Strata Technical Manual: Albert R. Kottke, Xiaoyue Wang, and Ellen M. Rathje October 16, 2019
Strata Technical Manual: Albert R. Kottke, Xiaoyue Wang, and Ellen M. Rathje October 16, 2019
1
https://www.gnu.org/licenses/gpl-3.0-standalone.html
i
Acknowledgements
ii
Contents
Abstract i
Acknowledgements ii
Contents iii
List of Figures v
1 Introduction 1
iii
CONTENTS iv
4 Using Strata 49
4.1 Strata Particulars . . . . . . . . . . . . . . . . . . . . . . . . . 49
4.1.1 Auto-Discretization of Layers . . . . . . . . . . . . . . 49
4.1.2 Interaction with Tables . . . . . . . . . . . . . . . . . . 51
4.1.3 Non-Linear Curves . . . . . . . . . . . . . . . . . . . . 51
4.1.4 Recorded Motion Dialog . . . . . . . . . . . . . . . . . 55
4.1.5 Output Widget . . . . . . . . . . . . . . . . . . . . . . 58
4.1.6 Command-line Operation . . . . . . . . . . . . . . . . 58
4.2 Examples . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
4.2.1 Example 1: Basic time domain . . . . . . . . . . . . . 61
4.2.2 Example 2: RVT and Site Variation . . . . . . . . . . . 64
Bibliography 67
Index 70
List of Figures
v
LIST OF FIGURES vi
3.2 Two variables with a correlation coefficient of: (a) 0.0, (b),
0.99, and (c) -0.7. . . . . . . . . . . . . . . . . . . . . . . . . . . 34
3.3 A layering profile consisting of 8 layers modeled by a ho-
mogeneous Poisson process with a rate of 1. . . . . . . . . . . 37
3.4 Toro (1995) layering model. (a) The occurrence rate (λ) as a
function of depth (d), and (b) the expected layer thickness
(h) as a function of depth. . . . . . . . . . . . . . . . . . . . . 38
3.5 Transformation between a homogeneous Poisson process
with rate 1 to the Toro (1995) non-homogeneous Poisson
process. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
3.6 A layering simulated with the non-homogeneous Poisson
process defined by Toro (1995). . . . . . . . . . . . . . . . . . 40
3.7 Ten generated shear-wave velocity (vs ) profiles for a USGS
C site class. (a) Using generic layering and median vs , (b)
using user defined layering and median vs . . . . . . . . . . . 46
3.8 Generated nonlinear properties assuming perfect negative
correlation. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
4.1 Soil profile at the Sylmar County Hospital Parking Lot site (Chang,
1996). The mean effective stress (σ0m ) is computed assuming
a k0 of 1/2 and a water table depth of 46 meters. . . . . . . . . 62
vii
Chapter 1
Introduction
1
https://www.gnu.org/licenses/gpl-3.0-standalone.html
1
Chapter 2
2
CHAPTER 2. SITE RESPONSE ANALYSIS 3
plex wave number (k∗ ) in equation (2.2) is related to the shear modulus (G),
damping ratio (D), and mass density (ρ) of the soil using:
ω
k∗ = ∗ (2.2)
vs
s
G∗
v∗s = (2.3)
ρ
√
G∗ = G 1 − 2D2 + ı2D 1 − D2 ' G(1 + ı2D) (2.4)
G∗ and v∗s are called the complex shear modulus and complex shear-wave
velocity, respectively. If the damping ratio (D) is small (< 10 − 20%),
then the approximation of the complex shear modulus in equation (2.4)
is appropriate. Strata uses the complete definition of the complex shear-
modulus, not the approximation, in the calculations.
Equation 2.1 applies only to a single layer with uniform soil proper-
ties and the wave amplitudes (A and B) can be computed from the layer
boundary conditions. For a layered system, shown in Figure 2.2, the
wave amplitudes are calculated using recursive formulas developed by
maintaining compatibility of displacement and shear stress at the layer
boundaries. Using these assumptions, the following recursive formulas
are developed (Kramer, 1996):
Am+1 = 21 Am 1 + α∗m exp ıkm ∗
hm + 21 Bm 1 − α∗m exp −ıkm
∗
hm
(2.5)
Bm+1 = 12 Am 1 − α∗m exp ıkm ∗
hm + 21 Bm 1 + α∗m exp −ıkm
∗
hm
CHAPTER 2. SITE RESPONSE ANALYSIS 4
1 A1 6?
B1 ρ1 h1 G1 D1
2 A2 6?
B2 ρ2 h2 G2 D2
m Am 6?Bm ρm hm Gm Dm
m + 1 Am+1 6?
Bm+1 ρm+1 hm+1 Gm+1 Dm+1
n An 6?
Bn ρn hn Gn Dn
where m is the layer number, hm is the layer height and α∗m is the complex
impedance ratio. The complex impedance ratio is defined as:
∗
km G∗m ρm v∗s,m
α∗m = ∗ = (2.6)
km+1 G∗m+1 ρm+1 v∗s,m+1
The transfer function for the site with the properties presented in Table 2.1
is shown in Figure 2.3. The locations of the peaks in the transfer function
are controlled by the modes of vibration of the soil deposit. The peak at the
lowest frequency represents the fundamental (i.e. first) mode of vibration
and results in the largest amplification. The peaks at higher frequencies
are the higher vibrational modes of the site. The first natural frequency of
a site is inversely related to the site period, where the site period is defined
as (Kramer, 1996):
4 · hsoil
Ts = (2.8)
vs
In equation (2.8), hsoil is the total height of the soil and vs average velocity
of the site.
For the example site (Table 2.1), the site period is calculated to be 0.57
s which corresponds to a natural frequency of 1.75 Hz. In the transfer
function (Figure 2.3), the peak with the highest amplification occurs at this
frequency. The amplitudes of the peaks are controlled by the damping
ratio of the soil. As the damping of the system increases, the amplitudes
of the peaks decrease, which results in less amplification at the surface.
The response at the layer of interested is computed by multiplying
the Fourier amplitude spectrum of the input rock motion by the transfer
function:
Ym (ω) = TFm,n (ω) · Yn (ω) (2.9)
where Yn is the input Fourier amplitude spectrum at layer n and Ym is the
Fourier amplitude spectrum at the top of the layer of interest. The Fourier
amplitude spectrum of the input motion can be defined using a variety of
methods and is discussed further in Sections 2.2.1 and 2.2.2.
One issue that must be considered is that the input Fourier spectrum
typically represents a motion at the ground surface, where the upgoing
and downgoing wave amplitudes are equal (A1 = B1 ), not at the base of
CHAPTER 2. SITE RESPONSE ANALYSIS 6
10
Surface / Within
7.5 Surface / Outcrop
|TF|
2.5
0
0 5 10 15 20 25
Frequency (Hz)
Figure 2.3: The input to surface transfer functions site in Table 2.1 consid-
ering different types of input.
a soil deposit, where the wave amplitudes are not equal (Figure 2.4). The
change in boundary conditions ( An = Bn for the surface, An , Bn at the
base of a soil deposit) must be taken into account. The motions at any
free surface are referred to as outcrop motions and their amplitudes are
described by twice the amplitude of the upward wave (2A). Equation (2.7)
can be modified to transfer an outcrop motion to a surface motion. This
result is obtained by multiplying equation (2.7) by a transfer function that
takes the outcrop motions and makes it a within motion at the base of the
soil column.
to an outcrop motion a second transfer function is required to translate
from an outcrop motion to a within motion. The combined transfer function
CHAPTER 2. SITE RESPONSE ANALYSIS 7
Figure 2.4: Outcrop describes upward and downward waves being equal,
within is used when the upward and downward waves are not equal.
is defined as:
An + Bn Am + Bm Am + Bm
TFm,n (ω) = · = (2.10)
2 · An An + Bn 2 · An
| {z } | {z } | {z }
outcrop→within within→layern outcrop→layern
0.02
0.01
Shear Strain (%)
−0.01
Figure 2.5: An example of the strain-time history and effective strain (γeff ).
where ρ is the mass density of the site, and vs is the measured shear-wave
velocity. Characterizing the nonlinear behavior of G and D is achieved
through modulus reduction and damping curves that describe the varia-
tion of G/Gmax and D with shear strain (discussed in the next section). Using
the initial dynamic properties of the soil, equivalent-linear site response
analysis involves the following steps:
1. The wave amplitudes (A and B) are computed for each of the layers
5. The strain compatible shear modulus and damping ratio are recal-
culated based on the new estimate of the effective strain within each
layer.
6. The new nonlinear properties (G and D) are compared to the previous
iteration and an error is calculated. If the error for all layers is below
a defined threshold the calculation stops.
After the iterative portion of the program finishes, the dynamic response
of the soil deposit is computed.
−10
−20
−30
−40
Depth (m)
−50
−60
−70
−80
−90
−100
0 100 200 300 400 500 600 700 800
Shear−Wave Velocity (m/s)
damping curves for soil are shown in Figure 2.7. These curves show a
decrease in the soil stiffness and an increase in the damping ratio with an
increase in shear strain.
The modulus reduction and damping curves may be obtained from lab-
oratory measurements on soil samples or derived from empirical models
based on soil type and other variables. One of the most comprehensive
empirical models was developed by Darendeli (2001) and is included with
Strata. The model expands on the hyperbolic model presented by Hardin
and Drnevich (1972) and accounts for the effects of confining pressure (σ00 ),
plasticity index (PI), over-consolidation ratio (OCR), frequency ( f ), and
number of cycles of loading (N) on the modulus reduction and damping
curves.
In the Darendeli (2001) model, the shear modulus reduction curve is a
hyperbola defined by:
G 1
= (2.13)
Gmax 1 + γ a
γr
where a is 0.9190, γ is the shear strain and γr is the reference shear strain.
CHAPTER 2. SITE RESPONSE ANALYSIS 12
25
20
0.8
Damping (%)
15
G/Gmax
0.6
10
0.4
0.2 5
0 0
−4 −3 −2 −1 0 −4 −3 −2 −1 0
10 10 10 10 10 10 10 10 10 10
Strain (%) Strain (%)
σ00
!0.3483
γr = 0.0352 + 0.0010 · PI · OCR0.3246 (2.14)
pa
where σ00 is the mean effective stress and pa is the atmospheric pressure
in atm. In the model, the damping ratio is calculated from the minimum
damping ratio at small strains (Dmin ) and from the damping ratio associ-
ated with hysteretic Masing behavior (DMasing ). The minimum damping is
calculated from:
Dmin (%) = (σ00 )−0.2889 0.8005 + 0.0129 · PI · OCR−0.1069 1 + 0.2919 ln f
(2.15)
where f is the excitation frequency (Hz). The computation of the Masing
damping requires the calculation of the area within the stress-strain curve
predicted by the shear modulus reduction curve. The integration can be
approximated by:
where: γ+γ
100
γ − γ r ln γr
r
Dmasing,a=1 (%) =
4 − 2 (2.17)
π γ 2
γ+γ
r
25
1
σ0‘ = 1.00 atm
20 OCR = 1.00
PI = 0
0.8
Damping (%)
15
G/Gmax
0.6
10
0.4
0.2 5
0 0
−4 −3 −2 −1 0 −4 −3 −2 −1 0
10 10 10 10 10 10 10 10 10 10
Strain (%) Strain (%)
Figure 2.8: The nonlinear soil properties predicted by the Darendeli (2001)
model.
25
1
σ0‘ = 1.00 atm
20 OCR = 1.00
PI = 0
0.8 Damping (%)
15
G/Gmax
0.6
10
0.4
0.2 5
0 0
−4 −3 −2 −1 0 −4 −3 −2 −1 0
10 10 10 10 10 10 10 10 10 10
Strain (%) Strain (%)
Figure 2.9: The mean and mean ±σ nonlinear soil properties predicted by
Darendeli (2001).
CHAPTER 2. SITE RESPONSE ANALYSIS 16
fNyquist 1
∆f = = (2.24)
N/2 − 1 2∆t (N/2 − 1)
After the FAS of the motion has been computed it is possible to perform
site response analysis with the motion. The following is a summary of the
CHAPTER 2. SITE RESPONSE ANALYSIS 17
steps to compute the surface acceleration time-series for the site described
in Table 2.1:
2. Compute the input FAS with the Fast Fourier transformation (FFT)
(Figure 2.10b, only amplitude is shown).
3. Compute the transfer function for the site properties (Figure 2.10c,
only amplitude is shown).
1
(a)
Accel (g)
−1
5 7.5 10 12.5 15 17.5 20
Time (s)
Fourier Amp. (g-s)
75
(b)
50
25
0
0 2.5 5 7.5 10 12.5 15
Frequency (Hz)
4
(c)
|T F |
0
0 2.5 5 7.5 10 12.5 15
Frequency (Hz)
Fourier Amp. (g-s)
75
(d)
50
25
0
0 2.5 5 7.5 10 12.5 15
Frequency (Hz)
1
(e)
Accel (g)
−1
5 7.5 10 12.5 15 17.5 20
Time (s)
Consider a time varying signal x(t) with its associated Fourier ampli-
tude spectrum, X( f ). The root-mean-squared value of the signal (xrms )
is a measure of its average value over a given time period, Trms , and is
computed from the integral of the times series over that time period:
s
Z Trms
1
xrms = [x(t)]2 dt (2.25)
Trms 0
Parseval’s theorem related the integral of the time series to the integral of
its Fourier Transform, such that Equation 2.25 can be written in term of the
FAS of the signal:
s Z ∞ r
2 m0
xrms = |X( f )| d f =
2 (2.26)
Trms 0 Trms
where m0 is defined as the zero-th moment of the FAS. The n-th moment
of the FAS is defined as:
Z ∞
n
mn = 2 2π f |X( f )|2 d f (2.27)
0
The peak factor (PF) represents the ratio of the maximum value of the
signal (xmax ) to its rms value (xrms ), such that if xrms and PF are known, then
xmax can be computed using:
xmax = PF · xrms (2.28)
Cartwright and Longuet-Higgins (1956) studied the statistics of ocean
wave amplitudes, and considered the probability distribution of the max-
ima of a signal to develop expressions for the PF in terms of the charac-
teristics of the signal. Cartwright and Longuet-Higgins (1956) derived an
integral expression for the expected values of the peak factor in terms of
the number of extrema (Ne ) and the bandwidth (ξ) of the time series (Boore,
2003): Z ∞
√ h iNe
E[PF] = 2 1 − 1 − ξ exp −z2 dz (2.29)
0
where the bandwidth is defined as:
s
m22
ξ= (2.30)
m0 m4
CHAPTER 2. SITE RESPONSE ANALYSIS 20
Boore and Joyner (1984) illustrated the need to modify the duration
used in the rms calculation when considering oscillator responses, and
they introduced the concept of an rms duration (Trms ). The rms duration
requires modification for spectral acceleration to account for the enhanced
duration due to the oscillator response. Generally, adding the oscillator
duration to the ground motion duration will suffice, except in cases where
the ground motion duration is short (Boore and Joyner, 1984). Boore and
Joyner (1984) recommend the following expressions to define Trms :
γn
!
Trms = T gm + To n (2.32)
γ +α
T gm
γ= (2.33)
Tn
Tn
To = (2.34)
2πβ
where To is the oscillator duration, Tn is the oscillator natural period, and
β is the damping ratio of the oscillator. Based on numerical simulations,
Boore and Joyner (1984) proposed n = 3 and α = 31 for the coefficients in
Equation 2.32.
Calculation of the duration for use in RVT analysis can be done using
seismological source theory or empirical models. Boore (2003) recom-
mends the following description of ground motion duration (T gm ) for the
Western United States:
1
T gm = + 0.05R (2.35)
f0 |{z}
|{z} Path duration, Tp
Source duration, Ts
For the Eastern United States, Campbell (2003) proposes that the path
duration effect be distance dependent:
0 R ≤ 10km
10km < R ≤ 70km
0.16R
Tp =
(2.36)
−0.03R 70km < R ≤ 130km
R > 130km
0.04R
Trms S2a, fn Z fn
1
|Y( fn )|2 ≈ R ∞
2
− |Y( f )|2
d f (2.37)
2 PF
2
0
|H f n
( f )| d f − fn 0
Within Equation 2.37, the integral of the transfer function is constant for a
given natural frequency and damping ratio (β), allowing the equation to
be simplified to (Gasparini and Vanmarcke, 1976):
Trms S2a, fn Z fn
1
|Y( fn )|2 ≈ 2
− |Y( f )|2
d f (2.38)
π 2 PF
fn 4β − 1 0
The peak factors in Equation 2.38 depend on the moments of the FAS
which is currently undefined. So the peak factors for all natural frequencies
are initially assumed to be 2.5. Equation 2.38 is applied first to the spectral
acceleration of the lowest frequency (longest period) provided by the user.
At this frequency, the FAS integral term in Equation 2.38 can be assumed
to be equal to zero. The equation is then applied at successively higher
frequencies using the previously computed values of |Y( f )| to assess the
integral.
After an initial estimate of the FAS is developed using Equation 2.38
with assumed constant peak factors, it would be possible to recompute
the peak factors for each period. The variable peak factors would provide
a better estimate of the FAS from the target response spectrum. This
approach was originally implemented, but the FAS based on the variable
peak factors still resulted in a response spectrum that was more than 10%
different than the target response spectrum at short periods.
To improve the agreement between the RVT-derived FAS (and associ-
ated response spectrum) and the target response spectrum by correcting
the FAS by the ratio of the two response spectra. This correction technique
is possible because of the narrow band property of the SDOF transfer func-
tion. Using an iterative process, the FAS from iteration i is corrected by the
CHAPTER 2. SITE RESPONSE ANALYSIS 23
SRVT
a (f)
|Yi+1 ( f )| = target
|Yi+1 ( f )| (2.39)
Sa (f)
1. maximum of 30 iterations,
This ratio correction works very well in producing a FAS that agrees with
the target response spectrum, but the resulting FAS may have an inappro-
priate shape at some frequencies, as discussed below.
To demonstrate the inversion process consider a scenario event of a
magnitude 6.5 earthquake with a strike-slip faulting mechanism at a dis-
tance of 20 km. The target response spectrum is computed using the
Abrahamson and Silva (1997) attenuation model (Figure 2.11). An initial
estimate of the FAS is computed using the Gasparini and Vanmarcke (1976)
method and then the ratio correction algorithm is applied. This method-
ology results in a good agreement – less than 5% relative error as shown
in Figure 2.13 – with the target response spectrum (Figure 2.11), but the
associated FAS tends to curl up at low and high frequencies (curve labeled
“Ratio Corrected” in Figure 2.12). The curling up at low frequencies can
be mitigated by extending the frequency domain.
CHAPTER 2. SITE RESPONSE ANALYSIS 24
0.5
Spectral Accel. (g)
0.2
0.1
0.05
Ratio Corrected
Ratio & Extrapolated
0.02 Ratio, Extrap., & Slope Forced
Target
0.01
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10
Period (s)
Figure 2.11: The comparison between the target response spectrum and
the response spectrum computed with RVT.
CHAPTER 2. SITE RESPONSE ANALYSIS 26
10−1
10−2
|FAS| (g-s)
10−3
Ratio Corrected
Ratio & Extrapolated
Ratio, Extrap., & Slope Forced
10−4
0.1 1 10 100 1000
Frequency (Hz)
4
Relative Error (%)
Ratio Corrected
−2
Ratio & Extrapolated
Ratio, Extrap., & Slope Forced
−4
0.01 0.02 0.05 0.1 0.2 0.5 1 2 5 10
Period (s)
Figure 2.13: The relative error between the computed response spectra and
the target response spectrum.
CHAPTER 2. SITE RESPONSE ANALYSIS 28
Table 2.2: The values of the RVT calculation for the input motion.
3. Compute the transfer function for the site properties (Figure 2.14b).
4. Compute the surface FAS by applying the absolute value of the trans-
fer function to the input FAS (Figure 2.14c). Using the surface FAS
the maximum expected acceleration can be computed as presented
in Table 2.3. The calculation shows that the site response increases
the expected peak ground acceleration by approximately 34%.
CHAPTER 2. SITE RESPONSE ANALYSIS 29
100
(a)
Fourier Amp. (g-s)
10−1
10−2
10−3
10−4
0.1 0.2 0.5 1 2 5 10 20 50 100
Frequency (Hz)
10
(b)
1
|T F |
0.1
0.01
0.1 0.2 0.5 1 2 5 10 20 50 100
Frequency (Hz)
0
10
(c)
Fourier Amp. (g-s)
10−1
10−2
10−3
10−4
0.1 0.2 0.5 1 2 5 10 20 50 100
Frequency (Hz)
Figure 2.14: Random vibration theory method sequence: (a) input Fourier
amplitude spectrum, (b) transfer function from input to surface, and (c)
surface Fourier amplitude spectrum.
Value
Parameter Input Motion Surface Motion Equation
Moments of FAS (m0 , m2 , and m4 ) 0.0280, 0.0635, 2.27
93.8435, and 39.6356, and
1.7382 ·107 1.6306 ·105
Bandwidth (ξ) 0.1346 0.3895 2.30
Number of extrema (Ne ) 623.3158 92.8944 2.31
Peak factor (PF) 3.1406 2.8568 2.29
CHAPTER 2. SITE RESPONSE ANALYSIS
Table 2.3: The values of the RVT calculation for the input and surface motions.
30
Chapter 3
3.1 Introduction
A soil profile consists of discrete layers that vary in thickness based on
the properties of the soil. The layers are typically discretized based on the
soil type, recorded from borehole samples or inferred from a shear-wave
velocity profile. In seismic site response analysis, each layer is charac-
terized by a thickness, mass density, shear-wave velocity, and nonlinear
properties (G/Gmax , and D). One of the challenges in defining values for
these properties is the natural variability across a site and the uncertainty
in their measurement. Because the dynamic response of a site is dependent
on the soil properties, any variation in the soil properties will change both
the expected surface motion and its standard deviation.
In a simple system, the variability of the components can be analyt-
ically combined to quantify the variability of the complete system, thus
allowing for the expected value and variability of the system response to
be computed. In seismic site response analysis, the nonlinear response of
the system does not allow an exact analytic quantification of the variability
of the site response. Instead, an estimate of the expected surface response
and its standard deviation due to variations in the soil properties can be
made through Monte Carlo simulations. Monte Carlo simulations estimate
the response of a system by generating parameters of the system based on
defined statistical distributions and computing the response for each set
of input parameters. The following chapter introduces Monte Carlo sim-
ulations as applied to site response analysis and presents the models that
31
CHAPTER 3. VARIATION OF SITE PROPERTIES 32
x = σx · ε + µx (3.1)
0.8
Cumulative Probability
0.6
0.4
0.2
0
−3 −2 −1 0 1 2 3
ε
random variables, it may not be used when the variables are related (i.e.
correlated).
In the case of random variables that are not independent, a more compli-
cated procedure is required for the generation of values. Before discussing
the technique for generating correlated random variables, the concepts of
correlation and linear functions of random variables must be introduced.
Consider random variables x1 and x2 . The covariance of the two random
variables is defined as:
where E is the expected value and µx1 and µx2 are the mean values of x1
and x2 , respectively (Ang and Tang, 1975). The covariance quantifies the
strength of relationship between x1 and x2 . If the variables are independent
of each other (Figure 3.2a), then the covariance is zero, however a covari-
ance of zero does not necessarily indicate the variables are independent.
Instead, it indicates that the variables do not have a linear dependence.
As the covariance becomes more positive, two variables have a greater
tendency to both differ from their respective mean values in the same
direction (Figure 3.2b). Conversely, as the covariance becomes more neg-
ative, variables have a greater tendency to differ in the opposite direction
CHAPTER 3. VARIATION OF SITE PROPERTIES 34
2 2 2
X2
X2
X2
0 0 0
−2 −2 −2
−4 −4 −4
−4 −2 0 2 4 −4 −2 0 2 4 −4 −2 0 2 4
X1 X1 X1
(a) (b) (c)
Figure 3.2: Two variables with a correlation coefficient of: (a) 0.0, (b), 0.99,
and (c) -0.7.
where σx1 and σx2 are the standard deviations of x1 and x2 , respectively, and
ρx1 ,x2 is the correlation coefficient, defined as:
cov(x1 , x2 ) E[x1 x2 ] − E[x1 ]E[x2 ]
ρx1 ,x2 = = (3.5)
σx1 σx2 σx1 σx2
The correlation coefficient can range -1 to 1.
Independent random variables from a normal distribution are gener-
ated using equation (3.1) independently for each random variable. By com-
bining the multiple applications equation (3.1) into a system of equations,
the generation of two independent variables is achieved by multiplying a
vector of random variables (~ ε) by a matrix ([σ]) and adding a constant (~
µ),
defined as:
σx1 0 ε1 µ1
( ) " #( ) ( )
x1
= + (3.6)
x2 0 σx2 ε2 µ2
CHAPTER 3. VARIATION OF SITE PROPERTIES 35
Here, the first random variable (x1 ) is calculated based on the value of ε1
alone, while the second random variable (x2 ) is a function of both ε1 and
ε2 . Note that ε1 and ε2 still represent random and independent variables
generated from the standard normal distribution.
Layering Model
The layering is modeled as a Poisson process, which is a stochastic process
with events occuring at a given rate (λ). For a homogeneous Poisson
process this rate is constant, while for a non-homogeneous Poisson process
the rate varies. Generally, a Poisson process models the occurrence of
events over time, but for the layering problem the event is a layer interface
and its rate is defined in terms of length (i.e., number of layer interfaces
per meter).
In the Toro (1995) model, the layering thickness is modeled as non-
homogeneous Poisson process where the rate changes with depth (λ(d),
where d is depth). Before considering the non-homogeneous Poisson pro-
cess, first consider the simpler homogeneous Poisson process with a con-
stant rate. For a Poisson process with a constant occurrence rate (λ), the
distance between layer boundaries, or layer thickness (h), is an exponential
distribution with rate λ. The probability density function of an exponential
distribution is defined as:
λ exp(−λh) for h ≥ 0
(
f (h; λ) = (3.8)
0 for h < 0
2.5
Cumulative Depth
7.5
10
12.5
0 2 4 6 8 10
Layer Number
The coefficients a, b, and c were estimated by Toro (1995) using the method
of maximum likelihood applied to the layering measured at 557 sites,
coming mostly from California. The resulting values of a be 1.98, 10.86,
and -0.89, respectively. The occurrence rate (λ(d)) quickly decreases as
the depth increases (Figure 3.4(a)). This decrease in the occurrence rate
increases the expected thickness of deeper layers. The expected layer
thickness (h) is equal to the inverse of the occurrence rate (λ(d)) and is
shown in Figure 3.4(b). The expected thickness ranges from 4.2 m at the
surface to 59 m at a depth of 200 m.
0.25 60
(a) (b)
Expected Thickness, h (m)
Occurance Rate, λ (1/m)
0.2
40
0.15
0.1
20
0.05
0 0
0 50 100 150 200 0 50 100 150 200
Depth (m) Depth (m)
Figure 3.4: Toro (1995) layering model. (a) The occurrence rate (λ) as a
function of depth (d), and (b) the expected layer thickness (h) as a function
of depth.
CHAPTER 3. VARIATION OF SITE PROPERTIES 39
Using Equations 3.11 and 3.12 the cumulative rate for the Toro (1995)
modeled is defined as:
Z d
(d + b)c+1
" #
bc+1
Λ(d) = a · (s + b) ds = a ·
c
− (3.13)
0 c+1 c+1
Using this equation the homogeneous Poisson process (Figure 3.3) can be
warped into a non-homogeneous Poisson process as shown in Figure 3.5.
The resulting depth profile is shown in Figure 3.6.
Velocity Model
After the layering of the profile has been established, the shear-wave ve-
locity profile can be generated. In the Toro (1995) model, the shear-wave
velocity at mid-depth of the layer is described by a log-normal distribution.
The standard normal variable (Z) of the ith layer is calculated by:
where Vi is the shear-wave velocity in the ith layer, Vmedian (di ) is the median
shear-wave velocity at mid-depth of the layer and σln Vs is the standard
deviation of the natural logarithm of the shear-wave velocity. Equation 3.15
is then solved for the shear-wave velocity of the ith layer (Vi ):
Equation 3.16 allows for the calculation of the velocity within a layer for
a given median velocity at the mid-depth of the layer, standard deviation,
and standard normal variable. In the model proposed by Toro (1995),
values for median velocity versus depth (Vmedian (d)) and standard deviation
(σln Vs ) are provided based on site class. However, in the implementation of
the Toro (1995) model in Strata the median shear-wave velocity is defined
by the user. The standard normal variable of the ith layer (Zi ) is correlated
with the layer above it, and this inter-layer correlation is also dependent on
CHAPTER 3. VARIATION OF SITE PROPERTIES 40
12.5
10
Unit Depth
7.5
2.5
0
0 50 100 150 200 250 300
Transformed Depth
100
Cumulative Depth
200
300
400
0 2 4 6 8 10
Layer Number
the site class. The standard normal variable (Zi ) of the shear-wave velocity
in the top layer (i = 1) is independent of all other layers and is defined as:
Z1 = ε1 (3.17)
where Zi−1 is the standard normal variable of the previous layer, εi is a new
normal random variable with zero mean and unit standard deviation, and
ρ is the inter-layer correlation.
Correlation is a measure of the strength and direction of a relationship
between two random variables. The inter-layer correlation between the
shear-wave velocities proposed by Toro (1995) is a function of both the
depth of the layer (d) and the thickness of the layer (h):
Designation Description
A Rock
Instrument is found on rock material (vs > 600 m/s) or
a very thin veneer (less than 5 m) of soil overlying rock
material.
Table 3.2: The USGS site categories, where vs,30 is the time weighted average
shear-velocity of the top 30 m (Toro, 1995).
CHAPTER 3. VARIATION OF SITE PROPERTIES 44
Site Classification
GeoMatrix USGS
Parameter A & B C & D A & B C & D A B C D
σlnV 0.46 0.38 0.35 0.36 0.36 0.27 0.31 0.37
ρ0 0.96 0.99 0.95 0.99 0.95 0.97 0.99 0.00
ρ200 0.96 1.00 1.00 1.00 0.42 1.00 0.98 0.50
∆ 13.1 8.0 4.2 3.9 3.4 3.8 3.9 5.0
d0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
b 0.095 0.160 0.138 0.293 0.063 0.293 0.344 0.744
Table 3.3: Coefficients for the Toro (1995) model, calculated by maximum
likelihood.
the difference between the site specific information and the generic USGS
site C median shear-wave velocity profile.
CHAPTER 3. VARIATION OF SITE PROPERTIES 45
Table 3.4: Median shear-wave velocity (m/s) based on the generic site
classification.
CHAPTER 3. VARIATION OF SITE PROPERTIES 46
Generic Layering and Velocity (USGS C) Site Specific Layering and Velocity
0 0
−10 −10
−20 −20
−30 −30
−40 −40
Depth (m)
Depth (m)
−50 −50
−60 −60
−70 −70
−80 −80
Figure 3.7: Ten generated shear-wave velocity (vs ) profiles for a USGS C
site class. (a) Using generic layering and median vs , (b) using user defined
layering and median vs .
CHAPTER 3. VARIATION OF SITE PROPERTIES 47
25
1 σ0‘ = 1.00 atm
20 OCR = 1.00
0.8 PI = 0
Damping (%)
15
G/Gmax
0.6
10
0.4
Input Median
0.2 5
Randomization Median
Randomization
0 0
−4 −3 −2 −1 0 −4 −3 −2 −1 0
10 10 10 10 10 10 10 10 10 10
Strain (%) Strain (%)
Using Strata
49
CHAPTER 4. USING STRATA 50
sublayers changes.
The maximum thickness (hmax ) of the sublayers of i-th velocity layer
is take as a fraction of the minimum wavelength to be captured by the
analysis:
vs,i
hmax,i = λfrac · λmin = λfrac · (4.1)
fmax
where λfrac is the wavelength fraction which typically varies between 1/10
and 1/5 (anything greater than 1/3 is not recommended), fmax is the max-
imum frequency of engineering interest which is typically around 20 Hz,
and Vs,i is the shear-wave velocity of the i-th layer. The actual thickness
of the sublayers is less than the maximum thickness such that the velocity
layer height divided by the sub-layer thickness is a whole number. These
parameters are defined on the General Settings tab. To prevent the lay-
ers from being auto-discretized the wavelength fraction can be increased
and the thickness velocity layers defined in the Soil Profile tab can be
reduced to an appropriate level.
In other site response programs, the location of the input motion or
the location of requested output (e.g., acceleration-time history) is gener-
ally referenced by a sublayer index. However, because the sublayers are
computed in Strata (and may change for each realization), the location is
defined in terms of the depth within the soil profile or at the top of the
bedrock. When the location is specified as Bedrock then the actual depth of
the location will change if the depth of the bedrock changes. The location
is specified with a drop down list shown in Figure 4.1 where the user can
specify the depth as Bedrock (Figure 4.1a) or a fixed depth (Figure 4.2c).
Figure 4.1: Location selection: (a) top of bedrock, (b) switching to a fixed
depth, and (c) fixed depth of 15.
CHAPTER 4. USING STRATA 51
Figure 4.2: By clicking on the button circled in red all rows in the table are selected.
52
CHAPTER 4. USING STRATA 53
Temporary Models
If you want to define a curve without adding it to the library of models,
simply select Custom from the drop-down list. Changing to the Custom
model does not clear the previous models data which allows for a model
to be modified.
CHAPTER 4. USING STRATA 54
1. Click on the File... button and select the acceleration time series
file. If the file is from the NGA database, then the remainder of
the form will be automatically completed as shown in Figure 4.5.
Regardless of the file type, the file is read and loaded into the preview
area.
3. The scale factor can be selected at this time or after the motion has
been loaded. The scale factor should result in the motion being in units
of gravity. After the motion has been loaded, the scale factor can also
be adjusted by setting the peak-ground acceleration.
4. After the form has been completed, the time-series can be viewed by
clicking on the Plot button.
Figure 4.6: Using the Output view to examine the results of a calculation.
CHAPTER 4. USING STRATA 60
be computed in Strata.
The following steps are recommended for running Strata from the
command-line:
1. Run Strata and create a model as you would typically with all of the
features that you are going to need. While it is possible to create an
input file from stratch, it is useful to have Strata create the skeleton.
2. Save the model as a JSON file. This is just a text file that uses the JSON
format, which is supported by lots of different languages (Matlab and
R incldued).
3. Use the JSON file as a template for the other profiles that you are
going to consider.
The results from the output are saved to the same file that was passed to
Strata and old results are silently replaced. Note that multiple file names
can be provided in the call to Strata. Additionally, Strata can be started
from your favorite programming language.
4.2 Examples
The following examples give a basic introduction to using Strata to perform
equivalent linear site response analysis. The examples files are found
within the examples folder in the installation path. The examples can be
opened by either double clicking on their file, or selecting them from the
Open... item from the File menu.
All examples use the deep alluvium Sylmar County Hospital Parking
Lot (SCH) site located in Southern California for the site profile. The soil
types and velocity layering of the site was proposed by Chang (1996). The
CHAPTER 4. USING STRATA 61
20
Depth (m)
40
60
80 Preferred
Minimum
Maximum
100
0 200 400 600 800
Shear-wave velocity (m/s)
Figure 4.7: The shear-wave velocity profile of the Sylmar County Hospital
Parking Lot site (Chang, 1996).
soil properties are listed in Table 4.1 with a water table at a depth of 46
meters. The nonlinear properties for each of the layers were computed
using the Darendeli (2001) empirical model with PI=0, OCR=1, and the
confining pressures listed in Table 4.1. The corresponding velocity profile
is shown in Figure 4.7. The time-averaged shear-wave velocity over the
top 30 meters was computed as 273 m/s which classifies the site as a USGS
site class C.
Table 4.1: Soil profile at the Sylmar County Hospital Parking Lot site (Chang, 1996). The mean effective
stress (σ0m ) is computed assuming a k0 of 1/2 and a water table depth of 46 meters.
62
CHAPTER 4. USING STRATA 63
Computing
After the project is fully defined, switch the Compute tab and click on the
Compute button. Information regarding the calculation will be displayed
in the window and once the calculation is complete the view will switch
to the output widget. For more information on using the output widget
see Section 4.1.5. The input widget can be selected by selected Input View
from the Window menu, or by pressing F2.
Figure 4.8: Settings to enable RVT site response and variation of the shear-
wave velocity.
hamson and Silva (1997) empirical model for a magnitude 7.0 earthquake
generated by a strike-slip fault with a distance of 20 km. The duration of
the event was computed uses the Abrahamson and Silva (1996) empirical
relationship for the duration using an normalized Arias intensity of 0.75.
Whenever a response spectrum is used for input it is important to check
that the resulting FAS looks appropriate. Strata allows the user to see both
the data and the plots of the data prior to the calculation.
Bibliography
J. N. Brune. Tectonic stress and the spectra of seismic shear waves from
earthquake. Journal of Geophysics Research, 75(26):4997–5009, 1970.
67
BIBLIOGRAPHY 68
S. W.-Y. Chang. Seismic response of deep stiff soil deposits. PhD thesis, Uni-
versity of California, Berkeley, December 1996.
H. B. Seed and I. M. Idriss. Soil moduli and damping factors for dynamic
response analyses. Technical Report EERC 70-10, Earthquake Engineer-
ing Research Center, University of California, Berkeley, CA, 1970.
G. R. Toro. Probablistic models of site velocity profiles for generic and site-
specific ground-motion amplification studies. Technical Report 779574,
Brookhaven National Laboratory, Upton, New York, 1995.
E. Vanmarcke. Random fields, analysis and synthesis. The MIT Press, Cam-
bridge, Massachusetts, 1983.
motion type
outcrop, 6
within, 7
Nyquist frequency, 16
shear modulus
complex, 3
maximum (small strain), 8
transfer function
acceleration, 5
strain, 8
wave equation, 2
70