You are on page 1of 15

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/301677668

EXPERIMENTAL STUDY OF A RESILIENT MOUNTING FOR MARINE DIESEL


ENGINES

Conference Paper · June 2012

CITATIONS READS

6 640

2 authors, including:

Lorenzo Moro
Memorial University of Newfoundland
31 PUBLICATIONS   71 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Container Stack Dynamics (CSD) View project

Transient torsional vibration analysis of propulsion shafting systems of ice-class ships View project

All content following this page was uploaded by Lorenzo Moro on 28 April 2016.

The user has requested enhancement of the downloaded file.


EXPERIMENTAL STUDY OF A RESILIENT MOUNTING
FOR MARINE DIESEL ENGINES

M. Biot1 and L. Moro1

ABSTRACT
In the paper, analysis procedures and design concepts for the selection of resilient mountings for
marine diesel engines are reviewed. After a description of the theoretical basis for studying resilient
mounting systems, an overview is given of approaches for their dynamic characterization by means of
laboratory tests, which are performed by implementing procedures given in the ISO 10846 standard.
The facilities of the Ship Noise and Vibration Laboratory (NVL) of the University of Trieste are then
presented. A comparative study carried out on two different configurations of a resilient element
designed by Vulkan Italia Srl to be used as vibration isolator for medium-speed marine diesel engines
is finally discussed.

KEY WORDS
Ship; Vibrations; Structure-borne noise; Resilient mounting; Laboratory test; Transmissibility; Dynamic transfer
stiffness.

INTRODUCTION
Design of comfort on board ships is mainly related to optimization of devices (i.e., the resilient mountings) able to reduce
vibrations transmitted by machinery to ship’s structure. Lack in vibration barriers may lead to high levels of vibrations
both at low and high frequencies, which are cause of different effects on comfort perceived by people on accommodation
decks. As for the high-frequency vibrations, they induce structural noise which in turn appears in cabins as noise
generated, for instance, by vibrating panels and ceilings (Biot, et al 2007). Such an issue is of particular interest on board
passenger vessels.

Medium-speed diesel engines, employed as gensets or propulsion engines, are among the main structure-borne noise
sources on board ships. Medium-speed diesel engines are the most popular engines on board cruise ships, where, due to
the high demand of comfort, are the objects of designers’ attention being they one of the primary sources of noise and
vibrations. Today, the most effective way to reduce the structure-borne noise on board ships is to decouple engine from
foundation by means of vibration isolators.

The decoupling between engine and ship structures produces both a significant reduction in the vibrational power
transmitted to the hull and a reflection of some vibrational energy back to the engine (Hynnä 2002). The latter must be
controlled in order to prevent excessive levels of vibrations to cause engine damages. In addition to engine decoupling
towards foundation, decoupling systems are also applied to connections between engine and engine devices and
auxiliaries, in order to avoid other propagation paths of structure-borne sound. Such decoupling systems serve also to the
purpose of preserving from engine high vibrational levels leading to possible breakage all that components which must
be rigidly constrained to the hull structure, as in the propulsion plants where a gearbox is joined to a medium-speed diesel
engine acting as prime mover. Within this frame, the optimization process of the vibration isolators for medium-speed
diesel engines is a topic of paramount importance.

Several types of resilient mountings are available in the market. They vary in shape, usually they have cylindrical,
conical or parallelepiped shape, in the connection interface to both engine feet and foundation, in the efficiency along the
different directions, and in the way they reduce the vibrational energy transmission by using rubber of different Shore
hardness or any presence of metal inserts.

Dynamic characterization of resilient mountings is performed to deepen knowledge of their isolation abilities. Special
studies are also performed in order to fit a resilient element for a specific engine and foundation structure. In general,

1
Department of Engineering and Architecture, University of Trieste, Italy
dynamic behaviour of resilient elements, even if in general easily predictable over the most part of the working frequency
range, is highly non-linear at certain frequencies. No analytical methods are today available to predict non-linear
behaviour of resilient mountings, so special laboratory tests need to be carried out, with the aim to identify isolators’
transmissibility over a suitable frequency range (Thompson, et al 1998).

SELECTION PROCESS FOR AN OPTIMAL RESILIENT MOUNTING


In the following, criteria used for studying resilient mounting systems are discussed, giving evidence to procedures and
practical methods to deal with their design at low frequency or audio-frequency vibrations. Main emphasis is given to the
concepts at the basis of the different approaches, with the aim to explain the role of the experimental study in the resilient
mounting system design.

Basic Approach
Engine designers may refer to not many manufacturers all around the world who produces resilient mountings of
sufficient size to support a marine engine. So, selection process of the resilient mounting is restricted to few possible
solutions, which means that one needs to take into account of the market availability just in the early design stage.

The resilient mounting design for a medium-speed marine diesel engine is carried out by different steps (Tao, et al 2000).
First, an initial sizing needs to be carried out from a static point of view. For this purpose, other than the engine weight,
the number of resilient elements to mount needs to be considered; the sizing should be performed also considering the
maximum pitch and roll angle of the ship. In this design phase, the maximum load to which each resilient mounting can
be subject and its static deflection have both to be taken into account. Data on static deflection versus applied load are
provided by manufacturers by different charts for any Shore hardness of the resilient element rubber.

Then, in order to ensure the proper functioning of the engine at different speeds, the rigid-body resonance frequency for
each degree of freedom has to be calculated. In effect, an engine supported by a resilient mounting system can be
considered as a rigid-body with six degrees of freedom. Once rigid-body resonance frequencies have been determined,
designer proceeds with checking that the free forces of the engine, which are considered applied to its mass centre of
gravity, do not excite engine to oscillate in its rigid-body modes. That should happen, the vibration level could be
dangerous for the proper functioning of the engine. This analysis is usually performed in a frequency range that generally
varies from about 4 Hz up to about 20 Hz.

Another fundamental aspect in the elastic coupling between engine and foundation, is the detailed design of the latter. As
a matter of fact, foundation should be designed not only to support the engine weight, but also to have a low mobility
level. Such an analysis is not commonly performed and a basic approach disregarding foundation mobility is applied. In
some cases, when a basic approach has been used, on site measurements are then performed to verify that foundation
doesn’t show high mobility levels at frequencies corresponding to those of the engine free forces, which would leads to
high structure-borne noise levels.

Controlling Structure-Borne Noise


Today, there are still no reliable methods for predicting the structure-borne noise due to marine diesel engines. As above
mentioned, the design and selection of a resilient mounting is carried out only by a static and dynamic point of view, the
latter analysis performed in a very low frequency range.

In order to make a complete analysis of the diesel engine structure-borne noise, the vibrational energy transmitted by an
engine in all directions and through the whole system of resilient mountings should be considered along with the mobility
of foundation at any given contact point between the engine and the single resilient mounting. That kind of analysis,
already quite complex, is even more complicated by the fact that the diesel engine resilient mountings are characterized
by a dynamic stiffness that changes with frequency as well as with direction. Furthermore, isolators’ behaviour deviates
much from the behaviour of an ideal spring in the high frequency range.

A simplified approach is based on the assumption of treating the problem as a one-dimensional problem in which the
structure-borne noise prediction is carried out by considering independent from each other all the transmission paths
through the resilient mountings. By means of this approach, resilient mountings interactions are neglected. A further
assumption is made of considering any resilient mounting to develop only the vertical translational velocity V and the
vertical force F, so neglecting the presence of moments and angular velocities. Under these assumptions, the system
composed by engine, isolators and foundation can be treated as a series of 2-DOF lumped parameter systems. Figure 1
shows a 2-DOF lumped model system, where parameters characterizing the system are given: the mass m of source and
receiver, the static stiffness k0 of the spring-like isolator, the input force F0 acting on the source, and forces and velocities
mutually exerted by the bodies.
The velocity level on the foundation is given by (Hynnä 2002):

LV , r ( f ) = LV , s ( f ) + LZ ,i ( f ) + LM , r ( f ) [1]

where the foundation velocity level LV,r (r for receiver) is a function of the source velocity level LV,s (s for source), of the
resilient mounting mechanical impedance level LZ,i and of the foundation mobility level LM,r. So, to be able to predict the
foundation velocity level, the three terms on the right hand of the equation have to be known.

Figure 1: The physical model for a single-point-connected system

The velocity level at the engine foot LV,s is known from measurements taken in the test room. It is a function of the
engine speed, of the resilient mounting impedance and of the foundation mobility. However, it can be shown that
changing the foundation and isolator dynamic characteristics, the velocity level measured at the engine foot varies of no
more than about ±3 dB within the audio-frequency range; hence, it can be considered independent from resilient
mounting and foundation dynamic characteristics. Anyway, due to the high number of variables which determine the
engine foot vibration level, its prediction by means of analytical or numerical methods is a very complex issue. In fact,
two kinds of dynamic forces have to be considered: on the one hand, there are the free forces, which are generally
restricted to a low frequency range, and, on the other hand, there is a system of forces acting at higher frequencies, due to
the cylinders’ pressure cycle, the crankshaft inertia and the impacts of the moving parts where there is a relative force
inversion, such as the bearings and pistons slaps.

With respect to the determination of the resilient mounting impedance LZ,i , it has to be emphasized that it cannot be
determined resorting to traditional FE methods. Non-linear FE analyses have been also performed for the calculation of
the resilient mounting transfer properties, but results so obtained generally lack in accuracy, predicting a resilient
mounting behaviour which does not agree with the actual one (Besio, et al 1997). So, the resilient mounting vibro-
acoustic transfer properties can be reliably determined only by means of measurements performed on laboratory test rigs.

As for the mobility level of the foundation LM,r , it can be determined by analysing the dynamics of the structure by FE
linear methods or by direct measurements.

Therefore, in order to reduce the structure-borne noise levels transferred by an engine foundation to surrounding
structure, actions on the individual terms of the right-hand side of Equation 1 are required. Lower engine vibration levels
of course reduce, at the source, the structure-borne levels transmitted to the structure. The same goal may be obtained by
lowering the mechanical impedance levels of the resilient mountings. Greater mobility levels of an isolator, obtained by
varying the isolator shape or the rubber stiffness, increase its ability to reduce the vibrational energy transmitted to
foundation. As already said, the latter should show mobility levels as low as possible, so as to have a very high contact
point stiffness between foundation and resilient mounting. That could be achieved by an optimization of the thickness of
the foundation plate below the mounting, of the frame or of the bedplate – and clearly, an optimization of the foundation
may be also done by improving the foundation structure shape (Ran Lin, et al 2009).

As previously mentioned, in order to determine the vibro-acoustic resilient mounting transfer properties, the only
possible way is the experimental one. By resorting to experimental tests carried out on a laboratory test rig, the resilient
mounting transfer function can be expressed as a dynamic inertia, and defined as the ratio between the force F2b(f)
measured on the blocked side (side 2) − that is the blocking force at the isolator output side that implies a null
displacement at the same side, and the complex acceleration a1(f) measured on the driven side (side 1). Once the resilient
mounting dynamic inertia curve has been achieved by measurements, the isolator impedance level is then given by:

F2 ( f ) F (f)
LZ ,i ( f ) = = 2π f 2 » 2π f T2b,1 ( f ) [2]
V1 ( f ) a1 ( f )

where the dynamic inertia is usually referred to as a dynamic transmissibility, and so denoted as T2b,1(f), in consideration
that it is an output to input force ratio where the input force is referred to a unit mass.

A typical transmissibility curve of a marine diesel engine resilient mounting is shown in Figure 2, where the curve is
represented in narrow band on a log-log scale. The solid line refers to measurements, carried out in a typical frequency
range from about 100 Hz to about 1.3 kHz with the resilient mounting properly preloaded, while the dashed lines refer to
extrapolated values. The transmissibility curve is divided into three different parts. The first part of the curve shows a
linear reduction of transmissibility, that gives evidence of an ideal spring-like behaviour of the resilient mounting. In the
second part, transmissibility deviates from ideal behaviour and shows two successive peaks where transmissibility
increases. At last, in a third part, it resumes its original trend. Such a behaviour is due to the presence of the resilient
rubber, which is the cause of two different phenomena: on the one hand, stationary waves propagate in the rubber due to
the natural vibration modes; on the other hand, the mass effect due to the rubber incompressibility implies a coupled
transversal movement (Besio, et al 1997). From Equation 2, it is clear that such an increase in transmissibility
corresponds to an increase of the resilient mounting impedance level and, therefore, to an increase in structure-borne
noise measured on the engine foundation.

Figure 2: A typical transmissibility curve of marine diesel engine resilient mountings

The method just described is known as the single-point-connected system method. Theoretically, this formulation is valid
for structure-borne noise sources that are coupled to the foundation by a single resilient mounting. A marine engine,
however, is always coupled to the foundation through various isolators (Figure 3).

Figure 3: A view of an actual multi-point-connected system


Fulford and Gibbs (1997) have extended the studies proposed by Mondot and Petersson (1987) to characterize a
structural noise source in the case of a multi-point-connected system. In general, diesel engine motion has 6 DOF, three
translational and three rotational, and it is coupled to the foundation at many points.

In the case of a 6 DOF system, using the source mobility matrix [Ys], the receiver mobility matrix [Yr] and the force ratio
vector { Fjm / Fi n}, where Fhk is a force acting on the foundation at point k in direction h, the transmitted power is given
by (Fulford, et al 1997):

(V ) n 2

Qin =
sfi
{ {
Re [Yr ] F jm Fi n }} [3]
{ F } + [Y ]{F }
2
2 [Ys ] F jm i
n
r
m
j Fi n

where Vsfin is the r.m.s. free source velocity measured at any point n (of a total of N points) in any direction i. Use of
Equation 3 in the evaluation of the transmitted power implies very difficult computations. The matrices are in fact large,
6N x 6N for both the source and the receiver. Moreover, by using such an approach, it is not straightforward to correlate
high power levels with relevant causes: the most important structure-borne sound transmission paths as well as the
relative significance of source and receiver mobility, are both unknown. Furthermore, knowledge of the force ratios
acting in the interface is required.

In order to help address this complex problem, the effective mobility is introduced as (Petersson, et al 1991):

N
Fim 6 F jn N 6 F jm
YiinnS = Yiinn + å Yiinm
Fin
+ å Yijnn
Fin
+ å å Yijnm
Fin
[4]
m=1, m¹ n j =1, j ¹i m=1, m ¹ n j =1, j ¹i

where the mobility Yijnm is a mobility where the force is acting at the point m and direction j to the point n and direction i.
The first term in the right hand side of the equation is the direct structure-borne sound transmission path, the second term
is the contribution of the other points in the direction n , the third term is the contribution of the other direction at the
point being considered and the fourth term is the contribution from other points in directions other than that being
considered.

If only a one-direction translational motion is considered, Equation 4 simplifies and, under the assumption that forces
have unity magnitude and zero phase, the effective mobility comes out to be (Petersson, et al 1982):

N
YiinnS = Yiinn + å Yiinm [5]
m =1,m ¹n

Formulation is very simplified compared to that of Equation 4, and therefore easier to apply in practice, even if the
contributions of the other points are still taken into account.

Although the multi-point-connected system approach allows evaluating more precisely the mobility level of the
foundation and therefore allows a more accurate assessment of the structural noise level, in the case of a marine engine it
is still not widespread, and the single-point-connected system approach is more used, because of its easiness and direct
application.

Fundamentals of Damper Characterization


As above discussed, characterization of resilient mountings is attained by means of measurements performed on
laboratory test rigs. The vibro-acoustic transfer properties of isolators by means of direct experimental tests are carried
out according to the ISO 10846 standards, which are widely accepted as a sound reference. In the same document, a
scheme is proposed for the simplified analysis of a suspended vibrating source. According to that approach (ISO 2008a),
the dynamic parameters of a resilient mounting system are defined by making reference to a three block system: the
vibration source (i.e., the diesel engine), the N isolators and the receiving structure (i.e., the engine foundation).
Application of the procedure is subject to the main assumptions, that only passive linear-response resilient elements may
be studied. It also supposes that the contact between the three blocks is a contact point and that by each contact point a 6
component force vector and a 6 component displacement vector are assigned (Figure 4).

According to ISO 10846-1:2008, to describe the vibro-acoustic characteristics of a resilient mounting the most significant
parameter is the dynamic transfer stiffness k2,1(f).
Figure 4: Schematic presentation of source-isolator-receiver system

The dynamic transfer stiffness is defined as the ratio between the dynamic force on the blocked output side of an isolator
F2b(f) and the complex displacement on the driven side u1(f). To prove the statement about the significance of the
dynamic transfer stiffness, the simple case of a unidirectional transmission of vibrations through an isolator is considered
in the following. If k1,1(f) and k2,2(f) are the driven-point stiffnesses when the resilient element is blocked at the opposite
side (u2 and u1 are null, respectively), the equilibrium equations of the isolator can be written for each frequency as:

ìï F1 = k1,1u1 + k1,2u2
í [6]
ïî F2 = k2,1u1 + k2,2u2

which, written in matrix form, becomes:

ì F1 ü é k1,1 k1,2 ù ì u1 ü
{F } = [ K ]{u} where {F } = í ý [K ] = ê ú {u} = í ý [7]
î F2 þ ë k2,1 k2,2 û î u2 þ

In the three block system, isolator excites the receiving structure with the force F2 which may be derived by the definition
of the driven-point stiffness of the receiving structure, that is kr = -F2 / u2. So, F2 may be expressed as:

k 2,1
F2 = u1 [8]
k
1 + 2,2
kr

From a practical point of view, if |k2,2| < 0.1 |kr| it comes out that F2 ≈ F2b = k2,1 u1, which is a very significant
relationship.

In general, displacements and forces are characterized by 6 orthogonal components, which, for the i-th connection point
(i = 1, 2) may be written as {ui } = {uix, uiy, uiz, γix, γiy, γiz}T and {Fi } = {Fix, Fiy, Fiz, Mix, Miy, Miz}T. Consequently,
Equation 7 can be expressed by:

ì {F1}ü é éë K1,1 ùû éë K1,2 ùû ù ì{u1}ü


í ý=ê ú í ý [9]
î{F2 }þ êë éë K 2,1 ùû éë K 2,2 ùû úû î{u2 }þ

where any [Kij] matrix is 6 x 6 and the global dynamic stiffness matrix is 12 x 12. So, to completely describe a resilient
mounting system of N elements, 144 x N transfer functions are needed. Owing of symmetry, many elements of the
stiffness matrix are equal to zero and some non-zero elements are equal in magnitude. In practical cases, few diagonal
elements are enough to describe the isolator behaviour in the translational directions (ISO, 2008a).

LABORATORY MEASUREMENT OF DYNAMIC TRANSFER STIFFNESS


As above shown, complete characterization of a resilient mounting may be achieved by knowledge of the dynamic
transfer stiffness k2,1 which may be correlated to the transmissibility by relationship T2b,1(f) = k2,1(f)/(2πf)2 and to the
dynamic impedance LZ,i (f) by Equation 2. Even if, by a theoretical point of view, dynamic transfer stiffness is the
reference parameter, usually dynamic behaviour of resilient mountings is discussed in terms of transmissibility.
In order to measure the vibro-acoustic transfer properties of resilient mountings, laboratory measurements are needed and
the most efficient method to evaluate such properties in the audio-frequency range is the so-called indirect method
discussed in the ISO 10846 standards (ISO 2008a; ISO 2002).

Experimental Approach: Principles and Standards


Depending on the frequency range to be investigated, three different methods are proposed in the ISO 10846 standards
“Acoustics and vibration − Laboratory measurement of vibro-acoustic transfer properties of resilient elements”, divided
in 5 parts.

To characterize resilient mountings at low frequencies the so-called direct method is the most appropriate (ISO 2008a;
ISO 2008b). This method assumes that, in the test rig, the resilient element is placed between a vibration exciter and a
rigid foundation. Between the isolator and the rigid foundation is placed a load cell for force dynamic measurement and a
sensor between the exciter and the isolator is placed to measure displacement, velocity or acceleration. Basing of the
measured data, the dynamic transfer stiffness is given by:

F2 F
k2,1 ( f ) = = -(2πf )2 2 [10]
u1 a1

The method is applicable in a limited frequency range from 1 Hz to an upper frequency limit that can vary between 300
Hz and 500 Hz. This range limitation is due, on the one hand, to the vibration actuator bandwidth, and, on the other hand,
to the first mode shape of the test rig chassis which, in the test rig used for testing large resilient mountings, such as those
for marine engines, is of about 300 Hz. The dynamic transfer stiffness of a resilient mounting in the low-frequency range
can be also obtained by the so-called driving point method (ISO 2008c). The test rig configuration is similar to that used
in the direct method: the resilient element is subject to the test static preload and is placed between the exciter and a very
stiff foundation. In this case, however, both dynamic load cell and accelerometers are placed between the exciter and the
input side of the isolator and therefore the measured parameters are F1 and a1. Assuming that the inertial forces are small
in comparison to the elastic forces, the dynamic transfer stiffness k2,1 becomes approximately equal to the driving point
stiffness k1,1:

F1 F1
k2,1 » k1,1 = = -(2πf ) 2 [11]
u1 a1
u2 = 0 a2 = 0

that is valid, in the case of very large resilient mountings, up to about 100 Hz. In fact, beyond this frequency the inertial
forces are generally no longer negligible compared to the elastic forces and the method is no longer valid.

In the audio-frequency range, the dynamic transfer stiffness of resilient elements is determined using the so-called
indirect method (ISO 2002). Unlike the direct method and the driving point method, in the indirect method the blocking
force is not directly measured and is derived from acceleration measurements performed on the blocking mass which is
dynamically decoupled from the test rig chassis. The resilient mounting is not directly coupled with the vibration source,
as a compact mass, called excitation mass, is interposed. The excitation mass function is to provide the condition of
contact point at the input side of the resilient mounting and, as the blocking mass, it is dynamically decoupled from the
test rig structure using auxiliary isolators. Under the resilient mounting, the so-called blocking mass is placed. It provides
a high-stiffness contact point at the isolator output side, so that the forces between the isolator output side and the
receiving mass are approximately equal to the blocking forces. The blocking mass must have a high inertia, both
translational and rotational, whereas its decoupling isolators should have a suitable low stiffness so as to keep low the
resonance frequencies of the 6 rigid motions of the mass. An actuator is used for applying a static preload, so that the
resilient mounting is tested in working condition.

In the indirect method test, the acceleration (or displacement or velocity) of the excitation mass and the acceleration (or
displacement or velocity) of the blocking mass are measured and the dynamic transfer stiffness is derived as:

F2 u a
= - ( 2πf ) m2 2 = - ( 2πf ) m2 2
2 2
k2,1 » [12]
u1 u1 a1

where the F2 force has been derived making use of the Newton’s law.

The approach is still valid if tests are performed in a frequency range which is not affected, at the low frequencies, by the
vertical motion at resonance frequency f0 of the blocking mass constrained between a soft isolator bed and the resilient
element, and, at the high frequencies, by the limit of the rigid-body like behaviour of the blocking mass at f3 frequency.
An approximate formulation for the frequency of the lowest internal resonance of the resilient mounting is the following:

k0
f e = 0.5 [13]
mel

where k0 is the low-frequency dynamic stiffness of the resilient mounting and mel is the mass of its elastic part.

According to the ISO 10846 standards, measurements are valid for frequencies higher than fe /3, but just if the resonant
frequency f0 is lower than fe /10. As for the upper limit, the ISO 10846-3 standard provides the frequency values f3 for
steel blocking masses of cylindrical or cubic shape. Upper frequency limit may be also determined by means of
experimental tests and referring to the concept of effective mass. When the effective mass m2,eff (f) has been determined,
the upper frequency limit f3 has to be chosen as the lower frequency value where the effective mass does not vary more
than 12 % from the mass value m2.

Test Facility at the NVL Laboratory


Research activity on elastomeric resilient mounting systems has been developed at the Ship Noise and Vibration
Laboratory (NVL) of the University of Trieste since the last five years, when a research program on structural-borne
noise caused by the main engines on board cruise ships was funded. Within that project, a test rig was designed and built
(Figure 5), and it is now working for the characterization of passive resilient mountings as far as both their vertical and
transverse transmissibility is concerned – the latter feature being under development. The parts are: (A) crossheads for
shaker support placed on the top of the (B) heavy and rigid main frame of the test rig, (C) trunnion hanged by isolators to
the crossheads, (D) electrodynamic vibration generator, (E) mobile crosshead and static load upper support plate, (F)
moving system between upper and lower isolator beds, (G) static load lower support discs and (H) static force
transducers and cylinder for hydraulic preload.

Figure 5: Test rig facility at the NVL

Characteristics of the test rig fulfil the guidelines given in the ISO-10846 standards for laboratory measurement of the
resilient element transmissibility. Specifically, experimental investigations may be carried out by applying the indirect
method for tests in the audio-frequency range, and the direct method may be also implemented in case of low frequency
tests.

The core of the test rig are the vibrating bodies (Figure 6), whose assembling is the so called moving system. The moving
system, being tightened by a hydraulic piston between an upper and a lower elastic bed of auxiliary isolators, may
vertically move under the action of an electrodynamic shaker. The moving system is made by the resilient element
supported by a blocking mass and holding on the top the excitation mass, by which the whole system is connected to the
elastically suspended shaker. In Figure 6 a moving system is shown, arranged for testing on steel springs.
Figure 6: Moving system of the test rig

Connection between moving system and shaker (Figure 7) is a very significant part and is made by a “stinger” rod.
Through the stinger, vibrational energy is transmitted to the excitation mass. Mechanical connection among such
elements has been designed to avoid local resonance frequencies in order to preserve measurements from such noise
sources. In Figure 7 a connection chain is shown, arranged for testing on low-stiffness elements.

Figure 7: Connection between the moving system and the shaker

Length and diameter of the stinger rod, inertia of the excitation and blocking mass and stiffness of the two soft isolator
beds (i.e., the auxiliary isolators) are all set according to the test type and the resilient element characteristics.

Laboratory tests can be developed on small to big resilient elements as test controlling parameters may be adjusted in a
wide range of values. The most distinctive feature of the test rig lies in its capacity of performing tests on the biggest
resilient mountings today used in the suspension systems of marine engines, with a static load capacity up to 150 kN and
a maximum dynamic load capacity of 4 kN up to a frequency of 2 kHz, which range largely covers the typical frequency
range in the investigations on medium-speed marine diesel engines. Dimensions of test rig are remarkable too, being
supporting frame for moving masses very large, so allowing tests to be carried out on resilient mountings 0,5 metres wide
and weighting about 1 kN placed on a foundation of up to 8 kN weight.

OPTIMIZATION OF A RESILIENT MOUNTING


To define the dynamic transfer properties of two different configurations of a conical-shaped resilient mounting, designed
by Vulkan Italia Srl for a medium-speed marine diesel engine of a rated power of about 16000 kW, a series of
experimental tests have been carried out at the Ship Noise and Vibration Laboratory.
In Figure 8 the resilient mounting is represented. Picture refers to a FE model used for numerical analyses on dynamic
characteristics of the resilient mounting parts. Such resilient mounting is made of three main parts: the resilient mounting
top, the rubber part, and the base. The resilient mounting top holds a flange for the coupling with the engine foot, which
flange allows the adjustment of resilient mounting in height, by means of a ring nut, for a better engine alignment. The
rubber part is a conical-shaped toroid body which is coupled by a conical-shaped seat with the resilient mounting base.
The base is bolted to the diesel engine foundation on the ship inner bottom.

Figure 8: Picture of the resilient mounting tested

The two tested configurations, the basic and the improved one, have in common the top and the base, while differ from
each other in the rubber element: a conical-shaped toroid body that, in the case of the improved configuration, contains a
steel thin sheet which is inserted at the middle-high of the rubber.

Experimental tests have been performed in a frequency range from about 100 Hz to 1000 Hz, as the interest of the engine
designer was to check if the improved configuration could led to a reduction in transmissibility just in this range. The two
resilient mountings have been both tested at their working load of 75 kN. A test campaign has been carried out on the
basic configuration to verify linearity in the dynamic response of resilient element under different static loads. Linearity
in dynamic response has been also verified by varying the levels of the applied dynamic loads. In all tests, shaker was
controlled in order to produce a flat acceleration autospectrum on the top of the excitation mass.

Special attention has been paid to the correct functioning of each element of the moving system of the test rig. A
comprehensive measurement of movement and deflection has been carried out on all the masses, and in particular on the
excitation mass and on its connection to the shaker by the stinger rod. Such an analysis, performed by the Operational
Deflection Shape (ODS) technique, shows that all the parts move perfectly in the direction of the driving force and that
they behave like rigid bodies within the investigation frequency range.

Setting of the Test Rig


Before to start with the experimental activity, a comprehensive prediction of the moving system behaviour has been made
in order to set a series of parameters according to the weight and static stiffness of the tested resilient mounting.

First, the more appropriate blocking mass in weight and shape has been chosen in order to properly perform the tests in
the required frequency range. The blocking mass has a cylindrical shape and is made by assembling a number of disks
that are bolted with each other. The experimental measurement of the effective mass has been carried out in order to
ensure that its deviation with respect to the inertial mass does not exceed 12%. Then, the choice of an excitation mass
that assures a contact point excitation has been made, so as to produce a uniform distribution of shaker excitation force
on the top of the resilient mounting. The experimental measurements of the effective mass have been also performed to
verify the frequency limit f3 of rigid-body behaviour of the excitation mass.

A preliminary analysis has been also performed on the blocking mass to control that its rigid body vertical resonance
frequency f0 satisfies the standard requirements. Background noise levels have been also checked. As required in the ISO
10846-3 standard, acceleration levels measured in vertical direction at the blocking mass when the vibration source is
turned on, must be 15 dB [ref. 10-6 mm/s2] higher than the corresponding levels measured while the vibration source is
turned off. Another check has been carried out to control the transverse acceleration levels (i.e. horizontal accelerations).
The acceleration levels measured at the excitation mass in orthogonal direction to the driving direction must be 15 dB
[ref. 10-6 mm/s2] lower than those measured in the driving direction. If that is not verified, the unwanted accelerations of
the excitation mass could cause on the blocking mass accelerations which are a linear combination of the excitation mass
transverse and vertical accelerations. Unwanted horizontal accelerations have been measured on the excitation mass
according to procedure given in the ISO standard (Figure 9).
Figure 9: Accelerometers layout for transverse measurements

Unwanted horizontal accelerations have been then compared with the vertical acceleration in Figure 10, where the flat
autospectrum of the latter is due to the procedure applied in the test, that is a random shaker control and a white-noise
excitation. Measured values comply with the standard limit values.

Figure 10: Measured acceleration autospectra on the excitation mass

Finally, the stinger rod length and diameter have been selected according to the dynamic characteristics of the moving
system. Environmental conditions were kept the same during all the test campaign. All test outcomes have been collected
in terms of accelerations acquired by piezoelectric accelerometers controlled by a dedicated data acquisition system,
while the shaker control has been made by means of separate measuring chains.

Discussion of the Experimental Results


A series of measurements has been carried out to investigate the linearity of the tested resilient mountings to static load.
Experiments have been executed on the basic configuration under the assumption that the improved one shows similar
behaviour. Static preloads have been set according to manufacturer’s suggestions. Three different static loads have been
applied (60 kN, 70 kN and 80 kN), characterizing a range of usual design working loads.

Figure 11 shows that the relevant transmissibility curves have the same trend and do not diverge significantly from each
other. Increment in dynamic stiffness due to increment in static load is clearly appreciable both in a higher
transmissibility at low frequencies and in higher resonance frequencies. Results are expressed in terms of dynamic
transmissibility levels LT, calculated as 20 log (T2b,1) = 20log (F2b/a1) and expressed in dB [ref. 1 kg] in consideration that
T2b,1 is an output to input force ratio where the input force is referred to a unit mass.
Figure 11: Measured transmissibility curves of the basic resilient mounting at different static loads

Comparison of the two resilient mounting configurations has been carried out at a static preload equal to 75 kN,
considered as an average design working load for that kind of resilient mountings. All other settings of the test rig were
the same for the two sets of tests. Figure 12 shows the transmissibility levels measured on the two resilient mounting
configurations.

Compared to that of the basic configuration, the transmissibility curve of the improved one shows a little bit higher
dynamic stiffness at the lower frequencies and an ideal behaviour in a shorter range, the latter due to the shifting of the
resonance peaks towards the lower frequencies. As the behaviour at the lower frequencies is about the same for the two
configurations, no differences are expected in the performance of the two configurations up to about 200 Hz. Anyway,
the first peak of the new resilient mounting curve may lead to critical situations in cases when resilient mounting will be
used on diesel engines showing high dynamic forces at about 300 Hz or in combination with a foundation having a high
mobility at the same frequencies, which fact is not so rare. On the other hand, behaviour of new solution is certainly
enhanced at the higher frequencies, where transmissibility shows a very different trend. Here, the resonance peak is very
low and measured levels are so low that they melt in the background noise.

Figure 12: Measured transmissibility curves of the basic and improved resilient mountings

Summing up considerations made upon the comparison of the two resilient mounting solutions, it may be concluded that
the improved one may have better impact when structure-borne noise is the main target of the resilient mounting design,
as its dynamic characteristics are enhanced in the audio-frequency range. On the other hand, as the improved
configuration amplifies transmitted vibrations in a range of typical high mobility of receiving structures, it may cause
possible critical situations. Anyway, such a localized lack of resilient mounting in controlling vibrations transmitted to
the ship structures should be avoided by a proper design of the foundation structure.

CONCLUSIONS
Behaviour of passive resilient mountings changes according to the frequency of the exciting force. In fact, the rubber part
of the resilient element does not show an ideal trend of the dynamic stiffness all over the extent of the engine force
spectrum. So, in a range of frequencies where typically engine transmits to support vibrations with high energy content,
resilient element loses effect in vibrations isolation due to a series of resonances developing inside the rubber part.
Resonances are the reason of a high vibration level below the resilient element, which in turn causes transmission to ship
structures of unwanted effects. As that takes place in the range of the audio frequencies, structure-borne sound propagates
along the light structures of the ship from the engine room to the accommodation decks causing there discomfort on
passengers and crew.

It is a common experience that the dynamic transmissibility uniformly decreases from the low frequencies to the first
resonance of the rubber part, where a series of high fluctuations in transmissibility takes place which gives rise to a
concentrated increment in transmission of structure-borne sound power. A medium size resilient element for the main
engines of a ship propulsion plant shows a series of resonances of the rubber part from about 300-500 Hz up to 1,0-1,5
kHz and studies based on laboratory experiments are necessary to find and identify such a phenomenon. In this frame,
comparative laboratory tests are essential in the selection process of the better resilient element solution to be used in
combination with a diesel engine and a foundation structure. In the paper, a test rig for the dynamic characterization of
large resilient mountings has been presented, showing theory and methods for performing accurate measurements of
dynamic transfer stiffness of resilient mountings. Then, a test campaign has been discussed, aimed to verify the ability of
an improved resilient element in isolating vibrations. Tested resilient mountings have been designed and produced by
Vulkan Italia Srl. Pros and cons of the two solutions are presented in order to explain how a resilient mounting may be
optimal within a given range of frequencies.

ACKNOWLEDGEMENTS
The research work reported in this paper was carried out with the technical support of Vulkan Italia Srl, and precisely in
close collaboration with Mr. Giovanni De Martini, who has promoted such a fruitful cooperation between the Ship Noise
and Vibration Laboratory (NVL) of the University of Trieste and Vulkan Italia Srl.

REFERENCES
BESIO G., LOREDAN V. and CONTENTO O. “Improved Noise Control on Board Ships. Optimization of Resilient
Mountings and Bed Structures”. Proceedings of the International Conference on Ship and Marine Research NAV’97,
Sorrento, Italy, 1997.
BIOT M. and DE LORENZO F. “Open Issues Regarding Noise and Vibrations on Board Cruise Ships: a Suggested
Approach for Measuring Comfort”. Proceeding of the Autumn Conference 2007 − Advanced in Noise and Vibration
Engineering, Institute of Acoustics, Oxford, UK, 2007.
FULFORD R.A. and GIBBS B.M. “Structure-Borne Sound Power and Source Characterization in Multi-Point-
Connected Systems, Part 1: Case Studies for Assumed force Distributions”. Journal of Sound and Vibration, 204(4)
(1997) 659-677.
HYNNÄ P. “Vibrational Power Method in Control of Sound and Vibration”. Research Report No. BVAL37-021229,
Technical Research Centre of Finland, Espoo, Finland, 2002. (available at: http://www.vtt.fi/inf/pdf/jurelinkit/BVAL37-
021229.pdf)
ISO 10846-1. “Acoustics and vibration − Laboratory measurement of vibro-acoustic transfer properties of resilient
elements − Part 1: Principles and guidelines”. International Organization for Standardization, 2008.
ISO 10846-2. “Acoustics and vibration − Laboratory measurement of vibro-acoustic transfer properties of resilient
elements − Part 2: Direct method for determination of the dynamic stiffness of resilient supports for translatory motion”.
International Organization for Standardization, 2008.
ISO 10846-3. “Acoustics and vibration − Laboratory measurement of vibro-acoustic transfer properties of resilient
elements − Part 3: Indirect method for determination of the dynamic stiffness of resilient supports for translatory
motion”. International Organization for Standardization, 2002.
ISO 10846-5. “Acoustics and vibration − Laboratory measurement of vibro-acoustic transfer properties of resilient
elements − Part 5: Driving point method for determination of the low-frequency transfer stiffness of resilient supports for
translatory motion”. International Organization for Standardization, 2008.
MONDOT J.M. and PETERSSON B. “Characterization of structure-borne sound sources: The source descriptor and the
coupling function”. Journal of Sound and Vibration, 114(1) (1987) 507-518.
PETERSSON B. and PLUNT J. “On Effective mobilities in the prediction of structure-borne sound transmission between
a source structure and a receiving structure, part 1: Theoretical background and basic studies”. Journal of Sound and
Vibration, 82 (1982) 517-529.
PETERSSON B. and GIBBS B.M. “Use of the source descriptor concept in studies of multi-point and multi-directional
vibrational sources”. Journal of Sound and Vibration, 168 (1991) 157-176.
RAN LIN T., PAN J., O’SHEA P.J. and MECHEFSKE C.K. “A Study of Vibration and Vibration Control of Ship
Structures” Marine Structures, 22 (2009) 730-743.
TAO .J.S, LIU G.R. and LAM K.Y. “Design Optimization of Marine Engine-Mount System”. Journal of Sound and
Vibration, 235(3) (2000) 477-494.
THOMPSON D.J., VAN VLIET W.J. and VERHEIJ W. “Developments of the Indirect Method for Measuring the High
Frequency Dynamic Stiffness of Resilient Elements”. Journal of Sound and Vibration, 213(1) (1998) 169-188.

View publication stats

You might also like