You are on page 1of 98

NATIONAL UNIVERSITY OF SINGAPORE

DEPARTMENT OF CIVIL AND ENVIRONMENTAL ENGINEERING

Geotechnical Centrifuge

USERS' MANUAL (2nd Edition)

by

R.F. Shen, C.Y. Wong and L.H. Tan

June 2011

Based on Original Version by

F. H. Lee (1991)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


ABSTRACT
ABSTRACT

This manual describes the operation of the National University of Singapore Geotechnical

Centrifuge. It is divided into three chapters. The first chapter describes the components of

the centrifuge and their functions. The NUS centrifuge has a capacity of 40,000 g-kg and

operates up to a maximum g-level of 200g. This implies that the allowable payloads at 200g

and 100g are 200 kg and 400 kg, respectively. The structure of the centrifuge is based on the

conventional dual swing platform design. The model is normally loaded onto one of the

platforms while the opposing platform either carry an appropriate counterweight or an

identical model. The distance from the axis of rotation to the base of the conterweight

platform is 1.871 m, while the distance from the axis of rotation to the base of the payload

platform is 2.021 m. The centrifuge is driven by a hydraulic motor which is capable of

delivering up to about 37 kW of power. In the present configuration, 40 differential channels

are available for data transmission. Three data acquisition systems are available for data

capturing; a general purpose system, a medium speed system, and a high speed system. The

maximum sampling rates of these three systems are 48,000, 120,000 and 1 million samples

per second, respectively.

Chapter two describes the categories of personnel who are involved in centrifuge testing.

Academic and research staff who wish to use the centrifuge must have been adequately

familiarised with the centrifuge before they are registered as centrifuge users. In addition,

test programmes proposed by registered users should be checked by an independent

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


checker whose function is to provide an expert opinion on whether the test will or will not

exceed the safe limits of operation. New model containers and major pieces of equipment

which will form part of the payload should be proof-tested before being used. The

recommended procedures for proof-testing consist essentially of stressing checks, 19-test

and finally high-g test.

The third chapter contains checklists for the centrifuge operator and registered user. The

operator checklists contain procedures for balancing, start-up, normal and emergency

stopping as well as routine maintenance of the centrifuge, which must be adhered to by the

operator. The user checklist contains procedures to assist the user in the preparation of

models and setting-up of the signal handling systems. The PID computer control system for

centrifuge operation and the cooling ventilation system for centrifuge enclosure are

explained in this chapter. Some guidelines for preparation of sand and clay model grounds

are also provided.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


CONTENTS
CONTENTS

ABSTRACT

CONTENTS

1. EQUIPMENT

1.1 Location

1.2 Structure

1.3 Motor drive and control system

1.4 Speed pressure and temperature readouts

1.5 Signal and power transmission between centrifuge and control room

1.6 Image Processing System

1.7 Further information

2. ORGANIZATION AND OPERATION

2.1 Introduction and overview

2.2 Classification of personnel involved centrifuge tests

2.3 Test programmes and authorization

2.4 Balance calculations

2.5 New model containers and equipment

2.6 Summary

3. OPERATING PROCEDURES

3.1 Introduction

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


3.2 Fixing of items onboard centrifuge

3.3 Preparation of sand and clay models

3.4 Checklist of procedures for centrifuge operator

3.5 PID computer control system for centrifuge operation

3.6 Cooling ventilation system for centrifuge enclosure

3.7 Items for routine maintenance

3.8 Checklist to assist registered users

FIGURES

Appendix A - Proposed Centrifuge Test Programme Form

Appendix B – Centrifuge Balance Calculation Worksheet

Appendix C - User Manual of Program Static Measurement for Strain gauge Measurement

Appendix D – Centrifuge Operator’s Checklist Before Starting Centrifuge

Appendix E - GeoPIV quick start user manual (courtesy of White D.J. White & W.A. Take)

Appendix F - Scaling Principles for Geotechnical Model Testing in Centrifuge Test (T.S. Tan)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


1. EQUIPMENT
1. EQUIPMENT

1.1 LOCATION

The centrifuge is located at one of the end bays of the Centrifuge Laboratory in Rm 01-01-01

of the Faculty of Engineering buildings, Fig. 1. 1. It is housed within a ferrocement enclosure

which is designed to protect users and equipment from objects which may have flown off

from the payload or any part of the centrifuge while the latter is rotating. As shown in Fig.

1.2, a set of steel doors provides entrance into the centrifuge enclosure from the outside.

The centrifuge always rotates in a clockwise direction so that any objects which cannot be

stopped by the steel door will be caught by the L-shaped extension, which is also

constructed of ferrocement. Figs. 1.3a and b show views of the centrifuge at rest and in

motion within the enclosure. There are two entrances into the centrifuge bay. The first is a

2.5-m wide entrance which provides access from within the laboratory. This is the entrance

which is normally used for loading and unloading of packages and for setting up test

equipment. The second entrance is a service entrance into the motor drive area from

outside the laboratory. Under normal circumstances, this entrance is to be used only for

loading, unloading and servicing of equipment in the motor drive area. Whenever a model

test is in progress, no personnel will be allowed into the centrifuge bay. To ensure this, all

entrances should be closed and the doors locked.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


1.2 STRUCTURE

The structure of the centrifuge is designed and fabricated by Harricks Steel Fabricators Pte.

Ltd. Fig. 1.4 shows the side and plan views of the centrifuge structure. The centrifuge

rotates about a vertical rotor shaft which is driven by a hydraulic motor located at its base.

The shaft has an outer diameter of 250 mm and a wall thickness of 50 mm. Driving torque is

transmitted from the shaft through a welded steel cage to the rotor arm. The centrifugal

force is taken by four parallel steel plates which form the main elements of the rotor arm.

These plates are held together by connecting and stiffener plates. A steel swing platform is

hinged onto each end of the rotor arm. Each platform has a working area of about 750 mm x

700 mm and a model headroom of about 1350 mm. When the platforms are fully swung up

during testing operation, the radial distance from the centre of rotation to the base of the

model container is approximately 2021 mm. The centrifuge is designed for a payload

capacity of 40,000 g-kg and a maximum working g-level of 200g. The maximum mass of the

payload can be computed from these two constraints. Hence. at 200g, the maximum

payload mass is 200 kg. At

100g, this quantity is increased to 400 kg, and at 50g, it is further increased to 800 kg. Under

normal testing conditions, it is very unlikely for the payload mass to exceed 800 kg given the

platform working area and headroom.

1.3 MOTOR DRIVE AND CONTROL SYSTEM

The driving torque for the centrifuge is supplied by a 50-hp (approximately 37.5 kW)

hydraulic drive system installed by Rexroth G.L. Pte Ltd. As shown schematically in Fig. 1.2,

the system consists essentially of a TECO 75-hp, 3-phase induction AC-motor driving a

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Mannesmann Rexroth A4V90HW1.0R001010 variable displacement hydraulic pump. High

pressure piping conducts the hydraulic oil into the Mannesmann Rexroth MR300 radial

piston hydraulic motor mounted onto the lower end of the rotor shaft. The flow of the soil

through the radial pistons causes the rotation of the rotor shaft. The stroke, that is the

displacement of the pump is controlled by the swivel angle of a swashplate inside the pump.

This angle can be adjusted by varying the electric current through a solenoid which swivels

the swashplate. This electric current is in turn adjusted by a VT200S4X control card housed

within the control box in the centrifuge control room. As the stroke of the pump Is changed,

so does the flow rate of the oil through the circuit, and therefore the RPM of the centrifuge.

Since the swivel angle is directly controlled and the AC induction motor rotates at virtually

constant speed, the design of the pump is such that the flow rate of the oil and therefore

the RPM is directly controlled whilst the resistance and thus the pressure gradient in the

circuit is allowed to equilibrate. This concept is different from that of an electric DC motor

which controls the driving torque and allows the RPM to vary until the resisting torque is

equal to the driving torque. In this latter system, the RPM is indirectly controlled by the

setting of the driving torque.

Fig. 1.5 shows the hydraulic circuit of the drive system. The outermost loop of the circuit is

the path which the oil would flow through when the centrifuge is being rotated. Along the

segment of the circuit downstream of the hydraulic motor, the pressure of the oil is

maintained fairly constant at about 35 bar. Upstream of the hydraulic motor, the oil

pressure (hereafter known as the system pressure) varies depending upon the driving

torque which the hydraulic motor is applying; the higher the torque, the higher is the

system pressure. In general. The system pressure will increase with RPM; it will also increase

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


momentarily when the centrifuge is accelerating. This is because the hydraulic motor has to

supply sufficient torque to overcome not only the resisting torque but also the inertia of the

centrifuge. Since the RPM is controlled directly by the swashplate swivel angle. the rate of

increase in RPM, that is the angular acceleration is related directly to the rate of wing of the

swashplate. In this pump, the swashplate can swing from zero stroke to maximum in 6

seconds. This implies that the centrifuge will accelerate from rest to full speed in the same

time. This is obviously unsafe as it will subject the centrifuge structure to very high

acceleration and the hydraulic circuit to such high system pressure that the high pressure

pipes may burst. To forestall this occurrence, pressure relief valves are installed in the

hydraulic pump. When the system pressure exceeds the maximum allowable pressure, in

this case 350 bars, the pressure relief valves will blow through, thereby allowing the oil to

short-circuit from the high pressure end to the low pressure end of the pump. With the

maximum allowable pressure thus set, the allowable acceleration of the centrifuge is also

kept below a maximum value.

As mentioned above. the swivel angle of the swashplate is controlled by changing the

electric current flowing through a solenoid connected to the swashplate. The centrifuge can

be decelerated by decreasing the swivel angle and therefore the oil flow rate. During the

deceleration phase, the inertia of the centrifuge causes the hydraulic motor to behave like a

hydraulic pump. In this situation, the downstream pressure will be higher than the upstream

pressure. The greater is the deceleration, the higher is the downstream pressure. To prevent

the downstream pressure from reaching dangerous levels, pressure relief valves set at 350

bars of blow-through pressure are also installed in the pump. Hence, the pressure relief

valves serve to limit the deceleration of the centrifuge to a safe level. Since the value of the

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


blow-through pressure is not adjustable from the control room, they cannot be used to

adjust the amount of braking deceleration. The braking deceleration is adjusted by changing

the blow-through pressure setting of a proportional pressure relief valve installed on a

segment of pipe across the hydraulic motor. The amount of pressure reduction across the

valve, and thus the deceleration, is adjusted by changing the electric current flowing

through a solenoid mounted onto the valve. This electric current is also controlled by a

VT2000S4X control card in the control box.

In addition to the speed control and braking facilities, there is also a pressure relief safety

valve in the drive system to prevent accidental start-up of the centrifuge. The valve can be

switched between open and close states by a solenoid, but unlike the proportional valve,

the amount of pressure reduction is not adjustable. When there is no current flowing

through the valve, the latter is in an open state. This allows hydraulic oil to flow through the

valve instead of the hydraulic motor, thereby preventing the centrifuge from being driven.

The valve can be closed by energising the solenoid from a 24Vdc supply. Once the valve is

closed, it will remain so up to a pressure difference of 350 bars before blowing through. An

analogue logic circuit in the control box ensures that the solenoid will not be energised

under any of the following conditions:

(a) The door leading into the centrifuge bay or the steel door into the centrifuge enclosure is

not properly closed. Microswitches on these doors are used to detect their closure.

(b) An emergency stop switch within the centrifuge enclosure has been activated. This

guards against accidental start-up of the centrifuge when there are personnel in the

centrifuge enclosure.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(c) The electric motor has not picked sufficient speed or the 3-phase power supply to the

motor fails. Two 3-phase supply circuits are fitted onto the starter box of the electric motor.

When the motor is first started, it is driven through a high-current, low-voltage star-circuit

which provides the necessary torque to the motor when the latter is starting up or running

at a low speed. Once sufficient speed has been attained, the starter system automatically

switches to a high-voltage, low-current ∆-circuit. Under no-load condition, the star-∆

switching occurs about 5 seconds after start-up. The logic circuit of the control box monitors

the voltages across the 3 phases of the ∆-circuit. As long as these voltages are less than 90%

of the rated value, the solenoid of the pressure relief valve is not energised and virtually no

load is applied to the electric motor. Thus by preventing centrifuge start-up before the

electric motor has acquired sufficient speed, current overloading of the 3-phase lines are

avoided.

(d) The power-on setting of the speed-control potentiometer is not at zero. When the

control box is switched on, the logic circuit immediately checks to ensure that the speed-

control setting is at zero.

These safety features are represented schematically in Fig. 1.5.

1.4 SPEED, PRESSURE AND TEMPERATURE READOUTS

As illustrated in Fig. 1.6, during centrifuge operation, parameters relating to the state of the

centrifuge are monitored by sensors and displayed on readouts in the control room. These

parameters include the following:

(a) The RPM of the rotor shaft is monitored by an OPTEX AIR-200 photoelectric retro-

reflective transmitter/sensor. The transmitter emits an infra-red beam which is reflected off

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


a reflecting strip attached to the rotor shaft. The reflected pulses are sensed by a sensor and

counted by a Koyo TC-4A tachometer.

(b) The system pressure in the hydraulic circuit. When the speed control is set to zero, there

is effectively no load on the hydraulic drive system and the system pressure should fall to

about 35 bars. When the RPM is being increased, the system pressure rises to a maximum of

about 350 bars to supply driving torque to the rotor shaft.

(c) Temperatures of the centrifuge bearings. hydraulic tank and the inner surface of the

centrifuge enclosure. The resistive temperature device (RTD) bridges used to measure

temperature are non-linear and bridge resistors are used to reduce the non-linearity to

about 2% over the temperature range of interest, that is from 30°C to 60°C. This means that

the accuracy of the temperature readings displayed on the readout units are only accurate

to about ±0.5°C. As the RTDs are only used to forewarn against overheating, this level of

accuracy is considered to be sufficient. The bearings are operational up to a temperature

rise of at least 200°C. whilst the hydraulic oil will retain its viscosity up to a temperature of

about 65°C. If the room temperature at the start of the test is about 30DC, this translates

into an allowable temperature rise of 35°C. From the tests conducted so far, seven hours of

continuous 6 running at about 60g produced a temperature rise of about 10°C in the bearing,

7°C in the oil and 2 to 3°C on the inner surface of the enclosure. The last parameter may also

be taken to represent the ambient temperature of the air in the enclosure.

1.5 SIGNAL AND POWER TRANSMISSION BETWEEN CENTRIFUGE AND CONTROL ROOM

Fig. 1.7 shows an overview of the signal handling system. As can be seen, signals are

transmitted between the centrifuge and the outside by 90 silver-graphite slip rings (ie. silver

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


rings with graphite brushes), termed hereafter as signal rings. Of these, 80 are used to form

40 differential transducer signal channels. The remaining 10 rings are currently unused and

can be utilised for any low current application (less than 1 A) as required by the experiment.

In addition to the 90 signal rings, there are also 10 copper-graphite (copper rings with

graphite brushes) slip rings for power transmission purposes. These power rings have a

maximum current rating of 15A. At present, 3 of the rings are used for the 230VAC live,

neutral and earth

lines while another 2 are used for video signals from the on-board close-circuit television

(CCTV) camera. The other 5 rings can be used for high current transmission up to 15 A as

required by the experiment.

The 230 VAC power supply for the centrifuge can be switched on from within the control

room. The mains power supply is sent through the power rings into a power distribution box

on the centrifuge arm. From here, power is tapped to drive on-board linear power supplies

supplying ±5, ±12 and ±24 Vdc. Linear power supplies are used as they have been shown to

generate less electromagnetic noise than the more common switching power supplies. In

general, the ±5 Vdc are used to power transducers whilst the ±12 and ±24 Vdc are used for

on-board amplifiers, relays and motors. The dc voltage supplies are transmitted via multi-

way connectors and Jump leads into junction boxes which are mounted on the model

container. At present. one standard design of junction box exists. In this design. power is

sent to individual transducers and signals are returned via 5-pin DIN connectors. As shown

in Fig. 1.8, three voltage options are available for powering transducers using this standard

box, viz +5 V. -5 V and 10 V dc. The last option is obtained by connecting the excitation leads

of the transducer across the +5 and -5 V terminals, leaving the 0 V terminal unconnected.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


The existence of a standard model of junction box should not be considered a restriction. It

is envisaged that as users' requirements become more varied, so does the variety of

Junction box designs. Indeed, users with highly specialised and specific requirements are

encouraged to develop specialised Junction boxes for their own use.

Signals from the transducers are also sent through the same junction boxes. From these

boxes, the signals are routed through multi-way jump leads, multi-way connectors, signal

rings and along screened twisted pair cables into the control room. Along the signal route,

cables are screened and star-earthed in order to minimize electromagnetic and ground

noise pick-up.

Within the control room, the signal cables are terminated inside a signal distribution cabinet.

The latter allows the signals to be conveniently monitored through meters or routed to

amplifiers and data acquisition systems as the user requires. As shown in Fig. 1.6, in the

present configuration, the signals are directed to three Signal conditioning and data

acquisition systems as listed below:

(a) General purpose system. In this system, the signals from the 40 differential channels are

routed to a bank of 40 Kyowa DA510B differential voltage amplifiers with adjustable gain up

to

1250. The bandwidth of these amplifiers is from 0 Hz to 10 kHz. Each amplifier also

incorporates an adjustable low-pass filter (10 Hz, 30 Hz, 100 Hz, 300 Hz, 1 kHz or out) to

filter out high frequency noise from signals in pseudo-static experiments. The amplified

signals are digitised and acquired using a PC-based Burr-Brown PCI-20098 multifunction

carrier board with two PCI-20031 analog expander/sequencer modules running on the

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


software LabTechNotebook. This system has a maximum sampling rate of 48 ksamples/sec.

The maximum sampling rate per channel is the above rate divided by the number of

channels in use.

(b) Medium speed system. As shown in Fig. 1.7. signals from the first 8 channels are also fed

into a medium speed system which is connected in parallel with the general purpose system.

In the medium speed system. signals are fed into 8 Fylde FE254GA differential voltage

amplifiers with adjustable gain up to 2500 and frequency bandwidth from 0 Hz to 40 kHz.

Each of these amplifiers are fitted with a FE299SF 4-pole (roll-off 24 dB/oct) low-pass

Butterworth filter with adjustable cut-off frequency (10 kHz. 20 kHz or out). The amplified

signals are digitised and acquired using a PC-based Metrabyte DAS20 analogue input board

configured for 8 differential channels. Digitised data is streamed directly onto ramdisk.

harddisk or floppy disk using a Metrabyte streamer driver software and an in-house

developed supervisory/pre-processor software. Calibration tests on this system shows that

its maximum data acquisition rate is about 90 ksamples/sec onto harddisk and 120

ksamples/sec onto ramdisk. The maximum sampling rate per channel is the above rate

divided by the number of channels in use. The capability of this system makes it suitable for

the low-frequency or steady state dynamic testing where data have to be streamed

continuously onto a large buffer such as a harddisk or ramdisk.

(c) High-speed system. Apart from the two systems outlined above, signals from the first 4

channels are also fed into a third high speed system. In this system. the signals are routed

into 4 Fylde FE351UA wide bandwidth (0 Hz to 100 kHz) differential voltage amplifiers with

adjustable gain up to 10,000. The amplified signals are acquired using a PC-based Metrabyte

DAS50 high speed analogue input card with 4 single-ended channels, driven by an in-house

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


developed software. The maximum sampling rate of this card is 1 Msamples/sec. The

maximum sampling rate for each channel is the above rate divided by the number of

channels in use. This rate is higher than the maximum data throughput rate of the PC bus.

Hence. the card incorporates an on-board buffer capable of storing 256k data points. This

buffer is expandable to 1M data points. Owing to its high sampling rate and limited data

storage capacity. this system is only suitable for very fast transient tests.

The latest PC data acquisition card adopted in the control room is 14-bit PCI DAQ with

maximum 800k samples per second per channel supporting voltage ranges up to ±10v. The

latest software is DASYLab version 11 with icon-based data acquisition, graphics, control and

analysis functions.

For instrumentation using strain gauges, like instrumented model piles, spudcans, etc, the

capturing of the strain readings from the strain gauges are preferably conducted within the

centrifuge enclosure without transferring the signals through the slip ring due to the weak

strain gauge signals which may pick up potential electric noise resulting in lower signal-to-

noise (S/N) ratio. To facilitate such purpose, a strain meter (Model TDS-303) is mounted on

board the centrifuge (see Fig. 1.9) and communicate with a PC running a program called

“Static Measurement” in the control room via the RS232 serial port. A total of 60 strain

gauge channels can be accommodated by using 2 onboard strain meters. Each strain

channel can be configured as single gauge, half-bridge gauge or full-bridge gauge in the

program and strain gauge readings will be saved directly to the control PC during a

centrifuge model test. A user manual of the program “Static Measurement” is compiled in

Appendix C for users’ reference.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Signals from the close-circuit TV camera are channeled into a monitor and time-lapse video

recorder via a remote control unit. This remote control unit allows up to 4 cameras to be

controlled simultaneously with remote illumination and electronic zooming adjustments.

1.6 Image Processing System

Traditionally, to carry out quantitative analysis of soil displacement field during a

geotechnical centrifuge model test, markers were placed on the soil surface in a regular grid

and then images of the marked soil surface are captured using photogrammetry or high-

resolution video. The displacement of the centroid of each marker is then determined by

tracking the movement of that marker through a series of sequential images. In recent years,

the use of the Particle Image Velocimetry (PIV) is gaining popularity due to its high precision

and automation of soil movement displacement analysis. The PIV technique is a texture-

based image processing method which allows displacements of small patches to be

measured to sub-pixel accuracy through statistical analysis with an appropriate algorithm.

GeoPIV is a MatLab module which implements Particle Image Velocimetry (PIV) in a manner

suited to geotechnical testing. More details of the PIV principle and use of the GeoPIV is out

of scope of this manual and centrifuge users are encouraged to refer to the following URL

for more details:

http://www-civ.eng.cam.ac.uk/geotech_new/publications/TR/TR322.pdf

The NUS centrifuge lab owns a copy of the software GeoPIV (courtesy of White D.J. & Take

W.A.) and a quick start of the user manual of the software is included in Appendix E.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


To order to fruitfully make use of the PIV technique, the requirement for an image

processing system with sufficiently high resolution camera which can operate under high g

condition with acquisition of the immense volume of image data accurately and without too

much noise is critical. The image processing system implemented in the NUS centrifuge lab

utilizes a CV-M2 1" CCD Progressive Scan High Resolution digital Camera mounted in front of

the model Perspex window (see Fig. 1.10). The camera possesses capturing rate of 17 full

frames per second with over a million pixels (1.6k x 1.2k) in 1" CCD format, providing

sufficiently high resolution for discerning the details of the images captured. In order to

eliminate the noise picked up via the slip rings, a compact PC with a size of 310 mm in width,

350 mm in length and 80 mm in height was mounted directly onboard the centrifuge, as

shown in Figure 1.10. For the onboard PC to survive the enhanced gravitational level during

a centrifuge test, a special solid-state hard disk was used in the onboard PC. Solid-state hard

disks are collections of solid-state semiconductors which provide faster access time than a

conventional hard disk. Due to the fact that a solid-state hard disk does not rely on a

read/write interface head that is synchronized with a rotating disk, as is the case for a

conventional magnetic hard disk, it is more robust and withstand physical vibration, shock

and extreme temperature fluctuations as well as enhanced gravitational level much better.

The frame grabber adopted in the new image processing system is PC2Vision which is a half-

slot PCI bus image capture board capable of interfacing with standard interlaced (RS-170,

CCIR) or progressive scan (VGA) cameras. Unique to PC2Vision is its built-in 8 MB of VRAM

memory for buffering image data between the camera and host PC system. This feature

allows faster image transfer over the PCI bus as well as simultaneous acquisition and

processing of data.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


In order to control the onboard PC during a model test, a wireless router (model Netgear

RangeMax WPN824) is installed on the roof of the centrifuge enclosure and communicates

with the onboard PC via the wireless network connection, while another control PC in the

control room is wired directly to the router. The WPN824 router provides 802.11g wireless

access with continuous, high-speed 108 Mbps data throughput between wireless and

Ethernet devices. By configuring the IP address of the onboard PC and the control PC with

the same Subnet Mask and Using the “Remote Desktop” facility available in Windows, the

control PC in the control room can connect to the onboard PC. Once this connection was

established, the computer in the control room could virtually take over the onboard PC

during a model test. The desktop of the onboard PC was displayed on the control

computer’s monitor and users could manipulate the onboard PC virtually like sitting in front

of the onboard computer and easily controlling the image capturing process at the user’s

will. A custom made software SEQSNAP has been installed in the onboard PC which can

fulfill various image capturing functions with a maximum image capturing rate of 17 frames

per second.

Appropriate lighting is essential for good, clear image capturing. As shown in Fig. 1.10, two

50-Watt spot light halogen bulbs were installed in front of the model setup and adjusted to

achieve the best lighting effects. During a centrifuge test, the florescent light inside the

centrifuge enclosure was turned off. Thus the power and uniformity of light from these

florescent lights located on the roof of the centrifuge room which used to plague the image

captured was eliminated. Since the spotlight was close to the model container without any

obstructions, and with the elimination of fluorescent lights, the quality of the images

captured was greatly enhanced.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


1.7 FURTHER INFORMATION

The foregoing discussion is only a brief outline of the technical information related to the

NUS Geotechnical Centrifuge. Further information can be obtained from the folder entitled

Centrifuge Technical Information and other documents which are kept together in the

centrifuge laboratory E1-01-01.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


2. ORGANIZATION & OPERATION

2. 1 INTRODUCTION AND OVERVIEW

As a piece of laboratory equipment, the use of the centrifuge automatically comes under

the purview of the Geotechnical Laboratory supervisor who is responsible to the Head of

Department for the smooth, effective and safe operation of the Laboratory. However. the

centrifuge is an expensive and complex piece of equipment which can pose a danger to

personnel and other equipment in its proximity if it is not properly used. In view of this, it is

necessary to set out certain guidelines in addition to those currently set out in the

Laboratory Operations Manual to ensure that centrifuge tests are undertaken within safe

limits and that prospective users are familiar with these limits and guidelines before they

commence their centrifuge test programmes. This means that academic and research staff

who wish to use the centrifuge for their research must have been adequately familiarised

with the centrifuge before they are registered as centrifuge users, hereafter known simply

as registered users. In addition, test programmes proposed by registered users should be

checked by an independent checker to ensure that the test does not exceed the safe limits

of operation. These issues are addressed by the guidelines set out below. which are similar

to those initiated by researchers at

the Cambridge University Geotechnical Centre in the 1970s and have since been adopted by

many other centrifuge centres throughout the world.

2.2 CLASSIFICATION OF PERSONNEL INVOLVED IN CENTRIFUGE TESTS

Personnel involved in centrifuge tests are classified into three categories as listed below.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(a) Registered users. These are academic staff, research staff or research students who are

sufficiently familiar with the operations of the centrifuge to propose test programmes and

undertake them when approved, and to help train other users. During the start-up phase of

the centrifuge operation, staff or research students wishing to register as users will be

required to undergo a short course consisting of seminars and hands-on sessions.

Subsequently, other personnel wishing to register as a user will be required to follow

currently registered users on a few tests for the purpose of familiarisation. Technical staff

are excluded from this group.

(b) Centrifuge operators. These are academic, research or technical staff members who have

sufficient experience of centrifuge operation to advise and assist registered users, mount

packages, operate centrifuge as directed by a registered user within an agreed test

programme.

(c) Checkers. These are academic or research staff members who have sufficient experience

of the centrifuge to check test programmes for other users, train other users or operators,

operate the centrifuge or use the centrifuge as research workers themselves.

It should be mentioned that the above regulations merely reinforces and do not override

existing laboratory operations procedures. Hence, the approval to register new users,

operators and appointment of checkers comes under the purview of the Laboratory

supervisor and the Head of Division who are responsible to the Head of Department, and so

on. Similarly, the role of the checkers is not to control centrifuge usage but rather to provide

expert opinions to the Laboratory supervisor on whether a proposed test is safe, and if not,

then how the safety aspects may be improved. Hence, in the event of a dispute, the decision

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


on whether the test should be allowed to proceed as it is proposed would rest on the

Laboratory supervisor.

Before the commencement of each centrifuge test, the registered user who is undertaking

the test should complete a Proposed Centrifuge Test Programme form and a set of Balance

Calculations and forward these completed documents to a checker. In addition to these

documents, the registered user is also responsible for producing drawings and calculations

which the checker requires to assess the safety of the tests. Once the checker has certified

that the proposed test programme is safe to conduct, the documents will be forwarded to

the centrifuge operator who will then mount the package and the certified counterweight

on the swing platform and proceed as directed by the registered user in the proposed test

programme. The registered user should be present during the mounting of the package, and

will decide if and when various activities (such as test runs) will actually be undertaken

within the test programme. If a checker is using the centrifuge in the capacity of a registered

user, he/she must arrange for another checker to assess the safety of his/her proposed test

programme. The centrifuge operator will be responsible for the filing of these centrifuge

test documents and for

the proper logging of each test in the Centrifuge Test Record Book. If the registered user

wishes to undertaken activities which are not included in the approved test programme, he

should seek the agreement of the centrifuge operator. If the latter feels that the additional

activities may not be within the safe limits of the centrifuge, the registered user may seek

on-the-spot clearance from the checker.

The above division of responsibility is based on the recognition that different parties have

rather different interests in the centrifuge test. The principal concern of the registered user

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


as a research worker is the successful conduct of the test and the acquisition of good

centrifuge data. On the other hand, the principal concern of the checker is merely to ensure

that the test is being undertaken within safe limits. Once this safety criterion is met, the

checker may clear the test. It is not the responsibility of the checker to ensure that the

proposed test will produce good data or to assist in the interpretation of the result. This is

the responsibility of the registered user. The above guidelines also recognise that there may

be a potential source of conflict of interests if the registered user undertaking the test is also

a checker. In such instances, the registered user is not allowed to act as a checker for his

own tests. Instead, he should forward his proposed test programme and calculations to

another checker who is not involved in the tests. Following this philosophy of avoiding

conflict of interest, Laboratory supervisor who is acting in the capacity of a registered user

for a particular test should not override the opinion of the checker in the event of a dispute,

but should instead refer the case of a higher authority.

Members of the technical staff are not allowed to be registered users since they cannot

independently undertake research or consultancy projects. However, they can be registered

as centrifuge operators. The principal concern of the centrifuge operator is the normal

running of the centrifuge. The operator will be responsible for ascertaining that all tests are

properly certified by the checker, that the packages are assembled and secured in a proven

safe manner, and that each test is properly logged. The operator who starts a test will

ensure that all operations begin in the manner set out in the starting and operating

procedures in the next chapter. The operator need not be present continuously throughout

a test, but an operator should be present during activities which cause a significant change

of load, mass or position of the centre of mass within the package, for example excavation,

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


construction, changes in imposed load on the soil and changes in speed of the machine. For

sample consolidation when the centrifuge needs to be run for long periods to allow pore

pressure dissipation and when there is to be no change in the package weight, the machine

may be left running unattended only with the prior agreement of the checker. If at any time

an operator is not satisfied that the facility and the test activities are within the bounds

specified in the certified test programme or are within safe operating limits, the operator

can either refer the case to the checker or terminate the programme without the

agreement of the registered user.

2.3 TEST PROGRAMMES AND AUTHORIZATION

Registered users who wish to use the centrifuge should forward the following documents to

a checker for certification prior to commencement of the test:

(a) outline of the proposed test, on the standard Proposed Centrifuge Test Programme form,

(b) a set of balance calculations and

(c) whatever calculations is required to show that the test is safe.

In order to minimize postponement of tests, registered users should give the checker

sufficient time to assess the safety of the tests.

Furthermore, from time to time, circumstances may necessitate changes in the timing of the

programmes and users must be prepared for this. Each test requires a checker's certification

on the Proposed Centrifuge Test Programme form, a copy of which is shown in Appendix A.

Where the test involves a model container which has already been tested and approved, a

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


sequence of approved test activities can be undertaken by the registered user and the

centrifuge operator in the absence of the checker. Proof testing of new containers requires

a different set of procedures which is covered in Section 2.5.

2.4 BALANCE CALCULATIONS

In order to minimize out-of-balance forces on centrifuge bearings, centrifuge tests should be

conducted with a minimum amount of imbalance. This is achieved by placing an appropriate

amount of counterweight on the opposing swing platform. The main objectives of the

balance calculations is to ascertain that the mass of the package does not exceed the

allowable working payload shown in Fig. 2.1 and to determine the counterweight needed to

balance the centrifuge.

Consider the centrifuge rotating at an angular velocity of ω. The balance condition is

achieved when the centrifugal force of the model, container and testing accessories,

hereafter known as the payload, is equal to that of the counterweight. This can be

expressed as

∫ rω dm = ∫ rω
2 2
dm
payload counterweight
(2.1)

in which r is the radius from the centre of rotation and dm is the mass of an element. Since

ω is constant and is the same for the payload and counterweight, Eq. 2.1 reduces to

∫ rdm =
payload
∫ rdm
counterweight
(2.2)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Eq. 2.2 can be written as

MPRP = McRc (2.3)

in which Mp = mass of the payload,

Mc = mass of counterweight,

Rp = distance between the payload centre of mass of payload and the centrifuge

axis of rotation,

Rc = distance between the counterweight centre of mass and the centrifuge axis of

rotation.

The essence of the balance calculations is to verify that Eq. 2.3 is at least approximately

satisfied. The allowable imbalance is discussed below.

The normal way to locate the centre of mass of the payload and counterweight is to divide

the sum of the first moments of their components by their respective masses. During model

preparation, it is often much more convenient to locate centres of mass of components with

respect to the base of the container rather than the centre of rotation, and use these

measurements to locate the centres of mass of the payload and counterweight. The datum

of measurement is then transfer from the base of container/counterweight to the axis of

rotation to compute Rp and R. To assist users in performing the calculations, a EXCEL

worksheet has been developed to automate a large part of the calculations. The soft copy of

the worksheet is stored in the PCs in the control room. A sample balance calculation using

this worksheet is shown in Appendix B.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


The centrifuge structure is designed to take an ultimate imbalance force of about 400 kN.

The allowable working imbalance force is set to 2% of this value, that is 8 kN or 800 kgf.

Dividing this by ng gives the equivalent imbalance mass at the model centre of mass. Fig. 2.1

shows the allowable imbalance mass at various g-levels. Hence, at 200g, the allowable

imbalance mass is ±4 kg. At 100g, this value can be increased to 8 kg. In most instances, it is

possible to reduce the imbalance mass to a value well below this working limit. Hence, in

their balance calculations. users should aim to minimize the amount of imbalance instead of

just keeping within the working limit. Checkers may, at their discretion, request for a revised

set of balance calculations if it is evident that insufficient effort has been made to minimize

the imbalance.

In general, a set of balance calculations must be shown for the initial configuration of the

model and its accessories. It must be noted that the imbalance mass must be kept within

the allowable working imbalance at all times during a test. Thus in addition to the initial

configuration, a set of balance calculations should also be completed if there is any

significant change in the mass of the payload or its centre of mass. This may occur, for

example during consolidation in which the soil stratum compresses or in the event of a

slope failure. In this situation, balance calculations should also be performed to ensure that

the final configuration of the model is not such as to cause the imbalance to exceed the

allowable working limit.

A secondary objective of the balance calculations is to determine the eccentricity of the

payload centre of mass with respect to the centreline of the swing platform as shown in Fig.

1.4. As far as possible, packages should be configured so that their centres of mass lie

directly above the centreline of the swing. Where this is not possible, the allowable

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


eccentricity is determined by the criterion that the hangers are tension members fabricated

from high strength steel and

the factor of safety to yielding at the critically loaded points of the hanger should not fall

below 3.0 at any time. This leads to the relation

eallow = 50 Mmax / M (2.4)

in which e is the allowable eccentricity from the centreline of the swing, measured in mm,

Mmax is the allowable payload mass corresponding to the maximum g-level to be

used in the test, and M is the actual payload mass.

For example, if the test is to be conducted at 100g, then Mmax = 400 kg from Fig. 2.1. If M =

200 kg, then eallow = 100 mm. The relation between maximum payload, allowable

eccentricity and g-level is shown in Fig. 2.2.

2.5 NEW MODEL CONTAINERS AND EQUIPMENT

From time to time, users will need to fabricate new model containers or customise existing

ones for their experiments. New model containers, as well as those which have been

substantially modified, especially by the cutting of openings, have to be tested to ensure

that it can withstand the soil stresses at the required g-level. The procedure for proof

testing a container is a matter which can be worked out between the registered user and

the checker. However. the following activities are recommended as part of proving a

container.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(a) Stressing checks. The suggested basis for stressing checks of packages on swinging

platforms is as follows. The surfaces of the platforms operate at 1.871 m radius and

stressing calculations should either relate to acceleration of each mass at its actual radius or

to all masses at a nominal 1. 75 m radius. The container should be assumed to be filled with

saturated soil which will become fluidised and apply pressures equivalent to a fluid densIty

of 2100 kg/m3. All water vessels or hydraulic lines should be assumed to be filled with water

either back to the rotor axis and above or to their vent levels if they are vented into the

centrifuge enclosure.

Design calculations should be based on an accepted code of practice for that material and

loading situation and should have an acceptable factor of safety. Where a Perspex window

is required it should be secured by a metal frame with rounded edges not less than 6 mm

radius and is kept free from scoring. In this condition, the Perspex may be assumed to have

a specific gravity of 1.30 and may reach 7 MPa stress. The latter includes a safety factor of

2.5 and a stress concentration factor of 2. For smaller radii, the stress concentration factor

should be increased accordingly.

(b) 1g-tests. As far as possible, 1g testing of the container should be performed prior to

proof testing on the centrifuge at high-g in order to minimize the likelihood of catastrophic

failure of the container during high-g proof testing. During 1g-testing, the container should

be loaded to a level that is deemed to be at least comparable to the design load. One way to

achieve this is by strain gauging critically loaded points. Wherever possible, points which are

shown by the design calculations or analyses to be critically loaded should be strain-gauged.

The container is then brought to a load level at which the stresses at the strain-gauged

points are measured to be at least 25% higher than that designed for. Unloading-reloading

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


cycles are then performed to ensure that the container as a whole and those critically

loaded points in particular have not undergone any plastic deformation. Loading may be

applied by filling up the container with dry soil and then applying an overburden stress on

the soil surface either on the consolidation press or a load frame. Alternatively, the

container may be filled with water and then pressurised either by a hydraulic or a

pneumatic system.

(c) Proof tests. Tests should be run to prove that the container is indeed able to withstand

the applied stresses. Three alternative methods are recommended for proof testing. In the

first, locations which have been shown to be critically loaded in prior analyses are strain

gauged. The container is then filled with saturated soil to the maximum working level and

then brought gradually to full working g-level while the stresses at the strain-gauged points

are monitored. The container is considered to be adequate if the measured stresses do not

exceed the calculated values. If the critically loaded points cannot be located with

reasonable certainty or if these points cannot be strain gauged, the container should be

taken very slowly up to 1.25 times the maximum working g-level. For instance, if the

maximum working g-level of a container is to be 100g, it should be subjected to 125g. If the

maximum working g-level is greater than 160g, using this method would mean running the

centrifuge above the 200g limit. To avoid this, such containers can be tested at 200g using a

dense material such as barite gel with an appropriately scaled-up density. For example, a

container which is to be operated up to 180g should be tested up to 200g with a material

which has a density of 2360 kg/m.

To facilitate potential users in selecting containers, all proof-tested containers will be

inventorized together with their maximum working g-level, soil level, dimensions, location

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


of their centres of mass when empty and photographic records. Some of this information

will also be reflected on the container.

Owing to the varied nature of centrifuge experiments, it is likely that new equipment,

brackets, gantries and rigs will often be mounted onto the package. In the interest of safety,

it is desirable to test all these new model accessories. On the other hand, it is recognized

that doing so would make it unrealistically tedious for potential users to perform their

experiments. It is therefore recommended that only those items which may collapse, fall or

detach themselves during testing, and whose collapse, failure or detachment will cause a

substantial increase in the imbalance mass need to be proof-tested. Proof testing of these

items can follow similar approaches as those discussed above.

2.6 SUMMARY

(1) Categories of personnel involved in centrifuge tests:

(a) registered user

(b) centrifuge operator

(c) checker.

More details are given in Section 2.2.

(2) Documents to be prepared by the registered user for checking by checker before the

commencement of each centrifuge test:

(a) Proposed centrifuge test programme, outlining the test.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(b) Balance calculations. Ensure that the imbalance and eccentricity do not exceed the

allowable values given in Figs. 2.1 and 2.2.

(c) Any other calculations verifying the test will not exceed the safe limits.

More details are given in Sections 2.3 and 2.4.

(3) Recommended procedures for proof-testing new model containers and equipment:

(a) Stressing checks, incorporating design calculations and analyses, to show the

equipment is capable of functioning at its maximum working g-level with an acceptable

factor of safety.

(b) 1-g loading tests to ensure that the equipment is capable of sustaining at least 1. 25

times its design working stress at the critically stressed locations.

(e) Proof testing with the container filled with saturated soil at up to 1.25 times its

maximum working g-level.

More details are given in Section 2.6.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


3. OPERATING PROCEDURES

3.1 INTRODUCTION

In general, two persons are required for a centrifuge test. One, the centrifuge operator will

be responsible for safety and for starting and stopping the machine and the other, the

registered user, will be responsible for the conduct of the experiment. Both these persons

should be present during the starting and stopping phases, during any change of RPM and

during loading, unloading or significant movement of masses within the package.

3.2 FIXING OF ITEMS ONBOARD CENTRIFUGE

Items onboard the spinning centrifuge will experience the centrifugal acceleration which will

increase the self-weight of the item by N times whereby N is the centrifuge acceleration. For

example, an innocuous item of 1 kg located on the centrifuge platform will amount to an

enhanced weight of 100 kg at 100 g. Although items installed on the centrifuge arm will

experience less centrifugal force due to reduced distance from the rotation axis, the

increased force is still substantial, and if not secured properly, the test item could detach

and fly from the centrifuge leading to damage of equipment and posing safety hazards. As

such, it is important to ensure that test items onboard centrifuge are securely fastened and

bolted down.

For test items placed on the centrifuge platform like model strongbox, the centrifugal force

will be in the radial direction and act towards the base of the setup which will in most cases

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


be effectively supported by the platform base plate. As such, test items installed directly on

the centrifuge platform are generally not required to be bolted, except some unusual

circumstances like:

(1) centrifuge dynamic or cyclic tests with quite substantial shaking of the model setup;

or tests involving substantial shifting of center of gravity (CG) leading to unstable

configuration of the whole setup. In such unusual cases, the strongbox must be

bolted securely to the centrifuge platform before spinning up for test.

(2) the test item itself is not very stable even at 1g condition, such as an item resting on

pointed base, or an unsecured cantilever item protruding over the edge of the

strongbox, etc.

On the other hand, all the test items onboard the centrifuge arm will experience laterally-

directed centrifugal force and must be bolted to the centrifuge arm via the threaded holes

that are made available. Other means of fixing, like cable tie or string/rope is absolutely

prohibited to be used for such fixing purpose. It should be noted that centrifuge operator

will conduct the final round of checking with a complete checklist to ensure the proper

securing of test items onboard the centrifuge before the permission of centrifuge spinning is

granted. A complete list of “Centrifuge Operator’s Checklist before Starting the Centrifuge”

can be referred to in Appendix D.

3.3 PREPARATION OF SAND & CLAY MODELS

Preparation of sand model can generally be conducted by pluviation method with the dry

sand rained into the strongbox through a hopper mounted on top of the strongbox as

illustrated in Fig. 3.1. Hoppers for pouring dry sand can be categorized into three types,

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


namely spot, line and plane pouring types. In each case, multiple small holes with regular

patterns are drilled at the base plate of the hopper and the sand is rained into the strongbox

at a fixed height by gradual upward adjustment of the sand hopper as the sand bed within

the strong box raises evenly. In general, the density of sand model thus formed depends on

the diameter of the raining holes of the hopper, the spacing of the holes and the falling

height, and generally the smaller the diameter and spacing of the raining holes and larger

the falling height, the higher relative density of the sand model would result. The actual

sand density should be calibrated by embedding small cups/molds within the sand model as

the sand is rained into the strongbox and filling up the cups/molds whereby the density of

the sand can be determined accordingly.

In case that saturation of sand model is required, de-aired water could be introduced into

dry sand deposit from the bottom of the strongbox, and preferably with the aid of suction

applied to the enclosed strongbox as the de-aired water permeates the sand model from

the base. The opening of valve introducing the water into the sand model should be

controlled carefully so as to not causing any disturbance or even boiling of the sand bed. In

some dynamic centrifuge model tests when there exists a profound conflict between the

scaling effect of consolidation (scaled to 1/n2 of the prototype behavior) and dynamic

deformation (scaled to 1/n of prototype behavior), silicone oil of N cSt viscosity under Ng

centrifuge condition can generally be adopted to substitute water to resolve the scaling

conflicts.

For centrifuge model tests on cohesive soils (either commercially available kaolin clay or

Singapore marine clay sampled from construction site), the clay model ground is generally

formed by consolidating soil-water mixture from the slurry state with high water content of

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


say 1.5~2.0 times its liquid limit. The water-soil mixture should be thoroughly mixed in an

automatic soil mixer for a minimum of 3 hours and preferably with simultaneous vacuuming

system to maximize the saturation ratio of the clay slurry thus formed, as illustrated in Fig.

3.2 for the automatic soil mixer with simultaneous vacuum facility installed in the centrifuge

lab. The thoroughly mixed clay slurry will then be transferred to the model strongbox for

high g self-weight consolidation (for NC clay) or subjected to 1g pre-consolidation using

deadweight or pneumatic jack if higher pre-consolidation pressure is desired before placing

to the centrifuge platform for high g self-weight condition leading to a model ground with a

top layer OC clay.

3.2 CHECKLIST OF PROCEDURES FOR CENTRIFUGE OPERATOR

3.2.1 Balancing the centrifuge.

(1) Check that the balance calculations and the Proposed Centrifuge Test Programme form

for that test has been completed and certified. Refer to Appendix A and B for the

samples.

(2) Check that the actual weight of the test package is within 5 kg of that stated in the

balance calculations.

(3) Check that the actual weight of the test package will not cause the imbalance mass to

exceed the allowable imbalance mass shown in Fig. 2.1. If this condition is not met, the

user should change the counterweight accordingly and reflect this change in a new

balance calculations.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(4) Mount the counterweight on the opposing swing platform with the help of the user,

checking to ensure that the steel plates are bolted down tightly.

(5) Check to ensure that the correct counterweight has been mounted.

3.2.2 Before starting the centrifuge

(6) Check that the centrifuge bearings (4 grease nipples) and slip ring bushing (1 no.) are

adequately greased and oiled.

(7) Examine the package and ensure that all accessories on the model such as fittings,

cables, junction boxes, gantries, plugs, valves, LVDTs etc are properly secured and

verified by an inspector independently. Refer to Appendix D for the Checklist to be

completed by the Centrifuge Operator before starting the centrifuge.

(8) Check that the swing can rotate to a horizontal position without any parts of the

package hitting any part of the arm especially the end plate and without fouling any

cables, connectors, hydraulic pipes and fittings.

(9) Proceed to the control room to switch on the video remote control unit, video recorder

and monitor. Check that the camera is giving a clear view of the end of the arm. Since

the package is usually out of view in the swing-down position, a substitute object at the

approximate swing-up position may have to be used for this purpose.

(10) Switch on the power to the tachometer as well as the system pressure and

temperature readouts.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(11) Check that the tachometer is showing a RPM and the pressure readout is showing 1

bar.

(12) Zero the temperature readings on the temperature displays using the zeroing knobs;

subsequent readings on the temperature readout units will reflect the temperature

change from the initial temperature.

(13) Check to ensure that the speed and brake setting potentiometers on the control box

are turned fully anticlockwise.

(14) If the test is to be computer-controlled, boot up the control PC and invoke the control

program (refer to Section 3.3 for more details for the PID computer control system).

Following instructions in the program, set the test phases as directed by the registered

user in the certified Proposed Centrifuge Test Programme Form.

(15) Turn the key switch S1 on the control box followed by the red switch S2 at the lower

left hand side of the box to turn on the 230 Vac supply. Then switch on the 24Vdc

supply by the 24Vdc supply by the green switch S3 on the lower right hand side.

(16) Check that the relay supply switch (S6) is on, that the doors and master override

switches are off.

(17) Switch S4 on the control box to either manual or computer mode depending on

whether the test is manually or computer controlled.

(18) Check that the voltage and current readings for the speed and brake control displays

are almost at Zero. If the readings differ substantially from zero, the control box is

faulty and the test should be halted.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(19) Walk round the centrifuge enclosure, checking to ensure that no tools, equipment or

any loose items are left lying on the package, arm and enclosure floor.

(20) Close and securely bolt the steel door of the centrifuge enclosure.

(21) Check that the service door leading from the access road to the centrifuge motor is

locked. Remove the key from the lock.

3.2.3 Starting the machine

(22) Switch on the isolator for the 3-phase power supply for the centrifuge motor and the

cooling system.

(23) Turn on the cooling water system for the centrifuge motor.

(24) Depress the starter switch to start the centrifuge motor. Listen for 5 to 10 seconds to

ensure the star-∆ switch has been activated. At this point. the centrifuge should not

be rotating.

(25) Close and lock the access door to the centrifuge bay. Remove the key from the lock.

Display the sign board showing that test is in progress.

(26) Enter the relevant test parameters into the log book.

(27) Check that all the red lights along the top of the control box are off and that the green

lights for the brake and speed control circuits are Qn. This indicates that centrifuge is

now ready to start.

(28) Check with the registered user that he/she is ready to start the test.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(29) Determine the required RPM from the g-level and the radius to the package centre of

mass using Fig. 3.3.

(30) Using Fig. 3.4 and the required RPM, determine the voltage setting on the control box

display.

(31) Ensure that the switch S4 is on, then turn the speed control potentiometer slowly

clockwise until the required voltage setting is reached.

(32) Check that the RPM on the tachometer is stabilising to a value at least close to the

required RPM. Then, using the tachometer reading as a guide, fine-tune the speed to

the required RPM, if necessary.

3.2.4 During running

(33) In the first 30 minutes of the test, monitor the bearing temperatures to ensure that the

rise does not exceed 5°C. Also monitor the system pressure to ensure that it does not

significantly exceed 350 bars. During acceleration phases, system pressures may rise

to 350 bars, but should stabilise to a lower value once the speed has reached a steady

value.

(34) To change speed during a test, repeat steps 29 to 32.

3.2.5 Stopping the centrifuge

(35) Check that the brake control potentiometer is in a fully anti-clockwise position.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(36) Slowly turn the speed control potentiometer fully anticlockwise, then switch S4

to off. This will allow centrifuge to free swing to a halt.

(37) If braking is required to help decelerate the centrifuge, slowly turn the brake

control potentiometer clockwise. The tachometer readings should now reflect

the increased deceleration. Generally, It is not necessary to turn the brake

control potentiometer fully clockwise to get sufficient braking. A brake solenoid

current of 400 to 500 mA is usually sufficient to bring the centrifuge to a halt

within 2 to 3 minutes.

(38) When the centrifuge has come to a halt, turn the brake control potentiometer

fully anticlockwise to release the brake.

(39) If the test is over, switch off control box, readout units and video displays.

(40) Depress stop switch on the starter box of the centrifuge motor to stop the motor.

(41) Depress stop switch on the cooling system box to turn off cooling water.

(42) Turn off 3-phase power supply to centrifuge motor and cooling system.

3.2.6 In the event of a power failure

(43) Follow the normal procedures for stopping the centrifuge as listed in sub-Section

3.2.5.

(44) To re-start centrifuge after power recovery, follow the procedures listed in sub-

Section 3.2.3.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


3.2.7 Emergency stopping of the centrifuge

(45) Switch off 54.

(46) If rapid braking is required, turn the brake control potentiometer fully clockwise.

(47) While centrifuge is decelerating, turn the speed control potentiometer fully

anticlockwise.

(48) After the centrifuge has come to a halt, activate steps 38 to 42.

3.3 PID COMPUTER CONTROL SYSTEM FOR CENTRIFUGE OPERATION

Manual operation of centrifuge is mainly meant for some maintenance trial runs,

calibrations and certain emergency spinning down, etc, while normal centrifuge tests will be

controlled by an in-house software CFGCTRL (shorthand for Centrifuge Control) which

employs the Proportional-Integral-Derivative (PID) programming controller for an automatic

precision centrifuge spinning control with a closed-loop control scheme. CFGCTRL is a menu-

driven user interface program likened to the ‘brain’ of the automatic centrifuge control

system as it controls the entire flight operation right from the controllable spinning up of

centrifuge to final application of controllable braking of the centrifuge. It was written in

QuickBasic Version 4.5, compiled with the QB compiler and linked to a firmware library,

DASG16.LIB, supplied by Metrabyte Inc. This library contains all the essential command

procedures necessary for activating the various functions of the Metrabyte DAS16 data

acquisition card.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


3.3.1 Closed Loop Control Circuit

Fig. 3.5 shows the closed-loop control system. In implementing the control circuit, the

hardware consideration was straightforward because many of the available components for

manual control as elaborated in Section 3.2 could be used for the automatic control system

and the additional components are relatively simple to install. As explained earlier, the

hydraulic drive system employs proportional swash-plates for speed control and

proportional valves for deceleration. These proportional valves are electronically controlled

by proportional amplifier cards Rexroth VT2000S40 housed in the control box in the

centrifuge control room Each amplifier (connected to one DAC) receives voltage signals from

the DAC and converts them into slowly rising or falling output electric currents that feed to

the proportional solenoids of the hydraulic drive system.

The hardware of the computer controller system consists of a Metrabyte DAS16 data

acquisition card housed in a desktop PC. There are two Digital-to-Analog-Converters (DAC)

and eight Analog-to-Digital Converters (ADC) on the data acquisition card. Five of the ADC

channels are used for the monitoring of the temperatures and hydraulic system pressure

during centrifuge operations as illustrated in Fig. 3.5. On the other hand, the speed of the

centrifuge during flight is monitored directly by a tachometer which is a photoelectric retro-

reflective transmitter and sensor that is installed at the base of the centrifuge at a standoff

of approximately 300mm away from the rotor shaft. The transmitter is oriented such that it

generates an infra-red beam towards the centrifuge’s rotor shaft. At the surface of the shaft

is attached a reflective strip that reflects off the beam and sends it to the sensor. The sensor

converts each signal received to a pulse and is output to a frequency to voltage (f/v)

converter, as illustrated in Fig. 3.5.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


3.3.2 PID Controller Algorithm

The PID controller operates on an iterative scheme. As illustrated in Fig. 3.5 (enclosed within

the rectangular dash line box), the PID controller utilizes an iterative scheme whereby the

controller constantly adjusts the variables in real time to optimize the system’s performance.

Rn is the reference (i.e., desired) value stored in the computer. Mn is the manipulated

process input (in this case, the input into centrifuge). Cn is the process output (i.e., output

from centrifuge); En is the error, or steady state error, i.e.

En = Rn – Cn (3.1)

The PID controller is of the form

Mn = Mn-1 + K0En + K1En-1 + K2En-2 (3.2)

where K0 = Kp + KiT + Kd/T, K1 = -Kp - 2Kd/T and K2 = Kd/T.

Kp is the proportional parameter that performs correction to Cn, in proportion to the error

En. Ki is the integral parameter that performs correction to Cn proportionally to the teim

integral of En. The derivative parameter Kd performs correction ot Cn proportionally to the

derivative of En with respect to time. T is the sample period. Centrifuge testing runs give the

following empirical values of the parameters for the NUS centrifuge spinning automatic

control: K0 = 1.067, Ki = 1.001 and Kd = 0 (disabled).

3.3.3 Control Software

The implementation of the PID controller algorithm, the closed-loop real time controlling of

centrifuge spinning at specified angular velocity (thus the specified g level for the test

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


model), the monitoring of various operating conditions as well as the application of braking

to centrifuge are all controlled by the menu-driven user-interface program CFGCTRL written

in QuickBasic Version 4.5, compiled with QB compiler and linked to a firmware library

DASG16.LIB supplied by Metrabyte Inc.. This library contains essential subroutines for

driving the data acquisition card Metrabyte DAS16. Before a test commences, the operator

logs the essential flight and test information, i.e. test model effective radius and payload,

test g-levels, test duration and ramp-up acceleration rate into the program. All input data,

including PID control parameters are saved in the files created by CFGCTRL for future

retrieval. During flight, CFGCTRL performs real-time automatic control of the speed and

monitoring of temperature sensors on-bard the centrifuge and in the enclosure as well as

the system pressure of the hydraulic drive system. The test is terminated automatically

when any of these parameters exceed the desirable safety limits.

An interrupt service is available to allow the flight to be interrupted either during

acceleration or steady-state g-level and let he operator log in a new set of data to override

the current set. Multi-stage g-level execution capability (maximum 15 stages) within a test

run enables centrifuge users to simulate varying loading forces within specified scale time

intervals set for each stage. Options for prolonged gentle braking during normal centrifuge

spinning down to avoid potential surge of hydraulic system pressure and accelerated

braking in case of emergency are also available in the software.

3.4 COOLING VENTILATION SYSTEM FOR CENTRIFUGE ENCLOSURE

Long hours’ continuous spinning of centrifuge machine, especially for centrifuge tests

involving consolidation process of soft marine clay with very low permeability, will inevitably

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


cause the over-heating of the centrifuge machine. This is especially so for the rotating

components like the upper and lower bearing of the centrifuge, the slip rings, etc. The long

duration of over-beating will aggravate the detrimental effects of wear-and-tear of the

centrifuge machine. The Operator can easily experience the very warm ambience within the

centrifuge enclosure after long hours of centrifuge spinning. On the other hand, the

increasing temperature within the centrifuge enclosure as a centrifuge test proceeds will

also cause the adverse effect of thermal shift of sensors for a model test. As such, to

alleviate these unfavorable thermal effects during centrifuge tests, a air cooling ventilation

system has been incorporated with the centrifuge enclosure for temperature control. As

illustrated in Fig. 3.6, the air cooling ventilation system starts with two air conditioner units

installed outside the enclosure. The aircon unit is DAIKIN Model R125LUY15 (outdoor use)

with rated power supply 380~415Volts and rated maximum current of 12.5A. The cooling air

generated is channeled via the circulation duct to the entry point on the roof of the

centrifuge enclosure as illustrated in Fig. 3.6. The circulation duct is insulated by aluminum

foil for better thermal insulation; the centrifuge enclosure wall is a 0.5m thick composite

ferro-cement-granular wall which possesses excellent thermal insulation as well. In the

same time, an air suction machine is installed at the main entrance of the centrifuge

entrance to further facilitate the cooling air circulation system as shown in Fig. 3.6. During a

centrifuge model test with the centrifuge spinning, the cooling air will enter by the roof

entry point, circulates in clockwise rotating direction with the spinning centrifuge and exit at

the air suction machine point, which provides a very effective temperature control system

within the centrifuge enclosure and thus a very conducive model test ambience. Many years

of model tests in the NUS centrifuge lab reveals that the temperature within the enclosure

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


essentially never exceed 30 degree Celsius, even during some special tests with over 24

hours continuous centrifuge spinning.

3.5 ITEMS FOR ROUTINE MAINTENANCE

3.5.1 Hydraulic drive

(1) Hydraulic oil level. Top up if necessary.

(2) Oil filter. Change if soiled or clogged.

3.5.2 Cooling system

(3) Water pump filters. Change if soiled or clogged.

(4) Water pumps Check that both are functioning.

(5) Cooling water. Change if the water is dirty.

3.5.4 Centrifuge structure

(6) Bearings.

(7) Slip ring bushing.

(8) Swing platform hinges. Check that items 6 to 8 are properly greased and oiled.

(9) Swing platforms. Check for any signs of structural distress or overloading, such as

pronounced deformation, cracking and so on.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(10) Check for rust patches and take necessary remedial actions.

3.6 CHECKLIST TO ASSIST REGISTERED USERS

Registered users do not have to follow the procedures in the checklist strictly. They are only

set out to assist users.

(1) During model preparation, check that all components constituting the package are

properly weighed and their locations noted. This will greatly assist you in your

balance calculations and ensure that the predicted weight of the package is close to

the actual weight.

(2) Prepare a diagram of your model showing the transducers used and their locations.

Pin this up on the wall or a table in the control room so that you can quickly and

conveniently assess transducer performance during the test.

(3) Set your amplifier gain and cut-off frequency to the required values using the gain

controls on each of the amplifiers. Technical information on the amplifiers and

instructions for their usage can be found in the relevant amplifiers' manual in the

centrifuge laboratory.

(4) If your test is a long duration test, transducer drift caused by the warming effects of

power in the strain gauges as well as amplifier drift may become significant. In such

cases, leave the transducer and amplifier power supplies on for at least 12 hours

before starting the test to allow their temperatures to stabilise.

(5) Ensure that the signals from your transducers are set to zero or the appropriate

datum using the zero adjustment potentiometers on each of the amplifiers, and

monitoring the amplifier outputs with a voltmeter.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(6) Configure your data acquisition system properly well before the start of the test,

especially if you are using many data channels.

(7) If you are activating motors, pumps, hydraulic supplies or any other means of loading,

test out your actuating system at 1g before running the centrifuge test.

(8) Ensure that you are fully ready before giving the centrifuge operator the go-ahead to

start the centrifuge.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.1 Layout of centrifuge laboratory

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.2 Layout of centrifuge bay

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.3 The NUS geotechnical centrifuge

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.4a Plan view of NUS geotechnical centrifuge

Fig. 1.4b Side view of NUS geotechnical centrifuge

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.5 Hydraulic circuit of centrifuge drive system

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.6 Overview of centrifuge control room electrical communication systems

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.7 Overview of signal handling system

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 1.8 Junction box DIN plug pin assignment

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Strainmeter
DTS303

Fig. 1.9 Strainmeter DTS-303 mounted onboard centrifuge

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Onboard PC

Cameras &
Spotlights Spotlight
Frame

CCTV Camera CV-M2 Camera

Fig. 1.10 Setup of image processing system

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 2.1 Relation between Max payload and imbalance mass vs G-level

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 2.2 Relation between payload and eccentricity vs G-level

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 2.3 Relation between RPM vs G-level

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 3.1 Sand model ground prepared by pluviation method

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Pneumatic cylinder

Move-up/down
switches

Speed controller

Cylinder piston

Vacuum pipeline

Mixer motor assembly

Acrylic cover

Clay mixing container

Fig. 3.2 Clay mixer with simultaneous vacuuming

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 3.3 RPM setting vs G-level (Nominal Radii 1.55 - 1.871m)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Fig. 3.4 Control box voltage setting vs RPM

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Braking Controller DAS16 - DAC

CFGCTRL

Desired Signal Output Signal


Manipulated Signal
Rn Cn
Mn
+
PID Controller DAS16 - DAC
_

DAS16 - ADC f/v converter Tachometer

Oil tank thermal sensor


Upper bearing thermal sensor
Lower bearing thermal sensor
Enclosure thermal sensor
System hydraulic pressure

Fig. 3.5 Control box voltage setting vs RPM

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Cool air entry at enclosure
roof

Cool air
circulation duct Air Conditioner
Model Daikin-
Ferrocement R125LUY15
enclosure wall

Cool air flow as


centrifuge spins

Air suction
machine CENTRIFUGE
ENCLOSURE
Cool air entry viewed
within the enclosure

Fig. 3.6 Illustrative cool air circulation system for centrifuge enclosure during tests (not to scale)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Appendix A
ABSTRACT

Sample of Proposed Centrifuge Test Programme

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Appendix
NATIONAL UNIVERSITY OF SINGAPORE
Department of Civil & Environmental Engineering
PROPOSED CENTRIFUGE TEST PROGRAMME

Test Identifier: 100g-T1(S2)


User: Yang Yu
Test Date: 29-Mar-11

Estimate Duration
Activity / Event G-level Package Mass (kg)**
(Hours)

for example: Spud installation, etc. 100 8 332.2

** Including on-board equipment and accessories

Instrumentation:
LVDT
PPT
Loadcell
(others add accordingly)

Special Requirement (tick where applicable)

1 After office hour operation ( √ )


2 Compressed air facilities ( √ )
3 Hydraulic facilities ( √ )
4 High voltage (230 AC or higher) ( √ )
5 Special lighting ( √ )
6 Others (please specify):

Certified by ______________________________________ (Name/Signature/Date)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Appendix B
ABSTRACT

Sample of Centrifuge Balance Calculation

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


The spreadsheet for centrifuge balance calculation is separated into 4 portions. The first

portion is for the centrifuge user to fill in the basic information of the centrifuge test,

including the test identifier, name of the user, maximum g-level of the test, and nominal

radius (from the center of rotation to the targeted point of the model ground for the

targeted g level), and other relevant information. This portion provides the centrifuge

operator with an overview information of the test objective and for obtaining the g-level

and nominal radius required for input to the centrifuge control program.

NATIONAL UNIVERSITY OF SINGAPORE Version: 03-2011

Department of Civil & Environmental Engineering

Geotechnical Centrifuge Balance Calculations

(For New Bucket Only)

Test identifier: Objective:

User: Test date:

Maximum g-level: g Nominal Radii: mm

The second portion of the spreadsheet concentrates on the payloads at the swing platform

side where the model setup is located. In particular, the first part lists 6 on-board items

which has been permanently installed onboard the centrifuge arm at the payload side and

centrifuge users are not supposed to touch these items in most cases. The 2nd part is for the

user to fill in any test items installed on the centrifuge arm at the payload side. The

information needs to fill is the mass of each item and the distance of c.g. of the item to the

base of the old platform, which can be determined by minus from 1871mm the distance of

the item from the centrifuge rotation center. The third part is for the user to fill in all the

components of the centrifuge model setup mounted at the active swing platform. Again, the

information needs to fill is the mass of each component items and the distance of centre of

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


mass of the item to the base of the swing platform. Unlike the items mounted on the

centrifuge arm listed in part 2 described above, the eccentricity of the item from the center

of the platform must be measured and input into the spreadsheet to facilitate the

calculation of final overall eccentricity to check that is should be within the machine

tolerance limit, as will be further discussed below. The spreadsheet will sum up all the items

mounted on the swing platform and check to make sure that this total weight should not

exceed 40,000g-kg/Ng, or 400kg in a typical model test at 100g. Otherwise, a warning

message will be displayed to alert user’s attention. If all the items have been input properly,

the total “moment to center shaft” will then be summed up by the spreadsheet for all the

items on the payload side and displayed at the end of this 2nd portion of spreadsheet.

PAYLOAD
DIST. OF CG ABOVE ECCENTRICITY (mm) MOMENT (kg-mm)
CG above base of
S/N ITEM MASS (kg) new bucket (mm) OLD BASE z (mm) x y Mx My Mz
---------------------- -------------- ----------------------------- --------------- ----------------- ----------------- ----------------- ----------------- -----------------
1. Built-in items on the Centrifuge arm at payload side (don't touch):
1 On-board power supplies 25.70 1500 0 0 0 0 38550
2 On-board camera & frame 9.00 1472 0 0 0 0 13248
3 On-board side switch box 2.00 1572 0 0 0 0 3144
4 On-board junction box 2.00 1472 0 0 0 0 2944
5 On-board cable with clamp 1.80 1432 0 0 0 0 2578
6 On-baord camera control unit 3.50 1172 0 0 0 0 4102

2. User items mounted on the Centrifuge arm at payload side (Add here:)
1 (onboard) 0 0 0 0 0
2 (onboard) 0 0 0 0 0
3 (onboard) 0 0 0 0 0
4 (onboard) 0 0 0 0 0
5 (onboard) 0 0 0 0 0
3. User items mounted on active swing platform
1 Due to New Bucket 50.2 -1218 0 0 0 0 -61144

2 0 0 0 0 0
3 0 0 0 0 0
4 0 0 0 0 0
5 0 0 0 0 0
6 0 0 0 0 0
10
7 0
0 0
0 0
0 0
0 0
0
----------------------------- --------------- ----------------- ----------------- ----------------- ----------------- -----------------
Total weight on active swing platform (kg) (should be < 400kg) : 50.2
Total weight on active swing platform + centrifuge arm (kg) : 94.2 Z= 36 mm 0 0 0 0 3422

*** MOMENT TO CENTRE OF SHAFT = 172920 kg-mm

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


The third portion of the spreadsheet concentrates on the payload at the counterweight side

for balancing the payloads listed in the second portion of the spreadsheet described

previously. In particular, the first part lists 2 on-board items which has been permanently

installed onboard the centrifuge arm at the counterweight side and centrifuge users are not

supposed to touch these items in most cases. The 2nd part is for the user to fill in any items

user installs on the centrifuge arm at the counterweight side. The information needs to fill is

the mass of each item and the distance of c.g. of the item to the base of the old platform,

which can be determined by minus the distance of the item from the centrifuge rotation

center from 1871mm which is the distance from the centrifuge rotation center to the base

of the old platform. The third part is for the user to fill in all the components of the

centrifuge model setup mounted at the counterweight platform, for example, another

strongbox for simultaneous consolidation purpose while the model setup proper at the

payload platform side is undergoing consolidation or testing. For proper balancing purpose,

one single base counter weight of 50 kg and multiple counterweight steel plates in unit

weight of 40, 20, 10, 5, 2 and 1 kg are provided and users should adjusted the number of

counterweight steel plates until the imbalance displayed is minimized. Again, the

spreadsheet will sum up all the items mounted on the counterweight platform and check to

make sure that this total weight should not exceed 40,000g-kg/Ng, or 400kg in a typical

model test at 100g. Otherwise, a warning message will be displayed to alert user’s attention.

If all the items have been input properly, the total “moment to center shaft” will then be

summed up by the spreadsheet for all the items on the counterweight side and displayed at

the end of this third portion of spreadsheet.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


COUNTERWEIGHT
DISTANCE OF C.G. MOMENT
ITEMS MASS (kg) ABOVE BASE (mm) (kg-mm)
-------------- -------------------------- ----------------- -----------------
1. Built-in items on the Centrifuge arm at counterweight side (don't touch):
1 On-Board Manifoldblock 4.0 1466 5864
2 On-board oil container 4.4 1422 6257

2. User items mounted on the Centrifuge arm at payload side (Add here:)
1 0.0 0 0
2 0.0 0 0
3 0.0 0 0
4 0
5 0
6 0
3. Counterweights & other items on the conterweight platform:
(e.g. Concurrent consolidation container?) 0.0 0 0
Counterweight (kg) pieces
50 1 (1 piece only) 50 6.4 318
40 1 40 19.3 770
20 1 20 29.1 581
10 1 10 34.0 340
5 1 5 36.7 184
3 1 3 38.5 115
2 1 2 39.5 79
1 1 1 40.1 40

(Hint: May Reduce Counterweight !) -------------------------- ----------------- -----------------


Total counterweight on counterweight swing (should be < 400kg) 131.0
:
Total counterweight on counterweight swing + centrifuge arm, 139.4 Z= 104 Z= 14547

MOMENT TO CENTRE OF SHAFT = 246410 kg-mm

With all the information provided by the user above, the spreadsheet will calculate the

imbalance moment, namely the difference between the total payload moment and the total

counterweight moment, and derive the equivalent imbalance mass at the platform base by

dividing the distance from the center of rotation to the base of the platform 1871 mm,

which should be less than the allowable imbalance of 800/g level. In the meantime, the

overall eccentricity of the model setup is calculated which should be less than the allowable

eccentricity of 50Mmax/M whereby Mmax is the maximum allowable payload on the active

swing platform as explained earlier, while M is the actual payload in the current test which

is available from portion 2 of the spreadsheet. In general, users should try to minimize the

imbalance mass and overall eccentricity as much as possible instead of just satisfying the

allowable limits of the centrifuge machine.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


IMBALANCE MOMENT :

PAYLOAD MOMENT - COUNTERWEIGHT MOMENT = kg-mm

EQUIVALENT IMBALANCE MASS AT BASE DISTANCE = kg Allowable imbalance= kg

(=800 / g level)

OVERALL ECCENTRICITY OF PACKAGE (mm) = mm

Allowable eccentricity= mm

(= 50 Mmax / M)

A complete sample of centrifuge model test balance calculation is provided below for user’s

reference.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)
Appendix C
ABSTRACT

User Manual of Strain Gauge Program “Static Measurement”

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Procedure of Using the StrainMeter Software:

1. double the icon “Strain meter” to open the software.

2. From “Envi” menu, click “selecting data logger”. The setting should be as follows:

Data logger: TDS-303


Interface: RS-232C
Port: Com1
Bps: 19200
Parity: Even
Data Length: 7 Bit
Stop Bit: 1 Bit
Delimiter: CR/LF
HandShaking: XON/OFF
SW-BOX selection: click “Internal
Click “OK” to finish data logger selection.

3. From “Ch.ID Set” menu, click “Setting Channel ID Table”, The setting should be as
follows:

for “on/off” column, double-click to toggle on/off the channels. set “on” for those
channels to be used for your tests.
for “Meas. Mode” column, double-click to pop-up a submenu, select “Measure”.
For “SensorMode” column, double-click and select the type of strain gauges you used,
like 2 gauges or 4 gauges etc.
For all the other columns, keep them as they are.

4. From “Moni. Set”, click “Monitor Setting”, you can select how to display your strain
gauge reading during the test, say “digital Monitor”, and/or “T-Y Monitor”. Double click
the lower panel and set the channels “on” to those you have connected you strain gauges
to.

5. From “Trg. Set” Menu, click “Settings Interval”, set the date as the testing day, the time
should be about 2 to 3 minutes ahead of the present computer clock time (as displayed at
the lower-right corner of the computer screen). Set the Interval to your desired data
capturing frequency, typically 1 minutes per data recording during consolidation, 3 to 5
seconds during test. The “control” column should be set as “INFINITY”.

6. From “Meas” menu, click “run” command, the programme will enter the running status
and another screen will pop up. If you have not properly connected your cables to your
computer or your strain-meter is not power on, there will be an error “RS232 error”. In
this case, check that all the hardware connection should be ok.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


7. Click the icon of “Digital Monitor” or/and “T-Y Monitor” from the toolbar, click “Tile
Horizotal” from the “Window” menu.

8. Click the “Internal Trigger” icon from the toolbar, then the “Start” icon. The data capturing
procedure will begin.

9. When the test finish, click the “Internal Trigger” icon, the program will ask you internal
measurement. Answer “yes”.

10. From the “file” Menu, click “exit”, the program will ask you whether you want to “Store
data and come to end?”. Answer “yes” and enter a file name. Take note this file is now
saved in a binary format and is not recognisable by EXCEL, etc.

11. After saving the data file, the program will return to the setup screen. Click “Rec & Print”
menu and select “file conversion”. On the left table select the file name you have just
saved, on the right-hand side, select your personal folder and enter a file name. Click “ok”.
The binary datafile will be converted to an ASCII file in your personal folder which can be
compiled in spreadsheet like EXCEL.

12. At this stage, you should save your configuration so far to a file so that you don’t need to
setup again next time you use. Click the “save measurement file” from the “File” menu,
and enter a file name. Next time you do your test, simply open this measure file from the
“open measurement file” from the “File” menu, and all the setting will be restored
automatically. You can then proceed from step 5.

13. If you want to re-start your data recording again. You can start from step 5 to set your
“Trigger Setting”

14. In case of an unexpected errors happen, the programme may be automatically shut-down.
But a temporary hidden file named “~~NewTest.mdb” will be automatically saved by the
computer in the folder “c:\ProgramFile\Static_Software\test\”. quickly save this files to
your personal folder for data recovery. Don’t run the software again, otherwise the
temporary file will be overwritten and all the data will be lost. Please contact senior users
when such case happens.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Appendix D
ABSTRACT

Centrifuge Operator’s Checklist before Starting the Centrifuge

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Centrifuge CHECKLIST for Setting Up

Operator Inspector
A. Setting Up

1) Ensure test container properly secured.

2) Ensure no loose items on board the centrifuge machine.

3) Do not use string, rope to tie any accessories (e.g. electrical controller) on board.

Bracket with bolts and nuts used must be able to withstand the intended g level.

4) Ensure the counter weight is correctly placed and they must be tally with the required
mass stated on the balance sheet.

5) Inspect the surrounding of the centrifuge machine to ensure nothing is obstructing the
rotation of the machine.

* Inspector should come in to check independently after operator completed part A

B. Preparing and spinning of centrifuge machine

6) Ensure strong encloser door is properly closed and no personnel are inside the enclosure.

7) Switch on the blower pump and exhaust fan.

8) Switch on hydraulic pump ( 1 OR 2 ).

9) Switch on the cooler.

10) Ensure the two emergency switches are functional before the test.

11) Ensure the outer wooden door is properly closed and operation light is switch on.

NAME OF OPERATOR: _________________________ DATE:_________________

SIGNATURE: ___________________________________

NAME OF INSPECTOR: _________________________ DATE:_________________

SIGNATURE: ___________________________________

JOB TITLE: _______________________________________________________________

NAME OF USER:__________________________________________________________

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Appendix E
ABSTRACT

GeoPIV quick start user manual

(courtesy of White D.J. & Take W.A.)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)
NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)
NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)
ABSTRACT
Appendix F

Scaling Principles for Geotechnical Model Testing in Centrifuge Tests

by

Tan Thiam Soon

(November 1991)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


There are two main reasons for testing reduced scale models, namely:

a. to study qualitatively soil phenomena and

b. to use the reduced scale models to predict quantitatively the behaviours of prototype
structures.

To ensure that the behaviour is properly simulated, correct principles of similitude must be
established. To obtain these principles, one can use either similarity analysis or dimensional
analysis (if feasible).

Let's take a look at a simple example of what is needed to achieve correct scaling. Consider
a reduced scale model in which the linear macroscopic dimensions have been reduced by a
factor N. Further assume that the soil used in the model is the same as that in the prototype.
For the time being, let's assume that this implies that the constitutive model for the soil in
the model is the same as that in the prototype. This means that:

xp = N xm (1)

ρp = ρm (2)

where x is some linear dimension, ρ the density and the subscripts m and p refer to the
model and prototype respectively. Now consider the equation of motion of a body of soil:

~ ~ d 2 u~p
∇ p ⋅σ p + bp = ρ p (3)
dt 2p

~ d 2 u~m
∇ m ⋅ σ~m + bm = ρ m (4)
dt m2

The ∇ is the gradient of a function, σ~ ~


is the stress tensor, is the body force vector
b
u~pforce,
which in most instances is just the gravitational is the displacement vector and t is
the time. As the linear dimensions have been scaled according to (1), this means that

∇p 1
=
∇m N (5)

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


Let's consider what happen if the reduced scale model is now tested in a normal laboratory
condition, that is, in a 1g condition. In this case, the gravitational forces will be the same
implying that:

~ ~ (6)
b p = bm

For the time being, consider a static situation. that is the right hand sides of (3) and (4) are
zeros. (5) and (6) will then imply that for correct similitude,

σ~ p
=N (7)
σ~m

Thus if the gravitational effect is important, then (7) has to be satisfied which means that
the stress level in the model has to be N times smaller than that in the prototype. But as soil
is a highly non-linear material, the strains in the model will not be properly scaled and
therein lies the problem of testing in 1g.

~
Note that if the gravitational effect is not important, then b can be neglected from (3) and
(4), and (7) is therefore not a necessary requirement. To ensure that the same stress-strain
level are mobi1ised in these circumstances, the applied stress level in the reduced scale
model is kept the same as in the prototype. This is the basis for testing in an odeometer for
Terzaghi’s type consolidation and determining soil properties using triaxial apparatus.

Back to the problem where gravitational effect is indeed an important consideration. What
can you do? Let’s suppose that the gravitational force in which the reduced scale model is
going to be subjected to is increased by a factor N, that is

~ ~ (8)
bm = N ⋅ b p

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


With this change, it can be shown that in order for the left hand side to obey similitude, the
stresses at homologous points in the model and prototype must be the same, that is:

σ~ p
=1
σ~m (9)

To check this, consider a soil element at the base of a thick layer of soil and its
corresponding position in the model:

hp
=N
hm

g Hp

Ng Hm

Prototype Model

From the figure above, it can be observed that the overburden stress in the prototype is
given by

(σv)p = (ρ)p g hp (8)

For the corresponding point in the model, the stress level is

(σv)m = (ρ)m g hm (9)

As hp is N.hm, it can be seen from (2), (8) and (9) that (σv)p is the same as (σv)m. This thus
ensures that the stress level is the same and therefore similitude is satisfied for the left hand

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


sides of (3) and (4). As the soils in the model and prototype are the same, the strains
induced are also expected to be the same.

When the inertia is not negligible, then the right hand sides of (3) and (4) have to be
considered. As the strain has been established to be the same, (1) therefore implies that:

u~ p
=N (10)
u~
m

Together with (2), (3) and (4), the inertia (dynamic) time scaling can now be determined:

tp
=N (11)
tm

This means that dynamic event will occur N times faster in the model than that in the
prototype. With (1) and (11), the scaling for other dynamic parameters can now be
established, namely:

vp
=1 (12)
vm

ap 1
= (13)
am N

v and a are the velocity and acceleration respectively. Note that the scaling for the
acceleration is consistent with that in (6). Note too that these scalings are for dynamic
problems.

When the soil is saturated with a fluid, besides the inertia, another time parameter that
often has to be considered for geotechnical problem is the time taken in consolidation or
seepage flow. In the theory of consolidation. the inertia effect is often neglected. but the

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


time effect appears in the continuity equation and is linked to the equation of equilibrium
through Darcy's law which is related to the viscous drag of a fluid flowing through a porous
media. The key question is whether this viscous time scaling is the same as the dynamic
time scaling?

According to most texts in soil mechanics, the coefficient of permeability as we normally


used is:

γ
k= K
µ (14)

where γ is the unit weight, µ is the viscosity of the fluid and K is referred to as the specific or
absolute permeability by Lambe and Whitman (1979). If the same fluid and soil are used in
the model and prototype, then it is expected that the viscosity and the absolute
permeability will remain the same. In this case, using (2) and (6) in (12) will give:

kp γp 1
= = (15)
km γm N

Darcy·s law is given by:

v=ki (16)

where v is the seepage velocity and i the hydraulic gradient, a dimensionless parameter. If
Darcy's law is now assumed to be valid in both the model and prototype, then:

vp k p ip 1
= ⋅ =
vm k m im N (17)

A quick comparison with (12) will quickly show that the two scalings are different. Since the
linear scaling is fixed, the viscous time scaling can now be obtained from (17) and is:

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


(18)
tp
= N2
tm
This time scaling is sometimes also referred to as the diffusion time scaling and implies that
the diffusion process is N2 times faster in the model than in the prototype. It is also obvious
that this time scaling differs from the dynamic time scaling given by (11).

The existence of two different time scalings can often pose a problem. When a soil with fluid
is tested, there is only one complicated physical process going on. The separation into
dynamic time and viscous time is for our own benefits to differentiate the processes. In
other word, if both the dynamic and the diffusion effects (I have used it in place of viscous)
are important, similitude will break down as the two processes cannot be simultaneously be
correctly scaled. If only the dynamic effect is important, (11) to (13) will hold. Alternatively,
in a steady state seepage problem or a consolidation problem, then (17) and (18) will hold.

But what about dynamic problems involving the generation and subsequent dissipation of
pore pressures, for examples, in liquefaction problem? As a solution, some researchers have
used in the model a fluid with a viscosity N times higher, for example silicon oil or glycerin.
In this case,

µm (19)
=N
µp

kp γ p µm 1
and = ⋅ = N =1 (20)
km γm µp N

With this, it can now be shown that

vp k p ip
= ⋅ =1 (21)
vm k m im

and the scaling can be made to be consistent again. However, in using a different fluid, a
number of things have to be ascertained, for example is the density sti!l remains the same,

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


does Darcy's law still applies and will the constitutive relation of the soil changes? If the
Reynolds Number is now considered, then:

(Re) p d p vp ρ p µ p
= ⋅ ⋅ ⋅ = 1 ⋅1 ⋅1 ⋅ N = N (22)
(Re) m d m vm ρ m µ m

where d is a typical size of the pore space which is assumed to be the same in the model as
in the prototype since the same soil is used. This means that all the requirements of
similitude cannot be satisfied. In using the centrifuge to study such problems, this violation
can be contained provided it can be shown that Darcy's law is still valid in both the model
and prototype (Tan and Scott, 1985). The linearity of Darcy's law will then allow for
consistent scaling.

Note that if dimensional analysis is used, a violation like this will pose a problem. Thus in a
such situation, similarity analysis has to be conducted, and the severity of the violation
assessed.

Testing in a Centrifuge

In the discussion so far, the centrifuge has not been mentioned yet. This omission is
deliberate and is intended to demonstrate an important point, that in centrifuge testing.
conceptually two

simulations are involved (Tan and Scott, 1985). First, the model is assumed to be tested in a
Ng environment and the question of similitude has to be answered. After that, the
centrifuge is assumed to be Simulating the Hg environment and the accuracy of that
simulation has to be assessed.

In using the centrifuge to simulate the Ng environment, two common sources of errors are:

a. The centrifugal acceleration at a point is ω2r where r is the radius from the axis of
rotation and w is the rate of rotation.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


b. If the velocity of soil particles relative to the bucket is not insignificant, Coriolis forces are
present.

axis of rotation
Radius R from

aR

These two sources of errors have been checked by Schofield (1980). Following Schofield's
work, suppose the model and prototype both have zero total pressure on their surface. Let
R be the effective radius of the model under test, that is, there is a correct pressure σv’ in
the 1/N scale model at this radius R, which is at a depth aR below the soil surface. Note that
a is normally a small factor. In the prototype, the corresponding depth is NaR and the total
stress there is σv’ = NaRpg. In the centrifuge, the variation of acceleration with radius means
that the total stress is given by:

R ρϖ 2 (23)
σ v '= ∫ ρrϖ dr =
2
R 2 a (2 − a ) = NaRρg
R (1− a ) 2
So,

Ng 2−a (24)
=
Rϖ 2
2

Now consider a point away from this effective radius, say at a depth aR/2 from the surface.
The total stress at this point in the model is now given by

R (1− a / 2 ) ρϖ 2 3 (25)
σv = ∫ ρrϖ dr =
2
R 2 a (1 − a )
R (1− a ) 2 4

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


The stress at the corresponding point in the prototype is given by σv = P g N a R/2. Using (24)
and (25), the error involved is therefore given by:

 ρgNaR / 2  a
ε =   = (26)
 ρϖ R a (1 − 0.75a ) / 2  4 − 3a
2 2

If a = 1/10, the error is about 2.7%. This suggests that the error involved 1s rather small in
most instances. One situation where the error may become appreciable is when very long
piles need to be tested and the soil layer needed is much larger than 0.1R.

The other possible source of error is due to Coriolis acceleration which is given by 2vw
where v is the velocity relative to the bucket. The ratio of this acceleration to the centrifuge
acceleration is given by:

Coriolis _ acceleration 2vϖ 2v


= 2 = (27)
Centrifugeacceleration ϖ R ϖR

For NUS centrifuge, R ≈ 2m. If the centrifuge is spinned to 100g, that is, N = 100, then w =
22.1 rad/sec. Then for a seepage experiment with a fairly high seepage velocity of 0.5 m/s,
the error

involved is only 2.3%. However, this effect still has to be assessed if the experiment involves
free falling particle. The other effect that needs to be studied in such a case is the fact that
the direction of Coriolls force is perpendicular to the centrifugal acceleration which is an
extra component that does not exist in the prototype.

Earlier on, the assumption was made that if the same soil as in the prototype was used in
the model, then the constitutive relations of the soils in the two cases were the same. One
conceptual problem that people normally faced is this: To relate to the prototype, the linear
dimension in the model has to be scaled by a factor N. This means that the sol1 in the model
will actually be simulating in the prototype, a soil with particle size N times larger. Then how
can one be sure that the constitutive relation remains the same.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)


There are two answers to this. An empirical answer is that researchers have tested the soil
properties in the two situations and found that they are generally the same. The second
answer is more subtle and partly explains the first answer. Note that in our entire discussion
of similitude, we have use parameters assuming that the soil is a continuum. This is the key.
If indeed the soil is behaving like a continuum, then the individual particle size is not a
critical factor as long as all the relative dimensions ensure that this assumption does not
break down. However, if gravel size particles are used in the model, then some attention
has to be paid to this problem.

As a final conclusion, in this discussion we have shown that by testing a soil in a centrifuge,
we can ensure that in situations where gravitational stresses play an important role,
principles of similitude can be enforced with marginal error. However, one cannot run away
from the fact that in the model, a small sample of soil is used and this can never fully replace
the complexity of a real soil layer. However, if the stresses involved are of the correct
magnitude, then testing a model in the centrifuge would allow one to see a more correct
deformation and failure pattern. Other methods of research then have to be used to
complement these findings.

References:

Lambe, T.W. and Whitman, R.V., 1979, Soil Mechanics. 51 Version, John Wiley and Sons.
Schofield, A.N., 1980, "Cambridge Geotechnical Centrifuge Operations," Geotechnique, 30,
No.3, pp. 227-268.

Tan, T.S. and Scott, R.F., 1985, "Centrifuge scaling considerations for fluid-particle systems,”
Geotechnique, 35, No.4, pp. 461-470.

NUS Geotechnical Centrifuge Manual – Version 2 (June 2011)

You might also like