Skip to main content

Inerter-Based Isolation System

  • Chapter
  • First Online:
Book cover Inerter and Its Application in Vibration Control Systems
  • 668 Accesses

Abstract

This chapter is concerned with the problem of analysis and optimization of the inerter-based isolators based on a “uni-axial” single-degree-of-freedom isolation system. In the first part, in order to gain an in-depth understanding of inerter from the prospective of vibration, the frequency responses of both parallel-connected and series-connected inerters are analyzed. In the second part, three other inerter-based isolators are introduced and the tuning procedures in both the \(H_\infty \) optimization and the \(H_2\) optimization are proposed in an analytical manner. The achieved \(H_2\) and \(H_\infty \) performance of the inerter-based isolators is superior to that achieved by the traditional dynamic vibration absorber (DVA) when the same inertance-to-mass (or mass) ratio is considered. Moreover, the inerter-based isolators have two unique properties, which are more attractive than the traditional DVA: first, the inertance-to-mass ratio of the inerter-based isolators can easily be larger than the mass ratio of the traditional DVA without increasing the physical mass of the whole system; second, there is no need to mount an additional mass on the object to be isolated.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 129.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 169.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

References

  • Asami, T., Wakasono, T., Kameoka, K., Hasegawa, M., & Sekiguchi, H. (1991). Optimum design of dynamic absorbers for a system subjected to random excitation. JSME International Journal Series III, 34(2), 218–226.

    Google Scholar 

  • Carrella, A., Brennan, M. J., Waters, T. P., & Lopes, V, Jr. (2012). Force and displacment transimissibility of a nonlinear isolator with high-static-low-dynamic-stiffness. International Journal of Mechanical Sciences, 55(1), 22–29.

    Article  Google Scholar 

  • Chen, M. Z. Q., Papageorgiou, C., Scheibe, F., Wang, F. C., & Smith, M. C. (2009). The missing mechanical circuit element. IEEE Circuits and Systems Magazine, 9(1), 10–26.

    Article  Google Scholar 

  • Chen, M. Z. Q., Hu, Y., Li, C., & Chen, G. (2015). Performance benefits of using inerter in semiactive suspensions. IEEE Transactions on Control System Technology, 23(4), 1571–1577.

    Article  Google Scholar 

  • Chen, M. Z. Q., Hu, Y., Huang, L., & Chen, G. (2014). Influence of inerter on natural frequencies of vibration systems. Journal of Sound and Vibration, 333(7), 1874–1887.

    Article  Google Scholar 

  • Cheung, Y. L., & Wong, W. O. (2011a). H-infinity optimization of a variant design of the dynamic vibration absorber-Revisited and new results. Journal of Sound and Vibration, 330(16), 3901–3912.

    Article  Google Scholar 

  • Cheung, Y. L., & Wong, W. O. (2011b). \({H_{2}}\) optimization of a non-traditional dynamic vibration abosorber for vibration control of structures under random force excitation. Journal of Sound and Vibration, 330(6), 1039–1044.

    Article  Google Scholar 

  • Den Hartog, J. P. (1985). Mechanical Vibrations. New York: Dover Publications, INC.

    MATH  Google Scholar 

  • Doyle, J. C., Francis, B. A., & Tannenbaum, A. R., et al. (1992). Feedback Control Theory. Oxford: Maxwell Macmillan Int.

    Google Scholar 

  • Dylejko, P. G., & MacGillivray, I. R. (2014). On the concept of a transmission absorber to suppress internal resonance. Journal of Sound and Vibration, 333, 2719–2734.

    Article  Google Scholar 

  • Hu, Y., Chen, M. Z. Q., & Shu, Z. (2014). Passive vehicle suspensions employing inerters with multiple performance requirements. Journal of Sound and Vibration, 333(8), 2212–2225.

    Article  Google Scholar 

  • Hu, Y., Chen, M. Z. Q., Shu, Z., & Huang, L. (2014). Vibration analysis for isolation system with inerter. In Proceedings of the 33rd Chinese Control Conference (pp. 6687–6692). China: Nanjing.

    Google Scholar 

  • Inman, D. J. (2008). Engineering Vibration (3rd ed.). Upper Saddle River: Prentice-Hall Inc.

    Google Scholar 

  • Lazar, I. F., Neild, S. A., & Wagg, D. J. (2014). Using an inerter-based device for structural vibration suppression. Earthquake Engineering and Structure Dynamics, 43(8), 1129–1147.

    Article  Google Scholar 

  • Marian, L., & Giaralis, A. (2014). Optimal design of a novel tuned mass-damper-inerter (TMDI) passive vibration control configuration for stochastically support-excited structural systems. Probabilistic Engineering Mechanics, 38, 156–164.

    Article  Google Scholar 

  • Nishihara, O., & Asami, T. (2002). Closed-form solutions to the exact optimizations of dynamic vibration absorbers (minimizations of the maximum amplitude magnification factors). Journal of Vibration and Acoustics, 124(4), 576–582.

    Article  Google Scholar 

  • Ren, M. Z. (2001). A variant design of the dynamic vibration absorber. Journal of Sound and Vibration, 245(4), 762–770.

    Article  Google Scholar 

  • Rivin, E. I. (2003). Passive Vibration Isolation. New York: ASME Press.

    Book  Google Scholar 

  • Piersol, A. G., & Paez, T. L. (2010). Harris’ Shock and Vibration Handbook (6th ed.). New York: McGraw-Hill.

    Google Scholar 

  • Scheibe, F., & Smith, M. C. (2009). Analytical solutions for optimal ride comfort and tyre grip for passive vehicle suspensions. Vehile System Dynamics, 47(10), 1229–1252.

    Article  Google Scholar 

  • Smith, M. C. (2002). Synthesis of mechanical networks: The inerter. IEEE Transaction on Automatic Control, 47(1), 1648–1662.

    Article  MathSciNet  Google Scholar 

  • Smith, M. C., & Wang, F. C. (2004). Performance benefits in passive vehicle suspensions employing inerters. Vehicle System Dynamics, 42(4), 235–257.

    Article  Google Scholar 

  • Wang, F.-C., & Chan, H. A. (2011). Vehicle suspensions with a mechatronic network strut. Vehicle System Dynamics, 49(5), 811–830.

    Article  Google Scholar 

  • Wang, F.-C., Hsieh, M.-R., & Chen, H.-J. (2012). Stability and performance analysis of a full-train system with inerters. Vehicle System Dynamics, 50(4), 545–571.

    Article  Google Scholar 

  • Wang, K., Chen, M. Z. Q., & Hu, Y. (2014). Synthesis of biquadratic impedances with at most four passive elements. Journal of the Franklin Institute, 351(3), 1251–1267.

    Article  Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Michael Z. Q. Chen .

Appendix

Appendix

Proof of Proposition 3.1

Observing Fig. 3.6, it is shown that the curve horizontally passing through P indicates the optimal damping. This optimal damping can be obtained by solving the following equation:

$$\begin{aligned} \left. \frac{\mathrm {\partial } \mu ^2}{\mathrm {\partial } q^2}\right| _{q=q_P}=0. \end{aligned}$$
(3.46)

Denote \(\mu =\sqrt{\frac{n}{m}}\), where \(n=\delta ^2q^2+4(1-\delta q^2)^2\zeta ^2\), \(m=\delta ^2(1-q^2)^2q^2+4(1-(1+\delta )q^2)^2\zeta ^2\). Equation (3.46) can be written in another form as

$$ n^\prime m-m^\prime n=0, $$

where \(n^\prime =\mathrm {\partial } n/ \mathrm {\partial } q^2\) and \(m^\prime =\mathrm {\partial } m/ \mathrm {\partial } q^2\). For the invariant point P,

$$ \frac{n}{m}=\frac{1}{(1-q^2)^2}=\frac{(1-\delta q^2)^2}{(1-(1+\delta )q^2)^2}, $$

therefore,

$$ (1-q^2)^2 n^\prime -m^\prime =0. $$

Since

$$ n^\prime =-8(1-\delta q^2)\delta \zeta ^2+\delta ^2, $$
$$ m^\prime =-8(1-(1+\delta )q^2)(\delta +1)\zeta ^2+\delta ^2(1-q^2)(1-3q^2), $$

after substituting \(q_P\) into (3.14), one obtains

$$ \zeta _{opt}=\frac{1}{2}\sqrt{\delta (1+\delta -\sqrt{1+\delta ^2})}. $$

Proof of Proposition 3.2

Denote

$$ A=4\lambda ^2(1-\delta q^2)^2q^2,~B=(1-\delta (1+\lambda )q^2)^2, $$
$$ C=4\lambda ^2(1-(1+\delta )q^2)^2q^2,~D=(1-(\delta +1+\delta \lambda )q^2+\delta \lambda q^4)^2. $$

Then, \(\mu \) in (3.17) can be rewritten as

$$\begin{aligned} \mu =\sqrt{\frac{A\zeta ^2+B}{C\zeta ^2+D}}. \end{aligned}$$
(3.47)

To find the invariant points which are independent of damping, it requires

$$\begin{aligned} \frac{A}{C}=\frac{B}{D}, \end{aligned}$$

that is,

$$ \frac{1-\delta q^2}{1-(1+\delta )q^2}=\pm \frac{1-\delta (1+\lambda )q^2}{1-(\delta +1+\delta \lambda )q^2+\delta \lambda q^4}. $$

With the plus sign, after cross-multiplication, one obtains \( \delta ^2 \lambda q^6=0, \) which leads to the trivial solution \(q=0\). With the minus sign, after simple calculation, one obtains

$$\begin{aligned} \delta ^2\lambda q^6-2\delta (\lambda +\delta +1+\delta \lambda )q^4+2(2\delta +1+\delta \lambda )q^2-2=0, \end{aligned}$$
(3.48)

which is a cubic form in \(q^2\). Therefore, there are three invariant points for the configuration C3.

Denoting these three invariant points as P, Q, and R (\(q_P<q_Q<q_R\)), separately, one obtains

$$\begin{aligned} q^2_P+q^2_Q+q^2_R= & {} \frac{2}{\delta \lambda }(\lambda +\delta +1+\lambda \delta ),\end{aligned}$$
(3.49)
$$\begin{aligned} q^2_Pq^2_Qq^2_R= & {} \frac{2}{\delta ^2\lambda },\end{aligned}$$
(3.50)
$$\begin{aligned} q^2_Pq^2_Q+q^2_Pq^2_R+q^2_Qq^2_R= & {} \frac{2}{\delta ^2\lambda }(2\delta +1+\delta \lambda ). \end{aligned}$$
(3.51)

Since at points P and Q, the values of \(\mu \) are independent of \(\zeta \), then in the case of \(\zeta =\infty \), one obtains

$$ \left| \frac{1-\delta q^2_P}{1-(1+\delta )q^2_P}\right| =\left| \frac{1-\delta q^2_Q}{1-(1+\delta )q^2_Q}\right| . $$

It can be checked that

$$ \frac{1-\delta q^2_P}{1-(1+\delta )q^2_P}>0,~\frac{1-\delta q^2_Q}{1-(1+\delta )q^2_Q}<0. $$

Then, one obtains

$$ \frac{1-\delta q^2_P}{1-(1+\delta )q^2_P}=-\frac{1-\delta q^2_Q}{1-(1+\delta )q^2_Q}. $$

After cross-multiplication and simplification, one obtains

$$\begin{aligned} 2\delta (1+\delta )q^2_Pq^2_Q-(q^2_P+q^2_Q)(1+2\delta )+2=0. \end{aligned}$$
(3.52)

Substituting (3.50) and (3.51) into (3.52), one can obtain a quadratic equation with respect to \(q^2_R\) as

$$\begin{aligned} \delta \lambda (1+2\delta )q^4_R-2(\lambda +2\delta \lambda +3\delta +2\delta ^2+1+2\lambda \delta ^2)q^2_R+4(1+\delta )=0. \end{aligned}$$
(3.53)

Note that \(q_R\) is the same solution as both (3.48) and (3.53) for the same \(\delta \) and \(\lambda \). Solving \(\lambda \) from (3.48) and (3.53), separately, one obtains

$$\begin{aligned} \lambda= & {} \frac{2(q^4_R\delta (1+\delta )-(1+2\delta )q^2_R+1)}{\delta q^2_R(q^4_R\delta -2(\delta +1)q^2_R+2)},\end{aligned}$$
(3.54)
$$\begin{aligned} \lambda= & {} \frac{2((1+2\delta )(1+\delta )q^2_R-2(1+\delta ))}{q^2_R(\delta (1+2\delta )q^2_R-2(1+2\delta +2\delta ^2))}. \end{aligned}$$
(3.55)

Equating the solutions and simplifying the results, one obtains

$$\begin{aligned} \delta q^4_R-(2+3\delta )q^2_R+2=0. \end{aligned}$$
(3.56)

Then, one obtains \(q^2_R\) as shown in (3.18).

From (3.18), it is easy to show that \(q^2_R\ge 3\), which is relatively large compared with the natural frequency. This can explain why only invariant points P and Q are involved in the \(H_\infty \) tuning of C3.

In this way, the optimal \(\lambda \) can be obtained by substituting \(q^2_R\) in (3.18) into (3.54) or (3.55). After obtaining \(\lambda \), all the three invariant points can be obtained by solving

$$ q^4-\left( \frac{2}{\delta \lambda }(1+\lambda +\delta +\lambda \delta )-q^2_R\right) q^2+\frac{2}{\delta ^2\lambda q^2_R}=0, $$

which is obtained from (3.50) and (3.51).

The procedure of calculating the optimal damping ratio \(\zeta \) is similar to the procedure in appendix, where the optimal \(\zeta \) makes the gradients at invariant points P and Q zero. After calculation and simplification, one obtains (3.21). Taking an average of \(\zeta ^2_P\) and \(\zeta ^2_Q\), one obtains the optimal \(\zeta _{opt}\) as in (3.20).

Proof of Proposition 3.3

Denote

$$ A=4(1-\delta (1+\lambda )q^2)^2, B=\delta ^2q^2,$$
$$ C=4(1-(1+\delta +\delta \lambda )q^2+\delta \lambda q^4)^2, D=\delta ^2(1-q^2)^2q^2, $$

and \(\mu \) in (3.23) can be rewritten as

$$\begin{aligned} \mu =\sqrt{\frac{A\zeta ^2+B}{C\zeta ^2+D}}. \end{aligned}$$
(3.57)

To find the invariant points which are independent of damping, it requires

$$ \frac{A}{C}=\frac{B}{D}, $$

that is,

$$ \frac{1-\delta (1+\lambda )q^2}{1-(1+\delta +\delta \lambda )q^2+\delta \lambda q^4}=\pm \frac{1}{1-q^2}. $$

Again, with the plus sign, one obtains the trivial solution zero, and with the minus sign, one obtains

$$\begin{aligned} \delta (1+2\lambda )q^4-2(1+\delta +\delta \lambda )q^2+2=0. \end{aligned}$$
(3.58)

Then, one obtains the two invariant points P and Q (\(q_P<q_Q\)) as

$$\begin{aligned} q^2_{P,Q}=\frac{1+\delta +\delta \lambda \pm \sqrt{(1+\delta +\delta \lambda )^2-2\delta (1+2\lambda )}}{\delta (1+2\lambda )}. \end{aligned}$$
(3.59)

Letting the ordinates at invariant points P and Q equal, one has

$$ \left| \frac{1}{1-q^2_P}\right| =\left| \frac{1}{1-q^2_Q}\right| . $$

It can be checked that \(\frac{1}{1-q^2_P}>0\) and \(\frac{1}{1-q^2_Q}<0\). Then, one obtains

$$ \frac{1}{1-q^2_P}=-\frac{1}{1-q^2_Q}. $$

After cross-multiplication and simplification, one has

$$\begin{aligned} q^2_P+q^2_Q=2. \end{aligned}$$
(3.60)

Considering (3.58), one obtains

$$ \frac{2(1+\delta +\delta \lambda )}{\delta (1+2\lambda )}=2, $$

which leads to (3.24).

Similar to the method in appendix, the optimal \(\zeta \) can be obtained by making \(\mu \) to have zero gradients at invariant points P and Q. After calculation and simplification, one obtains

$$ \zeta ^2_{P,Q}=\frac{q^2_{P,Q} \delta ^2}{4\left( 1-\delta (1+\lambda )q^2_{P,Q}\right) \left( 1+2\delta +2\delta \lambda -\delta (1+3\lambda )q^2_{P,Q}\right) }. $$

After substituting (3.59) and (3.24), one obtains (3.26) and (3.27).

Taking an average of \(\zeta ^2_p\) and \(\zeta ^2_Q\), one obtains the optimal \(\zeta _{opt}\) as in (3.25).

Proof of Proposition 3.4

Denote

$$ A=4(\lambda +1)^2q^2,B=(1-\delta (1+\lambda )q^2)^2,$$
$$ C=4(\lambda +1-\lambda q^2)^2q^2,D=(1-(1+\delta +\delta \lambda )q^2+\lambda \delta q^4)^2. $$

Then, \(\mu \) in (3.28) can be rewritten as

$$\begin{aligned} \mu =\sqrt{\frac{A\zeta ^2+B}{C\zeta ^2+D}}. \end{aligned}$$
(3.61)

To find the invariant points which are independent of damping, it requires

$$ \frac{A}{C}=\frac{B}{D}, $$

that is,

$$ \frac{\lambda +1}{\lambda +1-\lambda q^2}=\pm \frac{1-\delta (1+\lambda )q^2}{1-(1+\delta +\delta \lambda )q^2+\delta \lambda q^4}. $$

Similarly, with plus sign, one obtains the trivial solution zero, and with minus sign, one obtains

$$\begin{aligned} 2\delta \lambda (\lambda +1)q^4-\left( 1+2\lambda +2\delta (1+\lambda )^2\right) q^2+2(\lambda +1)=0. \end{aligned}$$
(3.62)

Thus, one obtains the two invariant points P and Q (\(q_P<q_Q\)) as in (3.32).

Letting the ordinates at invariant points P and Q equal, one has

$$ \left| \frac{\lambda +1}{\lambda +1-\lambda q^2_P}\right| =\left| \frac{\lambda +1}{\lambda +1-\lambda q^2_Q}\right| . $$

It can be checked that \(\frac{\lambda +1}{\lambda +1-\lambda q^2_P}>0\) and \(\frac{\lambda +1}{\lambda +1-\lambda q^2_Q}<0\). Then, one obtains

$$ \frac{\lambda +1}{\lambda +1-\lambda q^2_P}=-\frac{\lambda +1}{\lambda +1-\lambda q^2_Q}. $$

After cross-multiplication and simplification, one has

$$ q^2_P+q^2_Q=\frac{2(\lambda +1)}{\lambda }. $$

Comparing with (3.62), one obtains

$$ \frac{1+2\lambda +2\delta (1+\lambda )^2}{2\delta \lambda (\lambda +1)}=\frac{2(\lambda +1)}{\lambda }, $$

which leads to

$$ 2\delta \lambda ^2-2(1-2\delta )\lambda +2\delta -1=0. $$

It can be checked that this equation has real solutions if and only if

$$ \delta \le 1/2. $$

Under this condition, the optimal \(\lambda \) can be obtained as in (3.29).

Note that if \(\delta =\frac{1}{2}\), from (3.29), one has \(\lambda =0\) or \(k=\infty \). In this case, C5 reduces to C1. Thus, the more reasonable assumption is \(\delta <\frac{1}{2}\) rather than \(\delta \le \frac{1}{2}\).

Similarly, the optimal \(\zeta \) can be obtained by making \(\mu \) to have zero gradients at invariant points P and Q. After calculation and simplification, one obtains \(\zeta ^2_P\) and \(\zeta ^2_Q\) as in (3.31).

Taking an average of \(\zeta ^2_p\) and \(\zeta ^2_Q\), one obtains the optimal \(\zeta _{opt}\) as in (3.30).

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Singapore Pte Ltd. and Science Press, Beijing

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Chen, M.Z.Q., Hu, Y. (2019). Inerter-Based Isolation System. In: Inerter and Its Application in Vibration Control Systems. Springer, Singapore. https://doi.org/10.1007/978-981-10-7089-1_3

Download citation

  • DOI: https://doi.org/10.1007/978-981-10-7089-1_3

  • Published:

  • Publisher Name: Springer, Singapore

  • Print ISBN: 978-981-10-7088-4

  • Online ISBN: 978-981-10-7089-1

  • eBook Packages: EngineeringEngineering (R0)

Publish with us

Policies and ethics