Robust-to-Uncertainties Optimal Design of Seismic Metamaterials
Robust-to-Uncertainties Optimal Design of Seismic Metamaterials
Robust-to-Uncertainties Optimal Design of Seismic Metamaterials
Seismic Metamaterials
Paul-Remo Wagner 1; Vasilis K. Dertimanis, Ph.D. 2;
Eleni N. Chatzi, Ph.D., A.M.ASCE 3; and James L. Beck, Ph.D., M.ASCE 4
Downloaded from ascelibrary.org by MICHIGAN TECHNOLOGICAL UNIVERSITY on 09/07/19. Copyright ASCE. For personal use only; all rights reserved.
Abstract: Metamaterials, which draw their origin from a special class of structured (periodic) materials characterized by a dynamic filtering
effect, have recently emerged as a prospective means for structural seismic protection. This paper explores such a periodic arrangement in the
form of local adaptive resonators buried in the soil, serving as a seismic protection barrier. As a starting point, a simplistic representation is
chosen herein that comprises chains of mass-in-mass unit cells. A robust-to-uncertainties optimization of such a chain, addressing uncer-
tainties at the level of the excitation, the system properties and the model structure itself, is conducted. The optimization problem is for-
mulated within the context of reliability assessment, where the objective function is the failure probability of the structure to be protected
against seismic input. The problem is solved through adoption of the subset optimization algorithm enhanced through the simultaneous
implementation of a stochastic approximation algorithm. It is demonstrated that not all parameters of the chain model require optimization,
because the failure probability proves to be a monotonic function of a subset of the parameters. A primary objective herein lies in optimizing
the internal unit-cell stiffness properties. It is further demonstrated that the effectiveness of the protection offered by the metamaterial is
improved for spatially varying unit-cell properties. The optimization procedure is carried out in the frequency domain, with an example
application confirming that a time domain optimization is expected to yield similar optimal configurations. A parametric study using a
nonlinear model is also presented, offering a starting point for more refined future investigations. DOI: 10.1061/(ASCE)EM.1943-
7889.0001404. © 2017 American Society of Civil Engineers.
Author keywords: Seismic isolation; Periodic lattice; Metamaterials; Optimal design; Reliability.
(a)
(a)
x g (t )
L
(b)
(a) (b)
of the associated frequency band gap. For lower mass ratios, the
Fig. 3. Comparison of the band gaps in the infinite chain (dispersion band gaps extend into higher frequencies and are wider overall.
relation) and the RRR in the finite chain [FRF HðfÞ] of model G1 : Implementing low mass ratios forms a major engineering chal-
(a) dispersion relation; (b) complex FRF HðfÞ lenge, as the large internal mass ought to be supported by a much
smaller external one.
In addition to the aforementioned parameters and in an effort to
produce wider band gaps in lower frequency ranges, which is criti-
rigid which is always true in practice), if this were the case the cal for earthquake engineering applications, the use of the so called
excitation movement would simply bypass the local resonators rainbow-traps is proposed (Zhu et al. 2013; Krödel et al. 2015),
and transfer the motion directly from the first element of the chain according to which nonidentical mass-in-mass unit cells with dis-
to the last. Similarly, the external stiffness has to be non-null, tributed natural frequencies are utilized. In this setting, the aim is
kext ≠ 0, because null external stiffness would hinder any energy to create a chain of unit cells with individually overlapping band
transport along the SM and again render the latter ineffective. gaps and to thereby increase the overall RRR. Using this approach,
The effect of N and kext on the FRF amplitude inside the RRR a response reduction in a range from 4–7 Hz has been reported
is depicted in Fig. 4. (Krödel et al. 2015), which is close (but not sufficient) to the target
Define now an equivalent natural frequency, ω̂, for a lattice of frequency ranges for civil engineering structures. The use of rain-
mass-in-mass unit cells as bow traps is also followed in the current study, however, instead of
(e)
Fig. 4. FRF of the absolute acceleration amplitudes ẍ in dB for different external stiffness kext and unit cell numbers N with identical unit cells in
model G1 ; the excitation frequency is given by f ex and the internal frequency by f int ; the output is truncated from (−100 to 0) dB to emphasize the
locations of the band gaps: (a) kext ¼ 10 s−2 , N ¼ 1, ω̂ ¼ 3.16 rad s−1 ; (b) kext ¼ 1,000 s−2 , N ¼ 10, ω̂ ¼ 3.16 rad s−1 ; (c) kext ¼ 100 s−2 , N ¼ 10,
ω̂ ¼ 1 rad s−1 ; (d) kext ¼ 10,000 s−2 , N ¼ 100, ω̂ ¼ 1 rad s−1 ; (e) grayscale for the FRF
(d)
Fig. 5. FRF of the absolute acceleration amplitudes ẍ in dB for different mass ratios r ¼ m mint and unit cell numbers N with identical unit cells in model
ext
G1 ; the excitation frequency is given by f ex and the internal frequency by f int ; the output is truncated from (−100 to 0) dB to emphasize the locations
of the band gaps; mint ¼ 0.15, mext ¼ 0.015: (a) r ¼ 0.01, N ¼ 10; (b) r ¼ 0.1, N ¼ 10; (c) r ¼ 1, N ¼ 10; (d) grayscale for the FRF
θs ¼ ½kext ; mint ; mext ; f SDOF ; ζ ð3Þ this end, the amplitudes of the discrete Fourier spectra of the simu-
lated time histories are subtracted [Figs. 6(c and d)]. A synthetic
for which a probability density function (PDF) pðθs Þ is assigned.
time history is then constructed, with its Fourier amplitude spec-
In Eq. (3), f SDOF refers to the natural frequency of the attached
trum derived from the previous step and its phase determined as
oscillator, whereas ζ corresponds to a damping ratio, herein intro-
the phase of the high frequency component’s discrete Fourier trans-
duced in the form of viscous Rayleigh damping through a Caughey
form (DFT) [Fig. 6(e)]. Finally, the near fault ground motion time
matrix (Chopra 2007), where applicable. The uncertain excitation ϵ
history is obtained as the superposition of this time history with the
is further elaborated on in the next subsection. The design variables,
time history of the low frequency component [Fig. 6(f)].
φi , correspond to the model parameters that optimization is in-
It has to be noted that the subtraction in the above procedure
tended for. Following Eq. (1), the internal-cell properties are the
[Fig. 6(d)] cannot be guaranteed to generally yield positive values.
defining parameters for the location of the band gaps, and are there-
This fact is well known by the authors of the original paper
fore chosen as design variables. Equivalent to the parameter vector,
(Mavroeidis and Papageorgiou 2003) but was observed to not have
the authors define a design vector, φ, containing all φi . The result-
a significant impact on the resulting time histories.
ing stochastic system model has the form
The stochastic excitation model is parameterized by the uncer-
ξ ¼ Gðϵ; θs ; φÞ ð4Þ tain vector, θϵ , defined as
where the uncertain output ξ may be any response quantity of the θϵ ¼ ½Mw ; R; Zw ; γ p ; ν p ; εf ; εv ð5Þ
SM model, either in its infinite (Bloch theory), or in its finite
in which the seismic magnitude M w , the epicentral distance R and
(classical vibration theory) format.
the white noise sequence Zw correspond to the high frequency com-
ponent [details in Boore (2003)], whereas the remaining parameters
Earthquake Excitation Model γ p , ν p , εf , and εv correspond to the low frequency component
[details in Mavroeidis and Papageorgiou (2003)]. The assigned
To establish a robust-to-uncertainties system design procedure for PDF pðθϵ Þ of θϵ is discussed in the following section. The uncertain
SMs, the probabilistic seismic ground motion model of Taflanidis excitation, ϵ, to be used in the SM is thus the output of the stochas-
(2007) is implemented. The method extends the models initially tic excitation model, that is
developed by Boore (2003) and Mavroeidis and Papageorgiou
(2003) for synthetic ground motion generation, and provides a ϵ ¼ Eðθϵ Þ ð6Þ
stochastic model for near-fault seismic excitations. It is well suited
for stochastic system design, because of the well-defined physical It is noted that the source spectra currently used for the sto-
meaning of its parameters and its relatively low computational cost chastic ground motion model is developed for earthquakes in
Fig. 6. Near fault ground motion model used to create acceleration time histories; the shown plots are from a simulation with seismic magnitude
M w ¼ 7 and a fault distance R ¼ 10 km; the sampling rate is Δt ¼ 0.01 s: (a) high frequency component ðM w ; R; Zw Þ; (b) low frequency component
ðγ p ; ν p ; εf ; εv Þ; (c) DFT of time histories; (d) subtraction of spectra; (e) synthetic time history; (f) superposition of Figs. 6(b and e) [ϵðtÞ]
The problem treated in this study pertains to the optimal design Given a certain design choice, φ, and a single realization of the
of the SM under the presence of uncertainties. In this setting, a uncertain parameter vector, θ, the evaluation of the model class
stochastic model class M is defined, which combines the stochastic M pertains to the determination of the value of the failure indicator
system model, Gðϵ; θs ; φÞ, the stochastic excitation model, Eðθϵ Þ, function, I F . The underlying process is illustrated for one design
and a performance evaluation model, jðξ; θw Þ. The class M incor- choice, φ, and one realization of θ ∼ pðθjφÞ in Fig. 7 for a fre-
Downloaded from ascelibrary.org by MICHIGAN TECHNOLOGICAL UNIVERSITY on 09/07/19. Copyright ASCE. For personal use only; all rights reserved.
porates all the individual model parameterizations into the compact quency domain evaluation. The stochastic excitation model produ-
parameter vector, θ ces a time–series ϵðtÞ ¼ Eðθϵ Þ, which is then transformed to the
frequency domain, ϵðfÞ. The stochastic system model is accord-
θ ¼ ½θs ; θϵ ; θw ð7Þ ingly evaluated to produce a FRF as ξðfÞ ¼ G½ϵðfÞ; θs ; φ that
is driven to the performance evaluation model j½ξðfÞ; θw where
with corresponding PDF, pðθjφÞ. Both the uncertain parameter vec-
tor, θ, and the design vector, φ, are bounded by admissible spaces, θw ¼ η ð10Þ
Θ and Φ, respectively, so that θ ∈ Θ ⊂ Rnθ and φ ∈ Φ ⊂ Rnφ .
In defining the objective function for the SM design problem with η introduced below.
considered herein, attention is focused on the class of reliability The evaluation model consists of an augmented limit state func-
objective problems (ROPs) where a failure probability PðFjφÞ is ~
tion gðφ; θÞ, which checks the system model output, ξðfÞ, to evalu-
used as the objective function [e.g., Eq. (9)]. The goal is to find a ate whether the performance is acceptable or not (Taflanidis and
design vector, φ, that results in the lowest failure probability while Beck 2008a)
being bounded by the admissible design space, Φ. The induced
~
0 gðφ; θÞ ≤ 0
optimization problem is thus defined as I F ðφ; θÞ ¼ ð11Þ
~
1 gðφ; θÞ > 0
φ ¼ arg minPðFjφÞ ð8Þ
φ∈Φ The augmented limit state function g~ is obtained by superimpos-
ing a modeling error, η, to a noise–free limit state function, g, as
~
gðφ; θÞ ¼ gðφ; θÞ þ ηðφ; θÞ ð12Þ
Robust-to-Uncertainties Design of Seismic
Metamaterials where η = zero mean Gaussian distribution with standard deviation,
ση , that retains a certain influence on the outcome of the optimi-
Failure Probability zation process (Taflanidis 2007; Beck and Katafygiotis 1998).
It must be noted that such a modeling error can be alternatively
The failure probability, i.e., the probability that the SM performs in
assigned at any level of the model class, M, as, for example, di-
an unacceptable way, is expressed as a weighted integral of the
rectly at the model response (Papadimitriou et al. 2001).
form
Whereas other choices for gðφ; θÞ are applicable, for the pur-
Z
poses of this study the limit state function is defined either as the
PðFjφÞ ¼ I F ðφ; θÞpðθjφÞdθ ð9Þ exceedance of a threshold value (frequency domain analysis), or as
Θ
the peak response threshold (time domain analysis)
where I F ðφ; θÞ = failure indicator function with Boolean range,
depending on whether the system performs below accepted levels ξ ðφ; θÞ maxjξðφ; θÞj
gðφ; θÞ ¼ log RMS gðφ; θÞ ¼ log
(value 1), or if all performance objectives are met (value 0). ϑRMS ϑ
In view of Eq. (8), it follows that the solution of the asso- ð13Þ
ciated optimization problem requires the evaluation of the integral
in Eq. (9). This is generally not possible analytically, except for In the case of a frequency domain analysis, the failure proba-
special cases (Papadimitriou et al. 1997), whereas numerical evalu- bility, PðFjφÞ, implies the probability that the root–mean-square
ation also becomes infeasible for high dimensional parameter (RMS) of the system response exceeds the set threshold value
spaces (Taflanidis 2007). A way to solve the integral in a reasonable ϑRMS . It is noted that the RMS value is directly related to the energy
computational time is through stochastic simulation techniques. content of the earthquake excitation, which is propagated through
Another approach is to use stochastic subset optimization to find the SM. As such, the left counterpart of Eq. (13) essentially refers to
a subset of the design variable space that has a high plausibility the overall value of a seismic event, rather than on a specific time-
for containing the optimal design point, φ (Taflanidis and Beck frequency interval of the typically nonstationary excitation, aiming
Fig. 7. Key steps for structural analysis performed in the frequency domain using a system model
Fig. 9. Surface and contour plots of the failure probabilities in a six-unit-cell chain with nonidentical unit cells with RMS threshold ϑRMS ¼ 0.4 in the
limit state function, where φ ≈ ð0.44; 0.54Þ; the dotted line represents the identical unit-cell case; the failure probabilities were estimated using
Downloaded from ascelibrary.org by MICHIGAN TECHNOLOGICAL UNIVERSITY on 09/07/19. Copyright ASCE. For personal use only; all rights reserved.
Monte Carlo simulations with N ¼ 1,000 independent samples; at the far right the FRFs for both systems in the optimal 1D and 2D design cases are
presented
This places all admissible design variable vectors, φ, in a space 3. Find optimal subset, I . This is the subset, I, with the mini-
where the individual design variables, φi , are arranged in ascend- mum sensitivity that contains a maximum of ρN samples:
ing order.
NI
I ¼ arg minHðIÞ ≈ arg min ĤðIÞ ¼ arg min ð24Þ
SSO I∈Φ I∈Φ I∈Φ VI
The SSO algorithm (Taflanidis and Beck 2008a), with an initial
design space, Φ, in the shape of an nφ -dimensional simplex, iter- 4. Delete all samples that lie outside this identified subset, I.
atively identifies subsets I i in the shape of hyperelliptical segments, Compute Ĥ and COV Ĥ [Eq. (23)].
D, (see Appendix 7), which are likely to contain the optimal design 5. If Ĥ and COV Ĥ do not exceed the set threshold values, re-
variable, φ . With respect to Fig. 10, the SSO consists of the fol- start at Step 1, otherwise stop the algorithm.
lowing steps: The SSO identifies ever smaller subsets in the initial design
1. Initialize the SSO by selecting a desired number of samples, space, Φ. As stated in Taflanidis and Beck (2008a), it is desirable
N SSO , and a constraint value, ρ. Specify a threshold value for the to switch to a different algorithm to pinpoint the exact optimal sol-
average relative sensitivity of pðφjFÞ to φ: Ĥ and COV Ĥ as stop- ution, φ . The efficiency of the SSO algorithm depends strongly on
ping criteria. The sensitivity at each step is estimated as the chosen subset shape and sample ratio, ρ. In accordance with the
guidelines presented in Taflanidis (2007), the sample ratio is chosen
N Ii =V Ii to ρ ¼ 0.15.
HðIÞ ≈ ĤðIÞ ¼ ð23Þ
N Ii−1 =V Ii−1 SPSA with CRN
The SPSA optimization, as introduced in Spall (2003) and extended
where N I designates the number of samples in the subset and, V I , by Kleinman et al. (1999) to include a variance reduction technique
the subset volume. known as common random numbers (CRN), takes the typical re-
2. Use a sampling procedure to simulate N SSO samples from cursive form
pðφjFÞ (e.g., MCS, or MCMC). This corresponds to simulating
failed samples in the design variable space. φkþ1 ¼ φk þ ak ĝk ðφk Þ ð25Þ
Fig. 10. SSO algorithm for a reliability objective problem of a SM consisting of six unit cells with two independent properties [φ ¼ ðφ1 ; φ2 Þ]; the last
step shown is just before the stopping criteria were fulfilled; the optimal solution is marked with a cross and was pinpointed using the SPSA algorithm:
(a) algorithm Step 1; (b) algorithm Step 2; (c) algorithm Step 3; (d) algorithm Step 3—magnification; (e) algorithm Step 1; (f) algorithm Step 2
with the same set of random numbers for the positive and neg- Number of Independent Unit Cells nφ
ative perturbation. In Spall (2003), numerous admissible proba- The aim herein is to investigate whether the failure probability can
bility distributions for the perturbation direction are presented be reduced if an increasing number of differing unit cells is used.
(e.g., segmented uniform, u-shaped and symmetric Bernoulli 1). The actual number of unit cells, N, is kept fixed, whereas the num-
Downloaded from ascelibrary.org by MICHIGAN TECHNOLOGICAL UNIVERSITY on 09/07/19. Copyright ASCE. For personal use only; all rights reserved.
Herein, the symmetric Bernoulli 1 distribution is adopted, result- ber of design variables is increased. The unit-cell properties are then
ing in possible values for the perturbation vector Δnφ merely deduced from the design variables as already discussed.
The results are visualized in Table 2 and Figs. 12 and 13. The
Δnφ ¼ ½δ 1 ; : : : ; δ nφ ; with δ i ¼ f−1; 1g ∀ i ð27Þ highest failure probability is observed in the unprotected case
(kext ¼ ∞ s−2 ), followed by the disconnected case (kint ¼ 0 s−2 ),
The algorithm is initialized at the center of the last identified whereas the lowest failure probability, and therefore optimal de-
subset and normalizes the search space to be a unit radius nφ ball. sign, is achieved when all unit-cell properties are chosen independ-
This is beneficial if the subsets extend differently in the various ently. Fig. 12 reveals why a higher number of different cells does
dimensions. not necessarily imply a positive influence on the overall response.
Four illustrative runs of the algorithm starting at the center of the Each resonator effectively reduces one peak response value. In this
last identified subset φ ¼ ½0.4; 0.6 are shown in Fig. 11. It is a context, the SM does not create effective band gaps, i.e., locations
direct continuation of the last identified subset in the preceding in the system’s FRF that are smaller than 0 dB, but acts more like a
paragraph’s SSO example. The algorithm takes different paths form of damping.
but converges to similar optima, φ . It is further not forced to re- The limit state function in this case considers only the absolute
main within the last identified subset. A boundary condition is en- acceleration as a design criterion. Whereas the estimated optimum
forced only if the initial design space, Φ, is left. design variables, φ , are not guaranteed to be the optimal design
choice if relative displacement were a design criterion, it is worth
noting that the relative displacements are also reduced by the present
Parametric Study variable choice. That is, the displacement response is reduced with
respect to the initial unprotected system rather than amplified for the
optimal acceleration design variables, φ , as seen in Fig. 13.
Frequency Domain Analysis
The SM chain with an attached oscillator (G2 ) is now investigated Symmetry of the Optimal Design Variables φ
in the frequency domain, where the attached oscillator serves as It was already observed that the optimal design variables lie at
a simplified model for a protected structure. The optimization is points that lead to the flattest FRF. A pattern that emerges in higher
carried out by the two-stage optimization framework previously design dimensions is how these points are sequentially located at
Fig. 11. Four runs of the SPSA algorithm starting at the center of the last identified subset φ ¼ ½0.4; 0.6 to pinpoint the optimal solution φ :
(a) P̂ðFjφ̂ Þ ¼ 3.42%; (b) P̂ðFjφ̂ Þ ¼ 3.46%; (c) P̂ðFjφ̂ Þ ¼ 3.52%; (d) P̂ðFjφ̂ Þ ¼ 3.49%
φ6 jf 6 Hz — — — — — 0.6081j0.65
P̂ðFjφÞ 59.2% 37.8% 24.8% 20.23% 18.03% 17.13%
Note: The failure probabilities, P̂ðFjφÞ, were estimated using Monte Carlo simulations with N ¼ 3,000 samples and a limit state function with a threshold of
ϑ ¼ 0.4 m s−1 as defined in Eq. (13).
(a) (b)
(c) (d)
(e) (f)
Fig. 12. Magnitude of the absolute acceleration FRF jHẍ ðfÞj for the optimal variables for the acceleration design criterion, found through the two-
stage optimization framework; the internal frequencies, f i , from the design variables are shown as black lines; the parameters, θ, stem from a sample
realization in the parameter space, Θ: (a) unprotected oscillator; (b) kint ¼ 0 s−2 ; (c) nφ ¼ 1; (d) nφ ¼ 2; (e) nφ ¼ 3; (f) nφ ¼ 6
peaks of the FRF obtained by the optimal design in nφ − 1 dimen- 1. The optimum in one design space, φ ∈ Φ, is not at the same
sions (Fig. 12). Assuming that the order in which the unit-cell prop- location in the mirrored design space, Φ ~ i.
erties are assigned is of secondary importance to the failure 2. The optimum is at the same location, but some variable ar-
probability, one can deduce that there are totally nφ local optima rangements are more favorable [with respect to the failure proba-
in nφ simplices that are symmetrical to our initial design space, Φ. bility, PðFjφÞ] than others.
The optimal solutions for this initial design space for an increasing Although this is no proof, the observations made in Fig. 9
number of independent unit cells are presented in Table 2. To fur- indicate that variables in the design space, Φ, do generally yield
ther explore the above assumption about the secondary importance slightly different failure probabilities from their mirrored counter-
of the property assignment order, a sensitivity analysis is performed parts. This would point to the adoption of the second explanation.
by changing the order of the design variables and observing the So, to identify the global optimum, it would be necessary to inves-
failure probabilities, as shown in Table 3. This effectively places tigate the failure probability in the mirrored design spaces as well.
~ The analysis is con- This would mean expanding the initial design space from a simplex
the solution in mirrored design spaces, Φ.
to a hypercube. The high number of local optima, in high dimen-
ducted for the nφ ¼ 2 and three-dimensional design spaces only,
sions, would, however, require the use of different types of optimi-
because the nφ ¼ 6 dimensional design space already has 6! − 1 ¼
~ i . The results show that the failure zation algorithms.
719 mirrored design spaces, Φ
probability is in fact not the same if the results lie in the alternate External Stiffness kext Influence on φ
design spaces, Φ ~ i , (even when considering the COV of 1%). Two The optimization of the previous section is repeated for different
possible explanations for this are external stiffness values (k~ ext ¼ 100 , 500, and 1,000 s−2 ), and
(c) (d)
(e) (f)
Fig. 13. Magnitude of the relative displacement FRF jHu ðfÞj for the optimal variables for the acceleration design criterion, found through the two-
stage optimization framework; the internal frequencies, f i , from the design variables are shown as black lines; the parameters, θ, stem from a sample
realization in the parameter space, Θ: (a) unprotected oscillator; (b) kint ¼ 0 s−2 ; (c) nφ ¼ 1; (d) nφ ¼ 2; (e) nφ ¼ 3; (f) nφ ¼ 6
Table 3. Symmetry of the Optimal Solution, φ , by Calculating the Failure Whereas, for example, P̂ðFjφÞ reduces by 4.57% between nφ ¼ 1
Probabilities for the Optimal Design Variables from Table 2 in Different and nφ ¼ 2, the reduction between nφ ¼ 3 and nφ ¼ 6 is only
Mirrored Design Spaces, Φ or Φ ~i
0.9%. This could be explained by looking again at the FRF in
Design space nφ ¼ 2 nφ ¼ 3 Fig. 12. The SM structure only has one predominant natural fre-
φ
∈Φ P̂½Fjðφ1 ; φ2 Þ ¼ 20.23% P̂½Fjðφ1 ; φ2 ; φ3 Þ ¼ 18.03%
quency, f0 , in the seismically active range, therefore the response
φ ~1
∈Φ P̂½Fjðφ2 ; φ1 Þ ¼ 19.90% P̂½Fjðφ3 ; φ2 ; φ1 Þ ¼ 20.77% reduction occurs mostly around this frequency. After nφ ¼ 3, the
φ ~2
∈Φ — P̂½Fjðφ3 ; φ1 ; φ2 Þ ¼ 19.97% FRF is already flattened out considerably, with no major peaks re-
φ ~3
∈Φ — P̂½Fjðφ2 ; φ3 ; φ1 Þ ¼ 19.77% maining. Allowing additional independent unit cells does not help
φ ~4
∈Φ — P̂½Fjðφ2 ; φ1 ; φ3 Þ ¼ 18.53% with increasing the protection because there are no more predomi-
φ ~5
∈Φ — P̂½Fjðφ1 ; φ3 ; φ2 Þ ¼ 18.50% nant peaks to smooth out. In the case of multiple response peaks
Note: The failure probabilities P̂ðFjφÞ were estimated using Monte Carlo [e.g., multidegree-of-freedom (MDOF) protected oscillator system],
simulations with N ¼ 3,000 samples. The COV therefore is ≈1%. further failure probability reduction can be expected.
Another observation is the behavior of the optimal variable
P
mean, f̄ ¼ n1φ f i . As seen in Fig. 15, the mean tends to follow
the resulting optimal design variables, φ , are summarized in the fundamental frequency of the disconnected case, f 0 . This is not
Tables 2–5. It is observed that the failure probability decreases with surprising because the fundamental frequency is where the pre-
a higher number of independent unit cells (i.e., nφ ). Fig. 14 gives a dominant response is expected, and the response reduction is there-
visual representation of this performance. fore most effective in its proximity. The optimal frequencies tend to
A first observation is that the protection gets worse for higher lie below f0 as Fig. 4 indicates, where the band gap locations are
external stiffness, kext . More interesting, however, is the fact that shown. The band gaps, or in the finite case effective reduction
with a higher external stiffness, increasing the number of indepen- zones, start at the natural frequency of the unit cells f 0;int , but
dent unit cells does not significantly decrease the failure probability. extend to higher frequencies depending on the band gap width.
Table 4. Optimal Parameters, φ , Using Model G2 with kext ¼ 500 s−2 Determined through the Two-Stage Optimization Framework, for Different Number
of Independent Unit Cells nφ ¼ 1, 2, 3, and 6 and the Disconnected Case (kint ¼ 0 s−2 ) and the Unprotected Case (kext ¼ ∞ s−2 )
k~ ext ¼ 500 s−2
φ and P̂ðFjφÞ kext ¼ ∞ s−2 kint ¼ 0 s−2 nφ ¼ 1 nφ ¼ 2 nφ ¼ 3 nφ ¼ 6
φ1 jf 1 Hz — — 0.6614j0.73 0.6148j0.66 0.6018j0.64 0.2029j0.25
φ2 jf 2 Hz — — — 0.6883j0.78 0.6329j0.68 0.4976j0.50
φ3 jf 3 Hz — — — — 0.7113j0.82 0.4981j0.50
φ4 jf 4 Hz — — — — — 0.5153j0.52
φ5 jf 5 Hz — — — — — 0.6022j0.64
φ6 jf 6 Hz — — — — — 0.6948j0.79
P̂ðFjφÞ 59.2% 54.77% 45.63% 43.10% 42.80% 43.12%
Note: The failure probabilities P̂ðFjφÞ were estimated using Monte Carlo simulations with N ¼ 3,000 samples and a limit state function with a threshold of
ϑ ¼ 0.4 m s−1 as defined in Eq. (13).
φ6 jf 6 Hz — — — — — 0.7289j0.85
P̂ðFjφÞ 59.2% 55.63% 50.27% 49.97% 49.93% 50.33%
Note: The failure probabilities, P̂ðFjφÞ, were estimated using Monte Carlo simulations with N ¼ 3,000 samples and a limit state function with a threshold of
ϑ ¼ 0.4 m s−1 as defined in Eq. (13).
Fig. 16. Results from an optimization run using a time domain analysis
Fig. 14. Failure probability estimators for the optimal design variables, up until the last converged solution; the optimal solution estimator, φ̂ ,
φ , for different numbers of independent unit cells, nφ ; the case of displayed is the one obtained from a frequency domain optimization;
nφ ¼ 0 corresponds to the nonconnection case (kint ¼ 0 s−2 ); the fail- the number of SSO samples used were N ¼ 1,000 with a sample ratio
ure probabilities were estimated using standard MCS with N ¼ 3,000 of ρ ¼ 0.15
samples and a limit state function with a threshold of ϑ ¼ 0.4 m s−1 as
defined in Eq. (13)
Fig. 15. Location of the optimal frequencies, f , depending on the external stiffness, kext , for different number of design variables, nφ ; the plots also
include the fundamental frequency for the disconnected case (kint ¼ 0 s−2 ) as a function of the external stiffness; additionally, the mean of the optimal
frequencies, f̄ , is shown: (a) nφ ¼ 1; (b) nφ ¼ 2; (c) nφ ¼ 3; (d) nφ ¼ 6
(a) (b)
(c) (d)
(e) (f)
(g) (h)
(i) (j)
(k) (l)
Fig. 18. Comparison of the relative displacements between the internal and external mass of each unit cell in the six cell SM using model G2 and
linear and nonlinear internal springs; for brevity, the relative displacements are numbered as follows: ui ¼ ui;2 − ui;1 ; the design variables for the unit
cells were chosen according to φ from M2 for the case of nφ ¼ 6: (a) uðtÞ ¼ u1;2 − u1;1 ; (b) R1 ; (c) uðtÞ ¼ u2;2 − u2;1 ; (d) R2 ; (e) uðtÞ ¼ u3;2 − u3;1 ;
(f) R3 ; (g) uðtÞ ¼ u4;2 − u4;1 ; (h) R4 ; (i) uðtÞ ¼ u5;2 − u5;1 ; (j) R5 ; (k) uðtÞ ¼ u6;2 − u6;1 ; (l) R6
Nonlinear Internal Springs probability, but also indirectly determines the location of the optimal
As a final investigation, the introduction of nonlinear springs that resonator frequencies. When considering soil as the medium be-
can potentially be used to obtain a more realistic system response is tween the individual unit cells, additionally the linearity assumption
examined. It must be noted that performing optimization for non- of the external springs is not justified, and their nonlinearity has to
linear SM structures would imply the inclusion of the nonlinear be taken into account.
parameters in the uncertain parameter vector, θ, but at a prohibitive Initially, the goal was to carry out the optimization in the time
computational cost for the time domain analysis. To this end, domain, owing to computational limitations, however, the fre-
the system model, G2 , with the optimized model parameters, φ , quency domain was chosen as a viable alternative. This is not nec-
for the nφ ¼ 6 independent unit-cell case with k~ ext ¼ 100 s−2 is essarily a bad choice, because the frequency reduction zones are
reexamined with the integration of nonlinear internal springs. A only visible in the frequency domain, it does however prohibit the
hysteretic Bouc-Wen model (Charalampakis and Koumousis 2008) use of nonlinear models and does not take into account the transient
is utilized, with post yield stiffness kpost ¼ 10%kint , β ¼ 1, response of the structure. For linear systems, the frequency domain
γ ¼ 0.5, n ¼ 3, and fy ¼ 0.01 m s−2 . approach delivers satisfying results.
Fig. 17 indicates how the absolute acceleration of the protected From an algorithmic standpoint, the optimization becomes more
linear oscillator changes depending on whether linear or nonlinear challenging with an increase in dimensionality, as for every added
internal oscillator springs are used. A slight reduction in the effec- independent resonator property one expects a factorial increase in
tiveness of the SM in protecting the linear oscillator is detected. local optimal solutions. To circumvent this problem, the authors
When the oscillator reaches its yield deformation, dy , its effective chose to optimize only for resonators that have an increasing inter-
stiffness is greatly reduced to 10% of its original value. This re- nal stiffness. This places the design space in a simplex with only
duces the oscillator’s natural frequency and moves it away from one expected optimum. This is a limitation that should be addressed
the optimized frequencies. If the frequencies had been tuned to in future works.
higher frequencies, this might have a positive effect on the SM ef- On a similar note, the arrangement of the resonators has to be
fectiveness. The internal springs are ordered with an ascending considered in greater detail. In this paper, these were arranged as
internal stiffness (Fig. 18), which is seen in the hysteretic curves. homogeneously as possible, to obtain a more symmetric objective
As expected, the stiffer oscillators attract more force, and therefore function and thus justify the reduction of the design space to the
deform more than the soft oscillators. This becomes especially ap- mentioned simplex. This assumption will be relaxed and form a
parent in the relative displacement, u6 , where the peak deformation further design parameter in future work.
is approximately twice the deformation in the linear case.
ϵ= seismic excitation;
where M ∈ Rnφ ×nφ is positive-definite symmetric matrix defining
η= model prediction error;
the size and orientation of the hyperellipse and c ∈ Rnφ is its center.
Note that the above can be equivalently expressed as ζ= damping ratio;
Θ= parameter space;
E‘ ¼ fφ ¼ ðφ1 ; : : : ; φnφ Þ ∈ Rnφ jjjL⊺ ðφ − cÞjj ≤ 1g ð30Þ θ= parameter vector;
ξ= system response;
where L = Cholesky decomposition of the matrix, M. ση = standard deviation modeling error;
Φ= design variable space;
Hyperellipse Segment Φ ~ = mirrored design space;
In the first step of the SSO algorithm, a subset inside the simplex is φ= design variable vector;
identified by performing the optimization in Eq. (24). This corre- φ = optimal design variable vector;
sponds to finding the subset with the largest volume, that contains a ω̂ = equivalent natural frequency;
high enough samples ratio, ρ. This challenging optimization gets ω0;int = internal fundamental frequency; and
increasingly harder in high dimensions. Already in low dimensions, ωint = internal natural frequency.
it is difficult for ellipse shapes to cover the whole simplex.
Problems arise especially in the corners, which cannot be reached
without reducing the samples ratio, ρ. For this reason, the chosen References
subset shape is a hyperellipse segment.
This segment is defined by Achaoui, Y., Ungureanu, B., Enoch, S., Brûlé, S., and Guenneau, S. (2016).
“Seismic waves damping with arrays of inertial resonators.” Extreme
D ¼ S ∩ E ≔ fφ ¼ ðφ1 ; : : : ; φnφ Þ Mech. Lett., 8, 30–37.
Au, S.-K., and Beck, J. L. (2003). “Subset simulation and its application
∈ Rnφ j½φi < φiþ1 ∀ i ∧ ½ðφ − cÞ⊺ Mðφ − cÞ ≤ 1g ð31Þ
to seismic risk based on dynamic analysis.” J. Eng. Mech., 10.1061
/(ASCE)0733-9399(2003)129:8(901), 901–917.
that is, it is defined by the intersection of points that lie within the
Beck, J. L., and Katafygiotis, L. S. (1998). “Updating models and
simplex, S, and the ellipse, E.
their uncertainties. I: Bayesian statistical framework.” J. Eng. Mech.,
10.1061/(ASCE)0733-9399(1998)124:4(455), 455–461.
Boore, D. M. (2003). “Simulation of ground motion using the stochastic
Acknowledgments method.” Pure Appl. Geophys., 160(3), 635–676.
Charalampakis, A., and Koumousis, V. (2008). “Identification of Bouc–
This paper is based on a master’s thesis carried out during a stay
Wen hysteretic systems by a hybrid evolutionary algorithm.” J. Sound
of one of the authors as a visiting student researcher at Caltech. Vibr., 314(35), 571–585.
The authors thank Prof. Dr. Alexandros Taflanidis for his support Cheng, Z., Shi, Z., Mo, Y. L., and Xiang, H. (2013). “Locally resonant
during the implementation of the subset optimization algorithm. periodic structures with low-frequency band gaps.” J. Appl. Phys.,
Additional thanks also to Prof. Dr. Apostolos Papageorgiou for 114(3), 033532.
his clarifying remarks on the authors’ questions about the synthetic Chey, M. H., Chase, J. G., Mander, J. B., and Carr, A. J. (2010). “Semi-
time histories algorithm. active tuned mass damper building systems: Application.” Earthquake
Eng. Struct. Dyn., 39(1), 69–89.
Chopra, A. K. (2007). Dynamics of structures: Theory and appli-
Notation cations to earthquake engineering, Prentice Hall, Upper Saddle River,
NJ.
The following symbols are used in this paper: Colombi, A., Colquitt, D., Roux, P., Guenneau, S., and Craster, R. (2016a).
BI = base isolation; “A seismic metamaterial: The resonant metawedge.” Sci. Rep., 6,
CRN = common random numbers; 27717.
Colombi, A., Roux, P., Guenneau, S., Gueguen, P., and Craster, R. (2016b).
E = stochastic excitation model;
“Forests as a natural seismic metamaterial: Rayleigh wave bandgaps
fex = excitation frequency; induced by local resonances.” Sci. Rep., 6, 19238.
G = stochastic system model; Dertimanis, V. K., Antoniadis, I. A., and Chatzi, E. N. (2016). “Feasibility
g = limit state function; analysis on the attenuation of strong ground motions using finite peri-
g~ = augmented limit state function; odic lattices of mass-in-mass barriers.” J. Eng. Mech., 10.1061/(ASCE)
ĝk = gradient estimator; EM.1943-7889.0001120, 1–10.
Finocchio, G., et al. (2014). “Seismic metamaterials based on isochronous
I F = failure indicator function;
mechanical oscillators.” Appl. Phys. Lett., 104(19), 191903.
j = stochastic performance evaluation model; Harvey, P. S., Jr., and Kelly, K. C. (2016). “A review of rolling-type seismic
k̂ = equivalent chain stiffness; isolation: Historical development and future directions.” Eng. Struct.,
kint = unit cell internal mass; 125, 521–531.
optimization with stochastic approximation using common.” Manage. chastically robust structural control.” Ph.D. thesis, California Institute
Sci., 45(11), 1570–1578. of Technology, Pasadena, CA.
Taflanidis, A., and Beck, J. L. (2008a). “An efficient framework for optimal
Krödel, S., Thomé, N., and Daraio, C. (2015). “Wide band-gap seismic
robust stochastic system design using stochastic simulation.” Comput.
metastructures.” Extreme Mech. Lett., 4, 111–117.
Methods Appl. Mech. Eng., 198(1), 88–101.
Makris, N., and Gazetas, G. (1992). “Dynamic pile-soil-pile interaction. II:
Taflanidis, A., and Beck, J. L. (2008b). “Stochastic subset optimization for
Lateral and seismic response.” Earthquake Eng. Struct. Dyn., 21(2),
optimal reliability problems.” Probab. Eng. Mech., 23(2–3), 324–338.
145–162.
Taflanidis, A., and Beck, J. L. (2009). “Stochastic subset optimization
Maldovan, M. (2013). “Sound and heat revolutions in phononics.” Nature, for reliability optimization and sensitivity analysis in system design.”
503(7475), 209–217. Comput. Struct., 87(5–6), 318–331.
Mavroeidis, G. P., and Papageorgiou, A. S. (2003). “A mathematical Taflanidis, A., Scruggs, J. T., and Beck, J. L. (2008). “Probabilistically
representation of near-fault ground motions.” Bull. Seismol. Soc. Am., robust nonlinear design of control systems for base-isolated structures.”
93(3), 1099–1131. Struct. Control Health Monit., 15(5), 697–719.
Miniaci, M., Krushynska, A., Bosia, F., and Pugno, N. (2016). “Large scale Wagner, P.-R., Dertimanis, V. K., Chatzi, E. N., and Antoniadis, I. A.
mechanical metamaterials as seismic shields.” New J. Phys., 18(8), (2016a). “Design of metamaterials for seismic isolation.” Conf. Proc.
083041. Society for Experimental Mechanics Series, Vol. 2, Orlando, FL,
Palermo, A., Krödel, S., Marzani, A., and Daraio, C. (2016). “Engineered Springer, New York, 275–287.
metabarrier as shield from seismic surface waves.” Sci. Rep., 6, 39356. Wagner, P.-R., Dertimanis, V. K., Chatzi, E. N., and Antoniadis, I. A.
Papadimitriou, C., Beck, J. L., and Katafygiotis, L. S. (1997). “Asymp- (2016b). “On the feasibility of structural metamaterials for seismic-
totic expansions for reliabilities and moments of uncertain dynamic sys- induced vibration mitigation.” Int. J. Earthquake Impact Eng., 1(1–2),
tems.” J. Eng. Mech., 10.1061/(ASCE)0733-9399(1997)123:12(1219), 20–56.
1219–1229. Zhu, J., et al. (2013). “Acoustic rainbow trapping.” Sci. Rep., 3(1728), 1–6.