Você está na página 1de 92

ANGLO TECHNICAL DIVISION

AA Best Practice Guideline BPG S002

GUIDELINES FOR THE VIBRATION


DESIGN OF STRUCTURES
AA BPG S002 Issue 2
October 2007

GJ Krige
Anglo Technical Division
Anglo Operations Limited

ANGLO TECHNICAL DIVISION

CONTENTS
1
2

SCOPE ...................................................................................................................... 5
DEFINITIONS, NOTATION AND PROPERTIES....................................................... 7
2.1
Definitions ........................................................................................................... 7
2.2
Notation and Units be Used................................................................................ 8
2.3
Section Properties ............................................................................................ 11
VIBRATION DESIGN PROCEDURE....................................................................... 13
3.1
Step 1: Necessary Data and Information......................................................... 13
3.2
Step 2: Clarify Details of the Structure to be Designed ................................... 13
3.3
Step 3: Build a Computer Model....................................................................... 13
3.4
Step 4: Assess the Results............................................................................... 13
3.5
Step 5: Prepare Structural Design Calculations and Drawings ........................ 14
GENERAL CONCEPTS AND THEORY .................................................................. 15
4.1
Dynamic and Harmonic Loads.......................................................................... 15
4.2
Dynamic Characteristics of Structures.............................................................. 17
4.2.1
The Single Degree of Freedom System. ................................................... 17
4.2.2
Response to Harmonic Excitation ............................................................. 18
4.2.3
Resonance and Tuning ............................................................................. 18
4.2.4
Damping .................................................................................................... 20
4.2.5
Multi Degree of Freedom Systems ............................................................ 22
4.2.6
Mode Shapes and More about Natural Frequencies................................. 22
LOADS..................................................................................................................... 24
5.1
Rotating Unbalance .......................................................................................... 24
5.1.1
Motors and Turbines.................................................................................. 24
5.1.2
Vibrating Equipment .................................................................................. 26
5.2
Loads Applied to the Structure ......................................................................... 27
5.2.1
Data Required From the Equipment Supplier............................................ 27
5.2.2
Calculation of Spring Stiffness................................................................... 28
5.2.3
Example 1.................................................................................................. 30
5.2.4
Example 2.................................................................................................. 31
5.3
Impact Loads .................................................................................................... 32
5.3.1
Types of Impact Loads .............................................................................. 32
5.3.2
Energy Equations ...................................................................................... 32
5.3.3
Moving Mass Hits Stationary Mass ........................................................... 34
5.4
Ground Motion from Blasting and Piling ........................................................... 34
5.4.1
Basic Equation........................................................................................... 34
5.4.2
Blasting...................................................................................................... 34
5.4.3
Piling.......................................................................................................... 35
STRUCTURAL MODELLING AND RESPONSE ..................................................... 36
6.1
Modelling the Structure..................................................................................... 36
6.1.1
Models ....................................................................................................... 36
6.1.2
Degrees of Freedom.................................................................................. 36
6.1.3
Modelling Structural Geometry .................................................................. 37
6.1.4
Modelling Mass.......................................................................................... 38
6.1.5
Modelling Materials.................................................................................... 39
6.1.6
Modelling Connections .............................................................................. 39
6.1.7
Modelling Floors ........................................................................................ 39
6.2
Composite Beams and Floors .......................................................................... 42
6.3
Confirming the Accuracy of the Model.............................................................. 43
6.4
Modelling Machinery on Structures .................................................................. 44
6.5
Calculation of Dynamic Response.................................................................... 46
6.5.1
General...................................................................................................... 46
6.5.2
Significant Modes ...................................................................................... 46
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
2 of 92

ANGLO TECHNICAL DIVISION

6.6
Dynamic Analysis Computer Programmes ....................................................... 47
6.6.1
ROBOT V6 ................................................................................................ 47
6.6.2
PROKON. .................................................................................................. 47
6.6.3
Common Errors with Computer Vibration analysis.................................... 48
7 VIBRATION LIMITS ................................................................................................. 49
7.1
Introduction ....................................................................................................... 49
7.2
Human Sensitivity ............................................................................................. 49
7.3
Equipment and Machine Sensitivity.................................................................. 53
7.4
Structural Sensitivity ......................................................................................... 53
7.4.1
Brittle Finishes ........................................................................................... 53
7.4.2
Fatigue Life................................................................................................ 54
8 DESIGN GUIDANCE FOR SPECIFIC EQUIPMENT AND STRUCTURES............. 55
8.1
Crushers ........................................................................................................... 55
8.1.1
Modelling Crusher Support Structures ...................................................... 55
8.1.2
Loads Applied by Crushers ....................................................................... 55
8.2
Rotating Tubes ................................................................................................. 57
8.2.1
Types of Load Generated.......................................................................... 57
8.2.2
Specific Equipment.................................................................................... 59
8.3
Vibrating Screens and Feeders ........................................................................ 59
8.3.1
Basic Requirements .................................................................................. 59
8.3.2
Design and Use of Sub-frames ................................................................. 60
8.4
Rock Breakers .................................................................................................. 62
8.5
Design of Grizzly Bars ...................................................................................... 63
8.6
Vessel Agitation ................................................................................................ 64
8.6.1
Applied Loads............................................................................................ 64
8.6.2
Design Requirements ................................................................................ 65
8.7
Wood Chippers ................................................................................................. 66
9 PRACTICAL GUIDeLINES FOR FOUNDATIONS................................................... 67
9.1
Traditional Rules of Thumb............................................................................... 67
9.2
Simple Rules..................................................................................................... 67
9.3
Modelling Foundations...................................................................................... 68
9.3.1
Soil Conditions........................................................................................... 68
9.3.2
Simplified Preliminary Calculations ........................................................... 70
9.3.3
Damping .................................................................................................... 72
10
PRACTICAL DETAILS FOR TERTIARY STRUCTURAL ELEMENTS ................ 73
10.1
Individual Members ....................................................................................... 73
10.1.1 Approximate Natural Frequencies of Individual Members ......................... 73
10.1.2 Limitation of Slenderness Ratio to 80........................................................ 73
10.2
Walkways and Hand Railing ......................................................................... 74
10.3
Sheeting Rails ............................................................................................... 75
10.4
Plating on Chutes, Bins and Underpans ....................................................... 75
10.4.1 Natural Frequencies of Rectangular Panels.............................................. 75
10.5
Bracing Systems ........................................................................................... 78
11
PRACTICAL DETAILS FOR CONNECTIONS ..................................................... 80
11.1
Bolted connections........................................................................................ 80
11.2
Welded Connections ..................................................................................... 80
11.3
Beam-to-beam Connections ......................................................................... 80
11.4
Bracing Connections ..................................................................................... 81
12
VIBRATION MEASUREMENTS........................................................................... 82
12.1
What Should be Measured?.......................................................................... 82
12.2
Measuring Equipment ................................................................................... 82
12.3
Recording Measurements ............................................................................. 83
12.4
Relating Measured Displacements to Implied Stresses................................ 83
12.5
Baseline Vibration Measurement Guide........................................................ 84
12.5.1 Baseline Measurements ............................................................................ 84
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
3 of 92

ANGLO TECHNICAL DIVISION

12.5.2 What Baseline Measurements Can and Cant Do ..................................... 87


13
TROUBLE-SHOOTING AND STRUCTURAL MODIFICATION ........................... 88
13.1
Interpreting and Using Measurements .......................................................... 88
13.2
Changes to Applied Loads ............................................................................ 88
13.3
Structural Modifications................................................................................. 88
13.4
Common Concerns of Site Personnel........................................................... 89
14
BIBLIOGRAPHY................................................................................................... 91
14.1
Standards and Specifications........................................................................ 91
14.1.1 SANS Standards ....................................................................................... 91
14.1.2 AAC Specifications .................................................................................... 91
14.2
Text Books .................................................................................................... 91
14.3
Journal Papers .............................................................................................. 91
14.4
AAC Reports ................................................................................................. 91
15
RECORD OF MODIFICATIONS .......................................................................... 92

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
4 of 92

ANGLO TECHNICAL DIVISION

1 SCOPE
The purpose of this guide is primarily to assist with understanding the concepts on which
structural dynamics is based, and provide guidance in the practical implementation of
dynamic design. The following concepts need to be understood:
(a)
(b)
(c)
(d)
(e)
(f)

Natural frequencies, resonance and damping.


Mass, force and inertia force.
Degrees of freedom and computer modelling.
Evaluation of member section properties and end constraints.
Dynamic loading, dynamic reactions from vibrating equipment.
Assessment of allowable amplitudes and fatigue life.
Warning: Do not attempt dynamic analysis and design of any structures
unless you understand these concepts. Too often, the powerful computer
analysis packages available today are simply thrown at a dynamic design
problem and it is assumed that a satisfactory structure will simply pop out.
This does not happen.

This guide covers the full spectrum of the design procedures to be adopted for general
structures carrying equipment that generates dynamic loads. It is not intended to cover
design to resist environmental dynamic loads such as wind or earthquake, nor is it
intended to cover the design of unusual structures such as tall masts or long bridges.
Dynamic analysis and design of structures is aimed at ensuring three important criteria.
(a) There should not be resonance.
(b) The amplitudes of vibration should not exceed predefined limits.
(c) The structure should have an adequate fatigue life.
In order to achieve this, the Designer must understand the dynamic behaviour of
structures, must know the dynamic loads acting on the structure, must be able to model
and analyse the structure, and finally must be able to understand and assess structural
behaviour against the predefined limits. The guide is thus divided into the following eight
sections:
(a)

(b)
(c)

(d)
(e)
(f)

A basic presentation of the theory of the response of structures to dynamic loads.


This is not intended as a detailed or comprehensive coverage of dynamic analysis
theory, but rather a simple treatment to assist in understanding some fundamental
concepts. For a fuller treatment of dynamic analysis theory it will be necessary to
consult one of many excellent books on the market, a few of which are listed under
the bibliography.
Discussion of the dynamic loads that may be applied to industrial structures.
Guidance with modelling of structures for computer analysis. This section includes
a brief presentation of the computer programs used by ATD Structural engineering
for dynamic analysis of structures. This is not intended to replace the respective
manuals, but to provide guidance in their appropriate use.
Description of the limits placed on vibration severity.
A step by step practical guide to the design of some structures that must often be
dealt with by ATD Structural Engineering, including supporting structures for
crushers, screens, feeders, etc.
Some guidance for the appropriate design of tertiary structural members.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
5 of 92

ANGLO TECHNICAL DIVISION

(g)
(h)

Discussion of how to deal with problems that arise on specific installations. This
includes some comments regarding vibration measurements, typical problems that
have been experienced, and guidance regarding fixing problems.
A bibliography which includes the relevant codes and specifications as well as
additional material for anyone wanting more detailed information.
Warning: Throughout the document, various warnings are given. It is
important to take note of these, as they are areas where experience has
shown that mistakes tend to be made.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
6 of 92

ANGLO TECHNICAL DIVISION

2 DEFINITIONS, NOTATION AND PROPERTIES


The notation used throughout this guidelines document is listed below for ease of
reference.

2.1

Definitions

algorithm

Logical arithmetic or computational procedure for solving a problem.

amplitude

The maximum value of any harmonic quantity, ie force, acceleration,


velocity, or displacement. It is equal to half the peak-to-peak value.

damping

Dissipation or absorption of energy. Damping is usually assumed to


be viscous damping which means that it is proportional to the
velocity.

dynamic
magnification
factor

This is the ratio of the dynamic displacement amplitude of a structure


to the static displacement if the same structure is subjected to a static
load equivalent to the applied dynamic load amplitude.

frequency

This is the rate at which a harmonic quantity varies with time. It is


important to distinguish between a cyclical frequency, f, which is
measured in cycles/second, and radial frequency, , which is
measured in radians/second. These are directly related by:
= 2 f

frequency ratio

The ratio of the frequency of an applied harmonic load to the natural


frequency of the structure, ie E/ N.

harmonic

Any quantity that varies with time according to: Q = Qo sin (t +).
Many of the loads applied by industrial equipment are harmonic loads.

inertia force

The force required to accelerate the mass of a specified portion of a


structure.
These terms describe the relationship between two or more harmonic
quantities. If the quantities are in the same direction, and attain their
maximum or minimum values at the same time, they are said to be inphase, and their phase angles are the same, i.e. 1 = 2. If the two
quantities act in opposite directions and one attains its maximum value
when the other attains its minimum value, they are said to be out-ofphase, and their phase angles differ by , i.e. 1 = 2 + .

in-phase
out-of-phase

model

A model is a representation of a complex object. In dynamic design a


model typically refers to a computer representation of a real structure.

mode shape

A natural shape in which a structure will vibrate.


mathematically referred to as an eigenvector.

natural
frequency

The frequency at which a structure will naturally vibrate in the absence


of any applied force.

This is

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
7 of 92

ANGLO TECHNICAL DIVISION

peak-to-peak

The difference between the maximum and minimum values of a


dynamic quantity. When the quantity is harmonic the peak-to-peak
value is twice the amplitude. The peak-to-peak displacement of
vibrating equipment is sometimes referred to as the throw.

periodical

Changing in time according to a regular pattern that repeats itself.

period

The period of a harmonic quantity is the time taken for one complete
cycle. Thus:
P=

1 2
=
f

ppv

This is the peak particle velocity, generally used to describe the


severity of ground motion generated by blasting or other disturbance.

rms

This is the root mean square value of a dynamic quantity.


mathematically defined as:
a RMS =

It is

1T 2
a ( t )dt
T0

When the quantity is harmonic, its RMS value is 0,707 times the
amplitude.

2.2

Notation and Units be Used

The recommended units are given below for each symbol. It is highly recommended that
these units are used for all calculations, or serious errors can be introduced.
Warning: If these recommended units are not used for any reason, it is
necessary to reduce all calculations to the basic units of mass, length, and time
in order to ensure the accuracy of calculations. This problem usually arises
because the relationship between force and mass is determined by the gravity
constant. In static design calculations this does not matter, because we are
always using forces, never mass, so whatever gravity constant we use we get
the same answer as long as we are consistent. In dynamic design calculations
we use mass and force, so the correct gravity constant is crucial. If we choose
millimetres as our length unit, then the gravity constant is 9810 mm/s2, and not
9,81 m/s2. As one example, consider the elastic modulus of steel.
(a)

If we are using metres as our unit of length:


E = 200x103 MPa = 200x109 Pa = 200x109 N/m2 = 200x109 kg.m/s2/m2.

(b)

If we are using millimetres as our unit of length, we should use:


E = 200x109 kg.m/s2/m2 = 200x106 kg.mm/s2/mm2
But, because we tend to forget the effect our choice of units has on the
gravity constant, we are inclined to incorrectly use:
E = 200x103 MPa = 200x103 N/mm2 = 200x103 kg.m/s2/mm2.
This gives a stiffness which is too low by a factor of 1000, leading to
completely wrong results.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
8 of 92

ANGLO TECHNICAL DIVISION

Symbol Units
a
m
aM
ao
bx

m/s2
m/s2
m

by

A
b
B
c

m2
m
m
m

C
CC
CQ
D
do
dC
d1,d2
dw
D

N.s/m
N/m3
N/m3
m
m
m
m
m
m

DMF
Dv
Dx

Nm2/m
Nm2/m

Dy

Nm2/m

E
E
ECstat
ECdyn

m
N/m2
N/m2
N/m2

F
fE

Hz
Hz

fN
f1
f2
F(t)
FA
FB
FI
FM

Hz
Hz
Hz
N
N
N
N
Nm

FP
Fo
FQ

N
N
Nm

Description
Linear dimension, distance from
ground impact
Linear acceleration
Linear acceleration amplitude
Spacing of x-direction stiffeners in
orthotropic plating (m)
Spacing of y-direction stiffeners in
orthotropic plating (m)
Cross sectional area
Linear dimension
Deflection of beam or post
Axial displacement of steel coil
spring under permanent loads only
Damping constant
Soil uniform compression modulus
Soil uniform shear modulus
Displacement
Amplitude of displacement
Depth of beam or slab
Width of plates
Diameter of steel coil spring wire
Diameter of steel coil spring or
rubber buffer
Dynamic magnification factor
Shear rigidity of stiffened plate
Flexural rigidity of stiffened plate
about x-axis
Flexural rigidity of stiffened plate
about y-axis
Eccentricity of rotating mass
Elastic modulus of material
Static elastic modulus of concrete
Dynamic elastic modulus of
concrete
General cyclical frequency
Exciting frequency, i.e. frequency
of applied dynamic force
Natural frequency
First natural frequency
Second natural frequency
Applied load
Axial load applied by equipment
Breaking load on rock breaker
Inertia force
Bending moment applied by
equipment
Pushing load on rock breaker
Amplitude of applied load
Torque applied by equipment

Comment

See SAISC Red Handbook

200x109 N/ m2 for steel


30x109 N/ m2
34x109 N/ m2

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
9 of 92

ANGLO TECHNICAL DIVISION

FV
G
G

N
m/s2
N/ m2

Shear load applied by equipment


Acceleration due to gravity
Shear modulus of material

hCG

Ib

kg.m2

Im
Ix
Iox

kg.m2/m
m4
m4

Iy
Ioy

m4
m4

J
K

m4

k
KH

N/m
N/m

KV

N/m

Nm/rad

L
m
M
MB
MR

m
kg/m
kg
kg
kg

MS

kg

Free (unloaded) height of steel coil


spring. Height through which a
falling body falls before impact
Height of machine above centre of
gravity of base
Mass inertia of machine and
foundation
Mass inertia per unit length
Moment of inertia about x-axis
Moment of inertia of stiffener and
plate about x-axis
Moment of inertia about y-axis
Moment of inertia of stiffener and
plate about y-axis
Torsion constant
Dieckmann
K
value
for
determination of vibration severity
in terms of human sensitivity
Stiffness of structure
Horizontal stiffness of steel coil
spring, rubber buffer, or machine
base
Vertical stiffness of steel coil
spring, rubber buffer, or machine
base
Rotational stiffness of machine
base
Length of the member
Mass per metre
Mass
Mass of machine base
Mass of rotating or moving part of
machine
Mass of stationary part of machine
or body impacted by moving body
Number
Period of harmonic quantity
Frequency ratio
Radius of gyration about x-axis
Radius of gyration about y-axis
Time elapsed
Thickness of plate
Linear velocity
Amplitude of linear velocity
Peak particle velocity (ppv)
Linear
velocity
after
impact
between a moving body and a
stationary body

N
P
R
rx
ry
T
t
V
vo
vP
vS

s
m
m
s
m
m/s
m/s
m/s
m/s

9,81 m/s2
78x109 N/ m2 for steel
12x109 N/ m2 for concrete

See section properties below


See SAISC Red Handbook
See SAISC Red Handbook
See SAISC Red Handbook

See SAISC Red Handbook

See SAISC Red Handbook


See SAISC Red Handbook

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
10 of 92

ANGLO TECHNICAL DIVISION

kg or J

x(t)

rad
kg/m3

rad/s
rad/s

rad/s

Mass of explosive per delay


ground impact energy
Displacement of a degree
freedom with time
Measured horizontal deflection
top of portal column
Measured vertical deflection
centre of beam of slab
Damping ratio
Phase angle
Density of the material

or
of
at
at

7850 kg/m3 for steel


2450 kg/m3 for concrete

Poissons ratio for the material


General radial frequency
Radial frequency of applied
harmonic load
Radial natural frequency of
structure

Note that the values for elastic modulus, shear modulus and density for concrete quoted
above are typical values only. The use of specific aggregates may lead to different
values.
Warning:
Drawings, particularly those from European equipment
manufacturers, often give dynamic forces in units of kgf. This must be
multiplied by 9,81 to convert to units of N, which can then be used in further
calculations.

2.3

Section Properties

Member section properties are generally obtained from standard handbooks of section
properties, but some necessary properties are not often listed by these handbooks. The
elusive properties are defined here, for simple hand calculation if required.
Im is the mass inertia per unit length of the member. This should not be confused with
the moment of inertia, nor the bending moment. This value is not always immediately
available, but it can be calculated from section properties
d2
quoted in the SAISC Red Handbook as:

Im =

m
(I + I )
A x y

or

Im = (I x + I y )

t2
d1

For open sections, J may be approximated by:

J=

t1

diti3
3

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
11 of 92

ANGLO TECHNICAL DIVISION

For closed rectangular sections with uniform wall thickness, J


may be approximated by:
b

2 2

J=

2ta b
a+b

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
12 of 92

ANGLO TECHNICAL DIVISION

3 VIBRATION DESIGN PROCEDURE


3.1

Step 1: Necessary Data and Information

Get all the relevant data (using any means necessary!). The data must include:
(a)
(b)
(c)
(d)
(e)

(f)

3.2

Shape, size and mass of the equipment.


If the equipment is supported directly on the floor or by means of some kind of
springs or dampers. The type, stiffness and damping constant, of any spring
supports or dampers.
Magnitude of the dynamic loads applied by the equipment. Dont always just
believe the Suppliers on this one. Do your own check calculations.
Direction of the dynamic loads applied by the equipment. A screen is generally
assumed to apply a vertical load and a horizontal load at each spring. A vertical
crusher applies a load that sweeps around 360 degrees in the horizontal plane.
The time dependence of the dynamic loads applied by the equipment. By time
dependence is meant whether the dynamic load varies harmonically, whether it is
suddenly applied and remains constant for a while before it is just as suddenly
removed, or whether it is a short duration impact load.
If there are any special conditions regarding the allowable amplitudes. For
example, is there is any vibration sensitive equipment in the area, or is there an
office on a floor structurally connected to the vibrating floor? Ensure that all
Suppliers of equipment to be located in the vicinity of vibrating equipment provide
written statements of the vibration their equipment can withstand, or ensure that
they are informed in writing what vibration their equipment is required to withstand.

Step 2: Clarify Details of the Structure to be Designed

This means both the obvious aspects, such as whether floors will be grating or concrete,
the exact positions where columns and beams are required, etc, and the not so obvious
aspects, such as where exactly is access necessary preventing the use of bracing. All
these may seem simple but it is always surprising how much time is wasted when this
type of information has not been obtained from the start.
This step also includes getting information about the supports, in order to calculate the
spring constants.

3.3

Step 3: Build a Computer Model

Make sure you have read, and digested Sections 6, 8, 9 and 10. You are now ready to
build a computer model. Preliminary sizing of members can be done using static design
for all members remote from the actual vibrating equipment. For members providing
immediate support to vibrating equipment use the equations given in Section 10 to select
members having a sufficiently high individual natural frequency.
Before proceeding, check that the model looks right. Check the frequencies, check the
mode shapes.

3.4

Step 4: Assess the Results

Make sure you have read, and properly understood Section 7 and the implications of (f)
above. You are now ready to assess the results being spat out by your computer model.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
13 of 92

ANGLO TECHNICAL DIVISION

3.5

Step 5: Prepare Structural Design Calculations and Drawings

Make sure you have read, and can rationally apply Sections 10 and 11. You are now
ready to start preparing structural design calculations and approving what has been put
onto structural design drawings.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
14 of 92

ANGLO TECHNICAL DIVISION

4 GENERAL CONCEPTS AND THEORY


4.1

Dynamic and Harmonic Loads

A load has the right to call itself dynamic if, and only if:
(a)
(b)

Its magnitude changes in time.


The acceleration is big enough to produce significant inertia forces.

FI = Ma M
Examples of dynamic loads, shown in Figure 4.1, are:
(a)
(b)

The load produced by a boulder hitting a grizzly bar.


The load applied by a screen on its supporting structure.

Harmonic
force

Motor
Screen

Height
of fall

Springs
Grizzly bar

Harmonic Loads
Impact Loads

Figure 4.1: Typical Dynamic Loads


A dynamic load is called "harmonic" if it varies in time according to:

F = Fo sin(t + )
The values defining this harmonic movement are shown in Fig. 4.2
The relationships between the values of the respective amplitudes used in the analysis of
harmonic loads are:
Displacement amplitude
Linear velocity amplitude
Acceleration amplitude
Applied load amplitude
Inertia force amplitude

do
do
ao = do2
Fo
Mao2

(m)
(m/sec)
(m/sec2)
(N)
(N)

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
15 of 92

Period

RMS

Amplitude

RMS

Time

Amplitude

Peak-to-Peak

ANGLO TECHNICAL DIVISION

Figure 4.2: Harmonic Load


Warning: Some people loosely use the term "amplitude of displacement"
when they mean peak-to-peak movement (or stroke). Many people also
think in terms of displacement amplitude only, but amplitude can refer to
any harmonic entity. It is prudent to check what the person you are talking
to means by the word amplitude.
See Figure 4.3 for an example of out of phase movements. (Note that in Figure 4.3 the
difference of phase is shown in terms of time).

Difference of phase

Figure 4.3: Oscillations out of phase

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
16 of 92

ANGLO TECHNICAL DIVISION

4.2

Dynamic Characteristics of Structures

4.2.1 The Single Degree of Freedom System.


The most basic dynamic system is one which can move in only one way, or only one
direction. This is called a single degree of freedom (SDOF) system. The SDOF system
is the only one that we completely understand. However, Murphy's law being what it is,
very few of the structures that we analyze can be accurately represented by a single
degree of freedom model. See Figure 4.4 for examples of models for any single degree
of freedom system (ignoring damping, and replacing mass and stiffness by the real
values).

M
k=

48 EI
L3

k
k
M

These three single degree of freedom


models are mathematically equivalent
Figure 4.4: Single Degree of Freedom Models
We shall establish a series of rules and laws for the single degree of freedom system.
Then we will extrapolate these data for multi degrees of freedom systems.
The fundamental characteristics of the single degree of freedom system are:
Mass, M
Stiffness, k
Damping, C

(kg)
(N/m)
(Ns/m)

Then the basic equation of motion for the single degree of freedom system can be
written as:
M&x& + Cx& + kx = F( t )

A structure subjected to an impact and then left alone will vibrate until the cumulative
effects of damping will stop it. Without damping the structure will go on shaking until the
end of time. This type of movement, vibration when the applied load is zero, is called
"free vibrations". If we initially consider free vibration, then the equation of motion is:
M&x& + Cx& + kx = 0

The solution for this equation will be left to keen students and textbooks. The
frequencies of the free vibrations are called natural frequencies. These frequencies are
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
17 of 92

ANGLO TECHNICAL DIVISION

what the structure sort of prefers to vibrate at. The basic information to be derived from
this equation is the natural frequency, given below for under-damped structures, i.e.
where is less than 1,0.
D =

k
C 2
-(
)
M 2M

D = 1 - 2
C
C CR

The damping ratio, , actually very seldom exceeds 0,1 for normal structures, and it may
be as low as 0,01 for fully welded steel structures. When = 0,1 (the maximum likely
value), the above equation gives N = 0,995. This means that for all practical purposes
the damped natural frequency may be taken to be the same as the . The final
important dynamic characteristic of the system, the natural frequency, is thus given by:
f=

1
=
2 2

k
M

4.2.2 Response to Harmonic Excitation


The solution of the equation of motion with zero applied force helps us to understand the
dynamic system, but our real aim is to understand how structures respond to applied
loads. The simplest applied load is one which varies harmonically.
M&x& + Cx& + kx = F0 sin t

The final solution is complicated enough, so we will not attempt here to show how it is
obtained. The value of x(t) can be shown, (see textbooks) is:
x( t ) = e -t (A cos t + B sin t ) +

F0
sin(t - )
k

(1 - r 2 )2 + (2r )2

The first part of this equation is only of interest during start-up or shut-down of the
machine because it dies away quickly with time, and we are generally not too worried
about these as they generally happen quickly, without causing significant fatigue damage
or psychological disturbance to personnel. So what we are interested in is usually only
the second part of the equation, i.e.:
x( t ) =

F0
sin(t - )
k

(1 - r 2 )2 + (2r )2

In most cases, we are only really interested in the amplitude of x, which is:
xo =

F0
k

(1 - r 2 )2 + (2r )2

4.2.3 Resonance and Tuning


In the technical literature one encounters the concept of a dynamic magnification (or
sometimes amplification) factor, DMF. An understanding of this factor is the key to
grasping the dynamic performance of any structure, even a very complex one. This
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
18 of 92

ANGLO TECHNICAL DIVISION

factor is a way of describing the response of the structure to a harmonic applied load.
When the system is resonant, i.e. when the exciting frequency, E, equals its natural
frequency, N, the magnification factor may be extremely high (of the order of 40 for
welded steel structures, or 12 for concrete structures), whereas it may also drop well
below 1,0 when the exciting frequency is much higher than the natural frequency.
DMF =

xo
=
Fo
k

(1 - r 2 )2 + (2r )2

10
9
8
7
6
5
4
3
2
1
0

Dampin
g Ratio
High
tuning

0.05
0.08
0.2
0.5

Low
tuning

Resonance

Dynamic Magnification Factor

The diagram in Figure 4.5 represents the dynamic magnification factor as a function of
the damping and the frequency ratio.

Frequency Ratio

Figure 4.5: Dynamic Magnification Factor


Tuning is the action by which a structure is designed in a way that would insure that its
natural frequency complies with certain conditions. The most usual condition is for the
natural frequency to be different from the exciting frequency. If the natural frequency is
lower than the exciting frequency, then the system is low tuned, and the frequency ratio
is greater than 1,0. A frequency ratio too close to 1,0 still gives a very high dynamic
magnification factor, so we generally only talk of a low tuned structure if the frequency
ratio exceeds 1,4. If the natural frequency is higher than the exciting frequency, then the
system is high tuned and the frequency ratio is less than 1,0. A frequency ratio too close
to 1,0 still gives a very high dynamic magnification factor, so we generally only talk of a
high tuned structure if the frequency ratio is less than 0,7.
The dynamic magnification graph shown in Figure 4.5 shows that for very low tuning the
dynamic magnification factor becomes very small. This does not automatically mean
that the amplitudes of a low tuned system will be necessarily smaller that the amplitudes
of a high tuned system. A low tuned system is a lot more flexible than a high tuned one.
The quantity that it amplifies is normally far bigger than in a high tuned system. That
quantity can be a displacement, an acceleration, a force, in fact almost anything.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
19 of 92

ANGLO TECHNICAL DIVISION

The single degree of freedom system is the only one for which the dynamic magnification
factor has an immediate physical meaning. The notion itself can be defined in multi
degrees of freedom systems, but its physical meaning will be buried under tons of
equations.
When the exciting frequency equals the natural frequency the factor above tends
towards infinity. This is referred to as resonance. The displacements and related
values also tend towards infinity. They would reach it, or more probably die trying that is,
if not for the effect of damping. Resonance is the situation in which the response of a
structure to a dynamic excitation has its maximum value, which for a damped system is
inversely proportional to the damping ratio. At resonance r = 1, so that:

DMF =

1
2

The effect of damping at resonance is shown in Figure 4.6. When there is no damping
present, the amplitude of vibration response to any applied harmonic load keeps
increasing indefinitely. When there is damping present, the amplitude of the vibration
response increases up to a specific maximum value, then remains constant at that value.
4.2.4 Damping
Damping is the property of materials to absorb energy by internal friction. Contrary to
conventional wisdom, damping is not an unconditional blessing; when close to
resonance it reduces the dynamic forces, but when far from resonance in the low tuning
range, damping actually increases forces. Its effect is like a slight stiffening of the
springs on which the equipment is supported.
There are many ways to model the damping. None is perfectly accurate. The most
popular seems to be to model the damping effect as a force proportional to the linear
velocity. This is called viscous damping. Accurate or not, it will have to do. We will use
this model from two reasons. First, for our purposes it is accurate enough. Second,
using a more accurate damping model would lead to horrifyingly complex mathematical
developments.
There is another significant value that has to be defined: the critical damping. It means
the highest damping value that allows the system to oscillate. At more than critical
damping the system becomes so sluggish that it is no more able to follow the oscillations
of the exciting force. It just tries to sort of slowly crawl back to its initial position. Critical
damping is not very important for us. Steel damping is very much smaller. For a single
degree of freedom system the critical damping is:
C R = 2 kM

Damping is primarily a characteristic of the material, and the connections. For each type
of structure the damping can be expressed as a fraction (or percentage) of the critical
damping. For steel the damping is between 1% and 6%. For concrete it is between 6 %
and 10%. For rubber it is about 30%, but then structures are not made of rubber.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
20 of 92

ANGLO TECHNICAL DIVISION

50
40
30
Response

20
10
0
-10 0

No damping
With damping

-20
-30
-40
-50
Time (seconds)

Figure 4.6: Resonance with and without damping


A widely spread misconception about damping is to assume that if a structure is made
from a material with high damping then that damping will automatically influence all
dynamic deflections of that structure. This is not necessarily true, as damping only
becomes effective when there is movement.
For example, consider a structure consists of a square concrete floor supported by four
concrete columns, one in each corner, as shown in Figure 4.7. It can be described as
some kind of upside down pendulum, with the columns representing the beam of the
pendulum and the whole mass of the floor as a lumped mass at the end of the pendulum.
This structure supports equipment that produces horizontal dynamic forces. The
horizontal amplitudes are unacceptable. What can be done about this?

Dynamic Force

Concrete Floor

Figure 4.7: Effect of Additional Mass


Conventional wisdom would suggest that some concrete could be added to the floor so
that the damping characteristics of the concrete can reduce the amplitudes. The part
about damping is complete nonsense. Damping is caused by the internal friction within
the material. For damping to act the element has to deflect. In the example in Figure 4.7
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
21 of 92

ANGLO TECHNICAL DIVISION

the floor and the additional concrete are moving horizontally like a rigid block. Since
there is effectively no deflection of the concrete floor, damping is almost completely
inactive, and it can hardly play any role.
This example intends just to clarify the concept of damping. It does not matter what
action has to be taken to improve the performance of the system nor if increasing the
mass has a favourable effect or not.
4.2.5 Multi Degree of Freedom Systems
A general structure can move in many different ways, and in many different directions.
This is called a multi degree of freedom (MDOF) system. The basic concepts described
for SDOF systems can be transported to MDOF systems as well. By writing the
equations of motion for each degree of freedom we obtain a system of linear differential
equations that describes the movement of the model under harmonic forces, and whose
solutions describe the motion at every degree of freedom. If we put the condition that the
exciting forces are all equal to zero then we end up with a system of equations giving us
the free vibration of the structure. The mathematical condition for this system to have
non-trivial solutions is that the determinant of the characteristic matrix be zero. If the
number of degrees of freedom is n then we have to solve an equation of the nth degree.
The solutions are the n eigenvalues. We replace the eigenvalues in the system and find
the n eigenvectors. The eigenvalues are the radial natural frequencies, the eigenvectors
are the mode shapes.
Remember that the exciting forces have been set to zero. The non-trivial solutions, the
eigenvectors, or mode shapes, are ratios of displacements, describing the shapes in
which the structure will vibrate when no force is applied. Thus the mode shapes do not
give actual displacements in any physical units.
There are as many natural frequencies as there are degrees of freedom. In free
vibration, the structure will tend to vibrate with the fundamental (i.e. the lowest) natural
frequency. If there are some applied dynamic forces then the structure will try to vibrate
according to the mode shape whose natural frequency is closest to the frequency of the
dynamic forces, and the frequency will always be the frequency of the exciting forces. A
structure subjected to a periodical dynamic loading will vibrate with the same frequency
as the loading. That frequency is called the exciting frequency, or the forcing frequency.
Harmonic loading is the most common example of periodic loading.
4.2.6 Mode Shapes and More about Natural Frequencies
It would be "comfortable" to express a MDOF model in terms of a number of SDOF
models, which are easy to solve, as shown in Figure 4.8. Then solve them, add up the
results and find the solution for the complete structure. This is where the mode shapes
come in. Each mode shape is such an imaginary single degree of freedom system.
Solve them, add up the results (amplitudes) from each one and you have the total results
for your structure! Is it truly that simple?
It is basically true but far from being simple. So put your thinking cap on and keep
reading. Consider a MDOF system or model. Imagine a set of links that will force the
displacements of each node to have always a fixed ratio to the displacements of each
other node. For example, if the displacement of node 5 is 2,0 and the displacement of
node 7 is 2,5 then the ratio (2 / 2,5 = 0,8) will stay true, no matter what the magnitude of
each displacement is. If node 5's displacement is, say, 1,7 then the displacement of node
7 will be 1,7 x 0,8 = 2,13. The displaced position of the transformed system can then be
defined by single parameter. As stated earlier, this means that the new system has only
one degree of freedom.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
22 of 92

ANGLO TECHNICAL DIVISION

9
8

Height (m)

7
6

Mode 1

Mode 2

Mode 3
Overall

2
1
0
Mode Shape

Figure 4.8: MDOF System Represented as Several SDOF Systems


The computer will calculate the ratios between displacements and will print the results.
This is done according to certain rules. Those rules are not relevant for us. The result
will be a number of shapes equal to the number of degrees of freedom. These are the
mode shapes, also known as the eigenvectors. Each mode shape comes with its own
unique, personal and confidential natural frequency (also known as an eigenvalue). The
intractable multi-degrees-of-freedom system has been magically transformed in a sum of
comfortable single degree of freedom systems. The natural frequencies associated to
each mode shape are the natural frequencies of the system.
The next step is to calculate the responses of each of the single degree of freedom
system and to sum them up. The result will be the response of the complete structure.
To apply this method manually is not as simple as it seems. However, at this stage the
problem is to understand the concepts, not yet to apply them.
The figure above is not really correct, since it gives the impression that the amplitudes
are the sums of the mode shapes. The components of the mode shapes are not
"displacements". They are just non-dimensional ratios.
The method just described is called modal analysis. It can be summarized as follows:
The big, ugly, mean and hairy MDOF system is broken down into a number of cute
SDOF systems. The SDOF systems are solved one by one and the results are summed
up. The messy job of calculating eigenvalues and eigenvectors and then summing them,
is done by the computer. The designer just types in the data in whatever format the
program requires.
By comparing the mode shapes with the pattern of the amplitudes one can grasp
intuitively which modes are important in the response and which are not. It also
becomes clear what shapes must be changed. It is true that a modification to a mode
shape will usually influence other modes as well. The structure will be tuned by a series
of successive approximations. It still beats groping in the dark.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
23 of 92

ANGLO TECHNICAL DIVISION

5 LOADS
The dynamic loads applied to structures are sometimes quite easy to obtain, from simple
equations or from the equipment Suppliers. Other dynamic loads must be dragged out,
kicking and screaming, from the equipment Suppliers or somewhere else. To get the
dynamic reactions from a Supplier is arguably the most frustrating action in structural
dynamic design. The amount of weird data that sometimes floods us in answer to
technical questions is hardly believable. The following conversation is absolutely
authentic:
Q: What forces does your equipment apply to the supporting structure?
A: Well, our machine works very quietly. You can put a glass of water on top of
it and you will hardly notice a few ripples on the water surface.
Q: So the dynamic reactions are so low that they could be ignored?
A: Well, we specify a vertical dynamic force of about 150 kN at 1000 rpm.
(It is doubtful that even Fort Knox could withstand this kind of loading. The ripples would
be on the surface of the planet, not on the water in a glass)
Warning: Do not simply accept the loads specified by Suppliers of
equipment. Local Suppliers are often only agents for equipment sourced
from overseas, so they may not be Engineers, there may be confusion of
units from overseas countries, etc. It is always prudent to check that the
loads specified make sense. Only then should they be used as if they are
accurate.
Some guidance to determination of dynamic loads is thus provided here.
The vibration induced by various types of machinery is frequently of concern in the
structural design of buildings, in particular industrial and mining buildings. The magnitudes
of these vibrations are determined by the nature of the machines themselves and how they
are supported on the structure. Both of these influences will be considered below.

5.1

Rotating Unbalance

Rotating machines are designed to run at a constant speed for a long period of time. In the
case of some machines the intention is that eccentricity should be eliminated if possible.
These include turbines, axial compressors, centrifugal pumps, generators, electric motors
and fans. In the case of other machines, eccentricity is deliberately introduced in order for
the machine to function. These include vibrating screens, and vibratory feeders.
5.1.1 Motors and Turbines
Theoretically, it may be possible to eliminate all unbalance, but in practice it is impossible.
Static unbalance occurs when the centre of mass of a machine rotor does not coincide with
the axis of rotation. The term "static" refers to the fact that static forces, eg gravity, can
pinpoint this condition. Gravity will usually cause the out-of-balance rotor to rotate to a
position of static equilibrium in which the "heavy" side of the rotor is at the bottom.
Dynamic unbalance occurs when two or more masses in different planes on the rotor,
produce a moment when the rotor is rotating. In the simplest case of two masses at 180o to
each other, in different planes, the rotor may be statically balanced, but will tend to rock in
the bearings when rotating.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
24 of 92

ANGLO TECHNICAL DIVISION

Figure 5.1 : Rotor Unbalance to ISO 1940

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
25 of 92

ANGLO TECHNICAL DIVISION

Table 5.1 : Quality Grades for Different Machines to ISO 1940


Balance
Quality
Rotor Equipment Types
Grade, G
______________________________________________________________________
G 4000 Crankshaft drives of rigidly mounted slow marine diesel engines with uneven
number of cylinders.
G 1600 Crankshaft drives of rigidly mounted large two-cycle engines.
G 630 Crankshaft drives of rigidly mounted large four-cycle engines.
Crankshaft drives of elastically mounted marine diesel engines.
G 250 Crankshaft drives of rigidly mounted fast four-cylinder diesel engines.
G 100 Crankshaft drives of fast diesel engines with six or more cylinders.
G 40 Crankshaft drives of elastically mounted fast four cycle engines (petrol or
diesel) with six or more cylinders.
G 16 Drive shafts (propeller shafts, cardan shafts) with special requirements.
Crankshaft drives of engines with six or more cylinders, under special
requirements.
G 6.3 Marine main turbine gears.
Fans.
Flywheels.
Pump impellers.
Normal electrical armature.
Individual components of engines under special requirements.
G 2.5 Gas and steam turbine, including marine turbines.
Medium and large electrical armature with special requirements.
Small electrical armature.

Both static and dynamic unbalance manifest themselves as vibration at the running speed
of the rotor. The reason for this is simply that in both cases the centrifugal force due to the
eccentric mass is rotating at the running speed. The actual amount of unbalance present
may be difficult to ascertain, as manufacturers are often reluctant to admit that their
machinery has any unbalance. ISO 1940 provides some guidance on the balance quality of
rotating machines, giving the residual unbalance mass as a function of speed, as shown in
Figure 5.1. Different curves in this figure are appropriate for different quality grades, which
apply to different types of machine, as listed in Table 5.1. The eccentricity to be used in a
particular design may be obtained from the Supplier of a rotating machine, but failing this,
the ISO approach, using Table 5.1 and Figure 5.1, can be used to give an appropriate
design value.
The forces due to this unbalance are given by:
o

Fx ( t ) = Me 2 sin(t ) 0 transverse to rotor axis

Equation 5-1

Fy ( t ) = Me 2 cos(t ) 90o transverse to rotor axis

Equation 5-2

These two forces will always act at the same time, as the rotor turns.
5.1.2

Vibrating Equipment
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
26 of 92

ANGLO TECHNICAL DIVISION

Vibrating screens and vibratory feeders


typically consist of a mesh or solid bed within a
rigid frame which is supported on a number of
springs. They rely on motion of the bed in
order to operate. Eccentricity, which has the
same effect as unbalance, is thus deliberately
introduced on motors mounted below the bed,
usually in the form of two semi-circular
masses, as shown in figure 2. These masses
can be set at varying positions to give different
values of eccentricity, and hence different
force amplitudes.
Most commonly, two electrically coupled
motors are used, running in opposite
directions. This has the result that the forces
applied transverse to the axis of the machine
by the two motors exactly oppose each other,
giving a nominally zero resultant. The forces
applied parallel to the axis of the machine by
the two motors reinforce each other, doubling
their effect. Thus:
Fl ( t ) = 2Me 2 sin(t )

(longitudinally)

Ft ( t ) = 0

(transversely)

5.2

Figure 5.2 : Eccentric Masses

Loads Applied to the Structure


Warning: The loads defined here are in general NOT the loads applied to
the structure.
They are applied to the machine, which responds
dynamically. Go on to the next section before assuming you know all about
vibration loads.

5.2.1 Data Required From the Equipment Supplier


An assumption as popular as it is wrong is that the dynamic reactions of a screen
supported on springs or buffers are parallel to the exciting force produced by the
screen's vibrator. This means that if the exciting force is applied at, say, 30 degrees then
the dynamic reaction will also be applied at 30 degrees. This is usually not true because
the springs supporting the screen have different stiffnesses in the vertical and horizontal
directions, leading to different dynamic magnification.
Another popular superstition is that the dynamic reactions equal the weight of the
machine times some safety factor up to 10. This nonsense comes from confusing
something called "equivalent static load" with a dynamic reaction. An equivalent static
load is used mainly cover up ignorance or laziness, or sometimes to check a foundation
for overturning. A dynamic reaction is a periodical quantity, more often than not
harmonic. An equivalent static force is an imaginary STATIC loading.
So what do we need from the equipment Supplier in order to design the supporting
structure?

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
27 of 92

ANGLO TECHNICAL DIVISION

(a)
(b)

Either the direction, size and position of the forces applied by the equipment
Or appropriate data to enable us to calculate them.

For (a) above, we need the:


(i)
(ii)
(iii)
(iv)

working frequency of the equipment


dynamic reactions of the equipment
mass of the equipment
mass moment of inertia and position of the centroid

For (b) above, we need the:


(i)
(ii)
(iii)
(iv)
(v)
(vi)

working frequency of the equipment


mass of the equipment (kg)
mass moment of inertia and position of the centroid
static loads at each corner (N)
spring constants (vertical and horizontal) (N/m)
magnitude, position and direction of the exciting force (N)

In both cases, general data for the equipment may be required as well, not necessarily
Warning: It is extremely important to have the Supplier approve the loads
we have calculated. If the loads we use in design are not approved by the
Supplier, then we could be left holding the baby. As soon as that screen as
little as hiccups, for whatever reason, everybody could shrug and say "Not
my problem! I told you so. What do you intend to do to fix it?" We cannot
let them say this. This is our line.
for dynamic analysis.
5.2.2 Calculation of Spring Stiffness
But sometimes the Supplier is unable to provide the spring constants. They can be
calculated from the following equations:
(a) Steel springs
The axial stiffness for steel coil springs is given by:
kV =

Gd 4w

Equation 5-3

8D 3 n

The horizontal stiffness for steel coil springs is given by:


kH =

kV
h
0,385[1 + 0,77( ) 2 ]
D

Equation 5-4

where:
n is the number of free coils as shown in Figure 5.3
is a coefficient obtained from Figure 5.4
Other symbols are as defined in the symbols list.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
28 of 92

ANGLO TECHNICAL DIVISION

p = Pitch, d = Diameter of wire, n = Number of free coils


Plain ends Plain and Squared
Squared and
ground
ends
ground
Total coils
n
n
n+2
n+2
Solid length (n+1)d
nd
(n+3)d
(n+2)d
Free length
np+d
np
np+3d
np+2d
Figure 5.3: Steel Coil Springs

c/h=0,5
c/h=0,4

c/h=0,3
c/h=0,2

c/h=0,1

Figure 5.4: Values of for Equation 5-4


(b)

Rubber buffers
buffer height (mm)
diameter (mm)
characteristics of the rubber used

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
29 of 92

ANGLO TECHNICAL DIVISION

Note: It is good practice to check rubber buffer characteristics with the Supplier because
rubber properties are known to vary widely from batch to batch. The dynamic stiffness of
rubber buffers is significantly higher than the static one. A study done by Anglo
American Corporation in cooperation with VELMET showed that the dynamic stiffness
may be as much as 60% more than the static stiffness. (REPORT RAB/83/02 21 JULY
1983 "DYNAMIC STIFFNESS TESTS ON VELMET SCREEN SUPPORT SPRINGS:
VERTICAL STIFFNESS"). The horizontal stiffness of a rubber buffer is typically about
one third of the vertical stiffness.
(c)

Special springs
consult the spring Supplier

It also sometimes happens that the Supplier is unable to provide the magnitude of the
exciting force. In this case, the desired dynamic reactions can be calculated as
demonstrated in the following examples:
Warning: Even when the stiffness of the springs is given by the Supplier,
we cannot automatically assume that they are correct. The Suppliers often
underestimate the horizontal stiffness of a steel spring. It is prudent to
check the information supplied!
5.2.3 Example 1
Information provided by Supplier
Torque:
T
= 21 kgf.m = 206 Nm
Mass of screen:
M
= 3000 kg
Weight of screen:
W
= 3000x9,81 = 29430 N
Steel springs, two springs per corner
Spring wire diameter: d
= 0,020 m, height, h = 0,260 m
No of free coils:
n
= 7,5
Spring outer diameter
= 0,144 m
Spring diameter:
D
= 0,144 0,020 = 0,124 m
Exciting force at 65 to horizontal
(a)

Step 1: Calculate spring constants

The axial stiffness for one spring is:


kV =

78 x10 9 x0,02 4
8 x0,124 3 x7,5

= 109 x10 3 N/m

The static compression of a spring is:


W /8
29430/8
=
= 0,0338 m
KV
109 x10 3
c 0,0338
=
= 0,13
h
0,260
h
0,26
=
= 2,09
D 0,124

c=

The horizontal stiffness for one spring is:


kH =

109 x10 3
0,26 2
0,385 x1,5[1 + 0,77(
) ]
0,124

= 43,0 x10 3 N/m

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
30 of 92

ANGLO TECHNICAL DIVISION

(b)

Step 2: Calculate the stroke, S (the peak to peak displacement of the screen at 65
degrees)
S=

T
206
=
= 0,007 m
W 29430

The amplitude of displacement at 65 is:


The vertical amplitude of displacement is:
The horizontal amplitude of displacement:
(c)

d0
d0V
d0H

= S/2 = 0,0035 m
= d0 sin(65) = 0,00317 m
= d0 cos(65) = 0,00148 m

Step 3: Calculate the dynamic reactions per corner (2 springs):


Rv = 2x109x103x0,00317 = 691 N (vertical dynamic reaction)
Rh = 2x43x103x0,00148 = 127 N (horizontal dynamic reaction)

5.2.4 Example 2
Consider the same screen as that in Example 1. Assume that this time we know the
exciting force FO, and not the torque T. The other data stay the same. The information
from the Supplier is:
FO = 90456 N (at 65 )
The exciting screen runs at 900 rpm.
E = 94,25 rad/sec
Spring data is the same as above.
(a)

Step 1: Calculate the vertical and horizontal components of FO.


FOV = FOsin(65) = 81981 N
FOH = FOcos(65) = 38228 N

(b)

(c)

Step 2 : Calculate vertical and horizontal natural frequencies of the screen on


springs. Work with total mass of screen and with all 8 springs. Do not attempt to
split these they all form one single screen. KV and KH have been calculated in
Example 1 above.
NV =

8k V
=
M

8 x109 x10 3
= 17,05 rad/s
3000

NH =

8k H
=
M

8 x 43 x10 3
= 10,71 rad/s
3000

Step 3: Calculate the dynamic reactions on corner (2 springs)


RV =

RH =

FOV / 4
E 2 2
) ]
[1 (
NV
FOH / 4
E 2 2
) ]
[1 (
NH

81981 / 4
94,25 2 2
) ]
[1 (
17,05
38228 / 4
94,25 2 2
) ]
[1 (
10,71

= 693 N

= 125 N

The results are practically the same as the results of Example 1.


Note that the angle of the total dynamic reaction (i.e. the resultant of RV and RH) to the
horizontal is quite different from the angle of the exciting force to the horizontal, because
of the different dynamic magnification in the two directions.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
31 of 92

ANGLO TECHNICAL DIVISION

Angle of exciting force


Angle of dynamic reaction

= 65
= atan(RV/RH) = 79,8

What if the Supplier doesnt know the applied torque, the exciting force, or the mass of
the equipment? The Supplier has to know something!

5.3

Impact Loads

Impact loads are defined as those loads which are applied to structures for a short time
only, or which are suddenly applied to the structure.
5.3.1 Types of Impact Loads
The first possible source of impact is motion. The motion may be the movement of some
vehicle, which causes impact for example when a train collides with a station stopping
device. Alternatively, it may be due to a mass falling onto a structure below, such as for
example, when a conveyor belt breaks and its tensioning counterweight falls.
The important variables in determining the magnitude of this force are:
(a)
(b)

The impact velocity, v. This may be well known in certain instances, but in other
cases it may be necessary to make reasonable assumptions.
The distance over which the moving body is stopped, D. This is determined either
by the spring stiffness of the structure or buffer, or, in the case of plastic
deformation of either the structure or the moving body, by the extent of the plastic
deformation.

5.3.2 Energy Equations


It is usually appropriate to use energy considerations to calculate motion impact forces.
The equations are given in Table 5.2. The information in Table 5.2 gives the deflection
under impact and the maximum impact force. It is also often convenient to define an impact
factor, as:
=

F
Mg

If the impact factor, , is known then the impact force is:


F = Mg

Impact energy absorption may be either elastic or plastic strain energy. Where the energy
is absorbed by elastic strain energy, there is no permanent deformation of the buffer.
Where the energy is absorbed by plastic strain energy, the energy is absorbed mainly as
work done in causing the plastic deformation. There will also inevitably be a certain amount
of elastic strain energy, but this is usually small enough to be neglected. As the
deformation is plastic, it may be assumed that the force remains constant, at the yield
strength of the deforming member

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
32 of 92

ANGLO TECHNICAL DIVISION

Table 5.2: Energy Equations for Impact


Equations
Motion Impact
Impact energy
Kinetic Energy = 21 Mv 2

Falling Impact
Potential Energy = Mg(H + D)

Elastic Energy Absorption (i.e. no permanent deformation)


Absorbed energy Strain Energy = 1 kD 2
2

Conservation of
energy
during
impact event

1
2
2 Mv

D=v

M
k

Impact force

F = Dk

1
2
2 kD

D=

kD 2

Mg
2kH
1+ 1+
k
Mg

F = Dk

= v Mk

Impact factor

1
2

Mg(H + D) =

F = Mg 1 + 1 +

v k
g M

= 1+ 1+

2kH
Mg

2kH
Mg

= 1+ 1+

2H
D

Plastic Energy Absorption (i.e. permanent plastic deformation)


Absorbed energy Strain Energy = FD
Impact force
F is defined by the plastic strength of the buffer
Mg(H + D) = FD
Conservation of 21 Mv 2 = FD
energy
during
H
Mv 2
D=
impact event
D=
F
2F

Impact force
Impact factor

Mg

F=

Mv 2
2D

F = Mg 1+

F
v2
or
Mg
2Dg

H
D

F
H
or 1 +
Mg
D

This impact factor is plotted in Figure 5.5, for different ratios of the drop height to the static
deflection, H/D.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
33 of 92

ANGLO TECHNICAL DIVISION

Impact factor

5
4

Elastic impact
Plastic impact

3
2
1
0
0

Drop height to Deflection ratio, H/S

Figure 5.5: Impact Factors for Falling Bodies


5.3.3 Moving Mass Hits Stationary Mass
A final impact problem is the elastic impact force, and the resulting velocity when a
moving mass impacts a stationary mass, and both masses continue in motion at a
reduced velocity.
The new, reduced velocity is:
v S = 1,5

Mv
M + MS

The force applied at the point of impact is:


F = vS

5.4

MS k
2

Ground Motion from Blasting and Piling

5.4.1 Basic Equation


The ground motion generated by blasting, piling or other ground impact condition, is
approximately described empirically by the following equation:
v P = Ca (

W Cb
)
a

where:
a is the distance away from the point of blast or ground impact
Ca is a site constant defined below
Cb is a site constant defined below
vP is the ground peak particle velocity (m/s)
W is the mass of explosive per delay (kg) or the impact energy (J)
5.4.2 Blasting
The following values may be used for blasting in typical ground conditions:
W
Mass of explosive per delay (kg)
Ca
1,000
Cb
1,667
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
34 of 92

ANGLO TECHNICAL DIVISION

5.4.3 Piling
The following values may be used for piling in typical ground conditions:
W
Energy of falling pile hammer (J)
Ca
0,001
Cb
0,770

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
35 of 92

ANGLO TECHNICAL DIVISION

6 STRUCTURAL MODELLING AND RESPONSE


6.1

Modelling the Structure

The time has come to introduce a few new concepts related to taking a real, physical
continuous structure of steel and concrete, with paint, bolts, spillage and pigeons, and
turning it into something that a computer can deal with in discrete chunks. This is
generally referred to as modelling the structure.
6.1.1 Models
The first thing to recognise is that any real structure is an infinite continuum. It could thus
take quite a while, i.e. an infinity, to solve for each structure. That's where the model
kicks in. A model is a simplified mathematical representation of the real structure. The
use of models is not exclusive to dynamic analysis. They are used in static analysis as
well. When we calculate the bending moments in a straight beam we assume that the
beam only has one dimension, i.e. length. There is no such thing in nature. It is we who
create this imaginary entity, the model, in order to approximate stresses, displacements
etc in the real structure.
The only difference between a static model and a dynamic model is that the dynamic one
has to include the effect of mass and sometimes damping. It is a difference of detail, not
of essence. The concept is the same.
6.1.2 Degrees of Freedom
The deflected state of a structure is defined by certain parameters, usually the
displacements and rotations at joints, or nodes in the model. Each of these defined
parameters is called a degree of freedom. The simplest structure conceivable has one
degree of freedom. Real, useful structures may have hundreds or even thousands, of
degrees of freedom. The minimum number of independent parameters that completely
defines the deflected structure is the number of degrees of freedom.
Consider, for example, a pendulum with three masses lumped along its length, as shown
in Figure 6.1. If the beam of the pendulum is infinitely rigid then one parameter, the
rotation about the pinned support, will completely define the displaced position of all the
masses. The correct model will have one degree of freedom. If, however, the beam has
a finite stiffness then one parameter is not enough to completely describe the deflected
shape. Considering the pendulum in 2D, each mass can move sideways, and vertically,
and rotate. Each mass thus has three degrees of freedom, and the full model has nine
degrees of freedom. Generally, structural members are far more flexible in bending than
axially, so we may decide that for the purposes of analysing this pendulum, we can
justifiably assume that the links are infinitely rigid axially, but flexible in bending. In this
case, each mass has only two degrees of freedom, sideways movement and rotation, so
the complete model will have six degrees of freedom. If we believe that it is necessary to
analyse the behaviour of this pendulum in 3D, then each mass has six degrees of
freedom (i.e. displacement along, and rotation about, each of the three principal axes).
The complete model would then have eighteen degrees of freedom.
The first model has one degree of freedom. The second has up to eighteen. Both
represent the same structure. Which one to use is a matter of engineering judgement. It
is YOU who have to decide. There is no single reliable fool-proof rule that can solve this
problem. Having said this, it is only fair to add that in most cases a normal dose of
common sense should be enough to solve the problem.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
36 of 92

ANGLO TECHNICAL DIVISION

Infinitely
rigid links.
One degree
of freedom.

Flexible links.
Up to nine
degrees of
freedom.

Figure 6.1: Simple Pendulum Degrees of Freedom


A degree of freedom refers either to a joint displacement or to a joint rotation. If it refers
to a displacement then the quantities involved are length and mass. If it refers to a
rotation then the quantities are angle and mass moment of inertia. The bigger the mass,
the bigger the force required to impose a given linear acceleration. In a similar manner,
the bigger the mass moment of inertia, the bigger the moment required to impose a given
angular acceleration.
There is no mathematical or other difference between these two types of degrees of
freedom. They are conceptually identical. Do not confuse moment (force x length) with
mass moment of inertia (mass x length2). Although they are almost homonyms, they are
also completely different animals.
Warning: When modelling any structure, THINK! When you have finished
thinking, ask some questions, then THINK AGAIN!
6.1.3 Modelling Structural Geometry
Now that we understand what a degree of freedom is, we need to figure out where to put
them. Wise men (wise guys?) have conceived many mathematical algorithms and rules
to enhance the similarity between model and physical structure, and several packages
now automatically arrange the nodes within the structure. These in-built rules are a great
help, but they only apply after the Designer has decided upon questions such as which
parts of the structure should the model include, and whether to put nodes along the span
of each beam.
The first thing to remember is that a simple structure is easy to understand, whilst a
complex structure is complex. As the power of computers has increased, almost all
packages now run 3D models, rather than 2D models, and it is often assumed that this
gives better results. In general it doesnt, because models are inevitably much more
complex now, so the predicted behaviour may be difficult to understand. If your package
offers the possibility of doing a 2D model, in most cases this will give better results,
because you will understand them.
The second thing to remember is to sit back, relax, look at the structure and decide how
it is going to respond to the applied dynamic loads. If this doesnt help, then stick to
static analysis and design, and find someone who can visualise the response. Any part
of the structure that is moving and bending must be modelled as a member with accurate
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
37 of 92

ANGLO TECHNICAL DIVISION

mass and stiffness. Any part of the structure that is moving but not bending must be
modelled with accurate mass, but stiffness is irrelevant. Any part of the structure that is
neither moving nor bending is irrelevant, and it can be ignored, or modelled as a support
to the structure if necessary.
Generally, cladding and flooring on structures is not modelled, but remember that this
has implications. Floor beams and sheeting rails in a physical structure cannot move
sideways, because the cladding or the floor prevents this. However, if the cladding or
the floor is not modelled, the model allows these very laterally flexible members to flop
around at rather low frequencies, the structure cannot be high tuned, and so Designer
panic sets in. There is a simple solution to this dilemma. Dont model sheeting rails or
secondary floor beams either, but remember that the cladding, floors and secondary
steelwork do have some mass that must be modelled as lumped masses. When
creating a 3D model, also remember that a concrete or steel plate floor has very
significant diaphragm stiffness, whose omission may allow the model to develop bogus
modes of behaviour. It may be necessary to introduce some imaginary cross bracing
into the model to protect against this happening.
6.1.4 Modelling Mass
There are generally two approaches to modelling mass in structures. The simpler one,
the lumped mass method, simply calculates the mass of each element of the model, and
puts half of that mass at each end. It is no problem to model the lumped masses that
physically appear in the real structure.
This is accurate for blobs of material that physically occur at one place, but it is less
accurate for mass distributed along the element, mainly because the rotational effect of
the mass about the joints is ignored. This leads to the introduction of "consistent mass"
method, which is based on mathematical procedures that recognise the actual location of
mass throughout the structure, but are beyond the scope of this guide. This generally
gives the best model of the real masses in a structure, but it does take more computer
memory and more time for the analysis. However, with a modern computer, the time
required for the analysis will be similar for each of the three methods. Well, the
difference could be 2000%, but this would mean 40 ms (milliseconds) instead of 2 ms.
One must really be in a hurry for this to matter. Almost all commercial dynamic analysis
packages now use the consistent mass method.
The thoughtful Designer will now be realising that the imposed loads generally applied to
structures have an associated mass, and will be asking whether this must be included in
the model. Absolutely maybe! A sound dose of engineering judgement is required here.
First, remember whether the structure is high tuned or low tuned. Where a structure is
low tuned, much of the force causing the vibration is being resisted as an inertia force,
accelerating the mass. Extra mass is thus beneficial, and will reduce the dynamic
amplitudes. However, if the intention is to high tune a structure, extra mass will make it
more difficult to achieve high tuning. Then having achieved high tuning, extra mass will
move the structure towards resonance, leading to an increase in the dynamic
amplitudes. The normal recommendation is that if a structure is high tuned, then lumped
masses equivalent to approximately 20 % of the specified imposed load should be used,
but if the structure is low tuned, then no additional mass should be added. However, this
is not a hard and fast rule.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
38 of 92

ANGLO TECHNICAL DIVISION

6.1.5 Modelling Materials


The choice of characteristics for the material being used is the next issue to resolve.
Steel, being a reasonably agile material reacts fast to stimuli, so its dynamic properties
are generally very similar to its static properties.
Concrete, on the other hand is in a different category. Concrete is more of a laid-back
material, so it takes a harder push to get it going. The dynamic stiffness of concrete is
thus higher than its static stiffness, by about 10 % to 20 % for the frequency ranges
typically encountered in mining type structures. So always add VAT to the elastic
modulus of concrete to get the dynamic elastic modulus to use.
E Cdyn = 1,14xE cstat

Equation 6-1

6.1.6 Modelling Connections


There are six possible movements at each connection, i.e. displacement along, and
rotation about each of the three axes, as shown in Figure 6.2. The fixity of the
connection relating to each possible movement must be established.
Z -axis

Y -axis

X -axis

Figure 6.2: Possible Movements at Each Connection


Generally the dynamic displacements we deal with are small, and under these conditions
most connections behave as if they are rigid connections, but not in all cases.
Recommendations for different types of connection are given in Table 6.1.
Table 6.1: Recommended Connection Fixity
Connection type
Simple 2 bolt shear connection, with thin
end plate free of flanges
Simple shear connection with more than 2
bolts, and thin end plate free of flanges
Simple shear connection, with stiffened end
plate
Full moment connection

Displacement
along:
X
Y
Z

Rotation about:
X

Free

Rigid Rigid Rigid

Free

Free

Free

Rigid Rigid Rigid Rigid

Rigid

Rigid Rigid Rigid Rigid Rigid

Rigid

Rigid Rigid Rigid Rigid Rigid

Rigid

6.1.7 Modelling Floors


Three types of approximation are routinely made by practically everyone when designing
a steel floor supporting screens, feeders, or other materials handling equipment:
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
39 of 92

ANGLO TECHNICAL DIVISION

(a)

(b)

(c)

The stiffness and mass moment of inertia of the underpans, chutes and similar
stuff are often ignored, but they should be considered. Logically, these items
should be modelled in the same way as other equipment, but this requires
awkward calculations such as establishing the mass moments of inertia and
stiffness of the chute. An acceptably accurate analysis will result if the chute mass,
with or without contents, is modelled as lumped at its centre of gravity and
connected to the support points by rigid links, pinned at their ends. If the structure
is high tuned, then the chutes and under pans should be modelled with their full
operational contents. If the structure is low tuned, then the chutes and under pans
should be modelled empty.
A floor may be modelled accurately as a plane grid. What a computer program
means by plane grid may sometimes be subtly different from what a mere human
assumes. Here is a reminder of what the machine is doing:

Any horizontal translations are ignored

No horizontal loads can be modelled

Any moments about the vertical axis are ignored

The supports restrict vertical translations and/or rotations about the


horizontal axes
The model must use both the elasticity of the columns and the masses outside the
model but supported by those columns. The column stiffness (spring constant) is:
Warning: As far as the computer is concerned the quantities ignored are
NOT zero. They simply do not exist. If a designer thinks that horizontal
vibrations are significant for his structure then he must not use a grid but
some other type of structure, usually a 3D frame!
k = EA / H
where:

E is the elastic modulus


A is the column area
H is the column height

To ignore it is not acceptable, especially when using a modern computer.


To decide what masses to consider supported on that column may be
quite complex and involve more engineering judgement than rational
assessment.
For the initial sizing of a beam supporting vibrating equipment, the following formula can
serve as a first shot at the depth:
L2
L2
<D<
45 000
35 000

where: D is the depth of the beam (mm)


L is the span of the beam (mm)
Sectional properties are generally calculated as they would be calculated for static
analysis. The only exception to this may be where composite sections (concrete floors
on steel beams) are used. See Section 8.2.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
40 of 92

ANGLO TECHNICAL DIVISION

When building a computer model, the definition of member releases is crucial to its
success.
Figure 6.3: Typical Floor Grid Connections
If two beams 1 and 2 are framed at the same point into another beam 3 using pinned
connections as shown in section B-B in Figure 6.3, then only one of the two must be
released. The release must be in the direction of the relevant rotation. If no release is

Full deph end plate

3
A

A
1

2
B

Do not release

Moment connection
Do not release

Partial depth end plate


Do not release
Release
Block rotation of all
beams supporting floors

Section A-A

Release one side only


Section B-B

given then the machine will assume that the two beams are continuous over the support.
If both are released then the machine will assume that the beam they are framing into is
not supported against rotation, and it will put a rotational degree of freedom that is not
really there.
Beams 1 and 2 are physically identical, you may argue, so why is it correct to model
beam 2 as simply supported while beam 1 is fixed to beam 3? The answer is that, for
open section beams, the torsional moment of inertia is very much smaller than the
bending moment of inertia. Therefore the restraint imposed by beam 3 on beam 2 is
negligible. However, beam 2 does effectively restrain beam 3 against torsional rotation.
It does not matter whether the beam released is 1 or 2. If beam 3 were to be a closed
box section, then it has a much higher torsional stiffness, and its effect on the end
conditions of beams 1 and 2 would have to be considered.
A similar situation occurs when there are nodes placed along the span of a beam (to
provide amplitudes of vibration at those points), or when the computer model uses a
consistent mass formulation. The computer will insert a rotational degree of freedom
wherever it can, and unsupported beams will completely mess up the image of the
eigenvectors and eigenvalues, producing phantom modes that actually do not exist. This
is not a real situation, because the floor will physically restrain the beam from vibrating in
rotation about its own axis. The solution is to define fictional supports, restraining only
the direction of the appropriate rotation.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
41 of 92

ANGLO TECHNICAL DIVISION

6.2

Composite Beams and Floors

This is an uncertain area, because concrete is not as well behaved as steel, and the
connection between concrete and steel may, or may not, transmit vibration stresses.
Within ATD Structural Engineering, we use an adaptation of the SANS 10162-1 code of
practice, Section 17 Composite Beams. This has the advantage of being both userfriendly and accurate enough.
Warning: It can be bad news to use concrete in low tuned structures.
Concrete work is less accurate than steel. Contractors are required to
produce concrete slabs of at least a minimum thickness, and having a cube
strength of at least a certain amount. The mass increases in proportion to
the thickness, but the stiffness increases in proportion to the cube of the
thickness. The elastic modulus of concrete, unlike steel, increases with
increasing cube strength. At the low amplitudes typical of industrial vibration,
the friction between the concrete and the steel beams is likely to transmit
vibration strains, even in the absence of shear connectors, leading to an
effective stiffness that may be well above what was predicted and used in the
computer model. These factors mean that the actual frequency of the asbuilt beast may be well above the neat computer prediction. Think and plan
your modelling carefully!
The components of a general composite structure are shown in Figure 6.4

Internal slab
W1 W2

T
A

C1

Edge slab

C2

C3

Figure 6.4: General Composite Structure


In the analysis of a structure with composite beams, we first need to assess the effective
section properties to be used. This is done by the following three steps:
(a)

Determine the width of concrete that will act compositely with the steel. One of two
cases will apply:

Case 1 Internal Slab (slab extends on both sides of the steel beam)
(i)
The active slab width should not be taken to be more than a quarter of the
beam span (i.e. S/4)
(ii)
The active slab must not extend more than half the distance between the steel
beams on each side of the beam being considered (i.e. C1/2 or C2/2).
Case 2 Edge Slab (slab extends on one side of beam only)
The active slab width should not taken to be more than one tenth of the beam span (i.e.
S/10)
(iii)
The active slab must not extend inwards by more than half the distance to the
next steel beam (i.e. C2/2)
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
42 of 92

ANGLO TECHNICAL DIVISION

(b) Determine the effective thickness of the concrete slab.


The slab thickness is taken as the overall slab thickness T, provided that:
(iv)
The slab has a flat underside, or
(v)
The slab has corrugated steel forms where the height of corrugations does not
exceed of the slab thickness, or
(vi)
Where ribbed slabs are used, the rib width W1 is at least 125 mm, the rib
height A does not exceed 40 mm or 0,4T, and the width between ribs W2 does
not exceed 0,25T or 0,2W1.
In all other cases, the slab thickness is taken as the depth of the slab minus the height of
the ribs (i.e. T-A)
(c) Calculate the section properties for the composite section.
ROBOT and PROKON can both calculate the section properties for sections of a single
material, but not a composite material. In order to calculate the composite section
properties a spreadsheet has been written called Composite Beam Properties. This
spreadsheet is located on the network at G:/ENGINEERING/se/PRODUCTS AND
SERVICES/Design aids. The spreadsheet only gives section properties about the X-X
axis, because a beam supporting a concrete slab will not vibrate laterally (i.e. about the
Y-Y axis) nor will it vibrate torsionally. Any assumed large values of IY and J should thus
be entered.
The spreadsheet also assumes full shear connection between the concrete and the steel
beam, meaning that there is no slip at all. Where the composite slab is low tuned, it is
recommended that full shear connection should always be assumed, even if very few
shear connectors are used. Where the slab is high tuned, a reduced value of the
moment of inertia IXE should be calculated for the composite section, as specified in
SANS 10162-1 Section 17.3.1 (a), i.e.:

IXE = IS + 0,85p0,25 (IX IS )


where: IS is the moment of inertia of the steel beam only
IX is the moment of inertia of the composite beam as calculated from the
spreadsheet above
p is the fraction of full shear connection being used
This is a conservative assessment of the moment of inertia of the composite section, as
it will give a high estimate of the frequency where the floor is low tuned, and a low
estimate of the frequency where the floor is high tuned.

6.3

Confirming the Accuracy of the Model

How do we know that the response of the model to dynamic loadings will be reasonably
close to the response of the real structure? We don't! On the well established principle
of garbage in, garbage out a computer package will give us whatever follows from our
input. There are, fortunately some aids to help the careful Designer determine whether
the results from the model look reasonable or not.
(a)

Look at the first few mode shapes. Do they make sense, or are there members
flying off into space, at all sorts of crazy angles? If you understand what you are
looking at, the mode shapes will tell you most of what you want to know about the
performance of your model.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
43 of 92

ANGLO TECHNICAL DIVISION

(b)

A useful approximate method of obtaining the fundamental natural frequency of


any structure is the so-called Rayleighs method. In this method, the deflections
obtained by applying loads equivalent to the self weight of the structure, applied in
an appropriate direction (i.e. vertically for floor vertical vibration, horizontally for
overall building sway) are calculated.
W1=m1g
The loads are:
The deflections are:
W1 = m1g
y1
W2 = m2g
y2
W2=m2g
.
.
.
.
Wn = mng
yn
The fundamental frequency is then
given by:
N

1 =

g Wi y i
i =1
N

Wi y i

W1=m1g

W2=m2g

W3=m3g

W4=m4g

i =1

(c)

Calculate some key values by hand, or with a static structural analysis package
you understand well. Yes, believe it or not, in todays computer era there is still
value in doing some simple hand calculations! The first key value you can
calculate is the ratio of dynamic displacement to static displacement. If the
structure is high tuned (frequency ratio less than 0,7) then this ratio should be
between 1,0 and 2,0. If the structure is low tuned (frequency ratio more than 1,3)
then this ratio should be greater than zero, and less than 2,0. The second key
value you can calculate is the inertia force amplitude on the portion of the structure
directly supporting the machine. This is calculated by the formula:
FI =d o E2 M

Equation 6-2

If the structure is high tuned, the inertia force amplitude should generally be quite
small, and it must be less than the amplitude of the applied dynamic force. If the
structure is low tuned the inertia force amplitude should approach the amplitude of
the applied force, and it must be greater than the amplitude of the applied dynamic
force.
Warning: Build a computer model, and look at the mode shapes
before going any further. Mode shapes, with understanding, will tell
you more about the accuracy of the model than any other factor. Many
a stupid slip in modelling would have been identified early had the
interpretative value of the mode shapes been realised and utilised.

6.4

Modelling Machinery on Structures

A question frequently encountered when modelling structures supporting vibrating


machinery is whether the mass of the vibrating machinery itself must be included in the
model. The answer to this question is based on - you guessed it - frequency ratios. The
first frequency required may be called the machinery frequency, M. This is the lowest

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
44 of 92

ANGLO TECHNICAL DIVISION

natural frequency at which the equipment sitting on its springs or mounting pads will
vibrate, if its springs are rigidly supported. The second frequency required is the lowest
natural frequency of the structure without the equipment on it, S. The important
frequency ratio is the ratio between these two frequencies, RF = M / S.
Warning: Where machinery is supported on structures do not forget that it
has a significant mass. Where necessary, this mass must be considered in
the structural model.

Three different ranges of the ratio RF must be considered. The modelling of these three
ranges is illustrated in Figure 6.5
(a) RF < 0,25
In this low range, the machine has a low frequency relative to the supporting structure.
This means that the machine will tend to move on its supporting springs independently of
vibration of the structure. In this case the mass of the machine can safely be omitted
from the structural model, and the model is simply subjected to the loads applied through
the supports of the machine. Generally, where machinery relies on vibration for its
function, such as vibrating screens, vibratory feeders, etc, the supporting springs are
generally very flexible, so that this frequency ratio condition is easily met.
(b) 0,25 < RF < 1,50
In this intermediate range, the machine has a frequency similar to that of the supporting
structure. This means that there is a dynamic interaction between the machine and the
structure. In this case the mass of the machine must be modelled as one or more
separate degrees of freedom, connected to the structure through its mounts. Generally,
this condition does not occur because machinery is either mounted on flexible springs
(case (a)) or it is almost rigidly fixed to the structure (case (c)).
(c) RF > 1,50
In this high range, the machine behaves as if it is essentially rigidly fixed to the structure.
This means that the machine will tend to oscillate together with the structure with very
little relative movement. In this case the mass of the machine can safely be added to the
structural model as a lumped mass, and the model is then subjected to the loads applied
on the machine. Generally, where machinery does not rely on vibration for its function,
such as would be true for pumps, winders, crushers, etc, there is practically no flexibility
in the supports, so that this frequency ratio condition is easily met.
Warning: Where we are working in this high range, and machinery is
treated as a lumped mass, it is important to understand the influence of the
geometric location of the centre of gravity of the machine. Where the
centre of gravity is above the supporting structure, as is almost always the
case, the lumped masses must include lumped mass inertias to account for
the height of the centre of gravity, or the lumped mass must be added at a
node at the centre of gravity which is then connected to the structure by
means of rigid links pinned at their ends.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
45 of 92

ANGLO TECHNICAL DIVISION

Mass of machine
ignored

Mass of machine
modeled on supports

Mass of machine
added to structure

Low range

Intermediate range

High range

Figure 6.5: Modelling the Mass of Machinery

6.5

Calculation of Dynamic Response

6.5.1 General
Having built the model of the structure, it is generally necessary to use it to calculate the
response of the structure to one or more applied dynamic loads.

The required dynamic response is generally two different quantities.


(a) Displacement or acceleration amplitudes are required to assess human and
machine sensitivity levels.
(b) Stress amplitudes are required to calculate the fatigue life of the structure.
6.5.2 Significant Modes
The most important question faced by the Designer is how many modes should be
included in the response calculation. The default answer would typically be All of them,
but there may be times when the use of fewer modes gives a quicker result, which is
sufficiently accurate. The decision, as with much of vibration analysis, depends on
frequencies. Assuming that an accurate calculation of response is required, the
minimum number of modes that can be used is determined by including all modes with
natural frequencies up to at least 1,5 times the operating frequency of any vibrating
equipment supported on the structure.

Where a structure is high tuned, all natural frequencies are higher than the operating
frequency of the equipment supported on the structure. So, when a structure is high
tuned, a good prediction of the dynamic response will be given if only a small number of
modes are used in the analysis.
However, if a structure is low tuned, there may be many modes whose frequencies are
lower than the operating frequency of equipment supported on the structure. If the
equipment is a screen, say, running at 16 Hz, and the natural frequency of the 10th mode
is, say, 9 Hz, then the use of 10 modes in the response analysis will give completely
wrong results. The wrong results will always predict too low a response, so the error is
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
46 of 92

ANGLO TECHNICAL DIVISION

dangerous, and not conservative. For a low tuned structure with a screen operating at
16 Hz, say, it may be necessary to use 20, or 30, or 100 modes to ensure that all modes
with frequencies up to at least 24 Hz are included in the analysis.

6.6

Dynamic Analysis Computer Programmes

This section provides some advice regarding the use of the dynamic analysis
programmes available within ATD. This is not intended as a manual all the
commercially available packages have manuals, but rather it provides some pointers
regarding what works and what doesnt work, and what the programmes can, and cant,
do.
6.6.1 ROBOT V6
This is generally a very good program.
Warning: We have been advised by ROBOT V6 programmers that we could
modify the density of the elements supporting a distributed mass in such a
way as to include the added distributed mass.

Example.
If a 406x178x60 UB vibrates together with a distributed mass
of 100 kg/m then the total mass per metre would be 59,8 + 100 = 159,8 kg/m.
The density of steel is 7850 kg/m3. To get the new mass of 159,75 kg/m with
the same area 7,611x10-3 m2 requires a modified density of 20990 kg/m3 for
this particular beam.
One minor problem with this method: it does not work! The program does
not seem to understand it, and the results are erratic.
Lumped masses. In ROBOT the lumped masses are input in force units. The program
will make the necessary transformations. It does not matter if it is logical or not (it
probably isnt!), but we are stuck with this approach whenever we use ROBOT. Just
make sure that the forces (that are really masses) are given in all the directions in which
the respective joint can translate or rotate in the real structure, otherwise the result could
be wrong.
Distributed masses. ROBOT does not yet have the capacity to handle additional
distributed masses on the span of a beam.
6.6.2 PROKON.
PROKON is the one of the more common packages used within South Africa at present.
PROKON is written using several defaults or computer-specific settings, which must be
understood and altered if necessary. The following defaults must be noted:

(a) Application of Defaults


PROKON uses defaults set in the computer, not in the specific programme file. This
means that defaults set for a particular run will not be transferred if the data file is sent to
another computer, or even if it is later brought back into the same computer after the
defaults have been altered for different requirements on another project. Always check
settings when using PROKON.
(b) Mass
The default is that load case number 1 is taken as the self weight of the structure. Load
case 1 is thus assumed by default as the load case defining the structural mass. Any
lumped masses included in other load cases must be indicated.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
47 of 92

ANGLO TECHNICAL DIVISION

(c) Modes used to Calculate Response


By accepting the default settings, only the first 10 modes are used to calculate the
structural response. This is probably sufficient for high tuned structures, but it is very
Warning: The list below records common errors encountered when using
computer packages for vibration analysis. Take note! Be warned!

1.
Too many member/joint releases.
If too many releases are specified, a member or a node may end up
unrestrained in a particular direction. For example, if both ends of a member
have rotation about the member axis released, then the member is free to
spin. Or if all members framing into a particular node have their end rotation
about any global axis released, then the node is free to spin about that axis.
These will lead to zero divisions in the solution, which mathematically is not a
nice thing to do. The programme may refuse to work, or you may get very
strange results. Check your member and node releases.
2.
Wrong shear resistance.
Commonly, frame analysis packages are used where the structure has
concrete floors or shear walls. Because the package does not have any
finite elements to handle this slab construction, it is modelled as several
lumped masses. This is fine as far as mass is concerned, but these slabs
have very high shear stiffness. Do not ignore this. Wherever slabs are
modelled as lumped masses, some phantom stiff bracing members must also
be added in the appropriate directions to ensure adequate shear stiffness.
3.
Too few modes in response calculation.
Several computer packages allow the user the right to define how many
modes are used to calculate the structural response to any applied dynamic
load. If, say 10 modes are specified, only the lowest 10 modes will be
considered in the analysis. This is tempting, as it can speed up the analysis
significantly. Allow yourself to be tempted, but with care! ALWAYS USE ALL
MODES WITH FREQUENCIES UP TO, AND WELL BEYOND, THE
HIGHEST OPERATING FREQUENCY OF ANY EQUIPMENT ON THE
STRUCTURE.
4.
Model too complex
A computer model that is too complex is confusing. It has too many modes,
many of which are irrelevant, but which are calculated anyway. The
Engineer must always stay in control of the analysis, not let the computer
take over.
likely to be insufficient for low tuned structures. Refer to section 5.4.2, and adjust this
default accordingly.
6.6.3

Common Errors with Computer Vibration analysis

Warning: The files for a structural model may sometimes have to be


transferred from one computer to another for some reason. BEWARE when
this is done with PROKON. All default parameters revert to their default
values when files are transferred to another computer, because default
values are a function of the computer settings, and not the individual model.
All default values must be checked and reset where necessary.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
48 of 92

ANGLO TECHNICAL DIVISION

7 VIBRATION LIMITS
7.1

Introduction

Setting appropriate limits to vibration is one of the most vague and uncertain parts of
dynamic analysis and design. We need to consider how people respond to vibration,
how equipment and machinery are effected by vibration, and how the structure itself is
likely to suffer under the influence of continued vibration. The greatest degree of
uncertainty lies in the vibration limits which people can tolerate. It does not refer much to
the fatigue calculations, although they do have a high level of conservatism built into
them, so it is quite likely that a structure having a calculated fatigue life of, say, 5 years,
will survived unscathed for 10 years or more. When considering machines, the greatest
uncertainty is again people, this time how people forget to specify things, and then duck
and dive looking for scapegoats when something goes wrong.

7.2

Human Sensitivity

Human tolerance to vibrations varies not only from person to person, but the same
person may today be quite happy with a situation, complain bitterly tomorrow and wonder
the day after tomorrow why this structure is so bulky and heavy, since it does not vibrate
at all. The tolerance of the owner of a building who must foot the bill for remedial work,
or someone who thinks there must be an insurance claim, or a worker who is disgruntled
because salaries are too low, or a Consultant who can get paid for fixing the problem,
are all very different. Check out BS 6611 or SANS 2631 (ISO 2631), in which a Motion
Sickness Dose Value is defined for tolerance to low frequency vibration. This is related
to the percentage of people who will get seasick and vomit. The Designers decision is
whether that percentage of people vomiting is acceptable.
The codes of practice of the V.D.I. (Union of German Engineers) deny that there is such
a thing as allowable amplitudes and define something that represents the perceptibility of
the vibrations: almost perceptible, clearly perceptible etc.
However, we do need some criteria to determine the allowable amplitudes, so here goes.
There are three significant limits in human reaction to vibrations. These were defined by
earlier versions of ISO 2631 (which were probably more useful than the current version) as:
1.- Limit of comfort the Reduced Comfort Boundary
2.- Limit of efficiency the Fatigue Decreased Efficiency Limit
3.- Limit of health and safety the Exposure Limit
The Fatigue Decreased Efficiency Limits for vertical and horizontal vibrations are shown in
Figure 7.1 and 7.2 respectively. In both Figure 7.1 and 7.2 the limits are shown as a
function of frequency in (a), and as a function of exposure time in (b). In all cases, the
Reduced Comfort Boundary is obtained by dividing the Fatigue Decreased Efficiency Limit
accelerations by 3,15. The Exposure Limit is obtained by multiplying the Fatigue
Decreased Efficiency Limit accelerations by 2,0.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
49 of 92

ANGLO TECHNICAL DIVISION

RMS Acceleration (m/s2)

100

10

1 minute
1 hour
4 hours
8 hours
1 day

0.1
1

10

100

Frequency (Hz)

(a)

RMS Acceleration (m/s2)

100

10

1 Hz and 16 Hz
2 Hz and 11 Hz
4 Hz to 8 Hz
50 Hz

0.1
0.01

0.1

10

100

Exposure time (hours)

(b)
Figure 7.1: Fatigue Decreased Efficiency Limit for Vertical Vibration

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
50 of 92

ANGLO TECHNICAL DIVISION

RMS Acceleration (m/s2)

100

10

1 minute
1 hour
4 hours
8 hours
1 day

0.1
1

10

100

Frequency (Hz)

(a)

RMS Acceleration (m/s2)

100

10
1 Hz to 2 Hz
4 Hz
12,5 Hz
50 Hz
1

0.1
0.01

0.1

10

100

Exposure time (hours)

(b)
Figure 7.2: Fatigue Decreased Efficiency Limit for Horizontal Vibration
These limits are expressed in different ways by different codes. It is possible to become
very sophisticated about evaluating human sensitivity, as many of the more recent codes
do, but it is doubtful whether this is actually useful. ISO 2631 (earlier versions) and BS
6472 give specific numerical guidance regarding acceptable vibrations. Other earlier
codes use a factor typically called "K". K takes into account the direction (horizontal or
vertical), magnitude and frequency of the vibrations, position of the body and other
variables. The K factor is a function of the RMS value of the acceleration. K can be
expressed as a function of the amplitude (of displacement, acceleration etc). The
Specification AAC114001 adopts this K value approach. Table 7.1 gives the K values
used by Specification AAC114001. These limits are indicated in the spreadsheet
Design Aid DA6 Vibration Limits, located at G:/ENGINEERING/se/DESIGN AIDS.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
51 of 92

ANGLO TECHNICAL DIVISION

Table 7.1: K value Definition adopted by Specification AAC114001


Frequency (Hz)
Horizontal
Vertical
1 to 2
28aH
10aVf
2 to 4
56aH/f
4 to 8
20aV
160aV/f
8 to 80
In the above Table 7.1, aH and aV are the horizontal and vertical
mm/s2, and f is the cyclical frequency in Hz.

Undetermined
28a
33,5af0.25
160a/f
RMS accelerations in

Table 7.2 shows the K interpretation (limits of comfort) based on VDI 2057 Part 2 (1987)
Table 7.2: Interpretation of K Values
According to:
ISO 2631 (updated in 1982) &
VDI 2057 Part 2 (1986)
Reduced Comfort Boundary
1 minute ..
16 minutes ..
25 minutes ..
1 hour ..
2.5 hours..
4 hours ...
8 hours
16 hours
24 hours..

According to
VDI 2057 Part2 (1987)
Degree of perception

17.7
13.4
11.1
7.5
6.3
4.4
3.2
2.0
1.6
1.3
0.9
0.4
0.1

Very strongly perceptible

Strongly perceptible
Very well perceptible
Just Perceptible
Under limit of perception

The Specification AAC114001 allows the following K values:


(a)
(b)
(c)

Less than 4 hours exposure, i.e. where access is only required for short periods, K
7
Up to 8 hours exposure, i.e. where access is required for an entire shift, K 4,2
Up to 12 hours exposure, K 3,5.

An EXCEL spreadsheet giving a graphic representation of these limits is available at


g:/ENGINEERING/se/DESIGN AIDS Vibration limits.
Where there is vertical and horizontal vibration simultaneously, these must be combined
using the following equation:
K EQUIV = K H 2 + K 2V

where: KV = the K value for vertical vibration


KH = the K value for horizontal vibration
KEQUIV is then checked using the limits for vertical vibration.
As a matter of policy, if an outside Contractor feels that the allowable K should be
increased or decreased then he must contact ATD Structural Engineering for approval.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
52 of 92

ANGLO TECHNICAL DIVISION

7.3

Equipment and Machine Sensitivity

Believe it or not, equipment and machines are also sensitive to vibration. An interesting
fact of life seems to be that when discussing a purchase with Suppliers, the equipment
can accommodate almost any vibration that will be thrown at it, but once a purchase has
been made, and there is any malfunction of the equipment in service, the same Suppliers
claim that it is of course the ambient vibration that has caused the problem.
Warning: When negotiating the purchase of any equipment or machinery
that will operate in an environment that includes vibrating equipment,
ALWAYS insist that potential Suppliers specify (in writing, before any contract
is signed) the level of vibration that their equipment can tolerate. This must
then be checked against structural vibration analysis prior to any construction
work commencing. It can also be used afterwards if there is any dispute
regarding performance of the equipment.

In our experience, the most sensitive equipment likely to be located in a vibration


environment, is electrical switchgear, particularly soft start units, and area lighting.
Sensitivity of Spring Mounted Equipment
Experience has demonstrated that equipment supported on springs, such as vibrating
screens and vibratory feeders, may be sensitive to lateral vibration of the supporting
structure. The equipment is typically not designed to withstand lateral motion or lateral
forces, so fatigue cracking may result if the structure vibrates laterally.
Typically, the lateral vibratoin induced by this equipment is limited, because there is
nominally no lateral force. However, where the supporting structure is not symmetrical,
or other vibrating equipment also operates on the same structure, a lateral component of
vibration may be introduced. In order to ensure that there is little likelihood of damage to
this equipment, the lateral vibration must not exceed 10 % of the in-line vibration.

7.4

Structural Sensitivity

There are two aspects of structural sensitivity that require the Designers attention. The
first is brittle construction materials or finishes, where vibration at higher frequencies can
lead to cracking or dislodging of the material. The second aspect is the possibility of
fatigue damage or even failure due to the high number of stress cycles.
7.4.1 Brittle Finishes
Brittle finishes are not generally our concern in the mining environment. They include
things like tiled floors and walls, glazing, and poorly constructed brickwork. Research
into the likelihood of damage occurring to brittle finishes has tended to concentrate on
the ground motion leading to damage to building finishes. This work has determined that
the likelihood of damage is more closely related to ground velocity than to either ground
acceleration or ground displacement. A general, conservative rule of thumb, applied by
some design codes is that the ground velocity should not exceed 5 mm/s. This is
conservative, so if the ground velocity is less than 5 mm/s, think no further. Brittle
finishes will not be damaged. For more specific guidance, although still somewhat
conservative, use Figure 7.3.

The structural vibration velocities at which initial damage to brittle finishes may be
expected collected from various sources, and are given in Table 7.3. It should be noted
that observations of damage vary very widely. Damage is unlikely at velocities below
those in Table 7.3, but in many cases where the velocity was more than double these
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
53 of 92

ANGLO TECHNICAL DIVISION

values there was no observed damage. These values should thus not be taken as a
hard-and-fast rule, but as general guidance.

Peak ground velocity


(mm/s)

30
25
20
15
10
5
0
10

20

30

40

50

60

70

80

90

Frequency (Hz)
General masonry - Impact
Freestanding and brittle masonry - Impact
General masonry - Continuous
Freestanding and brittle masonry - Continuous

Figure 7.3: Ground Velocity Limits for Brittle Structures


Table 7.3: Structural Velocities at which Damage to Brittle Finishes may Occur
Type of Structure
Damage
Peak Structural
Velocity (mm/s)
Brick building
Appearance of first cracks
50
Brick building
Deep cracks
150
Brick building
Falling plaster
200
Concrete structure
Appearance of first cracks
200
Any
Appearance of cracks in tiles
30
7.4.2 Fatigue Life
Fatigue life is calculated in terms of structural design standards. This guide will not
attempt to teach the reader how to use SANS 0162 or BS 7608. This just serves as a
friendly reminder that, like it or not, fatigue exists. Wherever vibration is encountered,
the fatigue life of the structure must be calculated, to ensure survival. Vibration stresses
are generally low, but stress cycles quickly mount up to huge numbers.

There is just one word of warning! Computer analysis, and for that matter hand analysis,
give the maximum stress in one direction. Under vibration conditions, the vibration
stress varies between a positive maximum, and a negative minimum with the same
absolute value. The fatigue stress range is thus twice the stress calculated.
Warning: The fatigue stress range is usually twice the maximum
vibration stress calculated.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
54 of 92

ANGLO TECHNICAL DIVISION

8 DESIGN GUIDANCE FOR SPECIFIC EQUIPMENT AND


STRUCTURES
Warning: The loads given in this Chapter are given for guidance only.
Loads must always be obtained from Suppliers. If what you are given looks
suspicious, get the Supplier to check and explain!

8.1

Crushers

There are crushers and then there are CRUSHERS. Some produce negligible dynamic
loads. Others produce loads so big that it is impossible to support them on a
conventional structure. It is a big and crushing world out there.
8.1.1 Modelling Crusher Support Structures
Very few crushers are supported on springs. Most are supported directly on a
suspended floor or a concrete foundation. Others are supported not on proper springs or
buffers but on some funny little pieces of rubber or stuff. The single most important
difference between a crusher support model and a conventional model is that the mass
of the crusher must be included in that model, with mass moment of inertia and all the
trimmings.

The simplest way to model a crusher for a dynamic program is to define an element that
would have the same outline as the machine (normally a cylinder or a box-shape) and to
determine the mass such that the full crusher element would have exactly the same
mass as the real crusher. This automatically takes care of all the mass moments of
inertia.
8.1.2 Loads Applied by Crushers
The loads applied by crushers to the structures allocated the hazardous job of supporting
them depend on the type of crusher.

(a) Cone Crushers


Cone crushers consist primarily of a cone rotating and tilting about a
vertical axis inside a cylinder. See Figure 8.1. The cone is
mounted eccentrically with respect to the cylinder, so that as it
rotates it crushes rock falling between it and the cylinder. The
rotation speed is generally slow, in the region of 2 Hz to 4 Hz, but
the cone mass is relatively high because the cone is heavily
constructed, and the eccentricity is quite high or rock will not be
crushed.
The centrifugal load will be more-or-less horizontal and sweeping
360 degrees. The simplest (and acceptably accurate) way to
handle it is to analyze the structure under two separate horizontal
harmonic loads, applied at 90 degrees from each other. The results
will then be evaluated using engineering judgement. It is important
to recognise that the centre of gravity of crushers is above their
support point, so these forces will inevitably and unavoidably induce
dynamic moments (reflected as vertical support forces) as well.
Remember!

Figure 8.1: Cone


Crusher

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
55 of 92

ANGLO TECHNICAL DIVISION

Example
Crusher power:
Throughput:
Feed size:
Discharge size:
Total crusher mass:
Cone size:

520 kW
500 to 600 tons/hour
- 250 mm
- 25 mm
68 000 kg
3,1 m high x 2,0 m diameter

Forces:
m1
15111 kg
m2
474 kg
m3
122 kg
r1
0,013 m
r2
0,0318 m
r3
0,4286 m
Speed
220 rpm, i.e. 220/60 = 3,667 Hz
Speed ()
3,667x2 = 23,04 rad/s
F1
15111x0,013x23,042 = 104280 N
F2
474x0,0318x23,042 = 8001 N
F3
122x0,4286x23,042 = 27757 N
F1 + F2 - F3
Total
horizontal
= 104280 + 8001 27757 N
force
= 84524 N
Distances below bearing:
Crusher
2566,5 mm
support
m1
1085,3 mm
m2
2784,8 mm
m3
2717,3 mm
Total
F1(2,5665-1,0853)
+
F2(2,5665overturning 2,7848) - F3(2,5665-2,7173) Nm
=
104280(2,5665-1,0853)
+
moment
8001(2,5665-2,7848)

27757(2,5665-2,7173) Nm
= 156899 Nm
Figure 8.2: Cone Crusher
(b) Jaw Crushers
Jaw crushers consist primarily of a fixed steel plane and a
moving steel jaw. See Figure 8.3. The moving jaw is pivoted
at its base, and is thrust towards and away from the fixed
steel plane by an eccentric mass or an eccentric shaft. The
motion is primarily horizontal. Rock falls between the jaw and
the fixed plane as the jaw moves away from the fixed plane,
and it is crushed as the jaw moves back towards the fixed
plane. The speed is generally slow, in the region of 1 Hz to 4
Hz, but the moving jaw mass is relatively high because it is
heavily constructed, and the eccentricity is quite high or rock
will not be crushed.
The dynamic load generated by the action of jaw crushers is
essentially horizontal, and may be idealised as a single
harmonic load, applied in the direction of jaw motion.
Figure 8.3: Jaw Crusher
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
56 of 92

ANGLO TECHNICAL DIVISION

(c)

Roller Crushers and Mineral Sizers

Roller crushers and mineral sizers


consist primarily two contra-rotating
rollers, with rotation about a horizontal
axis. See Figure 8.4. Rock falling
between the rollers is crushed as it
passes through the narrowest passage
between the two rollers. The rotation
speed is generally slow, in the region of
0,5 Hz to 3 Hz, but the roller mass may
be relatively high because the rollers are
heavily constructed.
The rollers are
nominally concentric to their axes.

Figure 8.4: Roller Crusher

The dynamic load applied by a roller crusher to its supporting structure is small, because
the operation of the crusher does not rely on any eccentric motion of heavy components.
Because the rollers are contra-rotating, any dynamic loads from the two rollers will tend
to compensate horizontally, but be additive vertically. A relatively small vertical dynamic
load should be anticipated on the support structure for roller crushers. It may
conservatively be assumed that the eccentricity due to construction tolerance and wear
will be of the order of 1 % of the roller radius. This results in a dynamic load amplitude
that is generally less than 10 % of the weight of the rollers.
(d) Flywheels
Remember also that crushers use heavy flywheels. These are rotating masses, where
the intention is that the mass is concentric. However, due to manufacturing tolerances,
this will not be the case, so flywheels should be treated in the same way as motors and
turbines.
(e) Dynamic Loads due to Breaking Rocks
Breaking rock causes additional random loads to be applied to the crusher supporting
structure. However, because these loads are random, and they are generally relatively
small, they are usually not considered.

8.2

Rotating Tubes

A range of equipment is essentially composed of rotating tubes, supported by rollers at


each end or along the sides. The equipment includes mills, scrubbers, trommels,
cement kilns, and peletizers.
8.2.1 Types of Load Generated
In the ideal world (wherever that might exist!) these pieces of equipment are circular, and
just turn neatly about their axis, without imposing any vibration loads onto their
supporting structure. However, in the real world, things are never so simple. There are
several sources of vibration loads that may be applied by rotating tubes.

(a) Ovalling of the tube


Where the construction of the rotating tube is fairly light, and allows a small distortion of
the circular tube into an oval or some other shape, there is a vibration load applied as the
longer axis transfers the weight of the equipment and contents from one roller to the
other, as shown in Figure 8.5.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
57 of 92

ANGLO TECHNICAL DIVISION

The frequency of this vibrating load is determined by the number of lobes in the imperfect
shape multiplied by the rotational speed of the equipment. See Figure 8.6. Generally,
the number of lobes will be determined by the construction details of the tube.
The amplitude may be as much as the nominal reaction at the relevant roller, because in
the extreme the load may vary from zero up to twice the nominal reaction. Our
experience, however, suggests that a realistic assumption is that the load varies between
50 % and 150 % of the nominal constant load. This leads to a dynamic load with an
amplitude of 50 % of the nominal load. It must be remembered that the load applied to
the support rollers must go through the centre of the bearings, so it has a vertical
component and a horizontal component. Dont think that just because this is a dynamic
load it will let you off easily! Resolution of forces still applies. If the loads on both rollers
are equal, the net horizontal load is zero. However, if the two loads are not equal, there
is a resulting horizontal dynamic load app-lied to the supporting structure.

Figure 8.5: Ovalling of the Tube

4 lobes

2 lobes

3 lobes

Figure 8.6: Different Numbers of Lobes on the Tube


(b) Material falling off lifters
Some mills and trommels have lifters mounted on their inside surface to lift and mix the
contents thus ensuring adequate processing. As material falls off the lifters, there is
some tendency to generate oscillatory loads on the tube.
The frequency of this load is well defined by the number of lifters and the rotation speed
of the tube. The amplitude of the load is more difficult to define. However, unless there
is resonance between the lifting frequency and some natural frequency, this is unlikely to
be a problem load, so avoiding resonance is the key design consideration.
(c)

Misalignment of girth gear or cutting errors in girth gear or drive gear

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
58 of 92

ANGLO TECHNICAL DIVISION

Misalignment of the girth gear or the motor drive shaft, or poor cutting of the teeth on
either the girth gear or the drive gear may lead to vibration at the frequency of gear teeth
intersections. This is commonly in the frequency range of 25 Hz to 60 Hz.
8.2.2 Specific Equipment
(a) Mills
Mills are typically heavy pieces of equipment that are usually mounted directly onto a
heavy concrete foundation, composed of a thick base slab and a thick plinth at each end
to support the mill.

Experience shows that as an approximate general rule of thumb, the fundamental natural
frequency of pitching (i.e. rocking in the mill axis longitudinal direction) and of rolling (i.e.
rocking transverse to the mill axis direction) should both not be less than 3,0 Hz, when
calculated with the mill carrying its normal full load of material.

8.3

Vibrating Screens and Feeders

8.3.1 Basic Requirements


Vibrating screens and vibratory feeders are simple eccentric mass machines. They are
supported on flexible springs to beams below, or on cables and flexible springs to beams
above, as they rely on vibration to function, and flexible springs enable vibration without
imparting large forces to their supporting structures. This means that vibrating screens
and vibratory feeders are low tuned, leading to large displacements when they are shut
down. The shut down displacements should be obtained from Suppliers, and adequate
clearances must be provided to avoid vibrating screens and vibratory feeders striking
surrounding objects. The clearance should never be less than 100 mm.
Warning: Always check shut down displacements with Suppliers and
ensure adequate clearance around vibrating screens and vibratory
feeders. Dont believe they cant give it to you!

As a rough guide, the peak-to-peak displacement of vibrating screens is typically in the


range from 6 mm to 10 mm, a little less for vibratory feeders. This means that the
vertical amplitude of motion is about 2 mm to 3 mm, and the horizontal amplitude is
about 2 mm to 4 mm. The total dynamic load (sum of loads at all four corners) applied
by vibrating screens and vibratory feeders to supporting structures is typically a few
percent of the weight of the screen or feeder. If the total dynamic load is given as less
than 1 %, dont believe it. If the Supplier shows that this is correct, then check the static
deflection of the springs, because they will have to be VERY flexible. If the total dynamic
force is greater than 10 % the screen or feeder will shake the teeth out of the structure
too quickly, although with rubber blocks or buffers the total dynamic force may actually
approach this level.
When working with vibrating screens or vibratory feeders, the following aspects may
cause difficulties, and should be checked:
(a)

The dynamic loads given on drawings of screens or feeders are usually given per
spring, or per corner. Check that you are satisfied which has been specified. If
there is any doubt, check with the Supplier. Small screens and feeders typically
have one spring at each corner. Large screens or feeders may have two, three, or
even more springs at some corners. Frequently, on large screens and feeders,
there are more springs at the feed end, where material drops onto the screen or
feeder, than at the discharge end.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
59 of 92

ANGLO TECHNICAL DIVISION

(b)

The excitation load to screens and feeders can usually be adjusted, by setting the
eccentricity of the eccentric masses. Increasing this load is a possible way of
improving throughput, or setting it correctly may be overlooked during
commissioning, so excessive vibration may result from incorrectly adjusted
eccentric masses.
Rubber blocks typically last much longer than steel coil springs, so coil springs may
be replaced by rubber blocks. As rubber blocks are much stiffer than steel coil
springs, this leads to much higher dynamic loads being applied to the structural
supports. On numerous occasions, this has been found to be at least in part
responsible for reported high vibration levels.

(c)

8.3.2 Design and Use of Sub-frames


(a) General
A sub-frame is a mechanical device that absorbs part of the energy transmitted from the
equipment to the supporting structure. The reason for using sub-frames is to reduce the
dynamic reactions. A typical sub-frame looks more or less as shown in Figure 8.7. The
physical shape of the sub-frame is often determined by process considerations, such as
underpans and chutes. This reduces the options of structural optimization.

Screen outline
dotted

Sub-frame

Springs supporting
sub-frame on structure
Fig 8.7: Schematic Layout of Typical Sub-frame
(b) Reactions On the Sub-frames and the Structure
The screen reactions on the sub-frame and the sub-frame reactions on the supporting
structure are determined by the dynamic behaviour of the screen and sub-frame
assembly. A model of the screen and sub-frame must be created to determine this
dynamic behaviour. The approach for this is generally the same as for a conventional
structure.
(i)

Use ROBOT or PROKON. The model must contain the screen, the springs
between the screen and the sub-frame, the sub-frame, and the springs
between the sub-frame and the support structure.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
60 of 92

ANGLO TECHNICAL DIVISION

(ii)

(iii)

The springs can be modelled as spring elements in ROBOT, or as bar


elements in PROKON. To model the springs as bar elements, the effective
properties must be calculated either by hand or using the spreadsheet Design
Aid DA11 Equivalent Spring Dimensions for Modelling Screens which may be
found at G:/ENGINEERING/se/DESIGN AIDS. Use of the spreadsheet is selfevident.
In order to model the screen itself one must:
Either approximate the position of its centre of gravity, and lump the whole
mass of the screen there, then calculate the mass moment of inertia of the entire
screen about its centre of gravity. A single node at the centre of gravity of the
screen is then given this mass and mass moment of inertia, and is connected to
the springs by using rigid links pinned at their ends.
Warning: There are some restrictions in the use of rigid links. The
restrictions depend upon the specific program. Consult the manual!

Or the screen can be modelled (using some good deal of engineering


nouse, i.e. good sense and judgement) as a grid of beams connected to the
springs. This grid of beams must be braced to represent the diaphragm action of
the screen, and their mass and mass of moment of inertia must accurately
represent the whole screen structure.
Warning: Check the implications of tolerances and spillage on the operation
of sub-frames. Get it wrong, and a sub-frame may lead to the vibration loads
applied to the supporting structure being substantially larger, not smaller.

The forces in the springs between the sub-frames and the supporting structures are the
loads finally applied to the structure.
When using sub-frames, it is crucial to investigate the influences of tolerances and other
effects. The steel from which sub-frames are generally constructed has rolling
tolerances of up to 4%. There is also a strong likelihood of a certain amount of spillage
accumulating on the sub-frame quite quickly. The springs supplied have tolerances in
their stiffnesses. Now, the point is this. The effective operation of sub-frames depends
all of these factors. It is recommended that the Designer should check the effects of the
following what ifs:
(i)
(ii)

Mass of the sub-frame oversize and accumulated spillage increasing the subframe mass by 20 %, and the springs having a stiffness of 10 % less than
specified.
Mass of the sub-frame is unlikely to be undersize, so do not reduce the subframe mass, but use springs having a stiffness of 20 % more than specified.

The structure must then be designed for the highest loads arising from the nominal
design conditions or either of these what if scenarios.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
61 of 92

ANGLO TECHNICAL DIVISION

8.4

Rock Breakers

Rock breakers are mean beasts! The loads they can exert are big. Their main purpose
is to break rocks that are too big to fit through grizzlies, but their Operators will use them
to push large rocks around on the grizzly as well. The Operators also try to push large
rocks into a corner of the bin, so that they can be broken by hammering them where they
cant escape! It is thus important to know the magnitude of the loads (FB the breaking
force, FP the pushing force, and FQ the slewing torque) that they can apply in the various
different directions shown in Figure 8.8. The breaking load FB is also made up of two
different components, a quasi-static load and a hammer load, as shown in Figure 8.9.
(Note that the actual magnitude of the loads given in Figure 8.9 only applies to one
specific rock breaker. The actual values must be established in each particular case).

FQ

FB

FP

Figure 8.8: Loads Applied by a Typical Rock Breaker

70

Rock breaker force (kN)

60
50
40

Quasi-static force
Ham m er force
Total breaking force FB

30
20
10
0
0

0.5

1.5

Tim e (seconds)

Figure 8.9: Breaking Load FB


So, when working with rock breakers, it is important to talk in detail to the Supplier. At
least the following information must be obtained and included in design considerations:
(a)
(b)

The hammer load.


The maximum push that can be exerted vertically and horizontally, from which the
quasi-static component of the breaking load and the push load can be obtained. If
the Supplier cannot provide this information, it can be calculated from the
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
62 of 92

ANGLO TECHNICAL DIVISION

(c)

maximum thrust of the hydraulic cylinders controlling the boom, which the Supplier
should be able to provide.
The maximum slewing moment.

It is not generally the Structural Designers task to check or ensure integrity of the rock
breaker, but all of these loads induce reactions onto the supporting structure, whose
integrity is the Structural Designers responsibility.

8.5

Design of Grizzly Bars

The function of grizzlies is to keep oversize


material out of crushers and off conveyor
belts. Grizzlies are one of the structures
Structural Designers have to cope with that
really take a pounding. Large rocks (the
how big question is often not satisfactorily
answered) fall from the back of haul trucks
or LHDs (Load Haul Dumpers used
underground), a few metres above the
grizzly with frightening amounts of energy.
This Section describes an approximate
evaluation of the maximum stress
produced by a mass that hits one beam at
mid-span with a known speed.
Figure 8.10: Large Rock destined for Grizzly!!
The stress is calculated by assuming that the kinetic energy of the hitting mass is
transformed into strain (deformation) energy. By equating the two energies it is possible
to calculate the force that, if applied statically at the point of impact, would produce strain
energy equal to the kinetic energy of the hitting mass. See Section 5.3.
An EXCEL spreadsheet is available at G:/ENGINEERING/se/DESIGN AIDS Grizzly
design to perform the necessary calculations. The operation of the spreadsheet is selfexplanatory. This spreadsheet allows the use of billets, or other structural sections for
the grizzly bars. It also allows for loss of energy due to fracturing of the rock during
impact. The spreadsheet allows any amount of energy loss, but it is recommended that
this should never be set at more than 10 % energy loss.
The most important thing with the design of grizzlies is that a lot (A LOT!) of engineering
judgement must be used when evaluating the results. The following assumptions are
made in the spreadsheet Grizzly design:
(a)

(b)
(c)

The equations used for the programs assume that the falling mass is completely
stopped by the grizzly bar, unless the grizzly is angled at more than 45 above the
horizontal. This is obviously not true, as can easily be proved by watching a grizzly
at work during about 10 seconds.
The grizzly bars do not collapse when the yield stress is reached. The Designer
must thus check the compactness of whatever structural section is chosen.
No overall lateral torsional buckling is possible under this type of loading. This is
generally true when the grizzly has a rectangular grid of bars, but may not be true if
the bars run in one direction only. The Designer must ensure an adequate
structural design.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
63 of 92

ANGLO TECHNICAL DIVISION

(d)

The conditions are known. There is always the requirement for interaction
between the Structural Designer and the Mine to get as close as possible to this
assumption, but there are still difficulties. Questions that still arise include:
(i)
What is the height of the lip of the haul truck bucket when rocks fall? Is it the
closed height or the fully tipped height, or somewhere in between?
(ii)
How big are the biggest rocks?
(iii)
Are rocks generally dumped onto the clear grizzly, or onto a pile of other rock
on the grizzly?

8.6

Vessel Agitation

8.6.1 Applied Loads


Agitation of fluids in circular vessels is used in numerous processes in mining and paper
production. Flotation cells in mining applications are a typical example. The rotation
speed of the agitators is usually quite slow, of the order of 60 rpm (1 Hz) or even less.
However, experience in ATD has shown that the need for various services below the
vessels leads to little bracing being used in support structures which are thus rather
flexible laterally. This, linked to the relatively high mass of the vessels contents leads to
quite low natural frequencies, and the distinct possibility of resonance. Light walkways
are also often provided for access to machinery above the vessels, which may well also
have quite low natural frequencies.
Warning: When taking measurements of vibration induced by vessel
agitation watch the minimum frequency range of the instruments used. The
RION VA10 instrument available in ATD Structural Engineering does not
give accurate measurements below 3 Hz because of built in high pass
filters. The RION SA78 instrument available in ATD Structural Engineering
measures down to 1 Hz with accuracy.

The agitators frequently have three or four blades (or paddles) and there are vanes, or
baffles around the perimeter of the tank. This may lead to a higher vane passing
frequency, which is given by the agitator rotation frequency multiplied by the number of
vanes. Experience in ATD, however, suggests that this higher frequency is seldom the
culprit in vibration problems related to vessel agitation.
A typical agitator shaft has a torque and an axial force applied, and it may well also have
a bending moment applied. All of these forces should be obtained from the Supplier of
the agitator, but the description below allows approximate values to be determined if
necessary.
(a) Torque
The torque FQ arises from the need to swirl and mix the liquid in the tank. Normally, it is
derived from the power of the drive motor. The maximum value of torque is about three
times the motor power divided by the agitator speed E, because electrical motors do
strange things on start-up.
FQ 3

Power
Power
Power
29
=3
E
2 f E
rpm

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
64 of 92

ANGLO TECHNICAL DIVISION

LB

Two
blades

Figure 8.11: Vessel Agitation


(b) Axial Load
The axial force FA arises from the length and angle of the blades, the applied torque and
turbulence in the fluid. The axial force will be directed upwards or downwards,
depending on the blade angle and the direction of rotation. An approximation for the
axial force may be obtained from the equation:
FA =

FQ
L B tan

where:

LB is the length of the blade


is the angle of twist of the blade

(c) Bending Moment and Shear Load


The bending moment FM applied by agitators to their support structure is difficult to
determine, because it is determined by the very complex fluid-blade interaction. Also,
broken or damaged blades, and misalignment of the agitator shaft will tend to
significantly increase the bending moment. Typically, the bending moment is a similar
order of magnitude to the applied torque. The shear force FV is equal to the bending
moment divided by the vertical distance between the blades and the underside of the
gearbox.
8.6.2 Design Requirements
The loads defined above are the maximum quasi-static loads to be used for ensuring
adequate strength of the agitator support structure. However, there are also varying
loads due to operation of the agitator. ATD has experience of several cases of fatigue
damage induced by these loads, and some experience of unacceptable low frequency
vibration of the tanks on their supporting structures.

(a) The following quasi-static loads should be considered for strength design:
Torque:
Maximum load specified by Supplier
Bending Moment:
Maximum bending moment specified by Supplier
Lateral Load:
Maximum bending moment divided by length of shaft from gearbox
to blades
Axial Load:
Maximum load specified by Supplier
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
65 of 92

ANGLO TECHNICAL DIVISION

(b) The following fluctuating loads should be considered for fatigue design:
Torque:
Actual peak startup torque, or 50 % of maximum load specified by
Supplier, applied once for each startup
Bending Moment:
50 % of maximum bending moment specified by Supplier, applied
once for each rotation of agitator shaft
Lateral Load:
Bending moment divided by length of shaft from gearbox to blades
Axial Load:
Operating axial load specified by Supplier, 50 % of maximum load
specified by Supplier, applied once for each startup
(c) The following dynamic loads should be considered for vibration assessment:
Bending Moment:
Load amplitude equal to 50 % of maximum bending moment
specified by Supplier, applied at the rotation frequency of agitator
shaft
Lateral Load:
Load amplitude equal to bending moment divided by length of
shaft from gearbox to blades, applied at the rotation frequency of
agitator shaft

8.7

Wood Chippers

An early part of the processing of logs into paper consists of reducing the logs to small
chips. See Figure 8.12. This is done by means of wood chipper machines. Wood
chippers are rotating machines with fairly heavy chipping heads which double as
flywheels, so there will be a component of dynamic excitation related to their rotation, as
for any other rotating machine. However, there is an additional impact force component
as the blades strike the log to reduce it to chips.

Log being chipped

Figure 8.12: Schematic End View of Wood Chipper Head

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
66 of 92

ANGLO TECHNICAL DIVISION

9 PRACTICAL GUIDELINES FOR FOUNDATIONS


This section deals with the design of concrete block foundations, used for vibrating
equipment mounted at, or near, ground level. Typical equipment in this category
includes large electricity generators, compressors and crushers.

9.1

Traditional Rules of Thumb

The most common traditional rules of thumb used to design the concrete foundations
for small equipment simply require that the concrete block foundation has a mass of
more than 10 times the machine mass for a reciprocating machine, and more than 5
times the machine mass for a rotating machine. These simple rules may be used for
small machines, with a mass of up to 500 kg and a power output not exceeding 50 kW,
but are not good enough for larger machines.

9.2

Simple Rules

If it is assumed that the soil supporting a machine foundation is very flexible, so that it
provides very little resistance to small amplitude vibration motion, then the dynamic
forces generated by the machine only accelerate the mass of the machine and
foundation. Under these conditions, the amplitude of base motion in various different
places and directions may be described by the simple equations given in Table 9.1. In
these equations symbols are as defined in Figure 9.1 and e is the machine eccentricity,
MR is the mass of the moving portion of the machine, MS is the mass of the static portion
of the machine, MB is the mass of the foundation, M is the total mass (i.e. MR + MS + MB),
and Ib is the mass moment of inertia of the machine and foundation.
These equations give conservative predictions for foundation motion provided the
foundation is not in resonance with the machine speed.
Table 9.1: Displacements Based on Simple Foundation Motion Equations
Location and direction
Displacement Equation
Vertical or horizontal linear motion
M
d= R e
M
Vertical at edge of base due to rotation
M h b
d = R CG e
2Ib
2
Horizontal at centre of gravity of machine due to rotation
MRhCG
d=
e
Ib

h
d

a
hCG
b

Figure 9.1: Schematic of Machine on Foundation

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
67 of 92

ANGLO TECHNICAL DIVISION

In addition, it is generally recommended that the width of the base should not be less
than 1,5 times the height of the centre of the machine, i.e.:
b 1,5(h + d C )

9.3

Modelling Foundations

Equipment mounted on concrete block foundations, but larger than that covered in
Section 9.1 or where resonance is a possibility so that the equations given in Section 9.2
cannot be used, must be designed giving due consideration to the mass distribution, and
the underlying soil stiffness. This should be done using a finite element programme, but
a first approximation may be obtained using simplified calculations based on the dynamic
behaviour of the machine on its base.
9.3.1 Soil Conditions
Any adequate model of the machine foundation must use as good a representation of the
underlying soil as possible. Soil, whether sand, clay or even soft rock is a granular
material, rather than a homogeneous continuum.

The analysis requires the soil stiffness properties as inputs. Two important parameters
must be distinguished. These are the elastic modulus E of the soil, and the modulus of
uniform compression CC of the soil. A finite element analysis will generally use E,
whereas most simplified equations (including the ones used here) use CC. These are
theoretically related by the equation:
C C = 1.13

1
E
2
1
A

Warning: Understand the parameters used. E is not equal to CC,


they are not even equivalent, they just sound similar.

The soil properties are generally obtained from one of the following two procedures:
(a) Plate bearing test
In this test, a plate of a specified size is placed on the soil and a specified load is applied.
The settlement of the plate is measured as the load is applied. The soil compression
stiffness CC is then calculated directly from the plate area, the load and the measured
displacement.
A typical plate bearing test result is shown in Figure 9.2. In this test several cycles of
loading, to increasing maximum load, were applied. It can clearly be seen that there is
both an elastic component and an inelastic component of the settlement. Vibration
characteristics are determined by the elastic component of the settlement only. Thus, all
the soil stiffness values that are used for vibration analysis are derived from this elastic
component.
Bearing tests are typically carried out using plates with an area of the order of 0,2 m2 to
1,5 m2, otherwise the loads that must be applied become huge.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
68 of 92

ANGLO TECHNICAL DIVISION

Load (kN)

200

Elastic

150
100
Inelastic

50
0
0

10

Settlement (mm)
Figure 9.2: Typical Plate Bearing Test
Warning: Check, then recheck, soil properties with the Geotechnical
Engineer. Geotechnical Engineers usually think simple total long-term
settlement, not such fancy intricacies as vibration.

(b) CSW (continuous shear wave) test


In this test, an impact is applied to the soil, and the speed of transmission of the resulting
shear wave through the soil medium is measured. The shear modulus of the soil G is
then calculated from the shear wave transmission speed. This test gives the elastic
stiffness of the soil, generally quoted as the shear modulus, G.
The CSW test involves very small shear strains. A higher level of soil strain is typically
experienced below completed structures. Under these higher strain conditions, the
actual shear modulus of granular soils is reduced. Figure 9.3 shows the shear modulus
reduction typically experienced for completed structures on granular soils. The shear
strain encountered will usually be in the range 0,01 % to 0,1 %. The appropriate shear
modulus for soil assessment will thus usually be in the range of 75 % to 50 % of the
CSW value.

Figure 9.3: Shear Modulus Variation with Strain

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
69 of 92

ANGLO TECHNICAL DIVISION

However, where the structure foundation is located on soft rock, this may be reversed.
The CSW shear modulus obtained for soft rock is likely to be determined primarily by the
shear modulus of granular soil lenses within the rock. As the nominal shear strain
increases, these lenses are likely to be compressed, so that the actual behaviour of the
soft rock is representative of a higher shear modulus. However, little is known of how
much this increase is likely to be.
So, CSW shear modulus values reduce with increasing shear strain in granular soils, but
are more likely to increase in soft rock.
(c) Soil Characteristics to Use
The soil characteristics must be obtained from a geotechnical report for the site, prior to
completing the vibration analysis for any foundation. However, as a preliminary
approximation, the values given in Table 9.2 may be used for soil stiffness.
Table 9.2: Approximate Soil Modulus of Uniform Compression Values for AT = 10 m2
Soil type
Allowable
bearing Soil modulus of uniform
pressure (kPa)
compression, CC (kN/m3)
Firm clay or sandy clay.
100
25 000
Medium dense sand or silty sand.
Stiff clay or sandy clay.
200
50 000
Compact poorly graded gravel.
Dense sand or silty sand.
Very stiff clay or sandy clay.
400
100 000
Compact well graded gravel.
Very dense sand or silty sand.

9.3.2 Simplified Preliminary Calculations


The simplified preliminary calculations are based on the following procedure:

(a) The soil uniform compression modulus is calculated.


The soil uniform compression modulus is calculated from the test results, using the
equation:
C C = C Ctest

A test
A

where: CCtest is the soil uniform compression modulus from the geotechnical test
Atest is the area of the geotechnical test
A is the area of the machine foundation being designed
(b)

The soil uniform shear modulus and uniform rotation modulus are calculated.

The ratio between the soil uniform shear modulus CV and the soil uniform compression
modulus of the soil CC, the shear stiffness ratio, is primarily a function of the Poissons
ratio of the soil, and to a lesser extent also of the geometry of the foundation. Based
on the assumption of elastic continuum behaviour of the soil, the ratio Q between the
soil uniform shear modulus CQ and the uniform compression modulus CC is given as
listed in Table 9.3.
The soil uniform shear modulus is given by:

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
70 of 92

ANGLO TECHNICAL DIVISION

CQ = QCC
The value of Poissons ratio for soil is generally approximately 0,2.
Table 9.3: Ratio Q between CQ and CC
Poissons ratio
a/b = 0,5
0,1
0,96
0,2
0,91
0,3
0,85
0,4
0,78

a/b = 1,0
0,95
0,89
0,82
0,75

a/b = 2,0
0,94
0,87
0,80
0,72

Due to the fact that under conditions of rotation of the foundation the soil pressures are
not uniform, but vary linearly across the base, an adjusted soil uniform rotational
modulus must be calculated. Based on the assumptions of a rigid foundation and elastic
continuum behaviour of the soil, the ratio between the soil uniform rotational modulus
C and the uniform compression modulus CC is given as listed in Table 9.4.
Table 9.4: Ratio between C and CC
a/b = 0,5
a/b = 1,0
1,58
1,88

a/b = 2,0
2,31

The soil uniform rotational modulus is given by:


C = CC
(c) The stiffness for vertical motion is calculated.
This is given by:

K C = CC A
(d) The stiffness for horizontal motion is calculated.
This is given by:
K Q = CQ A
(e) The stiffness for rotational motion is calculated.
This is given by:
a 3b
12
ab 3
K = C I = C
12
K = C I = C

or

for rotation about an axis parallel to side b


for rotation about an axis parallel to side a

(f)
The system is analysed to determine to frequencies.
For a symmetrical arrangement of machine and foundation, vertical vibration can be
approximated as a single degree of freedom system. Horizontal and rotational vibration
will always be coupled, because the centre of gravity of the foundation and machine will
always be well above the soil-foundation interface. This can be idealised as two, two
degree of freedom systems.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
71 of 92

ANGLO TECHNICAL DIVISION

The calculation procedure outlined above has been programmed in the EXCEL
spreadsheet
Design
Aid
DA13
Machine
bases,
located
at
G:/ENGINEERING/se/DESIGN AIDS.
9.3.3 Damping
There are two damping mechanisms applicable to soils. The first is structural damping
which is similar to the damping that occurs within any material. Internal friction between
the grains of soil causes energy losses in much the same way as internal friction
between molecules in other materials. However, because the soil extends more-or-less
infinitely laterally and downwards away from the foundation under consideration, there is
also dispersion damping. Here energy is lost to the finite system considered because
oscillations get the travel bug, and head off into the far distance.

Unfortunately, very little information is available regarding the magnitude of soil damping.
Generally, the design approach is to avoid resonance, in which case damping has little
influence, so little effort has been expended in determining this difficult quantity.
A conservative design will assume zero damping. A more realistic design will assume
modal damping of, say, 5 %.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
72 of 92

ANGLO TECHNICAL DIVISION

10 PRACTICAL DETAILS FOR TERTIARY STRUCTURAL


ELEMENTS
10.1 Individual Members
The local vibration of individual members of structures is often not very well predicted by
computer models of the entire structure, so it is not uncommon to experience local lateral
or torsional vibrations of individual members in structures. When assessing structural
vibrations, it is thus necessary to evaluate the frequencies of individual members within
structures. This is probably more easily and more accurately done by hand using simple
formulae, rather than by using a computer package such as ROBOT or PROKON.
10.1.1 Approximate Natural Frequencies of Individual Members
The lowest two natural frequencies, in Hz, of individual structural members may be
estimated from the following formulae. A spreadsheet to calculate these frequencies is
located at g:/ENGINEERING/se/DESIGN AIDS Design Aids DA12 Beam Frequencies.

(a) Flexural frequencies


Members may be either simply supported, fixed ended, or cantilevered.
(a)
f1 =

1,57 EI
L2 m

(b)
f1 =

Simply supported beams


6,28 EI
m
L2

Fixed end beams

3,56 EI
m
L2

(c)

f2 =

f2 =

9,82 EI
m
L2

f2 =

3,51 EI
L2 m

Cantilevers

0,56 EI
f1 = 2
m
L

(b) Torsional frequencies


Members are assumed to be fixed against rotation about the axis of the member at both
ends.
f1 =

0,54 GJ
L
Im

f2 =

1,13 GJ
L
Im

(c) Axial frequencies


Members are assumed to be fixed against axial movement at both ends.
f1 =

0,54 EA
L
m

f2 =

1,13 EA
L
m

10.1.2 Limitation of Slenderness Ratio to 80


It has been fairly common practice to limit the local vibration of steel members by simply
ensuring that the slenderness ratio does not exceed 80. The flexural frequencies above
can be re-written for steel members to give:

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
73 of 92

ANGLO TECHNICAL DIVISION

(Simply supported)
f1 =
=
=
=

1,57
2

EI
A

1,57 E
L2
1,57
L2
7925r

(Fixed ends)
f1 =

I
A

200 x10 9
r
7850

=
=

L2
7925
=
L
L
r

3,56
2

EI
A

3,56 E

L2

I
A

200 x10 9
r
7850
L2
17969r
3,56

L2
12578
=
0,7L
L
r

If the slenderness ratio is limited to not more than 80, these equations can be written as:
(Simply supported)
7925
f1
80L
99

(Fixed ends)
12578
80L
157

f1

This procedure thus gives fairly high frequencies, provided the individual member lengths
are fairly short. Generally, vibrating screens and other vibrating equipment operate at
frequencies up to about 20 Hz, so natural frequencies of 30 Hz and higher are desirable
to avoid resonance problems. If the slenderness ratio is kept below 80, the natural
frequencies will be above 30 Hz where a simply supported member is less than 3,3 m
long, or where a fixed ended member is less than 5,2 m long. This method is simple to
apply, because PROKON (and other packages) can perform a design based on limiting
the slenderness ratio, but PROKON does not calculate the natural frequencies of
individual members. However, for longer members, it may not avoid resonance, so it
should be used with care.

10.2 Walkways and Hand Railing


A frequent phenomenon in any structure with vibrating machinery of any type, is to see,
the hand railing shaking, or hear it rattling. This is due to the fact that hand railing is
usually attached to light walkway, or platform stringers, often only a 180 mm or 200 mm
deep channel section. These stringers have very little torsional stiffness, so with the
hand railing protruding upwards by 1 m or so, the whole arrangement twists very easily.
Typical hand railing layouts have a fundamental natural frequency in the range between
8 Hz and 16 Hz, which leads to frequent problems of unacceptable vibrations due to
resonance.
There are two requirements for ensuring that hand railing has a sufficiently high natural
frequency to avoid resonance. These are:
(a)
(b)

Hand railing standards should be larger than the normal 48 mm diameter,


preferably not less than 70 mm.
Torsional stiffness must be provided to the stringers to which the hand railing is
attached. This can be done either by boxing the stringers, or by ensuring adequate
stiffness of the cross members between walkway stringers. Details that may be
used are suggested in Figure 10.1.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
74 of 92

ANGLO TECHNICAL DIVISION

Provide access
holes for bolting

10.1(a): Possible Stringer Sections

10.1 (b): Possible Cross Member


D t il
Figure 10.1: Possible Hand Railing Details
An EXCEL spreadsheet located at g:/ENGINEERING/se/DESIGN AIDS Hand railing
dynamic design is available to check the approximate fundamental natural frequency of
proposed hand railing layouts, or PROKON or ROBOT can be used to create a simple,
local model for the hand railing.

10.3 Sheeting Rails


Sheeting rails should be treated as isolated beams, as described in Section 10.1, with
allowance made for the mass of the attached sheeting. The sheeting is, however,
flexible, so its mass will not all move exactly with the sheeting rail. It is recommended
that two cases be considered to provide upper and lower bounds on the natural
frequency of sheeting rails.
(a)
(b)

Lower bound on frequency. Add the entire mass of the sheeting associated with
the relevant sheeting rail in the frequency calculation.
Upper bound frequency. Add 30 % of the mass of the sheeting associated with the
relevant sheeting rail in the frequency calculation.

An important practical detail is that sheeting in vibrating structures must be fixed to the
sheeting rail in every trough, not in every second trough as is usually done.

10.4 Plating on Chutes, Bins and Underpans


10.4.1 Natural Frequencies of Rectangular Panels

(a) Single Flat Plate Panels


The fundamental natural frequency f1 (Hz) of a single flat rectangular steel panel where
all four edges are simply supported is given by Szilard as:
f1=

1
Et 3
1
( 2+ 2 )
2 a
b
12(1- 2 )m

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
75 of 92

ANGLO TECHNICAL DIVISION

The fundamental natural frequency f1 (Hz), of a single flat rectangular steel panel where
all four edges are fixed is given by Szilard as:
f1

12
2

1
Et 3
7 1 4 1
( 4+
+ 4 )
2
2
7a b
2 a
b
12(1- 2 )m

where: a and b are the two panel dimensions as shown in Figure 10.2 (m).
E is the plate elastic modulus (Pa).
t is the plate thickness (m).
is the plate Poissons ratio, usually taken as 0,3 for steel.
m is the mass of the plate and liners (kg/m2).

a (2)

a (3)

b (2)

b (1)

a (1)

Figure 10.2: Typical Plate with Several Panels


As discussed earlier in this guide, in order to avoid resonance, the natural frequency of a
structure or component of a structure should be at least 1,5 times the operating
frequency of the equipment supported. Assuming equipment with a maximum frequency
of 16 Hz, the natural frequency required to avoid resonance is thus about 25 Hz. The
maximum plate panel sizes to ensure natural frequencies of the plating exceeding 25 Hz
are shown in Figures 10.3 for simply supported plate panels, and Figure 10.4 for fixed
edge plate panels.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
76 of 92

ANGLO TECHNICAL DIVISION

Simply Supported Edges


2.2
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2

Aspect ratio b/a


1 without liner
2 without liner
10 without liner
1 with 8 mm liner
2 with 8 mm liner
10 with 8 mm liner
3

10

11

12

Plate thickness (mm)

Figure 10.3: Maximum Simply Supported Plate Sizes for 25 Hz Natural Frequency

Fixed Edges
2.2
2
1.8
1.6
1.4
1.2
1
0.8
0.6
0.4
0.2

Aspect ratio b/a


1 without liners
2 without liners
10 without liners
1 with 8 mm liners
2 with 8 mm liners
10 with 8 mm liners
3

10

11

12

Plate thickness (mm)

Figure 10.4: Maximum Fixed Edge Plate Sizes for 25 Hz Natural Frequency
(b) Stiffened (Orthotropic) Plating
The equations for stiffened plates (often referred to as orthotropic plates) are a little more
complex, but still manageable by hand for the simply supported case. The lowest natural
frequency for simply supported stiffened plating is given by:
Dx
f1

2D v
2 2

a b
m

Dy
b4

where:
Dv =

Et 3
2

G Jx Jy
(
+
) is the shear rigidity of the stiffened plating
6 a
b

12(1 - )
EI x
is the flexural rigidity of the stiffened plating about the x-axis
Dx =
bx

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
77 of 92

ANGLO TECHNICAL DIVISION

Dv =

EI y
by

is the flexural rigidity of the stiffened plating about the y-axis

a and b are the two overall panel dimensions (m) as shown in Figure 10.5.
bx and by are the spacing of the stiffeners in the x- and y directions (m) as shown
in Figure 10.5.
E is the plate elastic modulus (Pa).
G is the plate shear modulus (Pa)
t is the plate thickness (m).
is the plate Poissons ratio, usually taken as 0,3 for steel.
Ix is the moment of inertia of the x-direction stiffeners including the plating (m4)
Iy is the moment of inertia of the y-direction stiffeners including the plating (m4)
Jx is the torsion constant of the x-direction stiffeners excluding the plating (m4)
Jy is the torsion constant of the y-direction stiffeners excluding the plating (m4)
m is the mass of the plate and liners (kg/m2).
A spreadsheet, Design Aid DA14 Plate vibration, to facilitate calculation of the
frequencies for plating is available at G:/ENGINEERING/se/DESIGN AIDS. Use of the
spreadsheet is self-explanatory.

X
Y
A
N
ax
ay
by
by

bx
N

Figure 10.5: Typical Arrangement of Stiffened Plating

10.5 Bracing Systems


Structural bracing, particularly tension bracing members, may be long and slender, and
thus be prone to vibrating. Tubular sections are quite popular as they have the lowest
slenderness for a given cross-sectional area of bracing members, and thus a higher
natural frequency than other section shapes. Because bracing members run diagonally,
they also run past other members such as girts or other bracing members. This may
lead to high noise levels if vibration causes the bracing and other members to rattle
against each other.
The following steps will ensure that bracing in dynamically loaded structures remains
trouble free.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
78 of 92

ANGLO TECHNICAL DIVISION

(a)

The natural frequency of bracing members must be high enough to avoid


resonance. See Section 10.1.
(b) End connections for bracing members must be welded (but check for fatigue), or
made using slip-resistant connectors.
(c) Bracing members must be detailed so that:
(i)
either they are at least 20 mm clear of other members they pass.
(ii)
or they are flush with, and bolted to, other members they pass.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
79 of 92

ANGLO TECHNICAL DIVISION

11 PRACTICAL DETAILS FOR CONNECTIONS


11.1 Bolted connections
Where bolts are used for joining members in a structure carrying vibration loading, the
following points should always be observed:
(a)
(b)

Never use ordinary bolts in tension. Bolts carrying tension must always be
tensioned, so either high strength friction grip bolts, or swage lock fasteners should
be used
Bolts carrying shear should preferably be high strength friction grip bolts or swage
lock fasteners. If ordinary bolts are selected, they must use nuts that will prevent
loosening.

11.2 Welded Connections


Welded connections should be avoided as far as possible on structures carrying vibration
loading. Where welding is necessary, the following points must be observed:
(a)
(b)

Never use intermittent welding. Rather use a smaller continuous weld.


Watch fatigue details. The fatigue life of a welded connection may be reduced
spectacularly by poor welding details. If you dont properly understand the effects
of welding on the fatigue life of structures, then ask someone who does!

11.3 Beam-to-beam Connections


Experience has shown several commonly used connections to be bad news when they
are used in a structure carrying vibration loading. These connections are shown in
Figure 11.1.

Warning: Never use the connections shown in Figure 11.1 in a structure


supporting vibrating loads. They will turn around and bite you every time.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
80 of 92

ANGLO TECHNICAL DIVISION

Cracks from weld


if plate girder
Cracks from radius

Cracks from
bottom of T

Full depth T OK

Figure 11.1: Bad Commonly used Beam-to-Beam Connection Details

11.4 Bracing Connections


The most important aspects of bracing connections in structures supporting vibrating
equipment are:
(a)
(b)

The bolts must be slip-resistant bolts, i.e. either friction grip bolts or swage lock
fasteners.
Care must be exercised in detailing, to ensure that gussets are properly anchored
to the main structural members. This is illustrated in Figure 11.2.

Cycling load
leads to web
cracking

Figure 11.2: Some Bad Bracing Connections

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
81 of 92

ANGLO TECHNICAL DIVISION

12 VIBRATION MEASUREMENTS
12.1 What Should be Measured?
Vibration amplitudes may be described by acceleration, velocity, displacement, or even
strain amplitude. Usually the easiest parameter to measure is acceleration, which can
be integrated over short durations to provide velocity or displacement. Simple
transducers can be attached to any point on a structure and provide an absolute
measure of acceleration. Displacement can quite easily be measured, but it is generally
not possible to measure absolute displacement. The displacement transducer must be
mounted somewhere and measures the change of displacement to the desired point, so
the displacement obtained is a relative displacement between the mounting point and the
desired point on the structure. We generally dont have fixed mounting points around
structures, because the whole structure is floating about in space. Strain is not often
measured, as placement of strain gauges on structures is time consuming, and dynamic
strains are often small.
When accelerations are measured, velocities can easily be calculated by integrating the
accelerations with respect to time, and displacements can be calculated by integrating
the velocities with respect to time. Generally, the equipment used for vibration
measurements has the capability of performing these integrations automatically, so that
accelerations, velocities or displacements can be displayed. If the vibration is harmonic
in nature, there is a simple relationship between the amplitudes of acceleration, velocity
and displacement. This is:
aM = E v = E 2 d
v = E d
or
d=

aM
v
=
E E 2

v=

aM
E

where: aM is the linear acceleration amplitude (m/s2)


v is the linear velocity amplitude (m/s2)
d is the displacement amplitude (m/s2)
E is the radial frequency of the vibration (radians/s)

12.2 Measuring Equipment


In selecting measuring equipment for vibration measurements, it is important to ensure
that certain characteristics of the equipment are appropriate. Generally, two factors must
be considered:
(a)
(i)

Frequency Range
Minimum value. Some transducers can measure from 0 Hz (i.e. constant
acceleration such as gravity) upwards. Other types of transducers only
measure from a defined minimum frequency upwards. Where the operating
frequency is low, such as is likely to be true of floatation cells and crushers, it
is possible that the minimum frequency range may be a problem. For
example, the RION Vibration Analyser used in ATD Structural Engineering,
the minimum frequency is 3,0 Hz. This means that vibration measurements
on float cells operating at 2,0 Hz will not be correct.
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
82 of 92

ANGLO TECHNICAL DIVISION

(ii)

Maximum value. All transducers will have their own natural frequencies, and
will thus experience resonance at some fairly high frequency.
Most
transducers measure accurately up to at least 100 Hz, whereas in structural
engineering we are generally only concerned about frequencies up to about
50 Hz.
This is thus seldom a problem in structural engineering
measurements, but it must be borne in mind.

Many measuring systems also incorporate filters. A low pass filter eliminates the high
frequency content of measurements, and allows the low frequency content through.
Conversely, a high pass filter eliminates the low frequency content, and allows the high
frequency content through. All electrical measuring systems have some drift, by which is
meant they shift with time. In order to eliminate the effects of this drift, many systems
use a low pass filter set at between 1 Hz and 3 Hz. This will have to same effect as a
non-zero minimum measuring frequency, so users must be aware of what filters are used
in vibration measuring equipment. The RION Vibration Analyser VA10 used in ATD
Structural Engineering, has a low pass filter that can be set at 3 Hz or 10 Hz. It should
always be set at 3 Hz for structural engineering measurements. The RION Spectral
Analyser SA78 used in ATD Structural Engineering may be used to frequencies as low
as 1,0 Hz
(b) Sampling Rate
The sampling rate is the rate at which the analogue measured signal is converted into
digital numbers for computer storage and analysis, or display. In order to obtain good
measurements, the sampling rate must be at least six times the highest frequency of
interest. Thus, if a maximum frequency of 25 Hz is expected, the sampling rate must be
at least 150 Hz, i.e. samples must be read at intervals not exceeding 0,006 seconds.

12.3 Recording Measurements


There are numerous different aspects to be considered when vibration measurements
are made, and these should all be recorded to ensure complete records.
(a)
(b)
(c)
(d)
(e)

Measurements may be RMS or peak-to-peak.


Acceleration, velocity, or displacement may be measured and recorded.
The filter may be set at various frequencies.
Measurements may be taken in any one of three different directions.
The key frequencies in any measurement should always be noted.

A sample measurement sheet is shown in Table 12.4.

12.4 Relating Measured Displacements to Implied Stresses


An important aspect of interpretation of measurements is to know what stresses are
implied by measured displacements. A simple, conservative method (i.e. a method
giving a high estimate of the implied stress) is to ignore the damping, giving the same
relationship between displacement and stress as for static load. This gives the equations
in Table 12.1 to obtain the stress (N/m2) for beams, in Table 12.2 to obtain the stress
for portal columns (i.e. where bracing is not used), and Table 12.3 to obtain the stress for
slabs.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
83 of 92

ANGLO TECHNICAL DIVISION

Table 12.1: Conversion from Measured Deflection to Implied Stress for Beams
Simply
Supported Fixed Ended Beam
Beam
Centre of Beam
End of Beam
6Ed S V
12Ed S V
12Ed S V
Beam with High

=
2
2
Central Mass
L
L
L2
Beam with Uniformly
Distributed Mass

4,8Ed S V
2

8Ed S V

16Ed S V
L2

Table 12.2: Conversion from Measured Deflection to Implied Stress for Portal Columns
Fixed one end, pinned other end
Fixed both ends
=

3Ed S H
2L

3Ed S H
L2

Table 12.3: Conversion from Measured Deflection to Implied Stress for Slabs
Simply
Supported Fixed Edge Slab
Slab
Centre of Slab
End of Slab
6Ed S
12Ed S
12Ed S
Slab
with
High

=
2
2
Central Mass
L
L
L2
Slab with Uniformly
Distributed Mass

4,8Ed S
2

8Ed S
2

16Ed S
L2

In Tables 12.1 to 12.3:


dS is the depth of the beam, portal column, or slab on which the measurements
are taken (m)
E is the elastic modulus of the material of which the beam is made (N/m2)
L is the length of the beam on which the measurements are taken (m)
H is the measured horizontal deflection at the top of the portal column (m)
V is the measured vertical deflection at the centre of the beam (m)

12.5 Baseline Vibration Measurement Guide


From time-to-time it is necessary to obtain baseline vibration measurements, by which
is meant ambient vibration measurements on an existing structure, for later comparison
with other vibration measurement data. The most common reason for wanting to obtain
baseline vibration measurements is to define a contractual baseline against which
vibration can be checked following commissioning of a new building, or modification of
an existing building.
12.5.1 Baseline Measurements
(a) The first consideration when planning baseline measurements is what
measurements are necessary. The fundamental requirements are the vibration
magnitudes and the major frequencies.

(i)

(ii)

Vibration Magnitudes
Baseline measurements will typically be accelerations, unless there is a specific
Client request, or some other good reason, for using another parameter.
Vibration magnitudes may be described by peak values, or averaged values
(usually RMS values are used for averaging), or both. Consideration must be
given to the time-varying nature of the vibration in defining the most appropriate
way to describe the vibration magnitude.
Frequencies
Guidelines for the Vibration Design of Structures Issue 2
AA BPG S002
84 of 92

ANGLO TECHNICAL DIVISION

Human sensitivity to vibration, and the sensitivity of equipment to vibration, tends


to be frequency-dependent. It is thus necessary to include frequencies in
baseline measurements.
(b)

The second consideration when planning baseline measurements is where, and in


which direction, measurements should be taken. This is dependent on the reason
for requiring the vibration baseline measurements. If they are required because a
machine is to be replaced, and vibrations due to the new machine are to be
checked against vibrations caused by the existing machine, then measurements
should clearly be taken on the closest structural members that will remain in place.
However, if it is desired to ensure that vibrations in the vicinity of sensitive
equipment of office areas do not increase, then the measurements should clearly
concentrate on these areas.

(c)

When baseline measurements are taken, it is important that the actual conditions
must be recorded. This must include:
(i)
(ii)
(iii)

What vibrating equipment is operating, and what equipment is standing.


The throughput in the equipment that is operating.
Other ambient conditions, such as the extent of spillage on floors and adhering
to the inside of chutes.

A typical form that can be used for recording vibration baseline measurements is shown
in Table 12.2.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
85 of 92

ANGLO TECHNICAL DIVISION

Table 12.4: Typical From for Recording Vibration Measurements or Baseline


Measurements
Mine
Area

Anglo Technical Division


Date

Building
Measuring device
Sampling speed

Filter setting
Other

Machines running
Throughput
Sketch of Structure

Measurement Frequency
location
(Hz)

P-P Accelerations (m/s2)


X
Y
Z

RMS Accelerations (m/s2)


X
Y
Z

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
86 of 92

ANGLO TECHNICAL DIVISION

12.5.2 What Baseline Measurements Can and Cant Do

(a)

What they can do


(i)
(ii)
(iii)

(b)

Define ambient existing vibration under actual existing conditions and at


specified locations.
Provide a comparative baseline against which other vibrations can later be
evaluated or compared.
Be used to give a good indication of whether existing vibration is within
serviceability and strength limits for the structure.

What they cannot do


(i)
(ii)
(iii)

Define worst case vibrations under operational conditions not existing when
measurements are taken.
Predict the structural behaviour under new equipment running at different
frequencies to existing equipment.
Predict the structural response at locations other than where measurements
are actually taken.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
87 of 92

ANGLO TECHNICAL DIVISION

13 TROUBLE-SHOOTING AND STRUCTURAL MODIFICATION


13.1 Interpreting and Using Measurements
(a) Implication of Frequencies
The following information can generally be obtained from the measured frequencies:
(i)
(ii)
(iii)

Frequencies equal to the operational speed of equipment, identify that


equipment as the source of the vibration, because structural vibration will
always be at the exciting frequency.
Frequencies much higher than the operating speed of equipment may indicate
misalignment or other problems with gearboxes, motors, or tooth meshing of
ring gears.
If the vibration loading has a significant impact component, then measured
frequencies are likely to show natural frequencies of the structure.

(b) Implication of Amplitudes


The amplitudes, together with the frequencies, enables assessment of the structural
integrity by checking whether serviceability, strength or fatigue life limits are exceeded.

13.2 Changes to Applied Loads


Experience shows that vibration problems may result from modifications which lead to
changes in the applied vibration loads. Typical examples of this are the following:
(a)
(b)

Setting of the eccentric masses on vibration screens or vibratory feeders may be


changed in an effort to increase production throughput. This leads to a
proportional increase in the applied vibration loads.
Steel coil springs often have only a fairly short operational life. Frequent
replacements may cause Site personnel to replace coil springs with rubber blocks,
which generally have a significantly longer life. What may not be realised, is that
rubber blocks are much stiffer than steel coil springs, particularly in the vertical
direction. See Section 5.2. This means that significantly higher excitation is
required to obtain the specified throw of the screen or feeder, leading to
significantly higher loads being applied to the structure. These loads may be as
much as five times higher than where steel coil springs are used.

Where structural vibration results from changes to the applied loads, it is often possible
to simply, and very easily, revert to the original conditions. Where this is not possible, a
re-design of the structure using the new loads is necessary, leading to structural
modifications.

13.3 Structural Modifications


Where a structural vibration problem is identified, the preferable solution is generally to
increase the stiffness of relevant structural members. If the structure is already high
tuned, this is the only rational solution, and is probably fairly easily achieved. Where the
structure is initially low tuned, increasing the stiffness may worsen the problem, as it will
move the structure closer to resonance. Extra mass may be added, but this impacts
negatively on the static design of the structure. A large increase in stiffness to change
the structure to a high tuned structure may be possible, but this is likely to involve major
structural modifications. This is a much more difficult problem than the high tuned
structure, and it may require a combination of extra mass and extra stiffness.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
88 of 92

ANGLO TECHNICAL DIVISION

Warning: Careful analysis and thought must go into modifications to


solve vibration problems in low tuned structures.

13.4 Common Concerns of Site Personnel


Experience has shown that there are some common concerns expressed by Site
Personnel. These include:
(a)

Beat phenomenon. This occurs in every installation where separate pieces of


equipment having similar operating speeds are located close together. The most
common example of this is where a number of parallel process streams are used,
requiring several nominally identical screens, feeders, or other vibrating equipment
next to each other. Although the equipment runs at nominally the same speed,
unless there is electrical or mechanical coupling, there is always a small difference
in speed. This leads to two pieces of equipment running in phase for a short while
with a reinforcing effect on resulting vibration, then running out of phase for a while
with a destructive effect on resulting vibration. This is shown in the vibrations in
Figure 13.1, which shows the interaction of two machines which nominally run at
16 Hz, but where one runs 2 % fast, and the other runs 2 % slow.

(b)

Rattling, or large amplitude vibration of light fittings, sheeting, or other finishes. It is


impossible to model all of these minor elements in a structural vibration model.
Each case must be treated on its own merits. It may be possible to easily move
the fitting to a less sensitive location. Local stiffening may be possible. Insertion of
rubber isolation mounts may solve the problem.

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
89 of 92

ANGLO TECHNICAL DIVISION

15,6 Hz vibration
1.5
1
0.5
0
-0.5 0

10

15

20

-1
-1.5
Time (seconds)

16,4 Hz vibration
1.5
Amplitude 2

1
0.5
0
-0.5 0

10

15

20

-1
-1.5
Time (seconds)

2.5
2
1.5
1
0.5
0
-0.5 0
-1
-1.5
-2
-2.5

10

15

20

Time (seconds)

Figure 13.1: Vibration Histories showing the Beat Phenomenon

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
90 of 92

ANGLO TECHNICAL DIVISION

14 BIBLIOGRAPHY
14.1 Standards and Specifications
14.1.1 SANS Standards
SANS ISO 4866:1990 Mechanical Vibration and Shock Vibration of Buildings
Guidelines for the Measurement of Vibrations and Evaluation of their effects on
Buildings. First Edition 1999.
SANS ISO 2631-1:1997 Mechanical Vibration and Shock Evaluation of Human
exposure to whole-body Vibration. Part 1: General Requirements. First Edition
1997.
ISO 2631-1:1985 Mechanical Vibration and Shock Evaluation of Human exposure to
whole-body Vibration. Part 1: General Requirements.
14.1.2 AAC Specifications
AAC 114/1 Design Criteria for Steel Structures. Issue 9 1998.

14.2 Text Books


Barkan DD Dynamics of Bases and Foundations, McGraw-Hill, 1962.
Irish K and Walker WP Foundations for Reciprocating Machines, Concrete Publications
Limited, London, 1969.
Paz M Structural Dynamics. Theory & Computation, Van Nostrand Reinhold.
Szilard R Theory and Analysis of Plates. Classical and Numerical Methods. Prentice
Hall.
Warburton GB
Library.

The Dynamical Behaviour of Structures, Pergamon International

14.3 Journal Papers


Sen GS Blasting Vibration and Structural Damage, Civil Engineering, September 1981,
pgs 42-44.

14.4 AAC Reports


Report RAB/83/02 "Dynamic Stiffness Tests on Velmet Screen Support Springs: Vertical
Stiffness", 21 July 1983.
Report 2005-S-25 Anglo, Base Skorpion Mine.
Agitator Support, 26 December 2005.

Leach and Neutralisation Tank

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
91 of 92

ANGLO TECHNICAL DIVISION

15 RECORD OF MODIFICATIONS
Date
Jan 2006
Oct 2007

Description
Publication of Edition 1.0
Symbols updated for consistency
Loading information for crushers and agitators added
Publication of Issue 2
Some editorial errors corrected
Section 8 revised to include more information on mills
Section 9 expanded to provide more detailed guidance

Author
G.J. Krige
G.J. Krige

Guidelines for the Vibration Design of Structures Issue 2


AA BPG S002
92 of 92

Você também pode gostar