22
QUT Digital Repository: http://eprints.qut.edu.au/ Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of heavy vehicles – update on recent test programme and proof-of-concept. In Wrigley, Cara and Lawry, Simon, Eds. Proceedings Built Environment and Engineering Postgraduate Student Conference, QUT, Brisbane. © Copyright 2008 QUT

QUT Digital Repository:  · Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

  • Upload
    vokhanh

  • View
    215

  • Download
    0

Embed Size (px)

Citation preview

Page 1: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

QUT Digital Repository: http://eprints.qut.edu.au/

Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of heavy vehicles – update on recent test programme and proof-of-concept. In Wrigley, Cara and Lawry, Simon, Eds. Proceedings Built Environment and Engineering Postgraduate Student Conference, QUT, Brisbane.

© Copyright 2008 QUT

Page 2: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

1

Lloyd Davis1, Dr. Jon Bunker2

In-service suspension testing of heavy vehicles – update on recent test programme and proof-of-concept

ABSTRACT: One response to requests from the transport industry to allow more “freight efficient” heavy vehicles (HVs) onto the road network has been that road authorities have allowed higher axle loads in return for HVs being equipped with “road-friendly” suspensions. These suspensions (particularly those with air springs) are critically dependant on shock absorber health for proper operation. They are only certified, however, as “road-friendly” at the time of manufacture and this via a type-test. Once in service, the “road-friendliness” is determined solely by the maintenance regime of the transport operator. There is no in-service test for HV suspensions in Australia yet. Over 50% of Australian HVs do not meet at least one of the criteria for Australian requirements of “road-friendly” suspensions. The Australian Government and the State Governments of New South Wales and Queensland are funding a programme to develop an in-service suspension test for HVs. This paper examines some possible low-cost test methodologies for an in service HV suspension test and their results within the context of a “proof-of-concept” test programme. The results show that low-cost testing is possible and is as accurate as the high-cost methods used for the original “road-friendly” certification.

KEYWORDS: Heavy vehicle; suspension testing.

Introduction One response to requests from the transport industry to allow more “freight efficient” heavy vehicles (HVs) onto the road network has been that road authorities have allowed higher axle loads in return for different (“road-friendly”) heavy vehicle suspension designs. Characterisation of these types of suspensions is the subject of this paper.

Background The Australian transport industry is focused heavily on road transport. Extra heavy vehicle (HV) payloads are allowed by most road authorities if HVs are equipped with “road friendly” suspensions (RFS). The move to higher payloads using HVs with RFS has allowed the road freight industry to somewhat absorb the increasing demand for long-haul freight.

1 Queensland University of Technology & Queensland Main Roads 2 Queensland University of Technology

Page 3: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

2

To somewhat mitigate the conflicting demands of more payload vs. responsible consumption of the road network, road authorities have allowed HVs to operate at slightly increased loadings or “higher mass limits” (HML) if the HV has a “road-friendly” suspension (RFS) fitted. The Australian specification for RFS is VSB 11 (Australia Department of Transport and Regional Services, 2004b). The original mass limits review project report recommended allowance of such overloads. That report also made authoritative statements regarding the nexus between road network asset damage, damper health and friendliness of air suspensions (National Road Transport Commission, 1993; OECD, 1998). Despite this, Australia at present has no specific requirement for HVs operating at HML loads to have their RFS tested to the Australian specification once the HV is in service. Previous work has documented the following (Davis, 2006; Davis & Bunker, 2007; Davis, Kel, & Sack, 2007; Davis & Sack, 2004, 2006), sometimes in more detail, but it is reiterated below for context. In-service testing was supported in theory by international studies (Cebon, 1999; Gyenes, Mitchell, & Phillips, 1992) and at least two Australian studies (Starrs Pty Ltd, Ian Wright and Associates, & ARRB Transport Research Ltd, 2000; Sweatman, McFarlane, Komadina, & Cebon, 2000) after the implementation of HML in Australia. Sweatman et al., (2000) investigated in-service suspension testing techniques. That report to the NRTC (as it was then) covered:

� shock absorber testing machines and methods; � shock absorber characteristics; � shock absorber wear; � suspension performance; and � any concomitant degradation due to that wear.

A lower acceptable bound of 15% (0.15) for damping ratio on operational HVs as determined from in-service testing was recommended. Sweatman et al., (2000) also suggested an alternative method for maintaining RFS health: regular replacement of suspension dampers as part of the normal maintenance discipline in the transport operator’s workshop. That report went on to note that governments (i.e. regulators and road authorities) were requiring the development of low-cost in-service testing methods that were able to determine body-bounce frequency, damping ratio and load sharing. The frequency and damping ratio parameters defined in VSB 11 are determined for axle-to-body movement. Accordingly, Sweatman et al., (2000) discussed the requirement for any in-service test to have instrumentation which measured these parameters and eliminated (or reduced to an acceptable level) any contribution that tyre dynamics made to body bounce. Starrs (2000) took the outputs of the Sweatman et al., (2000) report and examined the economics of shock absorber replacement and the use of tests as defined by the final report of the DIVINE project (OECD, 1998). The Starrs work also explored HV suspension evaluation and maintenance in a general manner. Neither report proposed a definitive way forward (Starrs Pty Ltd et al., 2000; Sweatman et al., 2000).

Page 4: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

3

At the time of the Starrs and Sweatman studies, the estimated savings on pavement rehabilitation in the case of all RFS operating correctly was $14M across the Australian HV network. That quantum was not considered to be cost-effective within the context of total pavement rehabilitation costs and operational costs to test RFS HVs (Starrs Pty Ltd et al., 2000). It is noted that one of the authors of the Sweatman et al., (2000) report had already recommended, at the international level, type-testing of RFS using parametric or other means combined with annual in-service testing. That work detailed extensively the options for the associated test equipment needed for these measures in a regulatory framework (Cebon, 1999). Others (Potter, Cebon, & Cole, 1997; Woodroofe, 1995), agreed. Even earlier, the need to test new-generation HVs for characteristics which contributed to their “road-friendliness” was recognised as “probable” (Woodroofe, LeBlanc, & Papagiannakis, 1988). Despite these efforts and the case in favour of in-service testing there is still no requirement for in-service testing of HV RFS parameters and no concomitant in-service test recognised in Australia. Such is the concern surrounding the issue of network asset damage from air-suspended HVs with deficient suspension dampers that two Australian States (Australia Department of Transport and Regional Services, 2005a, 2005b) have, in conjunction with the Australian Government, included conditional clauses in their current Bilateral Infrastructure Funding Agreements (BIFAs). National activity on in-service HV suspension testing in Australia lapsed from the time of the Starrs and Sweatman studies until NSW’s and Queensland’s BIFA negotiations in 2004 – 2005. Previous papers describing Queensland Main Roads’ efforts into an in-service HV suspension test research programme have documented “proof-of-concept” research using impulse tests from driving an HV over a pipe and other results from impulse testing using a pipe vs. VSB 11-style step-down testing (Davis, 2007; Davis & Bunker, 2008; Davis et al., 2007; Davis & Sack, 2004, 2006). This paper examines in more detail the results of the most recent activities in Queensland’s research programme.

Theory for determination of parameters of 2nd order systems viz: HV suspensions Characterising heavy vehicle (HV) suspensions is central to the EU (and Australian) test for “road friendliness” (Australia Department of Transport and Regional Services, 2004a; European Council, 1996). Two measurements used to show that heavy vehicle suspensions are road friendly are the damping ratio, zeta (ζ) and the damped free vibration frequency (f). The damped free vibration body-bounce frequency is the frequency at which a truck’s body has a tendency to bounce on its suspension with the largest excursions whilst being damped by the suspension dampers (shock absorbers). The damping ratio is a measure of how fast a system reduces its oscillations (and returns to quiescent or steady state) after a disturbance. It is a measure of the reduction in excursions of subsequent amplitudes of the output signal from a 2nd order system. In HV suspensions, it is related to a measure of how quickly the shock absorbers and other components reduce body bounce and wheel hop after the truck

Page 5: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

4

hits a bump. The damping ratio, zeta (ζ), is a dimensionless number and is usually presented as a value under 1.0 (e.g. 0.3) or a percentage (e.g. 30%) denoting the damping present in the system as a fraction of the critical damping value (Doebelin, 1980). Chesmond (1982) showed that system parameters may be characterised in a number of ways. Amongst these were: application of a random input signal to a system. Random signals are sometimes known as “white noise” and contain all frequencies in equal proportion; Fourier (or other frequency domain) analysis of the output signal resulting from that random input may be used to determine the characteristics of the system transfer function. The damped free vibration frequency (f) of the system characterises that transfer function and will show up as the largest magnitude peak in the frequency spectrum of the output signal after the application of “white noise” as an input signal; or application of an impulse input signal to a system: a perfect impulse signal contains all frequencies in equal proportion. Again, Fourier analysis of the output signal may be used to determine the characteristics of the system transfer function. Similar to random input signal excitation, the dominant (in this case, damped free vibration) frequency will manifest as the largest peak in the frequency spectrum of the output signal for a given impulse input. Subjecting any system to an impulse signal and measuring the reducing excursions of the output signal enables the damping ratio of a system to be determined. The input signal used to excite the system is known as the forcing function. Doebelin, p79 (1980) said, on the subject of length of time over which the impulse function is applied and its shape: ‘We see that if [the input function’s] duration is “short enough”, the system responds in essentially the same way as it would to a perfect impulse of like area and that the shape of [the input function] makes no difference whatsoever.’ The damping ratio (ζ) may be determined by comparing the values of any two consecutive response peaks in the same phase (i.e. comparing the magnitudes of the 1st and 3rd excursions or the 2nd and 4th excursions) of the output signal of an underdamped 2nd order system after an impulse function input has been applied (Meriam & Kraige, 1993). Prem, et al., (2001) used the formula (Meriam & Kraige, 1993):

22 )2( πδδζ

+=

(1)

Where:

δ is the standard logarithmic decrement (Meriam & Kraige, 1993) given by the following formula:

Page 6: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

5

=

2

1lnA

(2) Where:

A1 is the amplitude of the first peak of the response; and A2 is the amplitude of the third peak of the response to determine damping ratio.

Alternatively, A1 and A2 may be described as the first two peaks of the response that are in the same direction (i.e. on the same side of the x-axis of the time-series signal of the response). Where it is desired to determine damping ratio from a signal with more than 2 peak values of the signal on the same side of the x-axis in a time series, Thomson & Dahleh (1998) provide:

=

nx

x

n0ln

(3) to substitute into (1): Where:

xn is the amplitude after n successive cycles have elapsed; and x0 is the amplitude when n = 0; or for the case where continuous successive peaks are present (Technical Committee ISO/TC 22, 2000):

+++

−=

+13

2

2

1 ......1

1ln

n

n

x

x

x

x

x

x

(4) Note: (1) may be derived by solving for ζ in the following equation (Meriam & Kraige, 1993) as shown in other work (Davis & Bunker, 2007):

21

2

ζπζδ−

=

(5)

Damping ratio formula for a half-cycle Thomson and Dahleh (1998), amongst others (Meriam & Kraige, 1993) provide the basis for the derivation of the formula for damping ratio in an underdamped 2nd order system. Already dealt with (Davis & Bunker, 2007) has been the derivation of the classical equation for the damping ratio zeta (ζ):

22 )2( πδδζ

+=

(6)

Page 7: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

6

from the equation:

21

2

ζπζδ−

=

(7)

which was derived from first principles of the equations of motion of 2nd order systems (Meriam & Kraige, 1993; Thomson & Dahleh, 1998). This derivation for:

=

2

1lnA

(8) Where:

A1 = amplitude of the first peak of the response; and A2 = amplitude of the third peak of the response; or A1 and A2 as the first two peaks of the response that are in the same direction (i.e. on the same side of the x-axis of the time-series signal of the response) as shown in FIG 1.

FIG 1. Illustrating the values used to derive damping ratio of a 2nd order system.

The starting point for this part of the exercise is (7) derived from the equation of motion for a 2nd order underdamped system. In deriving it, the value forδ, being the ratio of two peak values, is required. It, in turn can be derived from the same equations of motion used to derive (8), by restating (7) (Thomson & Dahleh, 1998):

21

2

ζπζτζϖδ−

= = dn

(9) Where:

ζ is the damping ratio;

Page 8: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

7

ωn is the undamped natural frequency; and

dτ is the damped natural period (FIG 1).

Where a half-cycle of the response of a 2nd order system to an impulse is available, ζ may be found by using: the first two peaks: A1, A1.5; and

half the damped natural period 2dτ

;

as shown in FIG 1, as follows:

the period between A1 and A1.5 will be half the damped natural period or2dτ

;

substituting this period into (9) and adjusting the other sides of the equation for equality:

22/112 ζπζτζϖδ−

= = dn

(10) So, for the derivation of the damping ratio for a half-cycle, Where:

=

5.1

12/1 ln

A

(11) equating only the first and last parts of (10) yields:

22/11 ζπζδ−

=

(12)

⇒ ⇒

⇒ ⇒ ⇒ ⇒

2

222

2/1 1 ζζπδ

−=

2222/1

2 )1( ζπδζ =−

2222/1

222/1 ζπδζδ =−

2222/1

222/1 ζπδζδ +=

( ) 2222/1

22/1 ζπδδ +=

222/1

22/12

πδδζ

+=

Page 9: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

8

⇒ ⇒

(13)

Experimental procedure

Previous papers discussing this research programme have documented impulse tests using a pipe vs. VSB 11-style step-down testing (Davis, 2007; Davis & Bunker, 2008; Davis et al., 2007; Davis & Sack, 2004, 2006). The testing procedure used for that work and this paper is reiterated here briefly for context. The results analysed in this paper are for the 2-axle school bus so the test procedure for that vehicle will be addressed with references, where required, also to the semi-trailer and the coach used for the tests. A 2-axle school bus was used for part of the tests. The instrumentation relevant to the context of this paper was accelerometers and air pressure transducers (APTs). These were installed on the drive axle of the school bus. Figs 2 to 5 show photos, diagrams and details of the bus that is the subject of this paper.

Equipment and instrumentation Air pressure transducers (APTs) were mounted in the air lines to the air springs as shown in FIG 3. They were used to measure the air pressure in each air spring and therefore the static and dynamic forces between the axle at that spring and the chassis. Accelerometers were mounted on the drive axle close to the hubs (FIG 4). Daily checks on the quiescent outputs of the instruments showed slight variations due to vehicle supply voltage fluctuations. The steady-state values were noted and the relevant calculations or calibration graphs adjusted accordingly.

222/1

22/1

πδδζ

+=

222/1

2/1

πδ

δζ+

=

Page 10: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

9

FIG 2. 2-axle school bus used for testing (LHS) and sacks of horse feed used to load it (RHS).

An advanced version of the TRAMANCO on-board CHEK-WAY® telemetry system was used to measure and record the dynamic signals from the outputs of the strain gauges and accelerometers. FIG 5 shows the CHEK-WAY® recording system used for the tests.

Sampling frequency The telemetry system sampling rate was 1 kHz giving a sample interval of 1.0 ms. Note that the natural frequency of a typical heavy vehicle axle is 10 - 15 Hz (Cebon, 1999) compared with a relatively low 1 - 3 Hz for sprung mass frequency (Davis & Sack, 2006; de Pont, 1999). Any attempt to measure relatively higher frequencies (such as axle-hop) using time-based recording would necessarily involve a greater sampling rate than when relatively lower frequencies (such as the body-bounce frequency) are to be determined (Houpis & Lamont, 1985). Since axle-hop was the highest frequency of interest for the analysis undertaken, the sampling frequency used by the telemetry system was more than adequate to capture the test signal data. This because its signal sample rate was much greater than twice any axle-hop frequency (e.g. 1 kHz sample rate vs. 15Hz [max.] axle hop). In theory, a sampling rate of 30Hz would have sufficed but the telemetry system with 1 kHz was available for the tests so was used. Accordingly, and to check the validity of the choice of sampling frequency, the Nyquist sampling criterion (Shannon’s theorem) was met (Houpis & Lamont, 1985).

FIG 3. Air pressure transducer (arrowed) used for measuring forces at air springs.

Page 11: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

10

FIG 4. Accelerometer (arrowed) mounted on top of bus axle

FIG 5. Data capture management computers (LHS) and data capture and storage telemetry units (RHS).

Quasi-static VSB 11-style testing The bus drive axle was loaded to the maximum legal load and driven at least twice off an 80mm step to replicate the VSB11 step test (Australia Department of Transport and Regional Services, 2004b). All wheels were rolled off a set of blocks simultaneously (Peters, 2003). The signals from the air pressure transducers on each air spring were recorded using the on-board telemetry system during this test procedure. FIG 6 to FIG 8 shows examples of these tests (these for the coach but the bus was subjected to the same treatment). Chains (top left, FIG 6) attached to the chassis were used to drag the blocks once the wheels had moved off them so that the wheels were not fouled as they rolled subsequent to the step-down action.

Page 12: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

11

FIG 6. Before: showing preparation for the step test.

FIG 7. During: ready for the step test.

FIG 8. After: the step test that was set up in FIG 6.

Impulse testing – the pipe test Each axle tested was driven at least twice over a 50mm nominal diameter heavy wall pipe. The pipe had bars welded to it as shown in FIG 9 to prevent rotation as the tyres moved over it. The speed requested from the driver for this exercise was, as well as the driver could manage, at or just above 5km/h. The resulting pulse up into the axle and air spring was intended to provide an approximation for an impulse function into the suspension of the axle of interest on each test vehicle. Since the speedometers of the test vehicles did not register at speeds below 5km/h, actual speeds for these tests were measured from elapsed time between two marks 10m apart on the test vehicles’ paths using an elapsed time stopwatch. Actual speeds ranged between 4.9km/h and 9.6km/h.

Page 13: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

12

FIG 9. The pipe for the impulse test, bottom left.

FIG 10. Close up view of wheel rolling over the pipe during impulse testing.

Results

Air spring data The dynamic chassis-to-axle (body-bounce) forces were recorded from the outputs of the APTs for the step test and the pipe test. Figs 11 and 12 show time-series plots of typical signals recorded during the impulse testing outlined in the previous section. FIG 11 shows an example of a time series recorded from the bus drive axle APT signals during a VSB 11-style step test. FIG 12 shows an example of the same APT signals for a pipe test. In system analysis terms, the two time-series in FIG 11 and FIG 12 could be taken as classical 2nd order system response to impulse inputs.

Page 14: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

13

Bus drive axle APT signal - VSB 11-style step test

1550

1650

1750

1850

1950

2050

3.0

3.1

3.2

3.3

3.4

3.5

3.6

3.7

3.8

3.9

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

7.1

7.2

7.3

7.4

7.5

7.6

7.7

7.8

7.9

8.0

Time (s)

AP

T o

utp

ut

(arb

itra

ry li

nea

r sc

ale)

LEFT RIGHT

FIG 11. Time series of bus axle APT signals for impulse testing using VSB 11-style

step test.

Bus drive axle APT signal - pipe test

1650

1700

1750

1800

1850

1900

1950

3.0

3.1

3.2

3.3

3.4

3.5

3.6

3.7

3.8

3.9

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

7.1

7.2

7.3

7.4

7.5

7.6

7.7

7.8

7.9

8.0

Time (s)

AP

T o

utp

ut

(arb

itra

ry li

nea

r sc

ale)

LEFT RIGHT

FIG 12. Time series of bus axle APT signals for impulse testing using pipe.

Page 15: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

14

Accelerometer data The dynamic axle forces were recorded from the outputs of the accelerometers at the hubs of each axle of interest for the step tests and the pipe tests. FIG 13 and FIG 14 show examples of time series recorded from the outputs of the accelerometer during VSB 11-style step tests and pipe tests.

Bus drive axle accelerometer signal - VSB 11- style step test

1900

1950

2000

2050

2100

2150

2200

2250

2300

2350

2400

2450

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

Time (s)

Acc

eler

om

eter

ou

tpu

t (ar

bitr

ary

linea

r sc

ale)

LEFT RIGHT

FIG 13. Time series of bus axle accelerometer signals for impulse testing using VSB

11-style step.

Bus drive axle accelerometer signal - pipe test

2050

2100

2150

2200

2250

2300

2350

2400

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

Time (s)

Acc

eler

om

eter

ou

tpu

t (a

rbit

rary

lin

ear

scal

e)

LEFT RIGHT

FIG 14. Time series of bus axle accelerometer signals for impulse using pipe.

Analysis The number of tests was low due to scheduling considerations. Accordingly, statistical analysis was not possible with the low number of test runs per test type and load condition.

Page 16: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

15

FIG 11 provides what may well be taken to be a very good example of the classical model of the expected output response of an underdamped 2nd order system to an impulse function. Regarding FIG 11 and using the variables illustrated in FIG 1 it was fairly straightforward to derive the ζ value for the single drive axle on the bus (Meriam & Kraige, 1993; Thomson & Dahleh, 1998). Using the values of A1, A1.5 and A2 from FIG 11 and FIG 12, and substituting into (6), (8) and (13), a comparison between the derived LHS and RHS values for ζ was made for:

� the two types of impulse forcing function; � ζ using full cycle values A1 and A2; and � ζ using half cycle values A1 and A1.5.

These are shown in Table 1.

Variable LHS RHS

VSB 11-style step test

Pipe test

VSB 11-style step test

Pipe test

Quiescent signal value 1812 1824 1777 1774 A1 180 78 178 75 A1.5 36 15 40 19 A2 34 16 34 16 ζ [from (6) where

=

2

1lnA

Aδ ]

0.2564 0.2342 0.2548 0.2388 ζ [from (13) where

=

5.1

12/1 ln

A

Aδ] 0.4559 0.4931 0.4292 0.4005

Table 1. Comparison between ζ values for the two types of impulse testing – bus. Error in the readings from the APTs used for these tests has been documented previously (Davis, 2006; Transport Certification Australia Limited, 2007) at approx. ±1%. This was typical of APT readings emerging from work being undertaken currently in a related project (Karl, 2007). We note that that experimental error (including the inaccuracy of the APT signals) was contained within the results for each impulse method. Clearly the ½ -cycle response was not 2nd-order, otherwise the two derivations of ζ would have been more similar. This will be explored in the discussion. Now comparing the full-cycle results for the two impulses as forcing functions, we note that, for the values for bus axle ζ derived from the APTs:

Page 17: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

16

� the worst-case difference across methods (i.e. between the pipe test

and the VSB 11-style step test method) was approximately ±0.9%; � the differences within the pipe test method methods (i.e. when

comparing the results of the pipe tests) were approximately ±0.3%; and

� the differences within the VSB 11-style step test methods (i.e. when comparing the results of the step tests) were approximately ±0.1%.

Therefore, the bus axle ζ results:

� across methods were similar but with twice the error compared with the experimental results within methods; and

� experimental error being somewhere between ±0.1% and ±1.0 of the error present, APT error would contribute; but

� APT error would likely cancel out in the calculations since it would be present in both A1 and A2 peak values.

Note that the bus manufacturer was unable to respond to requests to supply type-tested ζ values for these axles.

Discussion Characterising a suspension’s response can be performed using an impulse as a forcing function (Chesmond, 1982; Doebelin, 1980). Indeed, an idealised version of this for HV testing purposes would be an impulse falling to zero instantaneously from some arbitrary value. The ideal vertically falling impulse is approximated as an analogue in the step test used as a forcing function in VSB 11. This is by the tyre rolling off a step with the axle falling 80mm as the tyre reaches the ground. In reality, the time taken to fall 80 mm is finite, unlike the idealised impulse response able to be produced by signal generators and used to characterise control systems. Figs 13 and 14 show that this time was approximately 0.4-0.5 s regardless of input impulse method. Doebelin (1980) showed that impulse durations need not be instantaneous but found that, provided that the input impulse duration (iτ ) does not exceed the system time

constant ( dτ ) of the system under test, analysable responses should result. As a

guide, the iτ should range from a pragmatic minimum of 1/16th to no more than

about ¼ of the system time constant (Doebelin, 1980). Further, as Doebelin (1980) pointed out, provided the impulse used as an input signal is short enough, its shape is irrelevant and will suffice as an approximation to an ideal impulse for purposes of characterising a system response. Previous work provided a theoretical study of the duration of the impulse signal from the pipe test (Davis & Sack, 2006). That theory predicted an impulse duration at the axle of 0.44 s for an 11R22.5 tyre and a 50 mm pipe. It went on to show that the combination of:

Page 18: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

17

� 5 km/h approach speed; � over a 50 mm pipe

would provide an impulse of sufficient signal strength and duration to allow the resultant signal at a HV spring to be analysed for body-bounce frequency and ζ value. We see from FIG 14 that the pipe tests provided impulse signal durations of approximately 0.4-0.6 s, validating the theoretical work predicting the period of the impulse function for that speed and an 11R22.5 tyre transiting a 50mm pipe. The 0.4-0.6 s periods were the same order-of-magnitude as those for the VSB 11-style step test. FIG 15 and FIG 16 show that the impulse strength drops off greatly after the first ¼ cycle of the suspension response at the air spring. This elicited a body-bounce waveform from the APTs which was analysable and yielded consistent results, regardless of input method. This duration is probably at the upper limit for a valid excitation time since the positive-going and negative-going excursions will not be proportional for impulse durations greater than this.

Air pressure transducer and accelerometer signals vs. time - bus rear axle step test

2100

2150

2200

2250

2300

2350

2400

2450

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

Time (s)

Acc

eler

om

eter

ou

tpu

t (ar

bitr

ary

linea

r sc

ale)

1550

1600

1650

1700

1750

1800

1850

1900

1950

2000

2050

AP

T o

utp

ut (

arb

itrar

y lin

ear

scal

e)

LHS accel

RHS accel

LHS APT

RHS APT

FIG 15. Comparing the accelerometer signal (impulse function) with the resultant

response from the bus axle for the VSB 11-style step test.

Page 19: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

18

Air pressure transducer and accelerometer signals vs. time - bus rear axle pipe test

2100

2150

2200

2250

2300

2350

2400

3.0

3.1

3.2

3.3

3.4

3.5

3.6

3.7

3.8

3.9

4.0

4.1

4.2

4.3

4.4

4.5

4.6

4.7

4.8

4.9

5.0

5.1

5.2

5.3

5.4

5.5

5.6

5.7

5.8

5.9

6.0

6.1

6.2

6.3

6.4

6.5

6.6

6.7

6.8

6.9

7.0

Time (s)

Acc

eler

om

eter

ou

tpu

t (ar

bitr

ary

linea

r sc

ale)

1650

1700

1750

1800

1850

1900

1950

AP

T o

utp

ut (

arb

itrar

y lin

ear

scal

e)

LHS accel

RHS accel

LHS APT

RHS APT

FIG 16. Comparing the accelerometer signal (impulse function) with the resultant

response from the bus axle for the pipe test. Uffelmann & Walter (1994) and Prem et. al (1998) documented variations in the derived values for ζ, dependent on the excitation method due to, in part, the non-linearity of suspension dampers and their action non-symmetry, depending on direction of travel. A 60% variation was noted in derived ζ values when comparing a “lift and drop” test with what they described as a “bump test” which was our step test (Prem et al., 1998). This was borne out in the results for ζ when we used the ½ -cycle response formula. Response non-symmetry dependant on directionality of the damper action has been well-documented (Sweatman et al., 2000). Accordingly, ½ -cycle measurements likely won’t be valid for determining damping ratios on HV suspensions. Comparing the results in Table 1 across the two test methods indicates that a workable suspension test to find ζ using a lower cost method than the VSB 11 step test is feasible for a worst-case error of about ±1.0%.

Conclusion

The ζ values for the axles tested were derived from APT data by two different test methods to compare an innovative, low-cost testing method with a direct copy of the Australian VSB 11 RFS standard test. The ζ values differed little between test methods and were within ±1.0% (worst case). The results show that, as a “proof-of-concept”, the pipe test produces results at the air springs within an acceptable experimental error range and similar to the ζ value derived from a VSB 11-style step test. Given the potential for error of up to 60%, depending on method (Uffelmann & Walter, 1994), this was quite a positive outcome.

Page 20: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

19

Care needs to be used when choosing the direction of dynamic signal excursion for valid results. Non-symmetry of damper response due to directionality needs to be considered. Either positive-going or negative-going suspension excursion values, but not both, should be used for analysis. The pipe test provides valid and reliable APT outputs provided each axle of the suspension group is allowed to settle before the subsequent axle encounters the pipe. This needs to be balanced against the requirement to have a short enough impulse imparted to the wheels to excite the suspension into a measurable response. Further, the axles with different masses did not yield to classical derivation of ζ values when the APT signals were examined. More HVs are using on-board telemetry systems to measure payload. Greater use could be made of these systems to gather dynamic signals from impulse testing and either verify HV suspension health or schedule damper replacement when required. More research needs to be done on these points and will form part of the work in the QUT/Main Roads project Heavy vehicle suspensions – testing and analysis. This will involve, but not be limited by: � analysis of the signals induced into the chassis at the air springs by the pipe

test and whether this approximates an impulse input to the suspension via the wheels;

� determination of the damped natural frequency of the test vehicles’ body bounce;

� analysis of the signals induced into the chassis by the step-down VSB 11-style test and whether this approximates an impulse input to the suspension via the wheels;

� analysis of the signals induced into the axle by the pipe test to determine the forcing function provided by the pipe test (i.e. the input signal and whether its frequency spectrum and power at the appropriate frequencies is adequate) when compared with the VSB 11-style step-down test; and

� draw conclusions regarding the validity of the “Australian Standard” for determining road-friendliness.

References Australia Department of Transport and Regional Services. (2004a). Certification of

road-friendly suspension systems; Road-friendly suspension certification requirements. Canberra, ACT, Australia: Australia. Department of Transport and Regional Services.

Australia Department of Transport and Regional Services. (2004b). Vehicle Standards Bulletin No.11. Retrieved 6 Sept 2007, from http://www.dotars.gov.au/roads/safety/bulletin/vsb_11.aspx

Australia Department of Transport and Regional Services. (2005a). Bilateral agreement between the Commonwealth of Australia and the State of New South Wales 2004 - 2009. Retrieved 7 Sept, 2007, from http://www.auslink.gov.au/publications/policies/pdf/NSW_Bilateral.pdf

Page 21: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

20

Australia Department of Transport and Regional Services. (2005b). Bilateral agreement between the Commonwealth of Australia and the State of Queensland 2004-05 – 2008-09. Retrieved 7 Sept, 2007, from http://www.auslink.gov.au/publications/policies/pdf/Qld_bilateral.pdf

Cebon, D. (Ed.). (1999). Handbook of vehicle-road interaction. Lisse, South Holland, Netherlands: Swets & Zeitlinger.

Chesmond, C. J. (1982). Control system technology (2nd ed.). Caulfield, Victoria, Australia: Edward Arnold.

Davis, L. (2006). Heavy vehicle suspension testing: on-board mass measurement system accuracy & tamper-vulnerability. Brisbane, Queensland, Australia: Queensland Department of Main Roads.

Davis, L. (2007). Further developments in dynamic testing of heavy vehicle suspensions. Paper presented at the Australasian Transport Research Forum (ATRF), 30th, 2007.

Davis, L., & Bunker, J. (2007). Heavy Vehicle Suspensions – Testing and Analysis. A literature review. Brisbane, Queensland: Queensland Department of Main Roads; Queensland University of Technology.

Davis, L., & Bunker, J. (2008). Suspension testing of 3 heavy vehicles – methodology and preliminary frequency analysis. Brisbane, Queensland: Queensland Department of Main Roads; Queensland University of Technology.

Davis, L., Kel, S., & Sack, R. (2007). Further development of in-service suspension testing for heavy vehicles. Paper presented at the Australasian Transport Research Forum (ATRF), 30th, 2007, ATRF.

Davis, L., & Sack, R. (2004). Analysis of heavy vehicle suspension dynamics using an on-board mass measurement system. Paper presented at the Australasian Transport Research Forum (ATRF), 27th, 2004, Adelaide, South Australia, Australia.

Davis, L., & Sack, R. (2006). Determining heavy vehicle suspension dynamics using an on-board mass measurement system. Paper presented at the ARRB Conference, 22nd, 2006, Canberra, ACT, Australia.

de Pont, J. J. (1999). Suspensions or whole vehicles? Rating road-friendliness. International Journal of vehicle design, 6(1-4), 23.

Doebelin, E. O. (1980). System modelling and response - theoretical and experimental approaches. New York, New York, USA: Wiley.

European Council directive 85/3/EEC as amended by Council directive 92/7/EEC, (1996).

Gyenes, L., Mitchell, C. G. B., & Phillips, S. D. (1992). Dynamic pavement loads and tests of road-friendliness for heavy vehicle suspensions. In Cebon & Mitchell (Eds.), Heavy vehicles and roads: technology, safety and policy (pp. 243-251). London, United Kingdom: Thomas Telford.

Houpis, C. H., & Lamont, G. B. (1985). Digital control systems theory, hardware, software. New York, New York, USA: McGraw-Hill.

Karl, C. (2007). Project Plan: Technical Feasibility Assessment of On-Board Mass-Monitoring (OBM) Devices. Melbourne: Transport Certification Australia Major Projects Division.

Meriam, J. L., & Kraige, L. G. (1993). Engineering mechanics Vol 2 - Dynamics. New York, New York, USA: Wiley.

National Road Transport Commission. (1993). Regulatory impact statement: mass and loading regulations (Regulatory impact statement). Melbourne, Victoria, Australia: National Road Transport Commission (NRTC).

Page 22: QUT Digital Repository:  ·  Davis, Lloyd E. and Bunker, Jonathan M. (2008) In-service suspension testing of ... Doebelin, p79 (1980) said,

21

OECD. (1998). Dynamic interaction between vehicles and infrastructure experiment (DIVINE). (Technical report No. DSTI/DOT/RTR/IR6(98)1/FINAL). Paris, France: Organisation for Economic Co-operation and Development (OECD).

Peters, I. (2003). In-service testing of road friendly suspensions. University of Western Australia. School of Mechanical, Materials and Mechatronics Engineering, Crawley, Western Australia, Australia.

Potter, T. E. C., Cebon, D., & Cole, D. (1997). Assessing “road friendliness”: a review. Journal Automobile Engineering I Mech E, 211 C(6), 20.

Prem, H., George, R., & McLean, J. (1998). Methods for evaluating the dynamic-wheel-loading performance of heavy commercial vehicle suspensions. Paper presented at the International Symposium on Heavy Vehicle Weights and Dimensions, 5th, 1998, Maroochydore, Queensland, Australia.

Prem, H., Ramsay, E., McLean, J., Pearson, R., Woodrooffe, J., & de Pont, J. J. (2001). Definition of potential performance measures and initial standards - performance based standards. (Discussion paper). Melbourne, Victoria, Australia: National Road Transport Commission/Austroads.

Starrs Pty Ltd, M. M., Ian Wright and Associates, & ARRB Transport Research Ltd. (2000). Evaluation of in-service compliance of road friendly suspensions (Report No. 0642544670). Melbourne, Victoria, Australia: National Road Transport Commission (NRTC).

Sweatman, P. F., McFarlane, S., Komadina, J., & Cebon, D. (2000). In-service assessment of road-friendly suspensions: for information (Report No. 0642544522). Melbourne, Victoria, Australia: National Road Transport Commission (NRTC).

Technical Committee ISO/TC 22. (2000). Road vehicles - heavy commercial vehicle combinations and articulated buses - lateral stability test methods. Geneva, Switzerland: ISO.

Thomson, W. T., & Dahleh, M. D. (1998). Theory of vibration with applications. Upper Saddle River, New Jersey: Prentice Hall.

Transport Certification Australia Limited. (2007). Heavy Vehicle On-Board Mass Monitoring: Capability Review (Report). Melbourne, Victoria, Australia.

Uffelmann, F., & Walter, W. D. (1994). Protecting roads by reducing the dynamic wheel loads of trucks. Paper presented at the 1994 European ADAMS conference ADAMS user conference, Frankfurt, Germany.

Woodroofe, J. H. F. (1995). Evaluating suspension road-friendliness. Paper presented at the Dynamic loading of heavy vehicles and road wear, mid-term seminar, 1995, Sydney, New South Wales, Australia.

Woodroofe, J. H. F., LeBlanc, P. A., & Papagiannakis, A. T. (1988). Suspension dynamics - experimental findings and regulatory implications. In Society of Automotive Engineers (SAE) (Ed.), Special publication 765. Warrendale, Pennsylvania, USA: Society of Automotive Engineers (SAE).