15
Exact tuning of a vibration neutralizer for the reduction of flexural waves in beams Jimmy S. Issa a) Department of Industrial and Mechanical Engineering, Lebanese American University, P.O. Box 36, Byblos, Lebanon (Received 14 January 2019; revised 6 June 2019; accepted 26 June 2019; published online 26 July 2019) In this work, designs of vibration neutralizers for the reduction of flexural waves in beams are pro- posed. The system considered consists of an Euler-Bernoulli beam experiencing a harmonically travelling wave. The neutralizer, which consists of a mass, spring, and damper (viscous or hyster- etic), is attached at a point on the beam. Two designs are considered. In the first, the aim is the min- imization of the transmitted energy, and in the second, the aim is the maximization of the energy dissipated in the neutralizer. For minimum energy transmission, the undamped neutralizer can be tuned to reflect all the energy back to its source at the tuning frequency. The tuning stiffness ratio and the neutralizer bandwidth is obtained in an analytical closed form. For maximum energy dissi- pation, the neutralizer can dissipate up to 50% of the incident energy. The maximum energy dissi- pated, i.e., 50% of the incident energy, is constant and occurs at the tuning frequency. It is neither affected by the neutralizer mass nor by the tuning frequency. The tuning parameters are obtained analytically. Finally, in all the above cases, the increasing of the neutralizer mass enhances the neu- tralizer performance. All analytical findings are validated numerically. V C 2019 Acoustical Society of America. https://doi.org/10.1121/1.5116690 [BA] Pages: 486–500 I. INTRODUCTION Vibration suppression has always been a research topic of paramount importance because of the damaging effect of unwanted vibration on both structures and human beings. For example, vibration of machines reduces their efficien- cies, increases wear, and can lead to failure due to fatigue. Once initiated from a source, the vibrational energy will not necessarily stay confined to a part of a structure, since it can easily get transmitted from one part to another. In the pro- cess, the vibration of these parts will interact with the sur- rounding air which absorbs part of this energy in the form of noise resulting in annoyance in the surrounding area and in the extreme case to hearing impairments. The vibration absorber is one of the most commonly used passive techni- ques for vibration reduction. Since its invention in 1909, 1 this device has been properly designed for maximum perfor- mance 24 when coupled to discrete single degree of freedom main systems. It reduces the vibration of the main system by applying forces passively that counteract the forces that are transmitted to the main system from the excitation source. When the main system can no longer be modeled as a point mass, e.g., slender parts in structures, a different approach for vibration suppression should be adopted. A common example of slender parts is beam elements in struc- tures, which are responsible for channeling and spreading the vibrational energy initially generated at one location on the structure. Once the main guides of the energy transmitted in a structure are identified, vibration neutralizers can be attached to them and designed to reduce energy propagation across these guides. When a wave propagates in a beam, it will not get disturbed unless it encounters an impedance mismatch at a point, e.g., a point mass or a spring. 5,6 Part of the energy will get transmitted through this point and the remaining part will get reflected and directed back to its source. The first meticulous study of a vibration neutralizer for the reduction of flexural waves in beams was conducted by Brennan. 7 The neu- tralizer is assumed to be lightly damped and approximate ana- lytical expressions for the tuning frequency; bandwidth and attenuation ratio are derived in two different frequency regimes where these expressions can be simplified. El Khatib et al. 8 used a hysteretically damped vibration neutralizer to either reflect the energy transmitted at a point or to absorb it. The effect of both the near field and far field waves is considered. Salleh and Brennan 9 considered four different neutralizer con- figurations, namely, the force type, the moment type, the cou- pled, and the uncoupled force moment types. The transmission ratio was derived for each case; however, the study focused on the force and the uncoupled force moment types only because they can be easily realized. It is shown that the uncoupled force moment type can be more effective than the force type. Thompson 10 attached a continuous vibration neutralizer with- out bending stiffness to a beam and showed that a blocked fre- quency region exists where waves will no longer be able to freely propagate across the beam at these frequencies. Approximate formulas for the waves decay rate and neutralizer bandwidth are derived. The use of multiple vibration neutral- izers for the reduction of flexural waves in beams was also con- sidered, e.g., 1116 the key parameter in this case is the separation between the different oscillators. The reader is referred to the cited works therein for a more exhaustive cover- age of this part of the research. a) Electronic mail: [email protected] 486 J. Acoust. Soc. Am. 146 (1), July 2019 V C 2019 Acoustical Society of America 0001-4966/2019/146(1)/486/15/$30.00

Exact tuning of a vibration neutralizer for the reduction

  • Upload
    others

  • View
    8

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Exact tuning of a vibration neutralizer for the reduction

Exact tuning of a vibration neutralizer for the reductionof flexural waves in beams

Jimmy S. Issaa)

Department of Industrial and Mechanical Engineering, Lebanese American University, P.O. Box 36, Byblos,Lebanon

(Received 14 January 2019; revised 6 June 2019; accepted 26 June 2019; published online 26 July2019)

In this work, designs of vibration neutralizers for the reduction of flexural waves in beams are pro-

posed. The system considered consists of an Euler-Bernoulli beam experiencing a harmonically

travelling wave. The neutralizer, which consists of a mass, spring, and damper (viscous or hyster-

etic), is attached at a point on the beam. Two designs are considered. In the first, the aim is the min-

imization of the transmitted energy, and in the second, the aim is the maximization of the energy

dissipated in the neutralizer. For minimum energy transmission, the undamped neutralizer can be

tuned to reflect all the energy back to its source at the tuning frequency. The tuning stiffness ratio

and the neutralizer bandwidth is obtained in an analytical closed form. For maximum energy dissi-

pation, the neutralizer can dissipate up to 50% of the incident energy. The maximum energy dissi-

pated, i.e., 50% of the incident energy, is constant and occurs at the tuning frequency. It is neither

affected by the neutralizer mass nor by the tuning frequency. The tuning parameters are obtained

analytically. Finally, in all the above cases, the increasing of the neutralizer mass enhances the neu-

tralizer performance. All analytical findings are validated numerically.VC 2019 Acoustical Society of America. https://doi.org/10.1121/1.5116690

[BA] Pages: 486–500

I. INTRODUCTION

Vibration suppression has always been a research topic

of paramount importance because of the damaging effect of

unwanted vibration on both structures and human beings.

For example, vibration of machines reduces their efficien-

cies, increases wear, and can lead to failure due to fatigue.

Once initiated from a source, the vibrational energy will not

necessarily stay confined to a part of a structure, since it can

easily get transmitted from one part to another. In the pro-

cess, the vibration of these parts will interact with the sur-

rounding air which absorbs part of this energy in the form of

noise resulting in annoyance in the surrounding area and in

the extreme case to hearing impairments. The vibration

absorber is one of the most commonly used passive techni-

ques for vibration reduction. Since its invention in 1909,1

this device has been properly designed for maximum perfor-

mance2–4 when coupled to discrete single degree of freedom

main systems. It reduces the vibration of the main system by

applying forces passively that counteract the forces that are

transmitted to the main system from the excitation source.

When the main system can no longer be modeled as a

point mass, e.g., slender parts in structures, a different

approach for vibration suppression should be adopted. A

common example of slender parts is beam elements in struc-

tures, which are responsible for channeling and spreading

the vibrational energy initially generated at one location on

the structure. Once the main guides of the energy transmitted

in a structure are identified, vibration neutralizers can be

attached to them and designed to reduce energy propagation

across these guides. When a wave propagates in a beam, it will

not get disturbed unless it encounters an impedance mismatch

at a point, e.g., a point mass or a spring.5,6 Part of the energy

will get transmitted through this point and the remaining part

will get reflected and directed back to its source. The first

meticulous study of a vibration neutralizer for the reduction of

flexural waves in beams was conducted by Brennan.7 The neu-

tralizer is assumed to be lightly damped and approximate ana-

lytical expressions for the tuning frequency; bandwidth and

attenuation ratio are derived in two different frequency regimes

where these expressions can be simplified. El Khatib et al.8

used a hysteretically damped vibration neutralizer to either

reflect the energy transmitted at a point or to absorb it. The

effect of both the near field and far field waves is considered.

Salleh and Brennan9 considered four different neutralizer con-

figurations, namely, the force type, the moment type, the cou-

pled, and the uncoupled force moment types. The transmission

ratio was derived for each case; however, the study focused on

the force and the uncoupled force moment types only because

they can be easily realized. It is shown that the uncoupled force

moment type can be more effective than the force type.

Thompson10 attached a continuous vibration neutralizer with-

out bending stiffness to a beam and showed that a blocked fre-

quency region exists where waves will no longer be able to

freely propagate across the beam at these frequencies.

Approximate formulas for the waves decay rate and neutralizer

bandwidth are derived. The use of multiple vibration neutral-

izers for the reduction of flexural waves in beams was also con-

sidered, e.g.,11–16 the key parameter in this case is the

separation between the different oscillators. The reader is

referred to the cited works therein for a more exhaustive cover-

age of this part of the research.a)Electronic mail: [email protected]

486 J. Acoust. Soc. Am. 146 (1), July 2019 VC 2019 Acoustical Society of America0001-4966/2019/146(1)/486/15/$30.00

Page 2: Exact tuning of a vibration neutralizer for the reduction

In this work, a comprehensive study on the reduction of

flexural waves in a beam using a vibration neutralizer is pre-

sented. The neutralizer is assumed to be attached at one point

on the beam. A wave propagating from one side of the beam

will encounter an obstacle, i.e., the neutralizer. The aim is

vibration reduction, and it is achieved by either minimizing

the energy transmitted passed the attachment point, or by

maximizing the energy dissipated in the neutralizer. It is

shown how one can select the neutralizer damping and stiff-

ness properties to achieve optimal performance. In both

cases, and to reduce errors resulting from simplifications, the

optimal tuning parameters of the neutralizer are obtained in

an analytical closed form without making any kind of

assumptions on the system damping size or on the system

operating regime. The results are obtained for the two com-

mon damping models, namely, viscous and hysteretic. The

exactness of the solution is validated for both objectives con-

sidered by a direct numerical resolution of the coupled beam

neutralizer equation of motion. In Sec. II, the dimensionless

parameters are defined and the relationships between the differ-

ent wave components are derived. In Sec. III, exact tuning for

minimum energy transmission is performed. The transmission

and reflection ratios are defined, and the tuning neutralizer

parameters are determined analytically. In Sec. IV, design for

maximum energy dissipation is considered. The dissipation

ratio is defined, and the tuning neutralizer parameters are deter-

mined in closed form. Numerical verification of the analytical

findings is done in Sec. V. Concluding remarks and directions

for future works are given in Sec. VI.

II. TRANSMISSION AND REFLECTION RATIOS

The schematic of an infinite beam coupled with a vis-

cously damped neutralizer is shown in Fig. 1(a) and with a hys-

teretically damped neutralizer in Fig. 1(b). In these figures, mand k depict the neutralizer mass and stiffness constant, respec-

tively. w(x, t) is the transverse motion of the beam, and y(t) is

the neutralizer displacement. The viscous damping constant is

depicted by c in Fig. 1(a), and the loss factor of the hysteretic

damping by d in Fig. 1(b). It is assumed that the neutralizer is

attached at the reference origin x¼ 0 in the far field of the

energy source and, hence, the incident wave wi comprises of its

propagating component only. The transmitted wave wt has two

parts, namely, wtp and wte, which are the propagating and eva-

nescent components, respectively. Similarly, wrp and wre

denote the propagating and evanescent components of the

reflected wave wr, respectively. Assuming harmonic excitation,

the incident, transmitted and reflected wave components can be

written as follows:

wiðx;tÞ¼Wi ei xt�bxð Þ;

wtðx;tÞ¼wtpðx; tÞþwteðx; tÞ¼Wt ei xt�bxð Þ þWte e ixt�bxð Þ;

wrðx; tÞ¼wrpðx;tÞþwreðx;tÞ¼Wr ei xtþbxð Þ þWre e ixtþbxð Þ:

(1)

In Eq. (1), Wi, Wt, and Wr are the amplitudes of the propagat-

ing component of the incident, transmitted and reflected

waves, respectively. Wte and Wre are the amplitudes of the

evanescent component of the transmitted and reflected

waves at x¼ 0, respectively, and x is the excitation fre-

quency. Let q and E be the material density and Young’s

modulus, respectively, and A and I the cross-sectional area

and its second moment of area, respectively. Then, the flex-

ural wave number is defined as b ¼ ðx2qA=EIÞ1=4. In the

remainder of the paper, the time term in Eq. (1) will be

ignored to reduce verbosity. The next step is to solve for the

reflected and transmitted amplitudes in terms of the incident

amplitude and the system parameters. This is achieved by

ensuring continuity of the displacement and slope of the

beam at x¼ 0 and by applying Newton’s Law to an infinites-

imal beam element centered at x¼ 0. The continuity equa-

tions on the displacement and slope of the beam at x¼ 0 are

written as follows:

wið0; tÞ þ wrð0; tÞ ¼ wtð0; tÞ;@wi

@xð0; tÞ þ @wr

@xð0; tÞ ¼ @wt

@xð0; tÞ: (2)

An infinitesimal element of the beam at x¼ 0 is shown in Fig. 2.

V and M denote the shear force and bending moment, respec-

tively, and Fd the dynamic force applied by the neutralizer on

the beam. The minus and plus superscripts represent variables

evaluated right before and right after x¼ 0, respectively.

Applying Newton’s law on this element results in the

following two equations:

M�ð0; tÞ �Mþð0; tÞ ¼ 0;

FdðtÞ þ V�ð0; tÞ � Vþð0; tÞ ¼ 0: (3)

Let Kd be the dynamic stiffness of the neutralizer, hence

Eqs. (2) and (3) can now be simplified to

FIG. 1. Schematic of an infinite beam coupled with a vibration neutralizer (a) with viscous damping, (b) with hysteretic damping.

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 487

Page 3: Exact tuning of a vibration neutralizer for the reduction

Wi þWr þWre �Wt �Wte ¼ 0;

Wi �Wr þ iWre �Wt þ iWte ¼ 0;

Wi þWr �Wre �Wt þWte ¼ 0;

KdðWi þWr þWreÞ

� EIb3ðiWi � iWr þWre � iWt þWteÞ ¼ 0: (4)

Solving Eq. (4) yields

Wt

Wi¼ ð1� iÞða� 1Þ

2a� ð1� iÞ ;Wr

Wi¼ � ð1þ iÞa

2a� ð1� iÞ ;

Wte

Wi¼ Wre

Wi¼ � ð1� iÞa

2a� ð1� iÞ : (5)

In Eq. (5), a¼Kd/4EIb3 is a complex dimensionless ratio that

encompasses all the system parameters. It is the ratio of the

neutralizer dynamic stiffness to that of the beam, and hence it

will be referred to as the dynamic stiffness ratio. In the remain-

der of the paper, parameters with subscripts v and h refer to the

viscous and hysteretic damping cases, respectively. For exam-

ple, Kd is defined as the neutralizer dynamic stiffness, then Kdv

is the dynamic stiffness of the viscously damped neutralizer

and Kdhis that of the hysteretically damped neutralizer. Kdv

and Kdh take the form (see Appendix A)

Kdv ¼1

1

k þ icx� 1

mx2

; Kdh¼ 1

1

kð1þ idÞ �1

mx2

:

(6)

Now, the following dimensionless parameters are defined

and used instead of the dimensional ones to generalize the

results:

l ¼ m

4qA3=2; f ¼ kA3=2

4EI; f ¼ c

4ffiffiffiffiffiffiffiffiEIqp ;

r ¼ x

� ffiffiffiffiffiffiffiffiEI

qA3

s0@

1A

1=2

: (7)

In Eq. (7), l is the mass ratio which is the ratio of the neutral-

izer mass to that of a portion of this beam of length 4A1=2. f is

the stiffness ratio; it can be viewed as the ratio of the neutral-

izer stiffness to the static stiffness at the mid-span of a simply

supported portion of this beam of length 121=3A1=2. f is the

damping ratio, which is the ratio of the neutralizer viscous

damping constant to the damping constant 2ffiffiffiffiffiffiffiffiEIqp

. Finally, r is

the frequency ratio and is defined as the square root of the ratio

of the excitation frequency to the first natural frequency of a

simply supported beam made of the same material with a

length of pA1=2. It is important to note that these parameters

are conveniently defined in a way to ensure that each one

depends on only one of the design parameters. For example,

the effect of the neutralizer mass is taken into account in the

mass ratio l only, and hence varying m will change the lparameter only without affecting the remaining ones.

Similarly, f, f, and r separately account for the effect of k, c,

and x, respectively. The dynamic stiffness ratio can now be

written in terms of the dimensionless parameters for the vis-

cously and hysteretically damped cases as follows:

av ¼1

1

f

r3þ i

fr

� 1

lr

or ah ¼1

1

f

r3ð1þ idÞ

� 1

lr

:(8)

To conclude this section, the aim of this work is to propose

two different vibration neutralizer designs, i.e., find the most

appropriate mass, stiffness, and damping constants of the

neutralizer that will fit a given objective. The first design

aims at reducing the energy transmitted passed the attach-

ment point at some given excitation frequency. And for the

second design, the aim is to maximize energy absorption by

the neutralizer at some given excitation frequency. In Sec.

III, the first design is treated.

III. DESIGN FOR MINIMUM ENERGY TRANSMISSION

A. Transmission and reflection ratios

A harmonic flexural wave w(x, t) carries energy as it

travels across a beam from one point to another. The instan-

taneous power flowing past a point in an Euler-Bernoulli

beam in the positive x axis direction is defined in Ref. 17.

The average energy incident at a point can be calculated

as Pi ¼ kWik2 EIxb3. Similarly, the average transmitted

energy passed a point and that reflected at the point are

Pt ¼ kWtk2 EIxb3 and Pr ¼ kWrk2 EIxb3, respectively.

The evanescent components of the waves do not show up in

these equations simply because they do not carry energy.

The fraction of the incident energy that is transmitted

passed the attachment point is defined as the transmission

ratio and calculated as T ¼ Pt=Pi ¼ kWtk2=kWik2.

Similarly, the reflection ratio, R ¼ Pr=Pi ¼ kWrk2=kWik2,

is defined as the faction of the incident energy that is

reflected at the attachment point. When there is no energy

dissipation at the attachment point (zero damping), the

energy incident will partly get transmitted and the remain-

ing part will get reflected, hence in this case TþR¼ 1.

When damping is present, part of the incident energy will

get dissipated in the damper. The transmission and reflec-

tion ratios are calculated in terms of the dimensionless

parameters for the viscously and hysteretically damped

cases and take the forms

FIG. 2. Schematic of an infinitesimal element of the beam centered around

x¼ 0.

488 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 4: Exact tuning of a vibration neutralizer for the reduction

Tv ¼ðf þ frl� r4lÞ2 þ r4f2ð1þ rlÞ2

ðf þ frl� r3fl� r4lÞ2 þ r2ðrfþ f lþ r2flÞ2;

Rv ¼r6f2l2 þ r2f 2l2

ðf þ frl� r3fl� r4lÞ2 þ r2ðrfþ f lþ r2flÞ2;

or

Th ¼ðf þ frl� r4lÞ2 þ f 2d2ð1þ rlÞ2

f 2ðdþ rlþ rdlÞ2 þ ðr4lþ frdl� frl� f Þ2;

Rh ¼f 2r2l2 þ f 2r2d2l2

f 2ðdþ rlþ rdlÞ2 þ ðr4lþ frdl� frl� f Þ2:

(9)

Since the target in this section is the minimization of the

transmitted energy, the neutralizer parameters should be

properly chosen to reduce the transmission ratio T. The

undamped neutralizer design is treated in Sec. III B.

B. Undamped neutralizer

The transmission ratio of an undamped neutralizer is

referred to as T0 and is obtained from Eq. (9) from Tv (or Th)

by setting the damping parameter to zero, as

T0 ¼ðf þ frl� r4lÞ2

ðf þ frl� r4lÞ2 þ ðrflÞ2: (10)

In Fig. 3, T0 is plotted versus the frequency ratio for a given

set of parameters. The plot is first thoroughly analyzed and

examined in order to facilitate proposing optimized design

guidelines. The figure clearly shows that for low and high fre-

quencies, the transmission ratio tends towards 1, i.e., at these

frequencies, waves are completely transmitted and hence neu-

tralizers are useless. At some frequency ratio denoted by r0,

the transmission ratio is zero and hence a wave travelling

with a frequency ratio r0 will get completely reflected back to

its source. This frequency ratio is referred to as the tuning fre-

quency ratio. Furthermore, it is observed that waves with fre-

quency ratios falling in the vicinity of r0 will have a large

chunk of their energy reflected back to their source. It is con-

cluded that, if properly designed, these neutralizers are very

useful in reducing flexural waves with a narrow frequency

band defined by a main major frequency and some fluctuations

around it. In this case, one should choose the neutralizer

parameters such that r0 coincides with the major excitation fre-

quency and should ensure that the remaining frequency band

receives adequate amount of energy reflection. This is referred

to as tuning the neutralizer. The neutralizer bandwidth is

defined as the frequency range within which the transmission

ratio is less than �3 dB, i.e., T0 � 1=2. The boundaries of the

neutralizer bandwidth are depicted in the figure by the two key

frequency ratios r1 and r2. Often, the major excitation fre-

quency is well defined; however, its fluctuations, i.e., frequency

range is not, and hence the best design is achieved by maximiz-

ing the neutralizer bandwidth so as to encompass the largest

amount of fluctuations possible around the main major fre-

quency. The frequency ratios r0, r1, and r2 are first obtained in

terms of the neutralizer parameters, then design guidelines for

the undamped neutralizer are proposed.

Using Eq. (10), the tuning frequency r0 is calculated

from the equation T0¼ 0. This results in the following fourth

order polynomial in r:

f þ fl r � l r4 ¼ 0: (11)

This equation can be solved using Ferrari’s method and

yields only one real solution, which takes the form

c0 ¼ �27f 2l3 þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið27f 2l3 þ 128f Þ2 � ð128f Þ2

q;

b0 ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi128f 3

c0

!1=3

þ c0

54l3

� �1=3

vuut;

r0 ¼1

2b0 þ

1

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi2f

b0

� b20

s: (12)

The bandwidth boundaries r1 and r2 are calculated form the

equation T0 ¼ 1=2, which after simplification, reduces to

ðf � r4lÞðf þ 2frl� r4lÞ ¼ 0: (13)

Solving this equation yields eight roots out of which two are

real positive and define the bandwidth of the neutralizer as

r1 ¼f

l

� �1=4

;

c2 ¼ �54f 2l3 þffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffið54f 2l3 þ 16f Þ2 � ð16f Þ2

q;

b2 ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi64f 3

c2

!1=3

� c2

27l3

� �1=3

vuut;

r2 ¼1

2b2 þ

1

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi4f

b2

� b22

s: (14)

FIG. 3. Plot of the transmission ratio showing the tuning frequency ratio r0

and the boundary of the neutralizer bandwidth r1 and r2, for l¼ 1.25 and

f¼ 0.25.

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 489

Page 5: Exact tuning of a vibration neutralizer for the reduction

For a given neutralizer mass and stiffness, i.e., l and f, the

tuning frequency ratio r0 can be calculated from the exact

expression given in Eq. (12) and the normalized bandwidth

(r2� r1)/r0 can be obtained in an exact closed form using

Eqs. (12) and (14). Since in a real scenario the major excita-

tion frequency is known, the neutralizer is tuned to this fre-

quency and therefore, r0 is known. Tuning the neutralizer is

achieved by choosing its mass and stiffness such that Eq.

(11) is satisfied after substituting r0 for r. Hence, for a given

neutralizer mass ratio l, the tuned stiffness ratio is calculated

as follows:

ft ¼l r4

0

1þ l r0

: (15)

The effect of l on the neutralizer bandwidth is examined by

inspecting the sign of @r1/@l and @r2/@l after tuning the sys-

tem, i.e., after substituting ft for f in the expressions of r1 and

r2. This results in, after simplifications

@r1

@l¼ � r2

0

4ð1þ l r0Þ5=4;

@r2

@l¼ r2ð2r2 � r0Þ

2ð1þ l r0Þð2þ 3lr2Þ:

(16)

It is concluded from Eq. (16) that 8l; @r1=@l < 0, since

both l and r0 are positive. This means that for a fixed r0, r1

decreases with the increasing of l. Furthermore, it is con-

cluded from the same equation that 8l; @r2=@l > 0, since

both l and r0 are positive and since by definition r2 > r0.

This means that r2 increases with the increasing of l. Hence,

in summary, when the neutralizer is tuned to some fixed fre-

quency ratio r0 by setting f to ft, the neutralizer bandwidth

increases with the increasing of the mass ratio l. Hence, one

should use the largest neutralizer mass possible to widen its

bandwidth as much as possible.

The transmission ratio is plotted in Fig. 4 for different

mass ratios of a neutralizer tuned to the frequency ratio r0¼ 1.

The figure clearly shows that as l increases, the bandwidth

increases to reach its maximum when l tends towards infinity.

In this case, the neutralizer mass is infinitely large, and this

scenario reduces to the case of a spring attached directly to a

fixed ground. r1 tends towards zero and r2 tends towards

21=3r0. Furthermore, it is important to note that the tuning stiff-

ness ratio ft depends on l and increases with increasing l.

When l tends towards infinity, the tuning stiffness ratio tends

towards r30. This concludes the discussion of the study of the

undamped neutralizer design. In Sec. III C, the case of the

damped neutralizer will be considered.

C. Damped neutralizer

When damping is present in the neutralizer, the trans-

mission ratio will no longer exhibit zero transmission at

some frequency, instead it will pass through a minimum

value which is higher than zero. The transmission ratio is

plotted in Fig. 5 for different neutralizer parameters, namely,

Fig. 5(a) with viscous damping and Fig. 5(b) with hysteretic

damping. For both damping cases, the mass ratio is chosen

to be l¼ 3 and the stiffness and damping ratios are properly

chosen so that the system remains tuned to r0¼ 0.5. Tuning

when damping is zero yields a transmission ratio passing

though zero at r0¼ 0.5; however, when damping is present,

tuning is achieved when the minimum of the transmission

ratio occurs at r0¼ 0.5. Each figure shows two damped cases

tuned to r0¼ 0.5. In the first case, the minimum transmission

ratio, which is denoted by h0, is equal to 0.1, and in the sec-

ond it is equal to 0.2. For example, in Fig. 5(a), the minimum

transmission ratio will equal 0.1 at r0¼ 0.5 when (f¼ 0.086,

f¼ 0.0756) and 0.2 when (f¼ 0.156, f¼ 0.0748). This means

that in these cases, the best scenario occurs when r¼ r0, at

this point zero transmission is not achieved; instead, 10% or

20% of the energy will still get transmitted through the point

of attachment of the neutralizer. Therefore, one might think

that the addition of damping is detrimental on the neutral-

izer. This is not entirely true because even though the addi-

tion of damping deteriorates the neutralizer performance at

the tuning frequency, it increases its bandwidth specially on

the right side of r0. The effect of the increase of the neutral-

izer damping on increasing r2 is higher than that on increas-

ing r1. The same observations are made for the hysteretically

damped cases shown in Fig. 5(b). Hence, the addition of

damping yields some type of trade-off in the neutralizer

design.

These damped neutralizers are very useful for the reduc-

tion of travelling waves with relatively larger frequency

bands compared to those treated by undamped neutralizers

which are considered to have narrow frequency bandwidths.

In this case, and since there is a trade-off between the mini-

mum transmission ratio and the neutralizer bandwidth, in

addition to the major frequency r0, which is considered to be

known, the sought after minimum transmission ratio h0 is

assumed to be known. Then, for a given l, the neutralizer

can now be tuned to r0 by a proper choice of the stiffness

and damping ratios. The tuning neutralizer parameters sat-

isfy the below two equations,

FIG. 4. Plot of the transmission ratio when the neutralizer is tuned to r0¼ 1

showing the effect of the mass ratio l on the neutralizer bandwidth.

490 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 6: Exact tuning of a vibration neutralizer for the reduction

@T

@r

����r¼r0

¼ 0; T

����r¼r0

¼ h0: (17)

The above equations are solved analytically in Appendix B

and the tuning stiffness and damping ratios are obtained in

closed form in terms of r0, h0, and l. The effect of the mass

ratio on the transmission ratio is depicted in Fig. 6. In this

figure, the neutralizer is tuned to r0¼ 0.5 and h0¼ 0.1. The

tuning neutralizer parameters corresponding to each mass

ratio, are calculated using the analytical expressions derived

in Appendix B. It is clearly shown that, as expected, the neu-

tralizer bandwidth increases with the increasing of the mass

ratio. The limiting case when l tends towards 1, i.e., when

the spring and damper are directly connected to a fixed sup-

port is shown too. The latter case yields the utmost perfor-

mance; however, connecting a spring and damper to a fixed

support is most of the time not physically achievable.

IV. DESIGN FOR MAXIMUM ENERGY DISSIPATION

A. Dissipation ratio

In this section, the aim is to maximize energy dissipa-

tion in the neutralizer. When coupled to a beam, a damped

vibration neutralizer will dissipate part of the travelling

energy in the damping element. Hence, in this case, the addi-

tion of a damped neutralizer is always beneficial; however,

FIG. 5. Plots of the transmission ratio (a) viscously damped for l¼ 3.0 and three different sets of stiffness and damping ratios, namely, (f¼ 0.0, f¼ 0.075),

(f¼ 0.086, f¼ 0.0756), and (f¼ 0.156, f¼ 0.0748), (b) hysteretically damped for l¼ 3.0 and three different sets of stiffness and loss factors, namely, (d¼ 0.0,

f¼ 0.075), (d¼ 0.29, f¼ 0.073), and (d¼ 0.53, f¼ 0.068).

FIG. 6. Plots of the transmission ratio of a vibration neutralizer tuned to r0¼ 0.5 and h0¼ 0.1 showing the effect of the mass ratio l (a) viscously damped for

(l¼ 0.1, ft¼ 0.000524, ft¼ 0.00595), (l¼ 2, ft¼ 0.059, ft¼ 0.0628), (l¼ 5, ft¼ 0.124, ft¼ 0.0908), and (l¼1, ft¼ 0.258, ft¼ 0.133); (b) hysteretically

damped for (l¼ 0.1, dt¼ 0.022, ft¼ 0.00595), (l¼ 2, dt¼ 0.236, ft¼ 0.0613), (l¼ 5, dt¼ 0.345, ft¼ 0.0863), and (l¼1, dt¼ 0.505, ft¼ 0.118).

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 491

Page 7: Exact tuning of a vibration neutralizer for the reduction

the neutralizer performance can be maximized if the latter is

properly designed. The average energy incident, transmitted

and reflected at the attachment point, were calculated in

Sec. III as Pi ¼ kWik2 EIxb3; Pt ¼ kWtk2 EIxb3 and

Pr ¼ kWrk2 EIxb3, respectively. The relative motion of

the neutralizer with respect to its attachment point on the

beam is defined as z(t)¼ y(t)�w(0, t). The average energy

dissipated over one cycle in the damper element of the

neutralizer is calculated as Pdv ¼ 1=2cx2kzðtÞk2for vis-

cous damping and Pdh¼ 1=2kdxkzðtÞk2

for hysteretic

damping. The dissipation ratio is defined as the ratio of the

dissipated energy to the incident energy as, D¼Pd/Pi. The

neutralizer parameters will be obtained analytically with

the aim of maximizing the dissipation ratio D. Let yðtÞ¼ Y eixt be the neutralizer response, where Y is the neutral-

izer complex amplitude. y(t) can be directly related to the

motion w(0, t) of the attachment point. Assuming harmonic

motion, the two displacements are related through the

equation of motion of the neutralizer, which can be written

for the viscous and hysteretic damping cases, respectively,

as follows:

m€yðtÞ þ c _yðtÞ þ kyðtÞ ¼ c _wð0; tÞ þ kwð0; tÞ

m€y þ k 1þ idð ÞyðtÞ ¼ k 1þ idð Þwð0; tÞ: (18)

Assuming harmonic motion, wð0; tÞ ¼ W0eixt, the steady

state neutralizer response Y is written in terms of W0 by solv-

ing Eq. (18) as

Yv ¼kþ icx

k�mx2ð Þ þ icxW0; Yh ¼

k 1þ idð Þk�mx2ð Þ þ ikd

W0:

(19)

Now, the beam displacement at the attachment point w(0, t) can

be written using Eq. (1) in terms of the incident and reflected

wave as wð0; tÞ ¼ wið0; tÞ þ wrð0; tÞ ¼ ðWi þWr þWreÞ eixt

or in terms of the transmitted wave as wð0; tÞ ¼ wtð0; tÞ¼ ðWt þWteÞ eixt. For convenience, the second expression is

used resulting in W0¼WtþWte. This simplifies Eq. (19) to

Yv ¼k þ icx

k � mx2ð Þ þ icxðWt þWteÞ;

Yh ¼k 1þ idð Þ

k � mx2ð Þ þ ikdðWt þWteÞ: (20)

The relative neutralizer motion takes the form zðtÞ ¼ yðtÞ�wð0; tÞ ¼ ðY � ðWt þWteÞÞ eixt. Let Z¼ Y� (WtþWte) be

the complex amplitude of z(t), which takes the form

Zv ¼mx2

k � mx2ð Þ þ icxðWt þWteÞ;

Zh ¼mx2

k � mx2ð Þ þ ikdðWt þWteÞ: (21)

Using the expressions for Wt and Wte from Eq. (5), the dissi-

pation ratio is written in terms of the dimensionless parame-

ters as follows:

Dv ¼1=2cx2kZvk2

EIxb3 kWik2

¼ 2r7fl2

r2 flþ fr þ r2flð Þ2 þ f þ frl� r3ðr þ fÞl� �2

;

Dh ¼1=2kdxkZhk2

EIxb3 kWik2

¼ 2fr5dl2

f 2 dþ rlð1þ dÞð Þ2 þ r4lþ frlðd� 1Þ � f� �2

:

(22)

Alternatively, the dissipation ratio can be obtained using a dif-

ferent approach. Instead of directly calculating the energy dis-

sipated in the neutralizer, one can calculate it from the incident

energy Pi, transmitted energy Pt, and reflected energy Pr. It is

a fact that, as mentioned before, for an undamped neutralizer,

the incident energy is equal to the summation of the reflected

and transmitted energies, i.e., Pi¼ (PrþPt), and hence no

energy is lost. However, when damping is present, the energy

dissipated in the neutralizer can be obtained as Pi� (PrþPt).

This means that the dissipation ratio can be written in terms of

the reflection and transmission ratios as D¼ 1� (TþR).

Substituting the expressions of T and R in terms of the dimen-

sionless parameters from Eq. (9) into D¼ 1� (TþR) yields

expressions of the dissipation ratios as Dv¼ 1� (TvþRv) and

Dh¼ 1� (ThþRh) identical to those given in Eq. (22). The

objective is to maximize the dissipation ratios Dv and Dh by a

proper choice of the neutralizer parameters. It is important to

note that the dissipation ratios depend on the frequency ratio rand hence the optimization problem is formulated as follows.

For a given mass ratio l, find the optimal stiffness and damp-

ing ratios such that the dissipation ratio is maximized at the

tuning frequency r0. Each damping case is treated in a separate

section. In Sec. IV B, viscous damping is considered.

B. Viscous damping

In order to better understand the behavior of the shape

of the dissipation ratio curve, the latter is first plotted versus

the frequency ratio for three different sets of parameters in

Fig. 7. All curves start from zero and tend towards zero on

the limit as r tends towards infinity. In between, the curves

exhibit a maximum value at some frequency ratio. At this

frequency, the neutralizer reaches its maximum performance

by absorbing the largest amount of energy possible for a

given set of parameters. For example, in case I, the neutral-

izer reaches its maximum value, i.e., 0.37, at a frequency

ratio of 0.61. In case II, the maximum dissipation ratio is

0.45 and occurs at r¼ 1.03. And in case III, the maximum

value is 0.49 and occurs at r¼ 1.76. The aim is to maximize

the dissipation ratio at some r¼ r0 by a proper choice of f, fand l. First, the optimal solution is obtained for a given l,

then the effect of the mass ratio is studied.

At a given mass ratio l and tuning frequency ratio r0,

the dissipation ratio becomes function of two variables,

namely, f and f, and hence the tuning parameters ft and ft are

obtained from the below equations,

492 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 8: Exact tuning of a vibration neutralizer for the reduction

@Dv

@f

����r¼r0

¼ 0;@Dv

@f

����r¼r0

¼ 0: (23)

The solution of Eq. (23) will yield either a saddle point or

a local or global optimum (maximum or minimum). The goal is

to find first the set of parameters that yields a global maximum.

The two equations in Eq. (23) reduce to after simplifications

r40lð1þ r0lÞ � f ð1þ 2r0lþ 2r2

0l2Þ ¼ 0;

r80l

2 � 2fr40lð1þ r0lÞ þ f 2ð1þ 2r0lþ 2r2

0l2Þ

� r40f

2ð1þ 2r0lþ 2r20l

2Þ ¼ 0: (24)

Solving these equations results in a unique set of positive

stiffness and damping ratios as

ft ¼r4

0lð1þ r0lÞ1þ 2r0lþ 2r2

0l2; ft ¼

r30l

2

1þ 2r0lþ 2r20l

2: (25)

To ensure that at this set of solution the dissipation ratio

passes through a maximum, the determinant of the Hessian

matrix should be positive at this solution and the second

derivative of the dissipation ratio with respect to one of the

parameters, i.e., f or f, should be negative. Let H be the

Hessian matrix of the dissipation ratio function, it is defined

and evaluated at the tuning parameters as

H ¼

@2Dv

@f 2

@2Dv

@f@f

@2Dv

@f@f

@2Dv

@f2

266664

377775

¼�ð1þ 2r0lþ 2r2

0l2Þ2

4r100 l4

0

0 �ð1þ 2r0lþ 2r20l

2Þ2

4r60l

4

266664

377775:

(26)

It is clear from Eq. (26) that, 8r0 and 8l, the determinant of His positive, and @2Dv/@f2 (or @2Dv/@f

2) is negative. This means

that the tuning parameters results in a maximum. This maxi-

mum value is global because the solution of Eq. (24) is unique,

and hence it is the sole optimum of the dissipation ratio. The

maximum value of Dv when the system is tuned is obtained by

substituting the tuning parameters into Eq. (22), which yields

Dvmax¼ 1

2: (27)

It is clear that the maximum value of Dv, when the system is

tuned to r¼ r0, is constant and depends on neither l nor r0.

Hence, once tuned, 8l and 8r0 the neutralizer will dissipate

exactly 50% of the incident energy at the tuning ratio r0.

Finally, it can be easily shown that

@Dv

@r

����r¼r0;f¼ft;f¼ft

¼ 0: (28)

This means that at the tuning parameters, if the dissipation ratio

function is plotted versus r, the peak which is shown in Fig. 7

will occur at r¼ r0. It is concluded that after tuning the neutral-

izer, it reaches its utmost performance at r ¼ r0; 8r > 0.

Hence, 8r > 0, the maximum energy that can be absorbed by

a tuned neutralizer is 50% of the incident energy and occurs at

r¼ r0. Furthermore, it can be shown that at the tuning fre-

quency, Tv¼Th¼ 0.25 and Rv¼Rh¼ 0.25. Hence, at r¼ r0,

50% of the energy will get dissipated, 25% will get transmit-

ted, and 25% will get reflected back to its source.

The dissipation ratio is shown in Fig. 8 tuned to r0¼ 0.5

and r0¼ 2 for three different values of l in each case. The

figure clearly shows that the maximum energy dissipation is

50% at r0¼ 0.5 or r0¼ 2 for all the mass ratios used.

Furthermore, the slope of the curves is zero at the tuning fre-

quencies, which yields a dome like shape for the dissipation

ratio, with the dome peak occurring at the tuning frequency.

Even though the mass ratio has no effect on the maximum

value of Dv, its effect is tremendous on the dome size, which

is clearly depicted in the figure. The dome size increases

with increasing l; this means that the neutralizer perfor-

mance improves in the vicinity of the tuning frequency r0

when the neutralizer mass is increased. Furthermore, the

neutralizer performance enhances with the increasing of the

tuning frequency for a fixed mass ratio. This is clearly illus-

trated in the figure, for example, for l¼ 0.5, the dome size is

larger when the neutralizer is tuned to r0¼ 2.0 compared to

that when it is tuned to r0¼ 0.5. In Sec. IV C, tuning of the

hysteretically damped neutralizer is considered.

C. Hysteretic damping

The same procedure used in the determination of the

tuning parameters of the viscously damped neutralizer will

be followed. First, at a given tuning frequency r0 and for a

given mass ratio l, the tuning parameters are calculated

from the below equations,

@Dh

@f

����r¼r0

¼ 0;@Dh

@d

����r¼r0

¼ 0; (29)

FIG. 7. Plot of the dissipation ratio versus the frequency ratio for three dif-

ferent parameter sets, namely, (I) l¼ 0.5, f¼ 0.05, f¼ 0.01, (II) l¼ 0.5,

f¼ 0.3, f¼ 0.2, and (III) l¼ 0.5, f¼ 2, f¼ 0.4.

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 493

Page 9: Exact tuning of a vibration neutralizer for the reduction

which reduce after simplifications to the following forms:

f 2ðd2 þ 1Þð1þ 2r0lð1þ r0lÞÞ � r80l

2 ¼ 0;

2fr40lð1þ r0lÞ þ f 2ðd2 � 1Þð1þ 2r0lð1þ r0lÞÞ� r8

0l2 ¼ 0: (30)

The tuning parameters are obtained by solving the above

equations as

ft ¼r4

0lð1þ r0lÞ1þ 2r0lþ 2r2

0l2; dt ¼

r0l1þ r0l

: (31)

To ensure that the tuning parameters yield a global maxi-

mum of the dissipation ratio, the Hessian matrix is first cal-

culated as

H ¼

@2Dh

@f 2

@2Dh

@f@d

@2Dh

@d@f

@2Dh

@d2

266664

377775

¼�ð1þ 2r0lþ 2r2

0l2Þ3

4r100 l4ð1þ r0lÞ2

�ð1þ 2r0lþ 2r20l

2Þ4r5

0l2

�ð1þ 2r0lþ 2r20l

2Þ4r5

0l2

�ð1þ r0lÞ2

4r20l

2

2666664

3777775:

(32)

The determinant of the Hessian matrix is now calculated and

simplified to the below form:

ð1þ r0lÞ2ð1þ 2r0lþ 2r20l

2Þ2

16r120 l6

: (33)

It is clear that the Hessian determinant is positive, further-

more @2Dh/@f2 (or @2Dh/@d2) is negative; this means that the

tuning parameters derived in Eq. (31) correspond indeed to a

maximum of the dissipation ratio. This maximum is global

because the solution of Eq. (30) is unique. Similar to the

case of viscous damping, it can be shown that when the neu-

tralizer is tuned, the slope of the dissipation ratio at r¼ r0 is

zero, hence its peak occurs at the tuning frequency. Finally,

the maximum dissipation ratio of a tuned neutralizer is equal

to 1/2 and occurs at the tuning frequency r0. For the sake of

completeness, the same plot shown in Fig. 8 is generated

using hysteretic damping as shown in Fig. 9. Six cases are

shown, three tuned to r0¼ 0.5 and the remaining three to

r0¼ 2.0. At each tuning frequency, three mass ratios are con-

sidered, namely, l¼ 0.1, l¼ 0.5, and l¼ 2.0. The same dis-

cussions and observations made for the viscous damping

case hold for this case and hence they will not be repeated to

reduce verbosity.

V. NUMERICAL VERIFICATION

In this section, the previously derived analytical

expressions are verified numerically by a direct resolution

of the coupled beam neutralizer system. The partial differ-

ential equation that governs the motion in a Euler-

Bernoulli beam and the equation of motion of a viscously

damped neutralizer shown in Fig. 1(a) are written as

follows:

qA €wðx; tÞ þ EI w0000ðx; tÞ þ ddðxÞ k ðwðx; tÞ � yðtÞÞðþc ð _wðx; tÞ � _yðtÞÞÞ ¼ 0;

m €yðtÞ þ k ðyðtÞ � wð0; tÞÞ þ c ð _yðtÞ � _wð0; tÞÞ ¼ 0:

(34)

For hysteretic damping, the above equations become

qA €wðx; tÞ þ EI w0000ðx; tÞ þ ddðxÞ kð1þ idÞð� wðx; tÞ � yðtÞÞð Þ ¼ 0;

m €yðtÞ þ kð1þ idÞ ðyðtÞ � wð0; tÞÞÞ ¼ 0: (35)

In Eqs. (34) and (35), the over-dot depicts the derivative

with respect to time and (0) the derivative with respect to

space x. dd(x) is the Dirac delta function. At this stage, the

following dimensionless time and space parameters are

defined as: s¼xt and X¼bx. Using this change of variables

and the dimensionless parameters defined in Sec. II, Eqs.

(34) and (35) reduce to the following forms after some

simplifications:

FIG. 8. Plot of the dissipation ratio

versus the frequency ratio tuned to

r0¼ 0.5 and r0¼ 2. For each case,

three mass ratios, namely, l¼ 0.1,

l¼ 0.5, and l¼ 2.0 are used.

494 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 10: Exact tuning of a vibration neutralizer for the reduction

€wðX;sÞþw0000ðX;sÞþddðXÞ 4f=r3 ðwðX;sÞ�

�yðsÞÞþ4f=r ð _wðX;sÞ� _yðsÞÞÞ¼ 0;

l€yðsÞþ f=r4 ðyðsÞ�wð0;sÞÞþ f=r2 ð _yðsÞ� _wð0;sÞÞ¼ 0:

(36)

€wðX; sÞ þ w0000ðX; sÞ þ ddðXÞ 4f=r3ð1þ idÞ�

� wðX; sÞ � yðsÞÞð Þ ¼ 0;

l €yðsÞ þ f=r4ð1þ idÞ ðyðsÞ � wð0; sÞÞ ¼ 0: (37)

In these equations, the over-dot and (0) denote now the deriv-

atives with respect to the dimensionless time s and space X,

respectively. To prepare these equations for the numerical

resolution, first the space X is discretized using finite ele-

ments. Hermitian beam elements are used, which results in

two degrees of freedom at each node, namely, the displace-

ment w(X, s) and the slope @w(X, s)/@X. Let me and ke be the

mass and stiffness matrices of a Hermitian beam element,

respectively. The elementary matrices associated with the

partial differential equation given in Eqs. (36) or (37) are

given below, where le is the element length

me ¼le

420

156 �22le 54 13le

�22le 4l2e �13le �3l2e

54 �13le 156 22le

13le �3l2e 22le 4l2e

2666664

3777775;

ke ¼2

l3e

6 �3le �6 �3le

�3le 2l2e 3le l2e

�6 3le 6 3le

�3le l2e 3le 2l2

e

2666664

3777775: (38)

To ensure good representation of the solution using the

discretized space, and since using the dimensionless space

parameter fixes the dimensionless wave length to 2p, 60 ele-

ments per wave length are used. This ensures that for any

material or geometric properties, the wave response is well

represented in the space dimension. Hence, the length of the

beam element used is le¼ 2p/60. The beam is assumed to

have infinite length; however, for calculation purposes, some

length should be assumed. A beam length of 120 wave

length (i.e., 240p) is used. This ensures that the length of the

beam will accommodate up to 120 cycles along its span.

This yields a total number of 7200 beam elements and 7201

nodes, thus 14 402 degrees of freedom for the beam only,

since as mentioned before, there are 2 degrees of freedom

per node. The X¼ 0 node, which is the neutralizer attach-

ment point, is selected to be the beam mid-span. The neutral-

izer equation, i.e., the second equation in Eqs. (36) or (37),

is an algebraic equation and hence does not need any spatial

discretization. It is simply added to the set of equations

(here, 14 402 equations) that results from the finite element

discretization of the beam. This increases the number of

degrees of freedom by one to Nd¼ 14 403, i.e., the absorber

motion. After assembling all elementary matrices of the

beam and adding the neutralizer equation, the total discrete

set of ordinary equations of motion of the whole system will

take the form

M €X þ K X ¼ 0 : (39)

In Eq. (39), M and K are the mass and stiffness matrices of

the whole system with sizes of (14 403� 14 403). The inci-

dent wave is generated by setting a prescribed harmonic

motion to the left end of the beam. This means that one of

the degrees of freedom is known and will become the driving

force of this equation. Here, the displacement of the first

node is assumed to be harmonic for example to have a

motion of the form of sin(s). Now, the forced set of equa-

tions of motion of the system become

�M €X þ �K X ¼ F : (40)

In Eq. (40), �M is obtained from M by simply setting to zero

all values in its first line. �K is obtained from K by setting to

zero all values of its first line except for the first value, which

is replaced by 1. Now, the forcing vector F has zero entries

except for its first value, which is the prescribed motion of

the first node, i.e., sin(s). Finally, Eq. (40), which is a second

order ordinary equation in time, is reduced to first order by

the classical change of variables Z ¼ ½X _X�T . The reduced

FIG. 9. Plot of the dissipation ratio

versus the frequency ratio tuned to

r0¼ 0.5 and r0¼ 2. For each case,

three mass ratios, namely, l¼ 0.1,

l¼ 0.5, and l¼ 2.0.

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 495

Page 11: Exact tuning of a vibration neutralizer for the reduction

equation takes the form _Z ¼ AZ þ b where the (2Nd� 2Nd)

matrix A and (2Nd) vector b are given below

A ¼0ðNd�NdÞ IðNd�NdÞ

�M�1K 0ðNd�NdÞ

" #;

b ¼ cos ðsÞ 0 � � � 0 �sin s 0 � � � 0½ �: (41)

A is a (2Nd� 2Nd) matrix and vector b is a zero vector with

2Nd dimension except for the first entry, which is cos s and the

Ndþ 1 entry, which is �sin s. The set of ordinary differential

equations that will result from the finite element discretization

are solved using the Crank–Nicolson technique, which is an

implicit finite difference method. This technique is used to solve

the reduced equation _Z ¼ AZ þ b. Based on Crank–Nicolson,

the state vector Ziþ1 at the next time step is calculated from the

state vector at the previous time step Zi as

Ziþ1¼ I� dt

2A

�1dt

2biþ1þAZiþbið ÞþZi

: (42)

Time stepping is essential here in order to avoid misrep-

resentation of the solution in the time domain. Here, the

dimensionless time s fixes the dimensionless period of

motion to 2p for any material or geometric properties. The

time step is chosen to equal dt¼ 2p/60, this ensures that

each steady state cycle at a given node is described with 60

points. Now, test cases of the transmissibility ratios are con-

sidered to validate the analytical findings. In Fig. 10(a), the

undamped neutralizer tuned to r0¼ 1 is considered with

l¼ 3 and its corresponding transmission ratio plotted using

the analytical expression in Eq. (10). The tuning stiffness

ratio is calculated from Eq. (15) and is equal to f¼ 0.75. The

figure shows, as expected, that the wave will be completely

reflected at the tuning frequency and the neutralizer perfor-

mance deteriorates as the frequency moves away from r0. For

the same case, the transmission ratio was calculated numeri-

cally by solving either Eqs. (36) or (37) and setting the damp-

ing parameter to zero. The numerical solution was obtained at

nine different frequency ratios including the tuning frequency

ratio r0¼ 1. Each data point is obtained as follows: the beam is

given a prescribed harmonic motion at its left end. This will

initiate a wave that will propagate through the beam and will

encounter an obstacle at the attachment point of the neutral-

izer. Part of the wave will get reflected and part will get trans-

mitted. After ensuring that the steady state regime is achieved,

the amplitudes of the incident and transmitted waves are mea-

sured and the transmission ratio is obtained. The numerically

calculated transmission ratios are illustrated by the solid points

in Fig. 10(a). The impeccable match between the numerical

and analytical solution is clearly shown in the figure and stems

from the fact that the analytical solution obtained is exact since

no assumptions were made during the derivation process.

Similar validations are shown in Figs. 10(b) and 10(c), where

the neutralizer is hysteretically damped in Fig. 10(b) and tuned

to r0¼ 1 with h0¼ 0.1 and viscously damped in Fig. 10(c) and

tuned to r0¼ 1 with h0¼ 0.2. The tuning stiffness and damping

ratios used to plot these curves are obtained from the analyti-

cally derived expression given in Appendix B. For each case,

nine data points are obtained numerically by solving either

Eqs. (36) or (37). Again, for these two cases, an almost perfect

match between the analytical solution and numerical results is

achieved.

FIG. 10. Plots of the transmission ratio of a vibration neutralizer tuned to r0¼ 1 with a mass ratio of l¼ 3.0. (a) undamped neutralizer with tuning stiffness

ratio of f¼ 0.75. (b) hysteretically damped neutralizer with h0¼ 0.1 with tuning parameters f¼ 0.72 and d¼ 0.36. (c) viscously damped neutralizer, with

h0¼ 0.2 with tuning parameters f¼ 0.76 and f¼ 0.51.

496 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 12: Exact tuning of a vibration neutralizer for the reduction

Finally, the analytical expressions of the tuning parame-

ters pertaining to the dissipation ratio part of the analysis are

validated numerically. Four cases are considered, two for

each damping type. The dissipation ratio with viscous damp-

ing Dv is plotted in Fig. 11(a) for two mass ratios, namely,

l¼ 0.2 and l¼ 1. In this plot, the neutralizer is tuned to

r0¼ 1 using the analytical expressions in Eq. (25), which

yields the following tuning parameters (f¼ 0.16, f¼ 0.027)

for l¼ 0.2, and (f¼ 0.4, f¼ 0.2) for l¼ 1. In the same plot,

nine data points for each dissipation ratio curve are calcu-

lated numerically by solving Eq. (36). For each data point,

the dissipation ratio is calculated as follows: first the beam is

given a prescribed harmonic motion at its left boundary to

initiate a travelling wave. After reaching steady state condi-

tions, the amplitude kZvk of the relative neutralizer motion is

obtained along with that of the incident wave, i.e., kWik. The

dissipation ratio can now be calculated as follows:

Dv ¼ ð2f=rÞðkZvk2=kWik2Þ, which is deduced from Eq. (22).

For all cases considered, almost a perfect match between the

analytical and numerical results is achieved. The same vali-

dation is done for the analytical results derived for the

hysteretic damping case. The dissipation ratio is first

plotted in Fig. 11(b) for the two mass ratios l¼ 0.2 and

l¼ 1 tuned to r0¼ 1. The tuning parameters are calcu-

lated using the analytical expression in Eq. (31), which

yields (f¼ 0.16, d¼ 0.167) for l¼ 0.2 and (f¼ 0.4,

f¼ 0.5) for l¼ 1. Nine data points for each curve are

obtained numerically for this case from the resolution of

Eq. (37), as depicted by the solid points in the figure. The

dissipation ratio is calculated numerically for each data

point from a modified version of Eq. (22) as follows:

Dh ¼ ð2f d=r3ÞðkZhk2=kWik2Þ. Again, here, the figure

shows a great match between the numerical and analyti-

cal findings.

VI. CONCLUSION

In this paper, the use of vibration neutralizers for the

reduction of flexural waves in beam elements is investigated.

The neutralizer is attached to the beam and its parameters are

properly chosen to maximize either the energy reflected back to

its source or the energy dissipated in the neutralizer. The

undamped neutralizer is considered first and for a given mass

ratio, the neutralizer was tuned to a frequency at which energy

was completely reflected back to its source. The tuning stiffness

ratio is obtained in an analytical closed form along with the

neutralizer bandwidth. Then, the damped neutralizer is consid-

ered and tuned to a frequency by forcing the minimum value of

the transmissibility ratio to occur at that frequency. The tuning

stiffness and damping parameters are obtained analytically for

both the viscous and hysteretic damping cases. For both the

undamped and damped neutralizer, the increasing of its mass is

shown to widen its bandwidth. As for energy dissipation, the

bare fact that a neutralizer is attached to a beam will lead to

some energy dissipation in the damping element. However, this

energy can be maximized and can reach at most 50% of the

energy propagating in the beam at a given tuning frequency.

The tuning stiffness and damping ratios are obtained in a closed

form in terms of the system mass and tuning frequency ratios.

It is shown that, even though the neutralizer mass has no effect

on the maximum energy dissipated at the tuning frequency, its

effect is tremendous on the neutralizer performance in the

vicinity of the tuning frequency. Finally, the analytically

derived expressions are validated numerically where almost a

perfect match is achieved between the two solutions.

APPENDIX A

In this appendix, the dynamic stiffness of a damped neu-

tralizer is derived. The schematic of a damped neutralizer is

FIG. 11. Plots of the dissipation ratio of a damped vibration neutralizer tuned to r0¼ 1 for two different mass ratios, namely, l¼ 0.2 and l¼ 1.0. (a) viscously

damped neutralizer with the following tuning parameters (l¼ 0.2, f¼ 0.16, f¼ 0.027) and (l¼ 1, f¼ 0.4, f¼ 0.2), (b) hysteretically damped neutralizer with

the following tuning parameters (l¼ 0.2, f¼ 0.16, d¼ 0.167), and (l¼ 1, f¼ 0.4, d¼ 0.5).

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 497

Page 13: Exact tuning of a vibration neutralizer for the reduction

shown in Figs. 12(a) and 12(b), with viscous and hysteretic

damping, respectively. The free end of the neutralizer is

depicted with point D. Let u(t) be the motion of D, F(t) the

force applied at D, and y(t) the motion of the neutralizer

mass. The force F(t) is written in terms of u(t) and y(t) as

FðtÞ ¼ kðuðtÞ � yðtÞÞ þ cð _uðtÞ � _yðtÞÞ for the viscous damp-

ing case and FðtÞ ¼ kð1þ idÞðuðtÞ � yðtÞÞ for the hysteretic

case. The dynamic stiffness Kd at point D is defined as the

ratio between the force and displacement assuming harmonic

motion, i.e., uðtÞ ¼ Ueixt and yðtÞ ¼ Yeixt. Hence, Kd¼F(t)/u(t) and after substituting the expression of F(t) and u(t), the

dynamic stiffness of the viscous and hysteretic damping

cases reduces to the following forms, respectively:

Kdv ¼ðk þ icxÞðU � YÞ

U; Kdh

¼ kð1þ idÞðU � YÞU

:

(A1)

The motions u(t) and y(t) are related through the equation of

motion of the neutralizer. Assuming a prescribed harmonic

support motion uðtÞ ¼ Ueixt and solving for y(t) yields the

following solution for the viscous and hysteretic damping

cases, respectively:

Y ¼ k þ icxk þ icx� mx2

U; Y ¼ kð1þ idÞkð1þ idÞ � mx2

U:

(A2)

Finally, substituting the expressions of Y in Eq. (A2) into Eq.

(A1) yields, after simplifications, the below dynamic stiff-

ness of the viscously and hysteretically damped neutralizer,

respectively

Kdv ¼1

1

k þ icx� 1

mx2

; Kdh¼ 1

1

kð1þ idÞ �1

mx2

:

(A3)

APPENDIX B

In this appendix, the tuning neutralizer parameters are

obtained analytically for both viscous and hysteretic damp-

ing with the aim of maximizing energy reflection at some

tuning frequency. For a given mass ratio of the system l,

tuning the system to r0 means forcing the transmission ratio

curve to pass through a minimum value at r¼ r0 and to have

a minimum transmission ratio h0 at this point. This can be

achieved if the stiffness and damping ratios are chosen such

that the below two equations are satisfied

@T

@r

����r¼r0

¼ 0; T

����r¼r0

¼ h0: (B1)

In order to simplify these equations, the transmissibility ratio

T is first written as T¼N/D since it is the ratio of two poly-

nomial equations as given in Eq. (9). Now, the equations in

Eq. (B1) are rewritten as

N00D0 � N0D00D2

0

¼ 0;N0

D0

¼ h0: (B2)

Here, prime denotes the derivative of the function with

respect to r and the subscript zero means that the correspond-

ing function is evaluated at r¼ r0. The equations in Eq. (B2)

are further simplified to the following forms:

N00D00¼ h0;

N0

D0

¼ h0: (B3)

And finally they are written as follows:

N00 � h0D00 ¼ 0; N0 � h0D0 ¼ 0: (B4)

At this stage, two cases are considered, namely, viscous

and hysteretic damping. The tuning parameters, i.e., the

stiffness and damping parameters, are obtained in closed

form for each case by solving the system of equations in

Eq. (B4).

1. Viscous damping

The expressions for N0 and D0 are obtained from Eq. (9)

as follows:

N0 ¼ ðf þ fr0l� r40lÞ

2 þ r40f

2ð1þ r0lÞ2;D0 ¼ ðf þ fr0l� r3

0fl� r40lÞ

2 þ r20ðr0fþ flþ r2

0flÞ2:

(B5)

The two equations in Eq. (B4) are rewritten as

a0 þ a1 f þ a2 f 2 þ a3 fþ a4 f2 ¼ 0;

b0 þ b1 f þ b2 f 2 þ b3 fþ b4 f2 ¼ 0: (B6)

Where

a0 ¼ ð1� h0Þr80l

2 ;

a1 ¼ 2ðh0� 1Þr40lð1þ r0lÞ ;

a2 ¼ ð1þ r0lÞ2� h0ð1þ 2r0lð1þ r0lÞÞ ;a3 ¼�2h0r7

0l2 ;

a4 ¼ r40ðð1þ r0lÞ2� h0ð1þ 2r0lð1þ r0lÞÞÞ ;

b0 ¼ 8ð1� h0Þr70l

2 ;

b1 ¼ 2ðh0� 1Þr30lð4þ 5r0lÞ ;

b2 ¼ 2lð1� h0þ ð1� 2h0Þr0lÞ ;b3 ¼�14h0r6

0l2 ;

b4 ¼ 2r30ðð1þ r0lÞð2þ 3r0lÞ � h0ð2þ r0lð5þ 6r0lÞÞÞ:

(B7)

FIG. 12. Schematic of a vibration neutralizer (a) with viscous damping, (b)

with hysteretic damping.

498 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa

Page 14: Exact tuning of a vibration neutralizer for the reduction

Now, the stiffness ratio is written in terms of f using the

equations in Eq. (B6) as follows:

f ¼ a0b2 � a2b0 þ ða3b2 � a2b3Þfþ ða4b2 � a2b4Þf2

a2b1 � a1b2

:

(B8)

The stiffness ratio is eliminated from the first equation of

Eq. (B6) using Eq. (B8). This yields the below fourth order

polynomial in f

a0 þ a1 fþ a2 f2 þ a3 f3 þ a4 f4 ¼ 0: (B9)

Where

a0 ¼ a2ða22b2

0 � a1a2b0b1 þ b2ða21b0 � a0a1b1 þ a2

0b2Þþ a0a2ðb2

1 � 2b0b2ÞÞ ;a1 ¼ a2ða2a3ðb2

1 � 2b0b2Þ þ 2a22b0b3

� a2ða1b1 þ 2a0b2Þb3

þ b2ð�a1a3b1 þ 2a0a3b2 þ a21b3ÞÞ ;

a2 ¼ a2ðb2ð�a1a4b1 þ ða23 þ 2a0a4Þb2 þ a2

1b4Þþ a2

2ðb23 þ 2b0b4Þ þ a2ða4ðb2

1 � 2b0b2Þ;�2a3b2b3 � a1b1b4 � 2a0b2b4ÞÞ ;

a3 ¼ 2a2ða3b2 � a2b3Þða4b2 � a2b4Þ ;a4 ¼ a2ða4b2 � a2b4Þ2: (B10)

The tuned damping ratio is obtained analytically by solving

Eq. (B9). The roots of a fourth order polynomial can be

obtained using Ferrari’s method. The positive root of this

polynomial is the tuning damping ratio, which can be shown

to take the form

ft ¼ �a3

4a4

� Sþ 1

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiq

S� 4S2 � 2p

r: (B11)

Where

S ¼ 1

2

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi1

3a4

Qþ D0

Q

� �� 2

3p

s;

Q ¼D1 þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiD2

1 � 4D30

q2

!1=3

;

q ¼ a33 � 4a4a3a2 þ 8a2

4a1

8a34

;

p ¼ 8a4a2 � 3a23

8a24

;

D1 ¼ 2a32 � 9a3a2a1 þ 27a2

3a0 þ 27a4a21 � 72a4a2a0 ;

D0 ¼ a22 � 3a3a1 þ 12a4a0: (B12)

This concludes this section of the appendix. Once the tuning

damping ratio is calculated from Eq. (B11), the tuning stiff-

ness ratio can be calculated from Eq. (B8).

2. Hysteretic damping

The expressions for N0 and D0 for the hysteretic damp-

ing case are obtained from Eq. (9) as follows:

N0 ¼ f 2d2ð1þ rlÞ2 þ ðr4l� rfl� f Þ2;D0 ¼ f 2ðdþ rlþ rdlÞ2 þ ðr4lþ frdl� frl� f Þ2:

(B13)

The two equations in Eq. (B4) are then rewritten as

a0 þ a1 þ a2dð Þ f þ a3 1þ d2ð Þ f 2 ¼ 0;

b0 þ b1 þ b2dð Þ f þ b3 1þ d2ð Þ f 2 ¼ 0; (B14)

where

a0 ¼ ð1� h0Þr80l

2;

a1 ¼ 2ðh0 � 1Þr40lð1þ r0lÞ;

a2 ¼ �2h0r50l

2;

a3 ¼ ð1þ r0lÞ2 � h0ð1þ 2r0lð1þ r0lÞÞ;b0 ¼ 8ð1� h0Þr7

0l2;

b1 ¼ 2ðh0 � 1Þr30lð4þ 5r0lÞ;

b2 ¼ �10h0r40l

2;

b3 ¼ 2lð1� h0 þ ð1� 2h0Þr0lÞ: (B15)

The stiffness ratio is now written in terms of d using the

equations in Eq. (B14) as follows:

f ¼ a0b3 � a3b0

a3b1 � a1b3 þ a3b2 � a2b3ð Þd : (B16)

Substituting the above expression of f in the first equation of

Eq. (B14) yields the below second order polynomial in d2

a0 þ a1 dþ a2 d2 ¼ 0; (B17)

where

a0 ¼ ða3b1� a1b3Þ2a0þ ða0b3� a3b0Þða3b1� a1b3Þa1

þ ða0b3� a3b0Þ2a3 ;

a1 ¼ 2ða3b1� a1b3Þða3b2� a2b3Þa0þ ða0b3� a3b0Þ� ða3b2� a2b3Þa1þ ða0b3� a3b0Þða3b1� a1b3Þa2 ;

a2 ¼ ða3b2� a2b3Þ2a0þ ða0b3� a3b0Þða3b2� a2b3Þa2

þ ða0b3� a3b0Þ2a3: (B18)

The tuning damping ratio is the positive root of the second

order polynomial of Eq. (B17), which takes the form

dt ¼�a1 þ

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffia2

1 � 4a2a0

p2a2

: (B19)

Finally, for a given frequency ratio r0, minimum transmis-

sion ratio h0 and mass ratio l, the tuning damping ratio is

first calculated form Eq. (B19), then the tuning stiffness ratio

is calculated from Eq. (B16). It is important to note that sim-

ilar constant names are used in both Appendixes. These

J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa 499

Page 15: Exact tuning of a vibration neutralizer for the reduction

constants are defined locally in each section, and these defi-

nitions hold in the sections where they are defined only.

1H. Frahm, “Device for damping vibrations of bodies,” U.S. patent 989,958

(1911).2J. E. Brock, “A note on the damped vibration absorber,” in Transactionsof the American Society of Mechanical Engineers (ASME, New York,

1946), pp. A-284.3J. P. Den Hartog, Mechanical Vibrations (McGraw Hill, New York,

1940), pp. 115–129.4O. Nishihar and T. Asami, “Closed-form solutions of the exact optimiza-

tions of dynamic vibration absorbers (minimizations of the maximum

amplitude magnification factors),” J. Vib. Acoust. 124(4), 576–582 (2002).5B. R. Mace, “Wave reflection and transmission in beams,” J. Sound Vib.

97(2), 237–246 (1984).6L. Cremer, M. Heckl, and B. A. T. Petersson, Structure-Borne Sound, 3rd

ed. (Springer, Berlin, 2005).7M. J. Brennan, “Control of flexural waves on a beam using a tunable

vibration neutralizer,” J. Sound Vib. 222(3), 389–407 (1999).8H. M. El-Khatib, B. R. Mace, and M. J. Brennan, “Suppression of bending

waves in a beam using a tuned vibration absorber,” J. Sound Vib.

288(4–5), 1157–1175 (2005).9H. Salleh and M. J. Brennan, “Control of flexural waves on a beam using a

vibration neutraliser: Effects of different attachment configurations,”

J. Sound Vib. 303(3–5), 501–514 (2007).

10D. J. Thompson, “A continuous damped vibration absorber to reduce

broad-band wave propagation in beams,” J. Sound Vib. 311(3–5),

824–842 (2008).11Y. Gao, M. J. Brennan, and F. Sui, “Control of flexural waves on a beam

using distributed vibration neutralisers,” J. Sound Vib. 330(12),

2758–2771 (2011).12Y. Liu, D. Yu, L. Li, H. Zhao, J. Wen, and X. Wen, “Design guidelines for

flexural wave attenuation of slender beams with local resonators,” Phys.

Lett. A 362(5–6), 344–347 (2007).13Z. Shuguang, N. Tianxin, W. Xudong, and F. Jialu, “Studies of band

gaps in flexural vibrations of a locally resonant beam with novel

multi-oscillator configuration,” J. Vib. Control 23(10), 1663–1674

(2017).14Y. Xiao, J. Wen, G. Wang, and X. Wen, “Theoretical and experimental

study of locally resonant and Bragg band gaps in flexural beams carrying

periodic arrays of beam-like resonators,” J. Vib. Acoust. 135(4), 041006

(2013).15Y. Xiao, J. Wen, D. Yu, and X. Wen, “Flexural wave propagation in

beams with periodically attached vibration absorbers: Band-gap behavior

and band formation mechanisms,” J. Sound Vib. 332(4), 867–893

(2013).16C. Yang and L. Cheng, “Suppression of bending waves in a beam using

resonators with different separation lengths,” J. Acoust. Soc. Am. 139(5),

2361–2371 (2016).17H. Peter and D. Anirvan, Vibrations and Waves in Continuous Mechanical

Systems, 1st ed. (Wiley, New York, 2007).

500 J. Acoust. Soc. Am. 146 (1), July 2019 Jimmy S. Issa