scispace - formally typeset
Search or ask a question
Journal ArticleDOI

Wind-induced vibration mitigation in tall buildings using the tuned mass-damper-inerter (TMDI)

01 Sep 2017-Journal of Structural Engineering-asce (American Society of Civil Engineers)-Vol. 143, Iss: 9, pp 04017127
TL;DR: In this article, a parametric numerical study is performed involving a top-floor-TMD-equipped planar frame capturing accurately the in-plane dynamic behavior of a 74-storey benchmark building exposed to a quasi-stationary spatially-correlated wind-force field accounting for vortex shedding effects in the across-wind direction.
Abstract: In this paper the classical linear tuned mass-damper (TMD) is coupled with an inerter, a two-terminal device resisting the relative acceleration of its terminals, in various tuned mass-damper-inerter (TMDI) topologies to suppress excessive wind-induced oscillations in tall buildings causing occupants’ discomfort. A parametric numerical study is undertaken involving a top-floor-TMD-equipped planar frame capturing accurately the in-plane dynamic behavior of a 74-storey benchmark building exposed to a quasi-stationary spatially-correlated wind-force field accounting for vortex shedding effects in the across-wind direction. It is found that the TMDI reduces the peak top floor acceleration more effectively than the TMD by considering smaller attached mass values, and TMDI topologies in which the inerter spans more stories in linking the attached mass to the host structure. Moreover, the inclusion of the inerter reduces dramatically the TMD stroke while it was verified that the magnitude of the developing inerter forces can be readily accommodated by the host structure. Pertinent illustrative examples are included showcasing that the TMDI meets code-prescribed serviceability design requirements for new tall buildings using significantly smaller attached mass compared to the TMD, and that inerter devices can be used to upgrade the performance of existing TMD-equipped tall buildings without changing the attached mass.

Summary (3 min read)

INTRODUCTION

  • In fact, in many cases, vortex shedding induces higher peak floor accelerations in slender/tall buildings in the across-wind direction than those exhibited in the along-wind direction (Ciampoli and Petrini 2012; Bernardini et al. 2015).
  • The effectiveness of the TMD relies on “tuning” its stiffness and damping properties for a given primary structure and attached mass, such that significant kinetic energy is transferred from the vibrating primary structure to the TMD mass and eventually dissipated through the dampers.
  • In the meantime, in recent years, there is a World-wide trend towards the design and construction of ever-more lighter, slender, and, therefore, more susceptible to wind-induced vibrations tall buildings.

THE TUNED MASS-DAMPER-INERTER (TMDI) FOR MULTI-STOREY BUILDINGS

  • Conceptually defined by Smith (2002), the ideal inerter is a linear massless two-terminal mechanical element resisting the relative acceleration at its terminals through the so-called inertance coefficient, b, measured in mass units.
  • Treating the above system as the primary structure, the TMDI topology introduced by Marian and Giaralis (2013,2014) comprises a mass mTMDI attached to the top floor via a linear spring of stiffness kTMDI and a linear dashpot of damping coefficient cTMDI, and linked to the penultimate floor by an ideal inerter of inertance b.
  • Note that the inclusion of the inerter device influences only the mass matrix M of the controlled structure (i.e., the matrices C and K are the same for the TMD and for the TMDI).
  • These cross-terms alter the dynamics of the primary structure such that higher modes of vibration are damped besides the fundamental mode shape.

ADOPTED PRIMARY STRUCTURE AND WIND EXCITATION MODELLING

  • Benchmark 74-storey building description and detailed FE model.
  • To assess the potential of the TMDI in Fig. 1 for suppressing wind induced oscillations in tall buildings, a high-rise building previously considered for the development of a performance-based wind engineering framework (Spence and Gioffrè 2012; Petrini and Ciampoli 2012) is taken as a benchmark structure.
  • The adopted structure is a 74-storey steel frame building of 305m total height with a 50m-by-50m footprint.
  • For illustration, the first three normalized mode shapes obtained from the FE model, φ(FE)j (j=1,2,…,6) and from the reduced 74-DOF dynamical system φj (j=1,2,…,6) are superposed in Fig.2(b) shown to match very well.
  • For natural frequencies above this range, increasing damping ratios with natural frequency are assumed converging asymptotically to an arbitrarily taken 18% damping ratio.

ASSESSMENT OF TMDI VIBRATION SUPPRESSION CAPABILITIES VIS-À-VIS THE TMD

  • A parametric investigation is undertaken to quantify the effectiveness of different TMDI topologies to mitigate vibrations in the across-wind direction of tall buildings accounting for vortex shedding effects.
  • To this aim, the previously discussed 74-DOF structural system is taken as a paradigm of a primary structure exposed to wind loading modelled by the adopted 74FFS PSD matrix for top floor mean wind velocity Vref= 35m/s (Fig.3).
  • These RFs are obtained by normalizing the peak response quantities in Eq. (11) by the corresponding values of the uncontrolled structure, while the inertance ratio is let to vary in the range of 0% to 100% of the primary structure mass, with the special case of β=0 being the TMD.
  • Still, they do yield reasonable values for the stiffness and the damping properties of the TMDI, accounting for the mass amplification effect of the inerter, which suffice to draw valid and useful comparisons of the effectiveness of different TMDI topologies vis-à-vis the TMD (β=0) case.
  • Overall, the herein reported numerical data suggests that the incorporation of the inerter to the TMD is rather beneficial in reducing floor accelerations and that the smaller the attached mass is the more significant this reduction becomes.

SOME PRACTICAL DESIGN CONSIDERATIONS FOR TMDI-EQUIPPED TALL BUILDINGS

  • Having established the potential benefits of coupling TMDs with inerters to control wind-induced vibrations in tall buildings, it is herein deemed important to discuss certain practical design considerations for TMDI-equipped tall buildings related to the attached mass relative displacement (TMDI stroke), the inerter force, and the required size and weight of ideal inerter device realizations for the application at hand.
  • Evidently, the inclusion of the inerter device reduces dramatically the TMDI stroke (note the logarithmic scale of the y-axis in the considered plots), though the reduction rate reduces as the inertance value increases for the same attached mass.
  • This consideration can be further facilitated by using more than one inerter devices in parallel such the total force demand F is split into several inerter devices, each one transferring a significantly smaller than F force to the connection with the primary structure.
  • The input energy to be accommodated by the flywheel is computed as the product of the peak inerter force in Fig.4 times the peak relative displacement at the inerter terminals, which yields a conservative value as these two quantities do not peak simultaneously.

MEETING SERVICEABILITY CRITERIA IN NEW AND EXISTING TALL BUILDINGS

  • Existing wind-excited tall buildings prescribed by typical building regulations.
  • Focusing first on new tall buildings, suppose that it is sought to design a TMD(I)-based passive vibration control system to meet the serviceability design criteria in Fig.6 for the benchmark building in Fig.2 assuming Vref=35m/s and apartment occupancy.
  • 22 Turning the attention to the use of inerters to enhance the structural performance of existing TMD- equipped tall buildings, suppose that the site-specific design wind velocity, Vref, increases from 35m/s to 42m/s during the lifetime of the same benchmark building.
  • The required RF in terms of peak top floor acceleration to meet the users’ comfort criteria in Fig.6 reduces from 0.803 to 0.595 as indicated in Fig.7 by the horizontal red broken lines.
  • For “-3” TMDI topology the users’ comfort criteria for Vref=42m/s can be met even with the small attached mass of μ=0.1% by incorporating an inerter with sufficiently large inertance (β>0.69%).

CONCLUDING REMARKS

  • The advantages and effectiveness of coupling inerter devices with linear passive TMDs to suppress excessive wind-induced oscillations in tall buildings in the across-wind direction have been numerically established and discussed.
  • It was found that the TMDI reduces the peak top floor acceleration (i.e., the critical demand parameter in meeting serviceability criteria for occupants’ comfort) beyond a same-weight TMD more effectively by considering: (I) smaller attached mass values, and (II) TMDI topologies in which the inerter spans more stories in linking the attached mass to the host structure.
  • To further support the feasibility of the above practical applications of the TMDI, numerical evidence were provided to argue that the developing inerter forces are of reasonable amplitude and can be locally accommodated by properly detailed inerter-structure connections.
  • Moreover, it was illustrated through energy-based calculations that, even though inerter devices are not commercially available for control vibration applications in large-scale civil engineering structures, such inerters are physically and technologically feasible to manufacture by scaling-up current small-scale devices tailored for automotive applications.
  • It is emphasized that the herein reported numerical data are based on sub-optimal TMDI properties.

Did you find this useful? Give us your feedback

Figures (8)

Content maybe subject to copyright    Report

City, University of London Institutional Repository
Citation: Giaralis, A. and Petrini, F. (2017). Wind-induced vibration mitigation in tall
buildings using the tuned mass-damper-inerter (TMDI). Journal of Structural Engineering,
143(9), 04017127. doi: 10.1061/(ASCE)ST.1943-541X.0001863
This is the accepted version of the paper.
This version of the publication may differ from the final published
version.
Permanent repository link: https://openaccess.city.ac.uk/id/eprint/17190/
Link to published version: http://dx.doi.org/10.1061/(ASCE)ST.1943-
541X.0001863
Copyright: City Research Online aims to make research outputs of City,
University of London available to a wider audience. Copyright and Moral
Rights remain with the author(s) and/or copyright holders. URLs from
City Research Online may be freely distributed and linked to.
Reuse: Copies of full items can be used for personal research or study,
educational, or not-for-profit purposes without prior permission or
charge. Provided that the authors, title and full bibliographic details are
credited, a hyperlink and/or URL is given for the original metadata page
and the content is not changed in any way.
City Research Online

City Research Online: http://openaccess.city.ac.uk/ publications@city.ac.uk

Journal of Structural Engineering, ASCE (2017)
1
*
Corresponding author
Wind-induced vibration mitigation in tall buildings using the tuned mass-
damper-inerter (TMDI)
Agathoklis Giaralis
1*
, Francesco Petrini
2
1
Department of Civil Engineering, City, University of London, London, UK
2
Department of Structural and Geotechnical Engineering, Sapienza University of Rome, Rome, ITALY
ABSTRACT
In this paper the classical linear tuned mass-damper (TMD) is coupled with an inerter, a two-terminal device
resisting the relative acceleration of its terminals, in various tuned mass-damper-inerter (TMDI) topologies
to suppress excessive wind-induced oscillations in tall buildings causing occupants’ discomfort. A
parametric numerical study is undertaken involving a top-floor-TMD-equipped planar frame capturing
accurately the in-plane dynamic behavior of a 74-storey benchmark building exposed to a quasi-stationary
spatially-correlated wind-force field accounting for vortex shedding effects in the across-wind direction. It is
found that the TMDI reduces the peak top floor acceleration more effectively than the TMD by considering
smaller attached mass values, and TMDI topologies in which the inerter spans more stories in linking the
attached mass to the host structure. Moreover, the inclusion of the inerter reduces dramatically the TMD
stroke while it was verified that the magnitude of the developing inerter forces can be readily accommodated
by the host structure. Pertinent illustrative examples are included showcasing that the TMDI meets code-
prescribed serviceability design requirements for new tall buildings using significantly smaller attached mass
compared to the TMD, and that inerter devices can be used to upgrade the performance of existing TMD-
equipped tall buildings without changing the attached mass.
Keywords: wind-excited tall building, tuned mass damper, inerter, passive vibration control, occupants
comfort

Journal of Structural Engineering, ASCE (2017)
2
INTRODUCTION
Wind-excited slender high-rise buildings with rectangular floor plan are prone to excessive
oscillations in the across-wind direction (i.e., within the normal plane to the wind direction) due to vortex
shedding effects generated around their edges (Liang et al. 2002). In fact, in many cases, vortex shedding
induces higher peak floor accelerations in slender/tall buildings in the across-wind direction than those
exhibited in the along-wind direction (Ciampoli and Petrini 2012; Bernardini et al. 2015). In such cases,
ensuring that the across-wind floor accelerations remain below a certain threshold associated with users’
comfort becomes the critical serviceability design requirement for slender buildings (Kwok et al. 2009).
In this context, over the past three decades, tuned mass-dampers (TMDs) have been widely used in
practice, among other devices and configurations for supplemental damping, for vibration mitigation in
wind-excited tall buildings to meet occupants’ comfort performance criteria prescribed by building codes
and guidelines (Kareem et al. 1999; Tse et al 2012). In its simplest form, the linear passive TMD comprises a
mass attached towards the top of the building (primary structure), via linear stiffeners, or hangers in case of
pendulum-like TMD implementations, and supplemental damping devices (dampers). The effectiveness of
the TMD relies on “tuning” its stiffness and damping properties for a given primary structure and attached
mass, such that significant kinetic energy is transferred from the vibrating primary structure to the TMD
mass and eventually dissipated through the dampers. Focusing on the suppression of lateral wind-induced
vibrations in (tall) buildings, the TMD is tuned to the first natural frequency of the primary structure aiming
to control the fundamental (translational) lateral mode shape (e.g. Rana and Soong 1998; Li et al. 1999).
It is well-recognized that the vibration suppression capabilities of the TMD depends heavily on its
inertial property: the larger the attached TMD mass is, the more effective and robust to uncertainties in the
structural properties the TMD becomes (e.g. De Angelis et al. 2012). However, practical structural and
architectural constraints apply to the weight and to the volume of the TMD mass that can be accommodated
by the primary structure. These constraints are particularly critical for tall buildings for which the attached
mass rarely exceeds 0.5% to 1% of the total building mass. In the meantime, in recent years, there is a
World-wide trend towards the design and construction of ever-more lighter, slender, and, therefore, more
susceptible to wind-induced vibrations tall buildings. This is partly due to the increase cost of land in

Journal of Structural Engineering, ASCE (2017)
3
congested urban environments and partly due to requirements for more aesthetically pleasing and sustainable
structures. Furthermore, the existing TMD-equipped tall buildings risk increase downtime (and consequent
financial losses) in the foreseeable future due to an anticipated increase of the frequency of strong wind
fronts caused by climate change effects.
To address the above issues and concerns in an innovative manner, this paper explores the potential of
incorporating inerter devices to wind-excited TMD-equipped tall buildings in different tuned mass-damper-
inerter (TMDI) configurations (topologies) to achieve enhanced vibration suppression in the across-wind
direction without increasing the attached TMD mass. Specifically, the considered TMDI topologies, a
special case of which was originally introduced in Marian and Giaralis (2013) for seismic protection of
multi-storey shear framed structures, benefit from the mass-amplification and from the higher-modes-
damping effects of the inerter. The latter is a two-terminal device of negligible mass/weight resisting relative
acceleration in analogy to the linear spring and dashpot resisting relative displacement and velocity,
respectively (Smith 2002). Marian and Giaralis (2014) showed analytically and numerically that the TMDI
can be more effective than the TMD for the same attached mass in reducing the deformation variance of
stochastically base-excited linear primary structures due to the contribution of the inerter to the inertial
property of the attached mass (mass-amplification effect). Furthermore, Giaralis and Taflanidis (2015)
demonstrated that the inclusion of the inerter influences and can potentially suppress higher modes of
vibration in linear multi-storey primary structures (higher-modes-damping effect), as opposed to the TMD
which can only control a single vibration mode. The latter effect is rather important for wind-induced
vibration control in tall buildings since higher vibration modes have a significant contribution to floor
acceleration response (i.e., the critical design criterion for tall buildings at the serviceability limit state), as
opposed to floor displacements. It is noted in passing that a similar higher-modes-damping effect in linear
dynamically excited multi-storey buildings were also reported by Lazar et al. (2014) for the case of the tuned
inerter-damper (TID), which is a special case of a TMDI with no attached mass. In fact, this effect was
accounted for by Krenk and Høgsberg (2016) to achieve a simplified analytical approach for the optimal TID
design targeting a single vibration mode in multi-storey buildings.
Herein, a parametric numerical study is undertaken to assess the effectiveness of different TMDI
topologies vis-a-vis the TMD for fixed attached mass in reducing peak top floor displacements and

Citations
More filters
Journal ArticleDOI
TL;DR: The review clearly demonstrates that the TMDs have a potential for improving the wind and seismic behaviors of prototype civil structures and shows that the MTMDs and d-MTMDs are relatively more effective and robust, as reported.

263 citations

Journal ArticleDOI
TL;DR: Compared to the conventional TMD system, the static stretching of the spring due to gravity and the oscillation amplitude of the mass block in the TMDI system are significantly reduced, which makes the proposed T MDI system an attractive alternative for the VIV control of long-span bridges.

106 citations

Journal ArticleDOI
TL;DR: A novel optimal T MDI design formulation is proposed to address occupants’ comfort in wind-excited slender tall buildings susceptible to vortex shedding (VS) effects and to explore optimal TMDI’s potential for transforming part of the wind-induced kinetic energy to usable electricity in tall buildings.

95 citations

Journal ArticleDOI
TL;DR: In this paper, the authors presented a theoretical and numerical study of a structure-TID system subjected to earthquake ground motions, and derived simple design formulas for the TID when it attached to a single-degree-of-freedom (DofF) structure subjected to ground acceleration excitations.

94 citations

References
More filters
Book
01 Jan 1995
TL;DR: In this paper, the authors present a single-degree-of-freedom (SDF) system, which is composed of a mass-spring-damper system and a non-viscous Damping Free Vibration (NFV) system.
Abstract: I. SINGLE-DEGREE-OF-FREEDOM SYSTEMS. 1. Equations of Motion, Problem Statement, and Solution Methods. Simple Structures. Single-Degree-of-Freedom System. Force-Displacement Relation. Damping Force. Equation of Motion: External Force. Mass-Spring-Damper System. Equation of Motion: Earthquake Excitation. Problem Statement and Element Forces. Combining Static and Dynamic Responses. Methods of Solution of the Differential Equation. Study of SDF Systems: Organization. Appendix 1: Stiffness Coefficients for a Flexural Element. 2. Free Vibration. Undamped Free Vibration. Viscously Damped Free Vibration. Energy in Free Vibration. Coulomb-Damped Free Vibration. 3. Response to Harmonic and Periodic Excitations. Viscously Damped Systems: Basic Results. Harmonic Vibration of Undamped Systems. Harmonic Vibration with Viscous Damping. Viscously Damped Systems: Applications. Response to Vibration Generator. Natural Frequency and Damping from Harmonic Tests. Force Transmission and Vibration Isolation. Response to Ground Motion and Vibration Isolation. Vibration-Measuring Instruments. Energy Dissipated in Viscous Damping. Equivalent Viscous Damping. Systems with Nonviscous Damping. Harmonic Vibration with Rate-Independent Damping. Harmonic Vibration with Coulomb Friction. Response to Periodic Excitation. Fourier Series Representation. Response to Periodic Force. Appendix 3: Four-Way Logarithmic Graph Paper. 4. Response to Arbitrary, Step, and Pulse Excitations.Response to Arbitrarily Time-Varying Forces. Response to Unit Impulse. Response to Arbitrary Force. Response to Step and Ramp Forces. Step Force. Ramp or Linearly Increasing Force. Step Force with Finite Rise Time. Response to Pulse Excitations. Solution Methods. Rectangular Pulse Force. Half-Cycle Sine Pulse Force. Symmetrical Triangular Pulse Force. Effects of Pulse Shape and Approximate Analysis for Short Pulses. Effects of Viscous Damping. Response to Ground Motion. 5. Numerical Evaluation of Dynamic Response. Time-Stepping Methods. Methods Based on Interpolation of Excitation. Central Difference Method. Newmark's Method. Stability and Computational Error. Analysis of Nonlinear Response: Central Difference Method. Analysis of Nonlinear Response: Newmark's Method. 6. Earthquake Response of Linear Systems. Earthquake Excitation. Equation of Motion. Response Quantities. Response History. Response Spectrum Concept. Deformation, Pseudo-Velocity, and Pseudo-Acceleration Response Spectra. Peak Structural Response from the Response Spectrum. Response Spectrum Characteristics. Elastic Design Spectrum. Comparison of Design ad Response Spectra. Distinction between Design and Response Spectra. Velocity and Acceleration Response Spectra. Appendix 6: El Centro, 1940 Ground Motion. 7. Earthquake Response of Inelastic Systems. Force-Deformation Relations. Normalized Yield Strength, Yield Strength Reduction Factor, and Ductility Factor. Equation of Motion and Controlling Parameters. Effects of Yielding. Response Spectrum for Yield Deformation and Yield Strength. Yield Strength and Deformation from the Response Spectrum. Yield Strength-Ductility Relation. Relative Effects of Yielding and Damping. Dissipated Energy. Energy Dissipation Devices. Inelastic Design Spectrum. Applications of the Design Spectrum. Comparison of Design and Response Spectra. 8. Generalized Single-Degree-of-Freedom Systems. Generalized SDF Systems. Rigid-Body Assemblages. Systems with Distributed Mass and Elasticity. Lumped-Mass System: Shear Building. Natural Vibration Frequency by Rayleigh's Method. Selection of Shape Function. Appendix 8: Inertia Forces for Rigid Bodies. II. MULTI-DEGREE-OF-FREEDOM SYSTEMS. 9. Equations of Motion, Problem Statement, and Solution Methods. Simple System: Two-Story Shear Building. General Approach for Linear Systems. Static Condensation. Planar or Symmetric-Plan Systems: Ground Motion. Unsymmetric-Plan Building: Ground Motion. Symmetric-Plan Buildings: Torsional Excitation. Multiple Support Excitation. Inelastic Systems. Problem Statement. Element Forces. Methods for Solving the Equations of Motion: Overview. 10. Free Vibration. Natural Vibration Frequencies and Modes. Systems without Damping. Natural Vibration Frequencies and Modes. Modal and Spectral Matrices. Orthogonality of Modes. Interpretation of Modal Orthogonality. Normalization of Modes. Modal Expansion of Displacements. Free Vibration Response. Solution of Free Vibration Equations: Undamped Systems. Free Vibration of Systems with Damping. Solution of Free Vibration Equations: Classically Damped Systems. Computation of Vibration Properties. Solution Methods for the Eigenvalue Problem. Rayleigh's Quotient. Inverse Vector Iteration Method. Vector Iteration with Shifts: Preferred Procedure. Transformation of kA A = ...w2mA A to the Standard Form. 11. Damping in Structures.Experimental Data and Recommended Modal Damping Ratios. Vibration Properties of Millikan Library Building. Estimating Modal Damping Ratios. Construction of Damping Matrix. Damping Matrix. Classical Damping Matrix. Nonclassical Damping Matrix. 12. Dynamic Analysis and Response of Linear Systems.Two-Degree-of-Freedom Systems. Analysis of Two-DOF Systems without Damping. Vibration Absorber or Tuned Mass Damper. Modal Analysis. Modal Equations for Undamped Systems. Modal Equations for Damped Systems. Displacement Response. Element Forces. Modal Analysis: Summary. Modal Response Contributions. Modal Expansion of Excitation Vector p (t) = s p(T). Modal Analysis for p (t) = s p(T). Modal Contribution Factors. Modal Responses and Required Number of Modes. Special Analysis Procedures. Static Correction Method. Mode Acceleration Superposition Method. Analysis of Nonclassically Damped Systems. 13. Earthquake Analysis of Linear Systems.Response History Analysis. Modal Analysis. Multistory Buildings with Symmetric Plan. Multistory Buildings with Unsymmetric Plan. Torsional Response of Symmetric-Plan Buildings. Response Analysis for Multiple Support Excitation. Structural Idealization and Earthquake Response. Response Spectrum Analysis. Peak Response from Earthquake Response Spectrum. Multistory Buildings with Symmetric Plan. Multistory Buildings with Unsymmetric Plan. 14. Reduction of Degrees of Freedom. Kinematic Constraints. Mass Lumping in Selected DOFs. Rayleigh-Ritz Method. Selection of Ritz Vectors. Dynamic Analysis Using Ritz Vectors. 15. Numerical Evaluation of Dynamic Response. Time-Stepping Methods. Analysis of Linear Systems with Nonclassical Damping. Analysis of Nonlinear Systems. 16. Systems with Distributed Mass and Elasticity. Equation of Undamped Motion: Applied Forces. Equation of Undamped Motion: Support Excitation. Natural Vibration Frequencies and Modes. Modal Orthogonality. Modal Analysis of Forced Dynamic Response. Earthquake Response History Analysis. Earthquake Response Spectrum Analysis. Difficulty in Analyzing Practical Systems. 17. Introduction to the Finite Element Method.Rayleigh-Ritz Method. Formulation Using Conservation of Energy. Formulation Using Virtual Work. Disadvantages of Rayleigh-Ritz Method. Finite Element Method. Finite Element Approximation. Analysis Procedure. Element Degrees of Freedom and Interpolation Function. Element Stiffness Matrix. Element Mass Matrix. Element (Applied) Force Vector. Comparison of Finite Element and Exact Solutions. Dynamic Analysis of Structural Continua. III. EARTHQUAKE RESPONSE AND DESIGN OF MULTISTORY BUILDINGS. 18. Earthquake Response of Linearly Elastic Buildings. Systems Analyzed, Design Spectrum, and Response Quantities. Influence of T 1 and r on Response. Modal Contribution Factors. Influence of T 1 on Higher-Mode Response. Influence of r on Higher-Mode Response. Heightwise Variation of Higher-Mode Response. How Many Modes to Include. 19. Earthquake Response of Inelastic Buildings. Allowable Ductility and Ductility Demand. Buildings with "Weak" or "Soft" First Story. Buildings Designed for Code Force Distribution. Limited Scope. Appendix 19: Properties of Multistory Buildings. 20. Earthquake Dynamics of Base-Isolated Buildings. Isolation Systems. Base-Isolated One-Story Buildings. Effectiveness of Base Isolation. Base-Isolated Multistory Buildings. Applications of Base Isolation. 21. Structural Dynamics in Building Codes. Building Codes and Structural Dynamics. International Building Code (United States), 2000. National Building Code of Canada, 1995. Mexico Federal District Code, 1993. Eurocode 8. Structural Dynamics in Building Codes. Evaluation of Building Codes. Base Shear. Story Shears and Equivalent Static Forces. Overturning Moments. Concluding Remarks. Appendix A: Frequency Domain Method of Response Analysis.Appendix B: Notation.Appendix C: Answers to Selected Problems.Index.

4,812 citations

Journal ArticleDOI
TL;DR: The main contributions of the paper are the introduction of a device, which win be called the inerter, which is the true network dual of the spring, which contrasts with the mass element which, by definition, always has one terminal connected to ground.
Abstract: The paper is concerned with the problem of synthesis of (passive) mechanical one-port networks. One of the main contributions of the paper is the introduction of a device, which win be called the inerter, which is the true network dual of the spring. This contrasts with the mass element which, by definition, always has one terminal connected to ground. The inerter allows electrical circuits to be translated over to mechanical ones in a completely analogous way. The inerter need not have large mass. This allows any arbitrary positive-real impedance to be synthesized mechanically using physical components which may be assumed to have small mass compared to other structures to which they may be attached. The possible application of the inerter is considered to a vibration absorption problem, a suspension strut design, and as a simulated.

1,118 citations


"Wind-induced vibration mitigation i..." refers background in this paper

  • ...In fact, in the limiting case of 2 0u , the inerter behaves as a virtual mass equal to b added to the physical mass of the dynamic degree-of-freedom (DOF) corresponding to the displacement u1 (Smith 2002)....

    [...]

  • ...The latter is a two-terminal device of negligible mass/weight resisting relative acceleration in analogy to the linear spring and dashpot resisting relative displacement and velocity, respectively (Smith 2002)....

    [...]

  • ...The fact that the angular velocity of the flywheel can be arbitrarily regulated through, for example, mechanical gearing (e.g. Smith 2002), renders the inertance be practically independent from the device physical mass....

    [...]

  • ...Conceptually defined by Smith (2002), the ideal inerter is a linear massless two-terminal mechanical element resisting the relative acceleration at its terminals through the so-called inertance coefficient, b, measured in mass units....

    [...]

  • ...physical mass of the dynamic degree-of-freedom (DOF) corresponding to the displacement u1 (Smith 2002)....

    [...]

Journal ArticleDOI
TL;DR: In this paper, simple expressions for optimum absorber parameters are derived for undamped one degree-of-freedom main systems for harmonic and white noise random excitations with force and frame acceleration as input and minimization of various response parameters.
Abstract: In recent papers the author has shown that when determining optimum parameters for an absorber which minimizes the vibration response of a complex system, the latter may be treated as an equivalent single degree-of-freedom system if its natural frequencies are well separated. Emphasis was on minimizing the displacement response when the excitation was a harmonic force. In the present paper simple expressions for optimum absorber parameters are derived for undamped one degree-of-freedom main systems for harmonic and white noise random excitations with force and frame acceleration as input and minimization of various response parameters. These expressions can be used to obtain optimum parameters for absorbers attached to complex systems provided that optimization is with respect to an absolute, rather than a relative, quantity. The requirement that the natural frequencies should be well separated is investigated numerically for the different cases. The effect of damping in the main system on optimum absorber parameters is investigated also.

832 citations

Frequently Asked Questions (15)
Q1. What are the future works in this paper?

In this regard, accounting for the influence of non-linear inerter behavior to the effectiveness of TMDIs for tall building wind vibration mitigation warrants further research which need to be supported by prototyping and characterization of large-scale inerters. To this end, it is envisioned that this study will not only further familiarize the civil/structural engineering community to the potential benefits of considering inerter devices for vibration control in TMD-equipped structures, but also justify the pursue of experimental research to develop inerters suitable for large-scale civil engineering structures. 

In this paper the classical linear tuned mass-damper ( TMD ) is coupled with an inerter, a two-terminal device resisting the relative acceleration of its terminals, in various tuned mass-damper-inerter ( TMDI ) topologies to suppress excessive wind-induced oscillations in tall buildings causing occupants ’ discomfort. A parametric numerical study is undertaken involving a top-floor-TMD-equipped planar frame capturing accurately the in-plane dynamic behavior of a 74-storey benchmark building exposed to a quasi-stationary spatially-correlated wind-force field accounting for vortex shedding effects in the across-wind direction. 

an energy approach is used assuming no losses and linear device behaviour in which it is sought to determine the required flywheel radius given a pre-specified flywheel thickness and flywheel tangential velocity. 

the TMDI suppresses not only the portion of the structural vibration response corresponding to the fundamental mode shape, as the TMD strictly does, but all higher modes as well, which contribute significantly more to response acceleration than to response displacement. 

These cross-terms alter the dynamics of the primary structure such that higher modes of vibration are damped besides the fundamental mode shape. 

Note that the choice to consider such a large number of translational horizontal DOFs equally spacedalong the height of the building model was made to capture accurately the influence of the stiff outriggers to the local dynamics of the structure (see Fig.2b), and to facilitate a fine spatial discretization of the wind loading. 

the TMDI topology does affect the amplitude of the stroke (but not the trends for increasing inertance): the more stories the inerter spans the higher the peak stroke is for the same TMDI properties. 

The latter effect is rather important for wind-induced vibration control in tall buildings since higher vibration modes have a significant contribution to floor acceleration response (i.e., the critical design criterion for tall buildings at the serviceability limit state), as opposed to floor displacements. 

In fact, in many cases, vortex shedding induces higher peak floor accelerations in slender/tall buildings in the across-wind direction than those exhibited in the along-wind direction (Ciampoli and Petrini 2012; Bernardini et al. 2015). 

The significance of this higher-mode-damping effect of the TMDI is related to the position of the cross-terms in the matrix M: the further away they lie from the main diagonal, the more effective theTMDI becomes in damping the higher modes (see e.g. Fig.2(c)). 

More importantly from a practical viewpoint, it was numerically shown that: (i) the TMDI can meetcode-prescribed serviceability design requirements by considering significantly smaller attached mass24(depending on the TMDI topology and inertance coefficient) compared to the TMD, enabling more lightweight construction and efficient material usage in the design of new tall buildings, and (ii) inerter devices can be used to upgrade existing TMD-equipped tall buildings, without changing the attached mass, to meet more stringent serviceability design requirements than those considered in the initial design due to site-specific climate change effects or changes to the surrounding built environment (i.e., increased wind exposure). 

The input energy to be accommodated (stored) by the flywheel is computed as the product of the peak inerter force in Fig.4 times the peak relative displacement at the inerter terminals, which yields a conservative value as these two quantities do not peak simultaneously. 

The fact that the herein considered non-optimal TMDIs are better suited to suppress flooraccelerations rather than floor displacements, compared to the TMD, is readily attributed to the highermode-damping effect of the TMDI. 

From the first row of plots in Fig. 4, it is seen that the inclusion of the inerter is beneficial in terms ofreducing the peak top floor displacement compared to the TMD, only for relatively small attached mass values. 

Giaralis and Taflanidis (2015) demonstrated that the inclusion of the inerter influences and can potentially suppress higher modes of vibration in linear multi-storey primary structures (higher-modes-damping effect), as opposed to the TMD which can only control a single vibration mode.