<br />

The derivation of the above is given next, but it uses the standard formula given by<br />

d<br />

A d<br />

A A<br />

dt dt resolved<br />

This is in the<br />

inertial frame of<br />

reference<br />

This is the same A, but its components<br />

are with respect to the body fixed<br />

coordinates system,<br />

Cross product<br />

59


14.2.1 Derivation for F = ma in 3D<br />

F = d dt p<br />

= d dt (mv)<br />

= m d dt v<br />

⎡⎛<br />

⎞ ⎛ ⎞ ⎛ ⎞⎤<br />

a x<br />

ω x<br />

v x<br />

= m<br />

⎢⎜a y ⎟<br />

⎣⎝<br />

⎠ + ⎜ω y ⎟<br />

⎝ ⎠ ⊗ ⎜v y ⎟⎥<br />

⎝ ⎠⎦<br />

a z ω z v z<br />

⎡⎛<br />

⎞ ∣ ∣⎤<br />

∣∣∣∣∣∣∣∣∣ ∣∣∣∣∣∣∣∣∣ a x<br />

i j k<br />

= m<br />

⎢⎜a y ⎟<br />

⎣⎝<br />

⎠ + det ω x ω y ω z ⎥<br />

⎦<br />

a z v x v y v z<br />

⎡⎛<br />

⎞ ⎛<br />

⎞⎤<br />

a x<br />

ω y v z − ω z v y<br />

= m<br />

⎢⎜a y ⎟<br />

⎣⎝<br />

⎠ + ⎜− (ω x v z − ω z v x )<br />

⎟⎥<br />

⎝<br />

⎠⎦<br />

a z ω x v y − ω y v x<br />

⎛<br />

⎞<br />

a x + ω y v z − ω z v y<br />

= m<br />

⎜a y − ω x v z + ω z v x ⎟<br />

⎝<br />

⎠<br />

a z + ω x v y − ω y v x<br />

60


14.2.2 Derivation for τ = Iω in 3D<br />

Let A = Iω then using the rule<br />

( ) d<br />

τ =<br />

dt A ( ) d<br />

=<br />

dt A + ω × A<br />

resolved<br />

Then τ = Iω can be found for the general case<br />

⎡<br />

⎤<br />

A<br />

A<br />

{ ⎛ }} ⎞ ⎛ ⎞{<br />

⎛ ⎞ { ⎛ }} ⎞ ⎛ ⎞{<br />

τ = d I xx I xy I xz<br />

ω x<br />

ω x<br />

I xx I xy I xz<br />

ω x<br />

dt<br />

⎜I yx I yy I yz ⎟ ⎜ω y ⎟<br />

+<br />

⎝<br />

⎠ ⎝ ⎠<br />

⎜ω y ⎟<br />

⎝ ⎠ ×<br />

⎜I yx I yy I yz ⎟ ⎜ω y ⎟<br />

⎝<br />

⎠ ⎝ ⎠<br />

⎢<br />

⎣ I zx I yz I zz ω z<br />

⎥<br />

⎦ ω z I zx I yz I zz ω z<br />

⎛<br />

⎞ ⎛ ⎞ ⎛ ⎞ ⎛<br />

⎞<br />

I xx I xy I xz<br />

α x<br />

ω x<br />

I xx ω x + I xy ω y + I xz ω z<br />

=<br />

⎜I yx I yy I yz ⎟ ⎜α y ⎟<br />

⎝<br />

⎠ ⎝ ⎠ + ⎜ω y ⎟<br />

⎝ ⎠ × ⎜I yx ω x + I yy ω y + I yz ω z ⎟<br />

⎝<br />

⎠<br />

I zx I yz I zz α z ω z I zx ω x + I yz ω y + I zz ω z<br />

⎛<br />

⎞ ⎛ ⎞ ∣ ∣∣∣∣∣∣∣∣∣ I xx I xy I xz<br />

α x<br />

i j k<br />

=<br />

⎜I yx I yy I yz ⎟ ⎜α y ⎟<br />

⎝<br />

⎠ ⎝ ⎠ + det ω x ω y ω z<br />

I zx I yz I zz α z (I xx ω x + I xy ω y + I xz ω z ) (I yx ω x + I yy ω y + I yz ω z ) (I zx ω x + I yz ω y + I zz ω z ) ∣<br />

⎛<br />

⎞ ⎛ ⎞ ⎛<br />

⎞<br />

I xx I xy I xz<br />

α x<br />

ω y (I zx ω x + I yz ω y + I zz ω z ) − ω z (I yx ω x + I yy ω y + I yz ω z )<br />

=<br />

⎜I yx I yy I yz ⎟ ⎜α y ⎟<br />

⎝<br />

⎠ ⎝ ⎠ + ⎜ω x (I zx ω x + I yz ω y + I zz ω z ) − ω z (I xx ω x + I xy ω y + I xz ω z )<br />

⎟<br />

⎝<br />

⎠<br />

I zx I yz I zz α z ω x (I yx ω x + I yy ω y + I yz ω z ) − ω y (I xx ω x + I xy ω y + I xz ω z )<br />

61


14.2.3 Derivation for τ = Iω in 3D using principle axes<br />

The above derivation simplifies now since we will be using principle axes. In this case, all cross products of<br />

moments of inertia vanish.<br />

⎛<br />

⎞<br />

I xx 0 0<br />

I =<br />

⎜ 0 I yy 0<br />

⎟<br />

⎝<br />

⎠<br />

0 0 I zz<br />

Hence<br />

⎡<br />

⎤<br />

A<br />

A<br />

{ ⎛ }} ⎞ ⎛ ⎞{<br />

⎛ ⎞ { ⎛ }} ⎞ ⎛ ⎞{<br />

τ = d I xx 0 0<br />

ω x<br />

ω x<br />

I xx 0 0<br />

ω x<br />

dt<br />

⎜ 0 I yy 0<br />

⎟ ⎜ω y ⎟<br />

+<br />

⎝<br />

⎠ ⎝ ⎠<br />

⎜ω y ⎟<br />

⎝ ⎠ ×<br />

⎜ 0 I yy 0<br />

⎟ ⎜ω y ⎟<br />

⎝<br />

⎠ ⎝ ⎠<br />

⎢<br />

⎣ 0 0 I zz ω z<br />

⎥<br />

⎦ ω z 0 0 I zz ω z<br />

⎛<br />

⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞<br />

I xx 0 0<br />

α x<br />

ω x<br />

I xx ω x<br />

=<br />

⎜ 0 I yy 0<br />

⎟ ⎜α y ⎟<br />

⎝<br />

⎠ ⎝ ⎠ + ⎜ω y ⎟<br />

⎝ ⎠ × ⎜I yy ω y ⎟<br />

⎝ ⎠<br />

0 0 I zz α z ω z I zz ω z<br />

⎛ ⎞ ∣ ∣ ∣∣∣∣∣∣∣∣∣ ∣∣∣∣∣∣∣∣∣<br />

I xx α x<br />

i j k<br />

=<br />

⎜I yy α y ⎟<br />

⎝ ⎠ + det ω x ω y ω z<br />

I zz α z I xx ω x I yy ω y I zz ω z<br />

⎛ ⎞ ⎛<br />

⎞<br />

I xx α x<br />

ω y (I zz ω z ) − ω z (I yy ω y )<br />

=<br />

⎜I yy α y ⎟<br />

⎝ ⎠ + ⎜−ω x (I zz ω z ) + ω z (I xx ω x )<br />

⎟<br />

⎝<br />

⎠<br />

I zz α z ω x (I yy ω y ) − ω y (I xx ω x )<br />

⎛ ⎞ ⎛<br />

⎞<br />

I xx α x<br />

ω y ω z (I zz − I yy )<br />

=<br />

⎜I yy α y ⎟<br />

⎝ ⎠ + ⎜ω x ω z (I xx − I zz )<br />

⎟<br />

⎝<br />

⎠<br />

I zz α z ω x ω y (I yy − I xx )<br />

So, we can see how much simpler it became when using principle axes. Compare the above to<br />

⎛<br />

⎞ ⎛ ⎞ ⎛<br />

⎞<br />

I xx I xy I xz<br />

α x<br />

ω y (I zx ω x + I yz ω y + I zz ω z ) − ω z (I yx ω x + I yy ω y + I yz ω z )<br />

⎜I yx I yy I yz ⎟ ⎜α y ⎟<br />

⎝<br />

⎠ ⎝ ⎠ + ⎜ω x (I zx ω x + I yz ω y + I zz ω z ) − ω z (I xx ω x + I xy ω y + I xz ω z )<br />

⎟<br />

⎝<br />

⎠<br />

I zx I yz I zz α z ω x (I yx ω x + I yy ω y + I yz ω z ) − ω y (I xx ω x + I xy ω y + I xz ω z )<br />

So, always use principle axes for the body fixed coordinates system!<br />

62


15 Wheel spinning precession<br />

x<br />

MgL<br />

L<br />

y<br />

z<br />

Mg<br />

Wheel spinning around y-axis,<br />

hanging from ceiling<br />

Weight create a torque which<br />

changes the angular<br />

momentum<br />

Torque vector<br />

representation<br />

y<br />

<br />

Important: This<br />

analysis is valid<br />

only for LARGE<br />

y<br />

H Angular momentum<br />

x<br />

z<br />

y<br />

H y I yy y<br />

p<br />

y<br />

t<br />

Angular momentum<br />

change due to the<br />

torque applied<br />

After torque applied for t<br />

z<br />

This is what wheel will<br />

look like after del T<br />

time due to<br />

precesses, notice that<br />

SPIN angular<br />

momentum changes<br />

in the direction of the<br />

torque<br />

y<br />

Time it takes<br />

for the wheel<br />

to precesses<br />

one full cycle<br />

T p 2<br />

p<br />

H<br />

z<br />

Angular momentum<br />

after<br />

t<br />

t<br />

H aftert<br />

x<br />

H I yy y t<br />

x<br />

I yy y<br />

p<br />

y<br />

H t H<br />

t<br />

<br />

But<br />

H<br />

t<br />

I yy y p , hence I yy y p hence p <br />

<br />

I yy y<br />

MgL<br />

I yy y<br />

Nasser M.<br />

Abbasi<br />

Precesses.vsd<br />

6/1/2011<br />

Therefore, precession velocity p is MgL<br />

I yy y<br />

16 References<br />

1. Structural <strong>Dynamics</strong> 5th edition. Mario Paz, William Leigh<br />

63


17 Misc. items<br />

The Jacobian matrix for a system of differential equations, such as<br />

is given by<br />

x ′ (t) = f (x, y, z)<br />

y ′ (t) = g (x, y, z)<br />

z ′ (t) = h (x, y, z)<br />

⎛<br />

J =<br />

⎜<br />

⎝<br />

For example, for the given the following 3 set of coupled differential equations in n 3<br />

df<br />

dx<br />

dg<br />

dx<br />

dh<br />

dx<br />

df<br />

dy<br />

dg<br />

dy<br />

dh<br />

dy<br />

df<br />

dz<br />

dg<br />

dz<br />

dh<br />

dz<br />

x ′ (t) = −y (t) − z (t)<br />

y ′ (t) = x (t) + ay (t)<br />

⎞<br />

⎟<br />

⎠<br />

z ′ (t) = b + z (t) (x (t) − c)<br />

then the Jacobian matrix is<br />

⎛<br />

⎞<br />

0 −1 −1<br />

J =<br />

⎜ 1 a 0<br />

⎟<br />

⎝<br />

⎠<br />

z (t) 0 x (t) − c<br />

Now to find stability of this system, we evaluate this matrix at t = t 0 where x (t 0 ) , y (t 0 ) , z (t 0 ) is a point in<br />

this space (may be stable point or initial conditions, etc...) and then J become all numerical now. Then we can<br />

evaluate the eigenvalues of the resulting matrix and look to see if all eigenvalues are negative. If so, this tells us<br />

that the point is a stable point. I.e. the system is stable.<br />

If X is N(0, 1) distributed then mu + sigma ∗ X is N(mu, sigma 2 ) distributed.<br />

64


18 Astrodynamics<br />

18.1 Ellipse main parameters<br />

18.2 Table of common equations<br />

The following table contains the common relations to use for elliptic motion. Equation of ellipse is x2<br />

a 2 + y2<br />

b 2 = 1<br />

term to find<br />

conversion between E and θ<br />

position of satellite at time t<br />

Solve for E, then find θ. τ<br />

here is time at perigee and n<br />

is mean satellite speed.<br />

eccentricity e<br />

relation<br />

( θ<br />

tan =<br />

2)<br />

√ ( )<br />

1 + e E<br />

1 − e tan 2<br />

cos E = e + cos θ<br />

1 + e cos θ<br />

cos θ = e − cos E<br />

e cos E − 1<br />

E − e sin E = n (t − τ)<br />

√<br />

e = c a = ra−rp<br />

r a+r p<br />

= 1 + 2Eh2<br />

µ 2<br />

65


Maor axes a<br />

a = r p(1 + e)<br />

1 − e 2<br />

= r a(1 − e)<br />

1 − e 2<br />

= − µ<br />

2E<br />

= √ b 2 + c 2<br />

= p<br />

1 − e 2<br />

Minor axes b b = a √ 1 − e 2<br />

r p = a(1 − e2 )<br />

r p 1 + e<br />

= a(1 − e)<br />

r a = a(1 − e2 )<br />

1 − e<br />

r a = a (1 + e)<br />

p<br />

specific angular momentum h<br />

Total Energy E<br />

= h2<br />

µ<br />

1<br />

1 − e<br />

r p<br />

r a<br />

= 1−e<br />

1+e<br />

p = a ( 1 − e 2) = h2<br />

µ = r p (1 + e) = r a (1 − e)<br />

h = r p v p = r a v a = ⃗r × ⃗v = √ pµ<br />

h = √ µr<br />

(circular orbit)<br />

E = v2<br />

2 − µ r = − µ 2a<br />

√ ( 2<br />

v = µ<br />

r a)<br />

− 1<br />

(vis-viva)<br />

velocity v<br />

√<br />

2µ<br />

v escape = (escape velocity for parabola)<br />

r<br />

√ µ<br />

v radial =<br />

p e sin θ<br />

√ µ<br />

v normal = (1 + e cos θ)<br />

√<br />

p<br />

( )<br />

µ 1 + e<br />

v p =<br />

a 1 − e<br />

v perigee (closest)<br />

v apogee (furthest)<br />

v a =<br />

√ µ<br />

= (1 + e)<br />

p<br />

√ ( 2<br />

= µ − 1 r p a)<br />

√<br />

µ<br />

a<br />

( ) 1 − e<br />

1 + e<br />

√ µ<br />

= (1 − e)<br />

p<br />

√ ( 2<br />

= µ − 1 r a a)<br />

66


magnitude of ⃗r<br />

period T<br />

mean satellite speed n<br />

r = a ( 1 − e 2)<br />

1 + e cos θ = h2 1<br />

µ 1 + e cos θ<br />

r cos θ = a (cos E − e)<br />

r = a (1 − e cos E)<br />

√<br />

T = 2 h πab = 2π a 3<br />

µ<br />

n = 2π<br />

T = √<br />

µ<br />

a 3<br />

eccentric anomaly E tan θ 2 = √<br />

1+e<br />

1−e tan E 2<br />

area sweep rate<br />

dA<br />

dt = h 2<br />

equation of motion ¨⃗r +<br />

µ<br />

r 3 ⃗r = 0<br />

(eq 4.2-14 Bate book)<br />

spherical coordinates relation cos(i) = sin(A z ) cos(φ) where i is the inclination<br />

and A z is the azimuth and φ is latitude<br />

2<br />

Notice in the above, that the period T of satellite depends only on a (for same µ)<br />

In the above, µ = GM where M is the mass of the body at the focus of the ellipse and G is the gravitational<br />

constant. h is the specific mass angular momentum (moment of linear momentum) of the satellite. Hence the<br />

units of h2<br />

µ<br />

is length.<br />

To draw the locus of the satellite (the small body moving around the ellipse, all what we need is the<br />

eccentricity e and a, the major axes length. Then by changing the angle θ the path of the satellite is drawn. I<br />

have a demo on this here<br />

See http://nssdc.gsfc.nasa.gov/planetary/fact<strong>sheet</strong>/earthfact.html for earth facts<br />

This table below is from my class EMA 550 handouts (astrodynamics, spring 2014)<br />

2 Image is from my class notes, page 7-9, EMA 550, Univ. Of Wisconsin, Madison by professor S. Sandrik<br />

67


68


18.3 Flight path angle for ellipse γ<br />

69


√<br />

v = |⃗v| = |⃗v r | 2 + |⃗v n | 2<br />

√ ( 2<br />

= µ<br />

r a)<br />

− 1<br />

To find γ, if r is given then use<br />

If θ is given, then use<br />

√<br />

cos γ =<br />

√<br />

ap<br />

r (2a − r) =<br />

a 2 (1 − e 2 )<br />

r (2a − r)<br />

tan γ =<br />

e sin θ<br />

1 + e cos θ<br />

18.4 Parabolic trajectory<br />

This diagram shows the parabolic trajectory<br />

18.5 Hyperbolic trajectory<br />

This diagram shows the hyperbolic trajectory<br />

70


This diagram below from Orbital mechanics for Engineering students, second edition, by Howard D. Curtis,<br />

page 109<br />

71


18.6 showing that energy is constant<br />

Showing that energy E = v2<br />

2 − µ r<br />

is constant.<br />

Most of such relations starts from the same place. The equation of motion of satellite under the assumption<br />

that its mass is much smaller than the mass of the large body (say earth) it is rotating around. Hence we can<br />

use ν = GM and the equation of motion reduces to<br />

¨⃗r + µ r 3⃗r = 0<br />

In the above equation, the vector ⃗r is the relative vector from the center of the earth to the center of the satellite.<br />

The reason the center of earth is used as the origin of the inertial frame of reference is due to the assumption<br />

that M ≫ m where M is the mass of earth (or the body at the focal of the ellipse) and m is the mass of the<br />

satellite. Hence the median center of mass between the earth and the satellite is taken to be the center of earth.<br />

This is an approximation, but a very good approximation.<br />

The first step is to dot product the above equation with ˙⃗r giving<br />

˙⃗r · µ<br />

r 3 ⃗r = µ ṙ<br />

r 2<br />

˙⃗r · ¨⃗r + ˙⃗r · µ<br />

r3⃗r = 0 (1)<br />

( )<br />

ṙ2<br />

2<br />

and we also see that<br />

And there is the main trick. We look ahead and see that ˙⃗r · ¨⃗r = ṙ¨r but ṙ¨r = d dt<br />

but µ ṙ = d ( −µ<br />

)<br />

r 2 dt r Hence equation 1 above can be written as<br />

Hence<br />

(<br />

d v<br />

2<br />

dt 2 − r )<br />

= 0<br />

µ<br />

E = v2<br />

2 − r µ<br />

Where E is a constant, which is the total energy of the satellite.<br />

18.7 Earth satellite Transfer orbits<br />

18.7.1 Hohmann transfer<br />

This diagram shows the Hohmann transfer<br />

72


73


18.7.2 Bi-Elliptic transfer orbit<br />

74


18.7.3 semi-tangential elliptical transfer<br />

18.8 Rocket engines, Hohmann transfer, plane change at equator<br />

two cases: Hohmann transfer, 2 burns, or semi-tangential. All burns at equator.<br />

75


76


18.8.1 Rocket equation with plane change not at equator<br />

77


18.9 Spherical coordinates<br />

78


18.10 interplanetary transfer orbits<br />

18.10.1 interplanetary hohmann transfer orbit, case one<br />

The following are the steps to accomplish the above. The first stage is getting into the Hohmann orbit from<br />

planet 1, then reaching the sphere of influence of the second planet. Then we either do a fly-by or do a parking<br />

orbit around the second planet. These steps below show how to reach the second planet and do a parking orbit<br />

around it.<br />

79


The input is the following.<br />

1. µ 1 planet one standard gravitational parameter<br />

2. µ 2 planet two standard gravitational parameter<br />

3. µ sun standard gravitational parameter for the sun 1.327 × 10 8 km<br />

4. r 1 planet one radius<br />

5. r 2 planet two radius<br />

6. alt 1 original satellite altitude above planet one. For example, for LEO use 300 km<br />

7. alt 2 satellite altitude above second planet. (since goal is to send satellite for circular orbit around second<br />

planet)<br />

8. R 1 mean distance of center of first planet from the sun. For earth use AU = 1.495978 × 10 8 km<br />

9. R 2 mean distance of center of second planet from the sun. For Mars use 1.524 AU<br />

10. SOI 1 sphere of influence for first planet. For earth use 9.24 × 10 8 km<br />

11. SOI 2 sphere of influence for second planet.<br />

Given the above input, there are the steps to achieve the above maneuver<br />

1. Find the burn out distance of the satellite r bo = r 1 + alt 1<br />

2. Find satellite speed around planet earth (relative to planet) V sat =<br />

√<br />

µ1<br />

r bo<br />

3. Find Hohmann ellipse a = R 1+R 2<br />

2<br />

4. Find speed of satellite at perigee relative to sun V perigee =<br />

√<br />

µ sun<br />

(<br />

2<br />

R 1<br />

− 1 a<br />

)<br />

5. Find speed of earth (first planet) relative to sun V 1 =<br />

√<br />

µsun<br />

R 1<br />

6. Find escape velocity from first planet V ∞,out = V perigee − V 1<br />

7. Find burn out speed at first planet by solving the energy equation V 2<br />

bo<br />

2 − µ 1<br />

r bo<br />

8. Find ∆V 1 needed at planet one ∆V 1 = V bo − V sat<br />

9. Find e the eccentricity of the escape hyperbola e =<br />

√<br />

1 + V 2 ∞V 2<br />

bo r2 bo<br />

µ 2 1<br />

10. Find the angle with the path of planet one velocity vector η = arccos ( − 1 e<br />

11. Find the dusk-line angle θ = 180 0 − η<br />

)<br />

= V 2 ∞,out<br />

2<br />

− µ 1<br />

SOI 1<br />

for V bo<br />

The above completes the first stage, now the satellite is in the Hohmann transfer orbit. Assuming it reached<br />

the orbit of the second planet ahead of it as shown in the diagram above. Now we start the second stage to land<br />

the satellite on a parking orbit around the second planet at altitude alt 2 above the surface of the second planet.<br />

These are the steps needed.<br />

1. Find the apogee speed of the satellite V apogree =<br />

√<br />

µ sun<br />

(<br />

2<br />

R 2<br />

− 1 a<br />

)<br />

80


2. Find speed of second planet V 2 =<br />

√<br />

µsun<br />

R 2<br />

3. Find V ∞ entering the second planet sphere of influence V ∞,in = V 2 − V apogree<br />

4. Find burn in radius where the satellite will be closest to the second planet. r bo = r 1 + alt 2<br />

5. Find burn out speed at second planet by solving the energy equation V 2<br />

bo<br />

2 − µ 2<br />

r bo<br />

6. Find impact parameter b on entry to second planet SOI b = r boV bo<br />

V ∞,in<br />

7. Find the required satellite speed around the second planet V sat =<br />

√<br />

µ2<br />

r bo<br />

8. Find ∆V 2 needed at planet two ∆V 2 = V sat − V bo<br />

9. Find e the eccentricity of the approaching hyperbola on second planet e =<br />

10. Find the angle with the path of planet two velocity vector η = arccos ( − 1 e<br />

11. Find the dusk-line angle for second planet θ = 2η − 180 0<br />

18.11 rendezvous orbits<br />

18.11.1 Two satellite, walking rendezvous using Hohmann transfer<br />

)<br />

√<br />

= V 2 ∞,in<br />

2<br />

− µ 2<br />

SOI 2<br />

for V bo<br />

1 + V 2 ∞ V 2<br />

bo r2 bo<br />

µ 2 2<br />

81


Algorithm 1 Hohmann Walking Rendezvous Orbit, case 1<br />

1: function hohmann walking rendezvous(θ 0 , r, altitude, µ)<br />

2: θ 0 := θ 0<br />

π<br />

180<br />

⊲ convert from degrees to radian<br />

3: r a := r + altitude<br />

4: T := 2π√<br />

r 3 aµ<br />

⊲ period of circular orbit<br />

5: n := 1<br />

6: done:=false<br />

7: while not(done) do<br />

(<br />

8: TOF := N − θ 0<br />

9: a := solve<br />

10: r p := 2a − r a<br />

11: if rp < r then<br />

12: N = N + 1<br />

13: else<br />

14: done:=true<br />

15: end if<br />

16: end while<br />

2π<br />

)<br />

T<br />

(<br />

T OF = 2π<br />

√<br />

17: V befor := µ<br />

h<br />

√<br />

18: V after := µ ( 2<br />

h − 1 )<br />

a<br />

19: ∆V := 2(V after − V before )<br />

20: return (TOF, ∆V )<br />

21: end function<br />

√<br />

An example implementation is below<br />

)<br />

a 3<br />

µ<br />

for a<br />

1 hohmannRendezvousSameOrbit[\[Theta]00_, r_, alt_, mu_] :=<br />

2 Module[{\[Theta]0 = \[Theta]00*Pi/180, n = 1, delT, v1, v2, period, a,<br />

3 rp, ra, done = False, vBefore, vAfter},<br />

4 ra = r + alt;<br />

5 period = 2 Pi Sqrt[ra^3/mu];<br />

6 While[Not[done],<br />

7 delT = (n - \[Theta]0 /(2 Pi)) period ;<br />

8 a = First@Select[a /. NSolve[delT == 2 Pi Sqrt[a^3/mu], a], Element[#, Reals] &];<br />

9 rp = 2 a - ra;<br />

10 If[rp < r,(*we hit the earth, try again*)<br />

11 n = n + 1,<br />

12 done = True<br />

13 ]<br />

14 ];<br />

15 vBefore = Sqrt[mu/h];<br />

16 vAfter = Sqrt[mu (2/h - 1/a)];<br />

17 {delT, 2 (vAfter - vBefore)}<br />

18 ]<br />

And calling the above<br />

82


1 mu = 324859;<br />

2 alt = 1475.776;<br />

3 r = 6052;<br />

4 \[Theta]0 = 3.80562; (*degree*)<br />

5 hohmannRendezvousSameOrbit[\[Theta]0, r, alt, mu]<br />

6 {7123.89, -0.0467913}<br />

18.11.2 Two satellite, separate orbits, rendezvous using Hohmann transfer, coplaner<br />

83


Algorithm 2 Hohmann rendezvous algorithm, case 1<br />

1: function hohmann rendezvous 1(θ 0 , r a , r b , µ)<br />

π<br />

2: θ 0 := θ 0 180<br />

⊲ convert from degrees to radian<br />

3: a := ra+r b<br />

2<br />

⊲ Hohmann orbit semi-major axes<br />

√<br />

4: TOF := π<br />

5: θ H := π<br />

6: ω a :=<br />

a 3<br />

µ<br />

⊲ time of flight on Hohmann orbit<br />

( ( ) ) 3/2<br />

1 − ra+r b<br />

2r b<br />

√ µ<br />

r 3 a<br />

⊲ required phase angle before starting Hohmann transfer<br />

⊲ angular speed of lower rad/sec<br />

7: ω b := √ µ<br />

⊲ angular speed of higher satellite rad/sec<br />

rb<br />

3<br />

8: if θ 0 ≤ θ H then ⊲ adjust initial angle if needed<br />

9: θ 0 := θ 0 + 2π<br />

10: end if<br />

11: wait time := θ 0−θ H<br />

ω a−ω b<br />

⊲ how long to wait before starting Hohmann transfer<br />

12: wait time := wait time + TOF ⊲ now ready to go, add Hohmann transfer time<br />

13: return wait time<br />

14: end function<br />

An example implementation is below<br />

1 hohmann_rendezvous_1:= proc({<br />

2 theta::numeric:=0,<br />

3 r1::numeric:=0,<br />

4 r2::numeric:=0,<br />

5 mu::numeric:=3.986*10^5})<br />

6 local theta0,thetaH,TOF,a,omega1,omega2,wait_time;<br />

7 theta0 := evalf(theta*Pi/180);<br />

8 a := (r1+r2)/2;<br />

9 TOF := Pi*(sqrt(a^3/mu));<br />

10 omega1 := sqrt(mu/r1^3);<br />

11 omega2 := sqrt(mu/r2^3);<br />

12 thetaH := evalf(Pi*(1-((r1+r2)/(2*r2))^(3/2)));<br />

13 if theta0


18.11.3 Two satellite, separate orbits, rendezvous using bi-elliptic transfer, coplaner<br />

In this transfer, the lower (fast satellite) does not have to wait for phase lock as in the case with Hohmann<br />

transfer. The transfer can starts immediately. There is a free parameter N that one select depending on fuel<br />

cost requiments or any limitiation on the first transfer orbit semi-major axes distance required. One can start<br />

with N = 0 and adjust as needed.<br />

85


Algorithm 3 Hohmann rendezvous algorithm, case 2<br />

1: function hohmann rendezvous 2(θ 0 , r 1 , r 2 , N, µ)<br />

π<br />

2: θ 0 := θ 0 180 (<br />

⊲ convert from degrees to radian<br />

( ) ) 3/2<br />

3: θ H := π 1 − r1 +r 2<br />

2r 2<br />

⊲ Find Hohmann ideal phase angle before transfer<br />

4: if θ 0 = θ H and N = 0 then ⊲ adjust for special case<br />

5: a := r 2+r 1<br />

2 (√<br />

6: TOF := π<br />

)<br />

a 3<br />

µ<br />

7: else<br />

8: ω 2 := √ µ<br />

r2<br />

3<br />

9: t 2 := (2π−θ 0)+2πN<br />

ω 2<br />

10: a 1 := rt+r 1<br />

2<br />

11: a 2 := rt+r 2<br />

2<br />

12: TOF := π<br />

(√<br />

a 3<br />

1<br />

µ<br />

+ a3 2<br />

µ<br />

)<br />

⊲ angular speed of slower satellite in rad/sec<br />

⊲ find time of light of the slower satellite<br />

⊲ time of flight for the fast satellite<br />

13: r t := solve (t 2 = T OF ) for r t ⊲ Solve numerically for r t<br />

14: end if<br />

15: return TOF<br />

16: end function<br />

An example implementation is below<br />

1 hohmann_rendezvous_2:= proc({<br />

2 theta::numeric:=0,<br />

3 r1::numeric:=0,<br />

4 r2::numeric:=0,<br />

5 N::nonnegint:=0,<br />

6 mu::numeric:=3.986*10^5})<br />

7 local theta0,thetaH,TOF;<br />

8 theta0 := theta*Pi/180;<br />

9 thetaH := Pi*(1-((r1+r2)/(2*r2))^(3/2));<br />

10 if theta0 = thetaH and N = 0 then<br />

11 proc()<br />

12 local a:=(r1+r2)/2;<br />

13 TOF:= Pi*(sqrt(a^3/mu));<br />

14 end proc()<br />

15 else<br />

16 proc()<br />

17 local t2,a1,a2,rt,omega2;<br />

18 omega2 := sqrt(mu/r2^3);<br />

19 t2 := ((2*Pi-theta0)+2*Pi*N)/omega2;<br />

20 a1 := (rt+r1)/2;<br />

21 a2 := (rt+r2)/2;<br />

22 TOF := Pi*(sqrt(a1^3/mu)+sqrt(a2^3/mu));<br />

23 rt := op(select(is, [solve(t2=TOF,rt)], real));<br />

24 end proc()<br />

25 fi;<br />

26 eval(TOF);<br />

27 end proc:<br />

And calling the above for two different cases gives<br />

86


1 TOF:=hohmann_rendezvous_2(theta=0,r1=6678,r2=6878,N=0):<br />

2 evalf(TOF/(60*60)); #in hrs<br />

3 1.576892101<br />

4 TOF:=hohmann_rendezvous_2(theta=160,r1=6678,r2=6878,N=1):<br />

5 evalf(TOF/(60*60)); #in hrs<br />

6 2.452943266<br />

87


18.12 Semi-tangential transfers, elliptical, parabolic and hyperbolic<br />

88


89


18.13 Lagrange points<br />

90


18.14 Orbit changing by low contiuous thrust<br />

18.15 References<br />

1. Orbital mechanics for Engineering students, second edition, by Howard D. Curtis<br />

2. Orbital Mechanics, Vladimir Chobotov, second edition, AIAA<br />

3. Fundamentals of Astrodynamics, Bates, Muller and White. Dover1971<br />

91


19 Dynamic of flights<br />

19.1 Wing geometry<br />

length of mean<br />

aerodynamic chord<br />

projected on the<br />

root chord<br />

Leading edge<br />

sweep angle<br />

root chord<br />

centroid of area of<br />

one half of wing<br />

located on mean<br />

aerodynamic chord<br />

local chord<br />

wing tip chord<br />

wing1.vsdx<br />

Nasser M. Abbasi<br />

020914<br />

half wing span<br />

92


C r below is the core chord of the wing.<br />

A.C.<br />

tail<br />

control-fixed<br />

static margin<br />

apex of<br />

wing<br />

wing<br />

A.C.<br />

mac<br />

c.g<br />

positive for<br />

static<br />

stability<br />

wing_4_generic.vsdx<br />

Nasser M. Abbasi<br />

020914<br />

This is a diagram to use to generate equations of longitudinal equilibrium.<br />

This distance is called the stick-fixed static margin k m = (h n − h) ¯c Must be positive for static stability<br />

93


tail mean<br />

aerodynamic<br />

chord<br />

The whole Airplane<br />

neutral point (aft c.g.<br />

for static stability)<br />

Wing mean<br />

aerodynamic chord<br />

km<br />

forces.vsd<br />

updated 2/3/<br />

14<br />

Nasser M.<br />

Abbasi<br />

19.2 Summary of main equations<br />

This table contain some definitions and equations that can be useful.<br />

# equation meaning/use<br />

94


1<br />

C L = ∂C L<br />

∂α α<br />

= C Lα α<br />

= aα<br />

C L is lift coefficient. α is angle<br />

of attack. a is slope ∂C L<br />

∂α<br />

which<br />

is the same as C Lα<br />

2 C Lw = C Lwα α wing lift coefficient<br />

3 C D = C Dmin + kC 2 L<br />

drag coefficient<br />

4 C mw = C macw + (C Lw + C Dmin α w ) (h − h nw ) + (C L α w − C Dw ) z¯c pitching moment coefficient<br />

due to wing only about the<br />

C.G. of the airplane assuming<br />

small α w . This is simplified<br />

more by assuming C Dw α w ≪<br />

C Lw and (C L α w − C Dw ) ≪ 1<br />

5 C mw = C macw + C Lw (h − h nw ) simplified wing Pitching moment<br />

C mwb = C macwb + C Lwb (h − h nw )<br />

6<br />

= C macwb + ∂C L wb<br />

∂α wb<br />

α wb (h − h nw )<br />

= C macwb + a wb α wb (h − h nw )<br />

simplified pitching moment coefficient<br />

due to wing and body<br />

about the C.G. of the airplane.<br />

α wb is the angle of attack<br />

7 C Lt = Lt<br />

1<br />

2 ρV 2 S t<br />

C Lt is the lift coefficient generated<br />

by tail. S t is the tail area.<br />

V is airplane air speed<br />

8 L = L wb + L t total lift of airplane. L wb is lift<br />

due to body and wing and L t<br />

is lift due to tail<br />

9 C L = C Lwb + St<br />

S C L t<br />

coefficient of total lift of airplane.<br />

C Lwb is coefficient of lift<br />

due to wing and body. C Lt is<br />

lift coefficient due to tail. S is<br />

the total wing area. S t is tail<br />

area<br />

10<br />

1<br />

M t = −l t L t = −l t C Lt 2 ρV 2 S t pitching moment due to tail<br />

about C.G. of airplane<br />

11 C mt = Mt = − lt¯c<br />

S t<br />

1<br />

2 ρV 2 S t¯c S C L t<br />

= −V H C Lt pitching moment coefficient<br />

due to tail. V H = lt¯c<br />

S t<br />

S<br />

is called<br />

tail volume<br />

12<br />

V H = lt¯c<br />

S t<br />

S<br />

¯V H = ¯l t¯c<br />

S t<br />

S<br />

introducing ¯V H bar tail volume<br />

which is V H but uses ¯l t instead<br />

of l t . Important note. V H<br />

depends on location of C.G.,<br />

but ¯V H does not. ¯lt = l t +<br />

(h − h nwb ) ¯c<br />

95


13 C mt = − ¯V H C Lt + C Lt<br />

S t<br />

S (h − h n wb<br />

) pitching moment coefficient<br />

due to tail expressed using ¯V H .<br />

This is the one to use.<br />

14 C mp pitching moment coefficient<br />

due to propulsion about airplane<br />

C.G.<br />

15 C m = C mwb + C mt + C mp total airplane pitching moment<br />

coefficient about airplane C.G.<br />

16<br />

C m = C mwb + C mt + C mp<br />

[<br />

]<br />

= C macwb + C Lwb (h − h nw )<br />

C L<br />

+ [ − ¯V H C Lt + C Lt<br />

S t<br />

S (h − h n wb<br />

) ] + C mp<br />

{ ( }} ) {<br />

simplified total Pitching moment<br />

coefficient about airplane<br />

S t<br />

= C macwb + C Lwb + C Lt (h − h nw ) −<br />

S<br />

¯V H C Lt + C mp<br />

C.G.<br />

= C macwb + C L (h − h nw ) − ¯V H C Lt + C mp<br />

17<br />

∂C m<br />

∂α<br />

= ∂Cmac wb<br />

∂α<br />

C mα = ∂Cmac wb<br />

∂α<br />

18 h n = h nwb − 1 ∂C L<br />

∂α<br />

+ ∂C L<br />

∂α<br />

(h − h n w<br />

) − ¯V H<br />

∂C Lt<br />

∂α<br />

+ C Lα (h − h nw ) − ¯V H<br />

∂C Lt<br />

∂α<br />

( ∂Cmacwb<br />

∂α<br />

− ¯V H<br />

∂C Lt<br />

∂α<br />

)<br />

+<br />

∂Cmp<br />

∂α<br />

+ ∂Cmp<br />

∂α<br />

+<br />

∂Cmp<br />

∂α<br />

derivative of total pitching moment<br />

coefficient C m w.r.t airplane<br />

angle of attack α<br />

location of airplane neutral<br />

point of airplane found by setting<br />

C mα = 0 in the above<br />

equation<br />

19<br />

∂C m<br />

∂α<br />

= ∂C L<br />

∂α (h − h n)<br />

C mα = C Lα (h − h n )<br />

rewrite of C mα in terms of h n .<br />

Derived using the above two<br />

equations.<br />

20 k n = h n − h static margin. Must be Positive<br />

for static stability<br />

19.2.1 Writing the equations in linear form<br />

The following equations are derived from the above set of equation using what is called the linear form. The main<br />

point is to bring into the equations the expression for C Lt written in term of α wb . This is done by expressing the<br />

∂C Lwb<br />

tail angle of attack α t in terms of α wb via the downwash angle and the i t angle.<br />

∂α wb<br />

in the above equations are<br />

replaced by a wb and ∂C L t<br />

∂α t<br />

is replaced by a t . This replacement says that it is a linear relation between C L and<br />

the corresponding angle of attack. The main of this rewrite is to obtain an expression for C m in terms of α wb<br />

where α t is expressed in terms of α wb , hence α t do not show explicitly. The linear form of the equations is what<br />

from now on.<br />

# equation meaning/use<br />

96


1<br />

2<br />

3<br />

C Lwb = ∂C L wb<br />

∂α wb<br />

α wb<br />

= a wb α wb<br />

C Lt = a t α t<br />

C mp = C m0 p + ∂C mp<br />

∂α α<br />

α t = α wb − i t − ɛ<br />

ɛ = ɛ 0 + ∂ɛ<br />

∂α α wb<br />

C Lt = a t α t<br />

[ (<br />

= a t α wb 1 − ∂ɛ ) ]<br />

− i t − ɛ 0<br />

∂α<br />

a wb is constant, represents<br />

∂C L wb<br />

∂α wb<br />

and C m0p<br />

is propulsion pitching<br />

moment coeff. at zero<br />

angle of attack α<br />

main relation that associates<br />

α wb with α t . α wb<br />

is the wing-body angle<br />

of attack, ɛ is downwash<br />

angle at tail, and<br />

i t is tail angle with horizontal<br />

reference (see diagram)<br />

Lift due to tail expressed<br />

using α wb and<br />

ɛ (notice that α t do not<br />

show explicitly)<br />

(<br />

4 a = a wb<br />

[1 + at S t<br />

a wb S 1 −<br />

∂ɛ<br />

∂α) ] a defined for use with<br />

overall lift coefficient<br />

5<br />

C L =<br />

a wb α { }}<br />

wb<br />

{<br />

C Lwb<br />

+ St<br />

S C L t<br />

= a wb α wb + St<br />

S a t<br />

= aα<br />

= (C L ) αwb =0 + aα wb<br />

[<br />

αwb<br />

(<br />

1 −<br />

∂ɛ<br />

∂α<br />

)<br />

− it − ɛ 0<br />

]<br />

overall airplane lift using<br />

linear relations<br />

6 α = α wb − at S t<br />

a S (i t + ɛ 0 )<br />

overall angle of attack<br />

α as function of the<br />

wing and body angle of<br />

attack α wb and tail angles<br />

cl_apha.vsdx<br />

Nasser M. Abbasi<br />

021 214<br />

7<br />

C m = C m0 + ∂Cm<br />

∂α α = C m0 + C mα α<br />

C m = ¯C m0 + ∂Cm<br />

∂α α wb = ¯C m0 + C mα α wb<br />

overall airplane pitch<br />

moment. Two versions<br />

one uses α wb and one<br />

uses α<br />

97


8<br />

9<br />

( )<br />

C mα = a (h − h nwb ) − a t ¯VH 1 −<br />

∂ɛ ∂C<br />

∂α + mp<br />

∂α<br />

C mα = a wb (h − h nwb ) − a t V H<br />

(<br />

1 −<br />

∂ɛ<br />

∂α<br />

)<br />

+<br />

∂C mp<br />

∂α<br />

C m0 = C macwb + C mop + a t ¯VH (ɛ 0 + i t ) [ 1 − at S t<br />

a S<br />

¯C m0 = C macwb + ¯C mop + a t V H (ɛ 0 + i t )<br />

( )]<br />

1 −<br />

∂ɛ<br />

∂α<br />

∂C m<br />

∂α<br />

Two versions of<br />

one for α wb and one one<br />

uses α<br />

C m0 is total pitching<br />

moment coef. at zero<br />

lift (does not depend on<br />

C.G. location) but ¯C m0<br />

is total pitching moment<br />

coef. at α wb = 0<br />

(not at zero lift). This<br />

depends on location of<br />

C.G.<br />

10<br />

¯Cm0p = C m0p + (α − α wb ) ∂Cmp<br />

∂α<br />

h n = h nwb + at ¯V<br />

( )<br />

a H 1 −<br />

∂ɛ<br />

∂α −<br />

1 ∂C mp<br />

a ∂α<br />

11<br />

= h nwb +<br />

a t (<br />

a wb<br />

[1+ a t S t<br />

a wb S<br />

1− ∂ɛ<br />

∂α<br />

)] ¯VH<br />

(<br />

1 −<br />

∂ɛ<br />

∂α<br />

)<br />

−<br />

1<br />

a wb<br />

[1+ a t S t<br />

a wb S<br />

(<br />

)] ∂Cmp<br />

1− ∂ɛ ∂α<br />

∂α<br />

Used to determine h n<br />

19.3 definitions<br />

1. Remember that for symmetric airfoil, when the chord is parallel to velocity vector, then the angle of attack<br />

is zero, and also the left coefficient is zero. But this is only for symmetric airfoil. For the common campbell<br />

airfoil shape, when the chord is parallel to the velocity vector, which means the angle of attack is zero,<br />

there will still be lift (small lift, but it is there). What this means, is that the chord line has to tilt down<br />

more to get zero lift. This extra tilting down makes the angle of attack negative. If we now draw a line<br />

from the right edge of the airfoil parallel to the velocity vector, this line is called the zero lift line (ZLL)<br />

see diagram below.<br />

Just remember, that angle of attack (which is always the angle between the chord and the velocity vector,<br />

the book below calls it the geometrical angle of attack) is negative for zero lift. This is when the airfoil is<br />

not symmetric. For symmetric airfoil, ZLL and the chord line are the same. This angle is small, −3 0 or so.<br />

Depending on shape. See Foundations of Aerodynamics, 5th ed, by Chow and Kuethe, here is the diagram.<br />

98


2. stall from http://en.wikipedia.org/wiki/Stall_(flight)<br />

In fluid dynamics, a stall is a reduction in the lift coefficient<br />

generated by a foil as angle of attack increases.[1] This occurs when<br />

the critical angle of attack of the foil is exceeded. The critical<br />

angle of attack is typically about 15 degrees, but it<br />

may vary significantly depending on the fluid, foil, and Reynolds number.<br />

3. Aerodynamics in road vehicle wiki page<br />

4. some demos relating to airplane control http://demonstrations.wolfram.com/ControllingAirplaneFlight/<br />

http://demonstrations.wolfram.com/ThePhysicsOfFlight/<br />

5. http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

6. Lectures Helicopter Aerodynamics and <strong>Dynamics</strong> by Prof. C. Venkatesan, Department of Aerospace<br />

Engineering, IIT Kanpur http://www.youtube.com/watch?v=DKWj2WzYXtQ&list=PLAE677E56C97A7C7D<br />

7. http://avstop.com/ac/apgeneral/terminology.html has easy to understand definitions airplane geometry.<br />

”The MAC is the mean average chord of the wing”<br />

8. http://www.tdmsoftware.com/afd/afd.html airfoil design software<br />

99


19.4 images and plots collected<br />

These are diagrams and images collected from different places. References is given next to each image.<br />

Perpendicular to c.g. velocity vector V<br />

Body x-axes<br />

(chord of<br />

airplane<br />

Zero left line<br />

for airplane<br />

Current<br />

velocity<br />

vector of<br />

airplane c.g.<br />

Parallel to V<br />

vector originate<br />

from A.C. and not<br />

C.G.<br />

Angle of attack of thrust<br />

Absolute angle of attack<br />

Weight of<br />

airplane.<br />

Vertically down<br />

By Nasser M. Abbasi<br />

my_drawing.vsd<br />

1/24/14<br />

100


http://en.wikipedia.org/wiki/Flight_dynamics_%28aircraft%29<br />

Dihedral angle<br />

Reference: http://en.wikipedia.org/wiki/Dihedral_%28aircraft%29<br />

101


definitions<br />

From Performance, Stability, <strong>Dynamics</strong>, and Control of Airplanes<br />

By Baudu N. Pamadi<br />

I do not see viscosity here?<br />

http://en.wikipedia.org/wiki/Drag_coefficient<br />

This below from http://www.grc.nasa.gov/WWW/k-12/UEET/StudentSite/dynamicsofflight.html<br />

102


http://www.grc.nasa.gov/WWW/k-12/UEET/StudentSite/dynamicsofflight.html<br />

http://www.grc.nasa.gov/WWW/k-12/airplane/alr.html<br />

103


http://www.grc.nasa.gov/WWW/k-12/airplane/alr.html<br />

http://www.grc.nasa.gov/WWW/k-12/airplane/alr.html<br />

From http://en.wikipedia.org/wiki/Lift_coefficient and http://en.wikipedia.org/wiki/File:Aeroforces<br />

104


http://en.wikipedia.org/wiki/Lift_coefficient<br />

http://en.wikipedia.org/wiki/File:Aeroforces.svg<br />

svg<br />

from http://adg.stanford.edu/aa241/drag/sweepncdc.html<br />

http://en.wikipedia.org/wiki/Lift_%28force%29<br />

105


http://adg.stanford.edu/aa241/drag/sweepncdc.html<br />

Images from http://adamone.rchomepage.com/cg_calc.htm and Flight dynamics principles by Cook, 1997.<br />

106


http://adamone.rchomepage.com/cg_calc.htm<br />

From http://chrusion.com/BJ7/SuperCalc7.html<br />

107


From http://www.willingtons.com/aircraft_center_of_gravity_calcu.html<br />

108


From http://www.solar-city.net/2010/06/airplane-control-surfaces.html nice diagram that shows<br />

clearly how the elevator causes the pitching motion (nose up/down). From same page, it says ”The purpose of<br />

109


the flaps is to generate more lift at slower airspeed, which enables the airplane to fly at a greatly reduced speed<br />

with a lower risk of stalling.”<br />

Images from flight dynamics principles, by Cook, 1997.<br />

110


Page 184, flight dynamics principles, by Cook<br />

Page 41, flight dynamics principles, by Cook<br />

Images from Performance, stability, dynamics and control of Airplanes. By Pamadi, AIAA press. Page 169.<br />

and http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

111


From: Performance, stability, dynamics and control of<br />

Airplanes. By Pamadi, AIAA press. Page 169<br />

http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

Image from http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

112


Image from http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

113


The profile drag of a streamlined object held<br />

in a fixed position relative to the airflow increases<br />

approximately as the square of the velocity;<br />

thus, doubling the airspeed increases the drag four times,<br />

and tripling the airspeed increases the drag nine times.<br />

The amount of induced drag varies inversely<br />

as the square of the airspeed.<br />

From the foregoing discussion, it can be noted that parasite drag increases<br />

as the square of the airspeed, and induced drag varies inversely as<br />

the square of the airspeed.<br />

lift varies directly with the wing area<br />

a wing with a planform area of 200 square feet lifts twice as<br />

much at the same angle of attack as a wing with an area of 100<br />

square feet.<br />

http://www.americanflyers.net/aviationlibrary/pilots_handbook/chapter_3.htm<br />

Image from FAA pilot handbook and http://www.youtube.com/watch?v=8uT55aei1NI<br />

114


From FAA pilot’s handbook<br />

I do not understand the above<br />

From FAA pilot’s handbook<br />

http://www.youtube.com/watch?v=8uT55aei1NI<br />

Image http://www.youtube.com/watch?v=8uT55aei1NI and http://www.youtube.com/user/DAMSQAZ?<br />

feature=watch<br />

115


http://www.youtube.com/watch?v=8uT55aei1NI<br />

http://www.youtube.com/user/DAMSQAZ?feature=watch<br />

A=area of wing<br />

D=density of air<br />

V=wind speed relative to<br />

wing<br />

116


zero-lift angle<br />

The angle of attack at which an airfoil does not produce any lift. Its value is generally less than zero unless<br />

the airfoil is symmetrical.<br />

This from Foundations of aerodynamics, 5 th ed. By Chow and Kuethe<br />

http://www.answers.com/topic/zero-lift-angle<br />

These are from text :foundation of aerodynamics by Kuethe and Chow<br />

1. the center of pressure is at ¼ chord for all values of lift coeff.<br />

2. center of pressure (c.p.) of a force is defined as the point about which the moment vanishes.<br />

3. The geometric angle of attack is defined as the angle between the flight path and the chord line of the airfoil.<br />

When the geometric angle of attack is zero, the lift coeff. Is zero.<br />

4. The point about which the moment coeff. Is independent of the angle of attack is called the aerodynamic<br />

center of the section. (a.c.)<br />

5. a.c. is at the ¼ chord line point.<br />

6. The value of the angle of attack that makes the lift coeff. Zero is called the angle of zero lift (Z.L.L.)<br />

There are from Foundations<br />

of aerodynamics, 5 th ed. By<br />

Chow and Kuethe<br />

19.5 Some strange shaped airplanes<br />

Image http://edition.cnn.com/2014/01/16/travel/inside-airbus-beluga/index.html?hpt=ibu_c2<br />

117


Image from http://edition.cnn.com/2014/01/16/travel/inside-airbus-beluga/index.html?hpt=ibu_<br />

c2<br />

Image from http://www.nasa.gov/centers/dryden/Features/super_guppy.html<br />

118


Image from http://www.aerospaceweb.org/question/aerodynamics/q0130.shtml ”Boeing Pelican ground<br />

effect vehicle”<br />

19.6 links<br />

1. https://3dwarehouse.sketchup.com/search.html?redirect=1&tags=airplane<br />

19.7 references<br />

1. Etkin and Reid, <strong>Dynamics</strong> of flight, 3rd edition.<br />

2. Cook, Flight <strong>Dynamics</strong> principles, third edition.<br />

3. Lecture notes, EMA 523 flight dynamics and control, University of Wisconsin, Madison by Professor<br />

Riccardo Bonazza<br />

4. Kuethe and Chow, Foundations of Aerodynamics, 4th edition<br />

119

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!