You are on page 1of 54

Page 1 of 54

For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Shear wave velocity as a geotechnical parameter: an overview

Mahmoud N. Hussien1, 2 and Mourad Karray1


1
Department of Civil Engineering, Université de Sherbrooke, Sherbrooke (Québec) J1K 2R1, Canada
2
Department of Civil Engineering, Faculty of Engineering, Assiut University, Assiut, Egypt
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Mahmoud N. Hussien, Ph.D


1-Lecturer, Department of Civil Engineering, Faculty of Engineering, Assiut University, Assiut, Egypt
mahmoudnasser2002@aun.edu.eg

2-Post-doctoral fellow, Department of Civil Engineering, Université de Sherbrooke, Sherbrooke (Québec) J1K
2R1, Canada
Tel: +1-819-3423429
E-mail: Mahmoud.Nasser.Ahmed@USherbrooke.ca

Mourad Karray, ing., Ph.D.


Professor, Department of Civil Engineering, Université de Sherbrooke, Sherbrooke (Québec) J1K 2R1, Canada
Tel.: (819) 821-8000 (62120)
Fax: (819) 821-7974
E-mail: Mourad.Karray@Usherbrooke.ca

1
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 2 of 54

Abstract

Shear wave velocity, Vs is a soil mechanical property that can be advantageously measured in both the

field and laboratory under real and controlled conditions. The measured Vs values are customary used in

conjunction with other in situ (e.g., N-SPT and qc-CPT) and laboratory (e.g., effective confining pres-

sure, σ m′ and void ratio, e) measurements to establish abundant number of Vs-based correlations that
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

could later be utilized to augment (in some cases, replace) designated testing. An attempt is made here to

present the salient features of some existing widely-used correlations to provide the reader with a com-

prehensive understanding about the nature of these correlations and their applicability in geotechnical

engineering practices. It is recognized that the reliability of some of these empirical formulations, still in

general use today, has been questioned as they are characterized by their lack of dependence on stress

state and particle characteristics. A new Vs1-(N1)60 correlation accounts for grain sizes is highlighted by

combining a recent published Vs1-qc1 formulation and available (N1)60-qc1 relationships. The new formu-

lation is applicable to uncemented relatively young Holocene-age soil deposits. The estimated Vs1 values

based on the proposed correlation are compared with reliable laboratory and field measurements, and the

comparison shows that accounting for grain size of granular soils yields less conservative results regard-

ing the Vs values than when particle size is not considered. The prime effect of grain size was to change

the range of possible void ratios which in turn had a substantial impact on Vs values. Moreover, a new

Vs1-(N1)60 chart has been proposed allowing the practitioner to estimate Vs1 values based on a combina-

tion of data including N-SPT, e, grain size, and relative density.

Keywords: Shear wave velocity, grain size, SPT, CPTu, Id, correlation.

2
Page 3 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

1. Introduction

It is essential to properly estimate shear wave velocity, Vs of granular soils at strain levels less than 10-
3
% as its relevance for the analysis and design of various geotechnical structures is well acknowledged

(e.g., McGillivray and Mayne 2004; Choi and Stewart 2005; Patel et al. 2008; Lee et al. 2014). Exten-

sive research efforts have been carried out to study this property, mainly through well-controlled labora-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

tory experiments and field measurements. The in situ measurements of Vs, in particular have recently

been used more often due to: (i) the development of new measurement techniques, such as the seismic

cone penetration test (SCPT) (e.g., Campanella et al. (1986), and (ii) the accessibility of methods using

surface waves (spectral analysis of surface waves (SASW), modal analysis of surface waves

(MMASW)) (e.g., Karray 2010; Karray et al. 2011). These different methods provide very different res-

olutions for Vs values, and it is a difficult task to verify that the obtained shear wave velocities are cor-

rect and accurate due to the lack of reliable references for comparison. In actuality, direct determination

of Vs requires specialized equipment and technical expertise in order to ensure that the measured data are

properly obtained and interpreted.

It is helpful to compare in situ Vs values to those obtained from conventional laboratory experiments

using resonant column (RC) (e.g., Hardin and Richart 1963; Chien and Oh 2000; Fam et al. 2002; Yang

and Gu 2013), bender elements (e.g., Dyvik and Madshus 1985; Viggiani and Atkinson 1995; Brignoli

et al. 1996; Yamashita et al. 2009; Clayton 2011), and piezoelectric ring-actuator technique (e.g., Ben

Romdhan et al. 2014; Karray et al. 2015). These techniques may effectively be used to study and quanti-

fy the effects of various parameters on Vs. They proved their versatility, convenience, and usefulness

especially to augment limited field data and to develop more reliable Vs-based correlations. Field and

laboratory Vs values are routinely used in conjunction with other in situ (e.g., standard penetration test

blow count, N-SPT and cone penetration resistance, qc-CPT) and laboratory (e.g., effective confining

3
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 4 of 54

pressure, σ m′ and void ratio, e) measurements to establish a great number of Vs-based correlations that can

be incorporated into engineering studies to capitalize on the abundant data available and experience of

others (Sykora 1987). Although Vs values are always favored over estimates, correlations with penetra-

tion resistance can offer timely and economical contributions in the following situations (Gazetas 1991;

Andrus et al. 2007; Wichtmann and Triantafyllidis 2009; Brandenberg et al. 2010; Karray and Hussien
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

2015):

• For developing regional ground shaking hazard maps (site classification) based on an averaged

shear wave velocity to a depth of 30 m (Vs30). Correlation with penetration measurements can

augment Vs values as the range of Vs30 for each class is relatively large;

• To provide a check against measured Vs values in situations where a great precision in the calcu-

lation of the deposit response is required such as in liquefaction studies. In these cases, it is rec-

ommended to verify the coherence between geophysical and geotechnical measurements;

• As a screening tool for identifying where geophysical measurements would be the most benefi-

cial;

• For feasibility studies and preliminary design estimates or for final design calculations in low-

risk projects where the costs of an appropriate Vs testing are not justified;

• Calculated Vs profile based on correlations can be used to start the inversion process in Rayleigh

wave testing.

Although their importance is undeniable, it is recognized that the reliability of most correlations, cur-

rently used in engineering practice, has been questioned as they neglect some of the factors that can

strongly affect Vs values such as soil type variations as well as particle characteristics, geological age

and initial fabric of soil. For example, both experimental results and theoretical considerations have

shown that Vs is primarily a function of void ratio (e) and effective confining pressure (σ m′ ) and the

4
Page 5 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

widely used empirical formula for estimation of Vs (e, σ m′ ) is the one originally proposed by Hardin

(Hardin and Richart 1963; Hardin and Black 1966):

[1] Vs = AF (e)(σ m′ ) B

where F(e) is a void ratio function; A and B are material constants. Table 1 summarizes some experi-

mental data founded in the literature for different granular soils. Table 1 shows that the stress exponent,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

B, varies from 0.22 to 0.29; and a practical value of 0.25 was proposed by many researchers (Bui 2009).

However, classical contact mechanics solutions using the Hertz-Mindlin contact theory predict B = 0.16

(Santamarina et al. 2001). This is because the soil particles are assumed to be smooth elastic spheres. If

the contacts are considered to be an interaction of rough surfaces, the modification of theory leads to

increases in the exponent B to values that are closer to the experimental observations given in Table 1

(Yimsiri and Soga 2000).

The effect of effective stress is often removed using an overburden correction factor (e.g., Skempton

1986; Sykora 1987; Karray et al. 2011). In other words, Vs is routinely normalized for a vertical effective

stress, σ v′ , as in studies for the evaluation of the liquefaction potential (e.g., Youd et al. 2001).

0.25
P 
[2] Vs1 = Vs  a' 
σ v 

where Vs1 is the stress-normalized shear wave velocity, Pa is normal atmospheric pressure in the same

units as σ v′ (i.e., Pa ≈ 100 kPa if σ v′ is in kPa).

By substituting Eq. 1 into Eq. 2, the normalized shear wave velocity Vs1 can be expressed as:

0.25
P 
[3] Vs1 = AF (e)(σ m′ ) B  a 
 σ v′ 

when B = 0.25, Pa = 100 kPa, and σ v′ = σ m′ (3 / (1 + 2 K 0 )) , where, K0 is the coefficient of earth pressure

at rest, Vs1 can be expressed as:

5
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 6 of 54

0.25
 100 
[4] Vs1 = AF (e)  
 (3 / (1 + 2 K o )) 

Table 1 presents also the values of Vs1 calculated using Eq. 4, for two different values of K0 = 1.0

and 0.5. For the range of soils presented in Table 1, one can realize that the maximum Vs1 (i.e. Vs1 corre-

sponds to the minimum void ratio, emin) varies from 175 to 327 m/s depending on the soil type while the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

minimum Vs1 (i.e. Vs1 corresponds to the maximum void ratio, emax) varies from 114 to 214 m/s. The

values of Vs1max and Vs1min vary significantly with the soil type as there is a significant variation in parti-

cle characteristics and initial fabric as well as emin and emax values. It can be noted from Table 1 that the

variation of Vs1 between the loosest and the densest state of the same soil is not more than 132 m/s. In

other words, the normalized shear wave velocities of specimens of the same granular soil tested at dif-

ferent void ratios collapse onto a relatively narrow range. Similar note is reported by Karray et al. (2010)

based on experimental and field measurements of shear wave velocities on a wide range of granular

soils. On the contrary, some correlations found in the literature between Vs and either N-SPT or qc-CPT

give up to a 400 m/s-variation of Vs1 between the loosest and the densest state of the soil. In fact, these

empirical relationships contain no function taking into account the particle characteristics (including

particle size, particle shape, particle gradation and mineral composition) and/or they didn’t impose parti-

cle-size limits for the validity of the proposed formulation.

An attempt is made in this study to present the prominent features of some existing widely-used cor-

(emax − e)
relations between Vs and N-SPT, qc-CPT, relative density ( I d = ), mean grain size (D50), and
(emax − emin )

e to assist the reader in solving the question of applicability of these correlations to geotechnical engi-

neering practices. A new Vs1-(N1)60 correlation accounts for mean grain size, D50 is highlighted by com-

bining the recent published Vs1-qc1 formulation by Karray et al. (2011) and available (N1)60-qc1 relation-

ships, where (N1)60 and qc1 are the stress normalized penetration resistances. Similar to Karray et al.

6
Page 7 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

(2011) empirical correlation, the proposed relationship is applicable to uncemented relatively young

Holocene-age soil deposits. Particular emphasis is placed on the validation of the correlation by compar-

ing the calculated Vs1 values based on the proposed relationship with reliable laboratory Vs values on

different granular soils. The discussion is also supported by practical examples inspired from field

measurements and actual geotechnical projects. Furthermore, a new Vs1-(N1)60 chart has been proposed
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

allowing the practitioner to estimate Vs1 based on a combination of data including N-SPT, e, D50, and Id.

The primary findings as well as recommendations for everyday practice pertained to the current work

were summarized as conclusions.

2. Shear wave velocity as a geotechnical parameter

Body waves, which can travel through a continuum, are of two types: P-waves and S-waves. As depicted

in Fig. 1, P-wave, also known as primary wave tends to induce volume change and its propagation ve-

locity (Vp) in a saturated soil is practically identical to Vp in water (Richart et al. 1970). Thus Vp can be

regarded as a total stress parameter that doesn’t represent the true behavior of the soil structure. On the

other hand, shear wave (S-wave) imposes only shear deformation (Fig. 1) and its velocity (Vs) can be

considered as an effective stress parameter that is, in fact, a direct measure of the rigidity or stiffness of

the material:

[5] Gmax = ρVs2

where Gmax is the small strain shear modulus, ρ is the density of the soil. Figure 2 shows the normalized

shear wave velocity data determined, respectively from MMASW and SCPT at different sites in Québec,

Canada, and at Yarbro-Excavation, USA. It is interesting to see the increase of Vs1 values with both N

and D50. As shown in Fig. 2, Vs1 of gravely medium sand (Péribonka site) is about 260±25 m/s while that

of clayey silt (Sorel site) is about 155±10 m/s.

7
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 8 of 54

3. Basic state of knowledge on Vs-correlations

Unlike N-SPT and qc-CPT, one of the main advantages of Vs is that it can be measured using well-

controlled laboratory tests obtaining correlations with e and σ m′ . In the subsequent discussion, a synthe-

sis of the correlations established in the field (i.e. Vs-N and Vs-qc) and in laboratory (i.e. Vs-e and Vs-Id) is

presented. Rather than completely review the Vs-based correlations developed to date, only a few of the
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

more prominent studies will be examined in the subsequent sections to reveal the most important aspects

that may help in evaluating these correlations and ascertain their applicability towards an adequate de-

sign.

3.1 Field correlations between Vs and N-SPT

The Standard Penetration Test (SPT) is one of the oldest and most common in situ tests used for soil

exploration in soil mechanics and foundation engineering, because of the simplicity of the equipment

and test procedures (Anbazhagan et al. 2012). Many regression equations of N-SPT versus Vs are availa-

ble in the literature for different soils. A summary of some of these empirical relationships proposed at

different locations/countries for different soil types and using different methods of Vs values is presented

in Table 2. It should be noted that almost all of the empirical relationships listed in Table 2 use a power-

law relationship between Vs and N:

[6] Vs = aN b

where the constants a and b are determined by statistical regression of a data set. The N-values are typi-

cally not corrected for overburden stress, but sometimes are corrected for hammer energy, rod length,

and sampler inside diameter, in which case N is replaced by (N)60:

 ER 
[7] ( N ) 60 = N 
 60 

8
Page 9 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

where ER is the measured energy ratio (%). The Vs-values are also typically not corrected for overbur-

den stress. In the relationships presented in Table 2, it should be noted that the values of the correlative

constant (a) that controls the amplitude and the exponential parameter of N-value (b) which controls the

curvature of the relationship are not consistent, and these relationships can be categorized according to

the value of the correlative exponent (b) into three main groups: group I; where b < 0.45, group II; where
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

0.45 ≤ b ≤ 0.55, and group III where b > 0.55. This way of grouping may be of importance in discussing

the normalization of Vs and N with respect to effective stress. Effective confining stress, σ m′ , is known to

affect both Vs and N, and the effect is often removed using an overburden correction factor. The result-

ing stress-corrected quantities, Vs1 and N1, are routinely correlated with relative density for soils. Eqs. 2

and 8 are common overburden correction equations for Vs and N, respectively:

0.5
P 
[8] N1 = N  a' 
σ v 

where N1 is the standard penetration blow count corrected for the effective stress, Pa is normal atmos-

pheric pressure . Focusing on group II; where 0.45 ≤ b ≤ 0.55, Eq. 6 can be rewritten as:

[9] Vs ≈ aN b ≈ 0.5

Substitution of Eqs. 8 and 9 into Eq. 2 yields:

[10] Vs1 ≈ aN10.5

It is worth to note from Eq. 10 that the overburden corrections of Vs and N have no significant effect

on the constant (a) or the exponent (b ≈ 0.5) determined by statistical regression of data. Thus the func-

tional form adopted in group II is superior because it automatically removes the effect of overburden

since both Vs1 and N1 are theoretically independent of overburden stress.

3.2 Field correlations between Vs and qc-CPT

9
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 10 of 54

Shear wave velocity can also be correlated to CPT cone resistance, qc. Similar to the overburden stress

correction used for Vs, the normalized cone tip resistance for sandy soil, qc1 can be defined by (Robert-

son and Wride 1998; Robertson 2012):

0.5
P 
[11] qc1 = qc .  a 
 σ v′ 
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

where, Pa is normal atmospheric pressure in the same units as σ v′ .

A significant number of correlations have been presented over the years to relate Vs values to qc-CPT

(e.g., Baldi et al. 1986; Rix and Stokoe 1991; Mayne and Rix 1995; Karray et al. 2011; Cai et al. 2014),

among others. It is beyond the scope of this paper to discuss the details of Vs-qc correlations. The inter-

ested reader is referred to Andrus et al. (2007), Robertson (2009), and Karray and Hussien (2015).

3.3 Effect of soil state on N, qc and Vs

Of the factors identified as affecting the N, qc or Vs values, effective stress and soil state are recognized

to be the main ones. The latter factor can be expressed by either void ratio (e) or soil density index (Id).

Once normalized for the overburden pressure effect, the N, qc or Vs values are therefore likely to essen-

tially reflect the void ratio or the soil density. One of the very first efforts to investigate the effect of the

relative density of soil is what was done by Gibbs and Holtz (1957). Gibbs and Holtz (1957) performed

the first chamber tests of sands with SPT. These chamber tests were then reproduced 20 years later by

Marcuson and Bieganousky (1977). Referring to Fig. 3a, chamber testing indicates that at a constant

relative density, penetration resistance increases with increasing confining pressures. Skempton (1986)

proposed the following relationship between (N1)60-values and relative density, Id:

[12] ( N1 )60 / I d2 = Constant

where N1(60) is the blow count corrected for the effective stress and energy level used in the SPT. A con-

stant of 40 was recommended by Skempton (1986) for young, fine-grained sands and was selected by

10
Page 11 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

the CANLEX project as being representative of loose, young sand (Wride et al. 2000). Cubrinovski and

Ishihara (1999) reported that at a given relative density and overburden pressure, N1-values are higher

for sands with larger grain sizes (D50) as depicted in Fig. 3b. Their results are consistent with other re-

sults reported by Kulhawy and Mayne (1990):

[13] ( N1 )60 / I d2 = 60 + 25 log D50


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

12
[14] qc1 / I d2 =
(emax − emin )0.8

A number of studies (e.g., Muromachi 1981; Robertson et al. 1983; Seed and De Alba 1986; Andrus

and Youd 1989) have been presented in the literature to correlate N-SPT with qc-CPT, and they have

ascertained that the ratio (qc)/N increases with mean grain size. Mayne et al. (1995) suggested the fol-

lowing equation:

qc
[15] ≈ 0.544 D500.26
N 60

where qc is given in MPa and D50 in mm. Referring to Eqs. 13 and 15, one can realize that both N and qc

increase with grain size and qc is more sensitive to the change in particle size. Eq. 15 can be rewritten in

term of stress-normalized penetration indices as:

pa
qc
qc σ v′ qc1
[16] = = ≈ 0.544 D500.26
N 60 pa ( N1 )60
N (60)
σ v′

The variation of qc1/(N1)60 ratio with mean grain size, D50 can be also expressed according to Anag-

nostopoulous et al. (2003) as:

qc q
[17] = c1 ≈ 0.76429 D500.26
N 60 ( N1 )60
where qc and qc1 are given in MPa and D50 in mm.

11
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 12 of 54

3.4 Laboratory relationships

As presented in Table 1, the variation of Vs1 with soil density is often expressed in term of e. However,

Vs-Id correlations are also found in the literature. For example, in the early 1970s, Seed and Idriss (1970)

proposed a relationship between Id and Vs, later rearranged and normalized by Karray and Lefebvre
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

(2008) as:

[18] Vs1 = 25.8 I d + 25

where Vs1 is in m/s.

It can be shown, using this relationship, that if soil density index Id varies from 20 to 100% (i.e. ∆Id

=80%), the resulting Vs1 varies from 173 to 288 m/s (i.e. ∆Vs1=115). The same variation in soil density

(∆Id =80%) produces a variation in the stress-normalized SPT resistance (∆N1) of 42 according to Gibbs

and Holtz (1957) as depicted in Fig. 3b. The results of Kulhawy and Mayne (1990) and Cubrinovski and

Ishihara (1996), portrayed also in Fig. 3b, however show that not only ∆N1 is affected by the change in

Id but it is also closely related to the grain sizes. According to Kulhawy and Mayne (1990), at an ∆Id of

80%, a variation of D50 between 0.15 and 2.0 mm produces an ∆N1 between 36 and 62, respectively.

This variation in ∆N1 in turn produces a substantial variation in ∆Vs1 according to the correlations pre-

sented in Table 2. Furthermore, correlations presented in Table 1 confirm that the variation in soil densi-

ty between the loosest and densest state (i.e. ∆Id =100%) would produce different variations in the nor-

malized shear wave velocity (∆Vs1). These evidences lead us to conclude that using Eq. 18, which

doesn’t consider the variation of soil particle characteristics, would lead to inaccurate predictions of Vs1.

On the other hand, despite the enormous number of correlations relating Vs1 to e as presented in Ta-

ble 1, few recommendations were given on how to consider the influence of soil grain size and gradation

when estimating Vs. Some researchers believe that particle characteristics affect Vs through influencing

12
Page 13 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

the e (macro effect), while other believe that particle characteristics affect Vs at particle contact level

(micro effect). Along the following lines, the interrelation between e and particle characteristics will be

examined.

The void ratio e of a given soil specimen is comprised between some minimum and maximum val-

ues. These maximum and minimum possible values of the void ratio, emax and emin, can be obtained ex-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

perimentally using detailed laboratory standards (ASTM 2011a; 2011b). One of the appreciated efforts

to relate D50 to emax and emin is what was done by Cubrinovski and Ishihara (2002) based on laboratory

test data from natural silty soils, sands with clay, sands with fines, clean sands, gravelly sands, and grav-

el deposits in Japan.

[19] emin = 0.59emax + 0.035

[20] emax = 0.865 D50−0.18

where D50 is given in mm.

Menq (2003) collected data from the literature of granular soils tested in Japan and USA during the

1990s. He found that the values of emax - emin are almost constant and vary between 0.2 and 0.3 with an

average value of about 0.25 for Cu > 10, (Cu=D60/D10 is the uniformity coefficient of the grain size dis-

tribution curve). For Cu ≤ 10, the value of emax - emin increases as Cu decreases. Menq (2003) plotted also

the variation of emax - emin with D50, and pointed out that the values of emax - emin vary between 0.2 and

0.3, with an average of 0.25 for gravelly materials (D50 > 4.76 mm) and coarse sands (D50 > 2.0 mm).

On the other hand, Menq (2003) found that the values of emax - emin of medium and fine sandy materials

show higher values and a wider variation (between 0.25 and 0.65). These evidences illustrate that both

the mean grain size, D50, and uniformity coefficient, Cu, have an impact on the void ratio, and they may

have a direct or indirect effect on Vs value of soils. A summary of the effort done to investigate these

particular effects will be recounted in the following paragraphs.

13
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 14 of 54

Hardin and colleagues (Hardin and Richart (1963); Hardin and Drnevich (1972)), based on RC tests,

reported that the particle size affects Vs of sands only through influencing the void ratio function, F(e) in

Eq. 1. They classified particle size as relatively insignificant parameters for Vs of soils. In contrast, Iwa-

saki and Tatsuoka (1977), using RC tests on normally consolidated reconstituted sands with different Cu

and D50, reported that Vs of sand is strongly affected by the grain size distribution characteristics. More
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

specifically, they found that Vs decrease significantly with increasing Cu and with increasing the content

of fine soils, FC (i.e., grains smaller than 0.074 m), at constant, σ m′ and e. For poorly graded sands (Cu <

1.8, 0.16 mm ≤ D50 ≤ 3.2 mm) without FC, Iwasaki and Tatsuoka (1977) reported that the values of Vs

(e) did not depend on D50. For wider range of Cu, from 2 to 16, Chang and Ko (1982), using the RC

technique to measure Vs in 23 medium loose sand specimens with a diameter of 30 cm and D50 ranging

from 0.149 to 1.68 mm, noted that Vs does increase with D50.

Ishihara (1996) collected Vs values for various sands and gravels tested in Japan in the 1980s with dif-

ferent values of particle sizes and void ratios. He pointed out that there is a significant difference in Vs

values for different types of gravel (e.g. Vs (crushed rocked) > Vs (gravel) > Vs (sandy gravel)). Ishihara suggested

that Vs is possibly influenced by particle size. Moreover, Rollins et al. (1998), through tests in which the

content of gravel was increased from 0 to 60%, confirmed that Vs increases with gravel content. Howev-

er, the Cu-values of the tested materials were not given. The results of Menq and Stokoe (2003) from RC

tests performed to measure Vs values of 59 reconstituted specimens of natural river sands with 10 grain-

size distribution curves supported that Vs is mainly a function of D50. They noted also that the effect of

Cu on Vs is minor, and is mostly due to the way by which Cu affects void ratio. In contrast, Wichtmann

and Triantafyllidis (2009) conducted a structured program of 163 RC tests on 25 different grain-size

distributions to examine the effect of particle-size distribution on Vs of quartz sand. Their data showed

that for a constant value of e and in the investigated range (0.1 mm ≤ D50 ≤ 6 mm, 1.5 ≤ Cu ≤ 8), the val-

14
Page 15 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

ue of Vs decreased markedly as Cu of sand increased. The experimental data also suggested that Vs is

independent of D50. In the same year, based also on the results of RC tests conducted on granular mate-

rials (Glass Ballotini (GB)) with different particle sizes, Bui (2009) obtained opposite results. His results

showed that at the same void ratio and effective confining stress, Vs increases with particle diameter.

Patel et al. (2008) measured Vs in assemblies of glass beads of different sizes using the BE technique.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Their results indicated that Vs increased as D50 of glass beads decreased. At the confining stress of 100

kPa, the Vs of glass beads with D50 of 0.5 mm was determined to be 320 m/s, which was about 45%

higher than that of glass beads with D50 of 2.5 mm. The experimental work of Patel et al. (2008), show-

ing an appreciable size dependence of small-strain stiffness, does not agree with that of Iwasaki and Tat-

suoka (1977). It does, however, appear to be consistent with that of Wichtmann and Triantafyllidis

(2009) and Bartake and Singh (2007), whose work adds further uncertainty: Bartake and Singh (2007)

performed BE tests on three dry sands with similar gradation and found that the Vs value increased as

D50 of the sand decreased. Yang and Gu (2013) conducted series of BE and RC tests for three uniform

types of glass beads with D50 varying from 0.195 mm to 1.750 mm, for a range of confining stresses and

void ratios. The BE and RC tests both show a trend that Vs decreases slightly with D50, particularly for

glass beads at the loose state. Given that the variations in the measured stiffness values are generally

small, and given the uncertainty in the laboratory experiments, they conclude that Vs is size independent.

4. Forms of Vs-based correlations as a function of D50

Figure 4 shows the high level of variability in Vs values predicted by empirical correlations presented in

Table 2; this variability likely reflects the different characteristics of the soils from which each of the

empirical models were developed. Such high variability suggests that some site-specific Vs values may

be required in order to make accurate predictions of Vs from SPT results. For example, Ohta and Goto

(1978), Lee (1992), Hasancebi and Ulusay (2007) suggested that the use of soil type in correlations in-

15
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 16 of 54

volving Vs improves the accuracy because a certain range in void ratio is represented. Ohta and Goto

(1978) found that there is a systematic change of Vs with soil types where: Vs (gravel) ≥ Vs (sand) ≥ Vs (clay),

and they attributed this change to the corresponding change in void ratio where: e (gravel) ≤ e (sand) ≤ e (clay).

Therefore the use of soil type as a mean to group data then reflects the average void ratio of the soils. In

fact, particle characteristics have an impact on the void ratio as described in the previous section, and
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

can be used to represent soil type in the correlations of shear wave velocity as suggested by Karray et al.

(2011).

Karray et al. 2011 used the Péribonka dam data obtained on gravelly coarse sands (0.2 mm < D50 <

10 mm) in conjunction with the Canadian Liquefaction Experiment (CANLEX) project data obtained on

fine sands (0.16 mm < D50 < 0.25 mm) to establish a unique correlation between Vs1 and qc1 considering

the effect of mean grain size

[21] Vs1 = 125.5(qc1 )0.25 D500.115

where Vs1 is in m/s, qc1 in MPa, and D50 in mm. Eq. 21 is applicable to uncemented relatively young

Holocene-age soil deposits. Eq. 21 together with a reliable (N1)60-qc1 relationship can be used to estimate

Vs1 values from SPT results as will be provided below.

The variation qc1/(N1)60 with D50 according to Eqs. 16 and 17 is plotted in Fig. 5. Similar

qc1/(N1)60‒D50 empirical formulations proposed by Robertson et al. (1983), Andrus and Youd (1989),

and Seed and De Alba (1986) are also plotted in Fig. 5. Average data from borings at Péribonka dam site

with D50 of 1.8 mm and at CANLEX project with D50 = 0.2 mm, respectively on gravelly coarse sands

and fine sands are plotted as references. It is worth noting from Fig. 5 that Péribonka data lies on the

curve corresponding to Eq. 16 (Mayne et al. 1995) while CANLEX data lies on the curve corresponding

to Eq. 17 (Anagnostopoulos et al. 2003). All the correlations presented in Fig. 5 show similar trends in

16
Page 17 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

that the qc1/(N1)60 ratio increases with grain size, D50 with an upper bound that corresponds to Eq. 17 and

a lower bound that can be represented by:

qc1
[22] ≈ 0.483D500.19
( N1 )60

where qc1 in MPa and D50 in mm.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Substitution of Eqs. 17 and 22 into Eq. 21 yields, respectively:

[23] Vs1− max = 118(( N1 )60 )0.25 D500.18

[24] Vs1− min = 104.6(( N1 )60 )0.25 D500.1625

where Vs1-max and Vs1-max in m/s and D50 in mm.

Equations 23 and 24 are thought to represent the upper and lower bounds of the Vs1 variation with

(N1)60 or void ratio. The practicality of this type of correlations can be easily comprehended with the aid

of Fig.6. Experimental data in terms of Vs1‒e relationships for different granular soils with variable grain

sizes collected from different studies found in the literature are plotted in Fig. 6. Figure 6 indicates that

the presented experimental data generally fall between an upper and a lower bound of Vs1, and the range

of (Vs1-max-Vs1-min) varies between 125 m/s and 200 m/s, with an average value of 160 m/s for fine soils (D50 <

0.2 mm). On the other hand, (Vs1-max-Vs1-min) of gravelly materials (D50 > 4.75 mm) shows higher values and a

wider variation (between 200 m/s and 475 m/s). As shown in Fig. 6, the prime effect of grain size appears

to change the range of possible void ratios which in turn had a substantial impact on Vs values.

In the same context, Fig. 7 presents Vs1‒e correlations based on experimental data of Menq (2003) on

granular soils with different grain sizes and gradations. Similar to Fig. 6, Fig. 7 illustrates that Menq’s

correlations tend to predict Vs1-min and Vs1-max values that fall also within two bounds, and the predicted

Vs1 generally increases with the increase in D50 and the decrease in Cu.

17
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 18 of 54

The potential to use the upper and lower bounds of Vs1 expressed respectively, using Eqs. 23 and 24

to evaluate typical Vs1‒e correlations of different granular soils is assessed in Figs. 8‒11. To plot the

variation of both Vs1-min and Vs1-max with e, the maximum normalized penetration index, (N1)60-max is first

evaluated based on the mean effective diameter, D50 using Eq. 13 assuming Id = 100 %, while (N1)60-min

is selected at 1.6 and 2.3 for relatively fine and coarse-grained soils, respectively. These two values of
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

(N1)60-min are selected after initial trials and found to give consistent results with experimental data. The

test results on TS sand with (D50 = 0.34, Cu = 1.95) provided by Kokusho and Yoshida (1997) in term of

Vs1‒e data are compared in Fig. 8a to the variation of both Vs1-min and Vs1-max with e predicted by the cur-

rent study and those estimated by Seed and Idriss (1970) correlation (Eq. 18) as well as Seed et al.

(1983) correlation presented in Table 2. Figure 8a illustrates that the experimental data fall within Vs1

bounds estimated by Eqs. 23 and 24. On the other hand, Seed et al. (1983) correlation generally overes-

timates the Vs1 values of the TS sand at its densest state and underestimates them at the loosest state.

Seed et al. (1983) correlation gives up to a 325 m/s-variation of Vs1 values between the loosest and the

densest state of the TS sand which is different from that observed in the original data provided by Ko-

kusho and Yoshida (1997) (∆Vs1 ≤ 150 m/s as shown in Fig. 8a). The correlation suggested by Seed and

Idriss (1970), on the contrary, gives a ∆Vs1 ≤ 150 m/s that agrees well with the experimental data on TS

sand. However, this correlation (Eq. 18) tends to predict Vs1 values higher than the measured data. When

the estimated Vs1 values from Seed and Idriss (1970) correlation is adjusted (multiplied) by a factor of

0.82, a better agreement with experimental data is obtained as shown in Fig. 8b. Data from other exper-

imental tests on granular soils having comparable grain sizes and gradations are also plotted in Fig. 8b

for comparison. Figure 8b indicates that the new introduced data and the predicted values of Vs1 using

the adjusted version of Eq. 18 fall in the same range of Vs1 bounds suggested by the current study.

18
Page 19 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

The experimental data of Kokusho and Yoshida (1997) on G25 gravel with (D50 = 1.13, Cu = 5.65)

collapse also onto the range given by Eqs. 23 and 24 as shown in Fig. 9a. Figure 9a illustrates that the

use of Seed et al. (1983) correlation yields inconsistent results regarding the ∆Vs1 values than when Seed

and Idriss (1970) correlation is considered, but Vs1 values estimated from Eq. 18 are generally lower than

the measured values. The suggested correction factor in this case is therefore higher than unity (1.1) as
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

depicted in Fig. 9b. The adjusted Vs1 values are generally coherent with Vs1 values predicted by the cur-

rent study as well as those measured by Kokusho and Yoshida (1997) and Bui (2009) on glass beads.

Figures 10a and 11a confirm that Seed et al. (1983) correlation gives erroneous estimation of ∆Vs1 val-

ues, and the suggested correction factors for Eq. 18 are respectively, of 1.25 and 1.50 for G50 (D50 =

2.28, Cu = 11.3) and G75 (D50 = 7.30, Cu = 31.1) gravels. Vs1 values estimated using Eq. 18 and correct-

ed using appropriate correction factors are compared satisfactory with measured data available from the

literature in Figs. 10b and 11b. Figures 8‒11 indicate clearly that the magnitude of correction factors

applied to Seed and Idriss (1970) correlation (Eq. 18) increases with mean grain size. Figure 12 presents

the suggested values of the correction factor to be applied to Seed and Idriss (1970) correlation as a

function of D50 based on the test results provided by Kokusho and Yoshida (1997):

[25] Fact = 0.22 ln( D50 ) + 1.065

where D50 is in mm. Note that this correlation is applicable only to granular soils with D50 ranges from

0.2 to 10 mm. By combining Eq. 25 with Seed and Idriss (1970) correlation (Eq. 18) yields:

[26] Vs1 = 5.68[ln( D50 ) + 4.84] I d + 25

where D50 is in mm and Vs1 in m/s.

Based on the comparative results shown in Figs 8‒11, it is expected that the use of Eq. 26 (the adjust-

ed version of Eq. 18) in predicting Vs1 values of granular soils (0.2 mm ≤ D50 ≤10 mm) leads to more

refined results.

19
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 20 of 54

5. Evaluation of the proposed Vs1-(N1)60 correlation based on field data

The applicability of the proposed Vs1-(N1)60 formulations (Eqs. 23 and 24) is assessed by comparing the

data (Vs1 and D50 profiles) obtained from field investigations at different sites with those predicted by

Eqs. 23 and 24 as portrayed in Figs 13‒15. Three sites (Péribonka, CANLEX, and Ohio River) are se-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

lected for the comparison. At each site, Vs1 and (N1)60 profiles are derived from data of the same sound-

ings with D50 identified in the boring. Details of the three sites are given below.

The Péribonka dam, in Québec, Canada, is an embankment dam constructed on deposits found at the

bottom of the river and generally constituted of well graded gravelly sand to gravel. Shear wave velocity

profiles were determined by MMASW tests. The Details of the MMASW procedure and Vs results at

Péribonka dam are presented by Karray et al. (2010). A number of SPT tests were performed, allowing

comparisons between Vs and N. In addition, the sampling performed in the boreholes, was important

with respect to identifying the tested materials (D50). Figure 13a presents a comparison between the

MMASW normalized shear wave velocity (Vs1) profiles and the upper and lower bounds calculated from

(N1)60 using Eqs. 23 and 24, respectively. Figure 13b, on the other hand, presents a comparison between

mean grain size identified in the boring (TF-88-05) and the max and min D50 estimated from Eqs. 23 and

24, respectively. In general, on an overall profile basis, the Vs1-(N1)60 correlation proposed in the current

study for granular material appears to work well, producing good agreement between MMASW and SPT

results (Figs. 13a and 13b). It is important to note that the estimated D50 from Vs1-(N1)60 pairs (Fig. 13b)

can be more reliable since the grain-size distribution obtained from boring can suffer a lack of precision

due the small volume of samples especially in the case of gravelly soils.

The Canadian Liquefaction Experiment (CANLEX) project has involved detailed investigation of six

sites in Western Canada in the context of studying the phenomenon of soil liquefaction. Extensive in situ

testing was performed at each site, and it consisted of the following: CPT, SPT, geophysical (gamma–

20
Page 21 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

gamma) logging, pressuremeter testing, and downhole Vs values through SCPTs (Wride et al. 2000). A

comparison between Vs1 profiles and the corresponding upper and lower bounds estimated from (N1)60

using Eqs. 23 and 24 is presented in Fig. 14a. Figure 14b presents a comparison between the average

mean grain size identified at the site (D50 av. = 0.2 mm); and the bounds of D50 estimated from the current

study based on Vs1 and (N1)60 measured data. For KD9305 boring, the measured Vs1 and D50 profiles are
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

located within the upper and lower bounds provided by Eqs. 23 and 24. Eqs. 23 and 24 tend to predict

D50 somewhat higher than the average mean grain size obtained from field data in SPT/E boring at a

depth of about 15 m. This overestimation of D50 values may be attributed to the reduction in the meas-

ured (N1)60 value at the corresponding depth. Despite this overestimation of the D50 value, the overall

D50 profile is still in accordance with the average measured D50 value.

The third site (site C) is located on the floodplain of the Ohio River. Comprehensive subsurface in-

vestigation program was carried out at the site including SPT, CPT, SCPT, and laboratory testing (Liao

et al. 2011). The results of sieve analysis on SPT soil samples from the subsurface investigation is pre-

sented in Fig. 15b. These data combined with measured (N1)60 profile are used to plot the Vs1 bounds

using Eqs. 23 and 24. The estimated Vs1 bounds are then compared to the measured Vs1 profile in Fig.

15a. The mean effective diameter profile is back calculated using Eqs. 23 and 24 and plotted in Fig. 15b

for comparison. The comparison shown in Fig. 15 in addition to those shown in Figs 13 and 14 can lead

us to conclude that the Vs1-(N1)60 formulations suggested in the current study has a reasonable applicabil-

ity to predict both Vs1 and D50 profiles.

The field data at the three sites described above is further utilized in conjunction with Eqs. 23 and 24

to examine the applicability of some existing widely-used Vs1-(N1)60 correlations. Figure 16 shows a

comparison between Vs1-(N1)60 ranges estimated by Eqs. 23 and 24 at different grain sizes (D50 = 0.2,

0.8, 2, and 5 mm) and those predicted by Seed et al. (1983) and Seed and Idriss (1981) correlations pre-

21
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 22 of 54

sented in Table 2. Vs1-(N1)60 data measured at Péribonka, CANLEX, and Ohio River sites are plotted in

Fig. 16 . The limits of relative density, Id of 30% and 100% calculated based on Eq. 13 are also plotted

as references. The correlations proposed by Seed et al. (1983) and Seed and Idriss (1981) are selected for

the evaluation as they fall in group II characterized by a correlative exponent (b ≈ 0.5), thus they auto-

matically remove the effect of overburden pressure. Based on the comparative results shown in Fig. 16,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

the following trends should be highlighted:

1) In contrast to the correlation proposed by Seed et al. (1983) and Seed and Idriss (1981) that give

up to a 375 m/s-variation of Vs1 between the loosest and the densest state of the soil, Vs1 variation

using Eqs. 23 and 24 at the same mean grain size collapses onto a relatively narrow range and

this range increases with the mean grain size.

2) Correlation of Seed et al. (1983) and all similar correlations that have exponential parameter of

N-value equal or higher than 0.5 generally overestimate the normalized shear wave velocity for

sands especially at its densest state ((N1)60 > 25). This observation is fully consistent with that re-

ported by Sykora (1987).

3) Although empirical correlations, fall in group I (i.e. correlative exponent b < 0.45), don’t auto-

matically remove the effect of overburden pressure, they predict satisfactory the variation of Vs1

between the loosest and the densest state of the soil. In other words, the consistency between the

Vs1 values estimated based on the corrected version of Eq. 18 (Eq. 26) and Vs1 range calculated

using Eqs. 23 and 24 with field and laboratory measurements, presented in Figs. 8b to 11b, is in

part related to the compatibility between the correlative exponent of the proposed relation (b =

0.25) and the laboratory relationship (Eq. 26) as can be demonstrated by the substitution of Eq.

13 into Eq. 26.

22
Page 23 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

100.( N1 ) 60
[27] Vs1 = 5.68[ln( D50 ) + 4.84] + 25
0.6 + 0.25 log( D50 )

where D50 is in mm and Vs1 in m/s.

Similar to Fig. 16, Fig. 17 presents the upper and lower bounds of Vs1-(N1)60 ranges estimated from

Eqs. 23 and 24 at different grain sizes (D50 = 0.2, 0.5, 1, 2, and 5 mm), respectively. The ranges of rela-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

tive densities calculated from Eq. 13 for Id of 30%, 50%, 65%, 85%, and 100% are also plotted in Fig.

17. In additions, void ratios calculated based on Eqs. 19 and 20 are plotted in the same figure. Figure 17,

based primarily on Eqs. 23 and 24, represents a suggested chart that can be applied over a wide range of

uncemented, Holocene-age soils with the separation of Vs1-(N1)60 correlation trends according to mean

grain sizes, void ratios, and relative density indexes and compared favorably with field as well as relia-

ble laboratory tests as presented in Figs 8-16. Figure 17 shows that at a constant e, the variation of (N1)60

between the loosest and densest state of the soil is accompanied by a smaller change of Vs1. On the con-

trary, a great variation of Vs1 is accompanied by a smaller variation of (N1)60 at the same state of density.

This observation indicates that Vs1 is more sensitive to void ratio, e, while (N1)60 is more sensitive to

relative density, Id. In other words, Vs1 and (N1)60 can be regarded as complementary geotechnical pa-

rameters that can be used together, for example, to obtain a good idea of grain-size distribution of a rep-

resentative volume of soil.

6. Conclusions

This paper presents an overview of the salient features of some existing widely-used correlations be-

tween Vs and basic, more common geotechnical parameters especially N-SPT, e, D50, and Id so that the

applicability of correlations to geotechnical engineering practice can be ascertained for each individual

project. The following observations have been made while analyzing these correlations to assist in solv-

ing the question of applicability and are based primarily on results of comparisons made in this study:

23
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 24 of 54

1. The variation of Vs1 between the loosest and the densest state of the soil falls onto a relatively

narrow range and this range increases with grain size.

2. Practitioners should exercise caution when using Vs1-(N1)60 correlations having exponential pa-

rameter of (N1)60-value equal or higher than 0.5 as they produce higher values of Vs1 at larger

values of (N1)60 (>25) and a large ∆Vs1 (up to 400 m/s) between the loosest and the densest state
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

of the same soil, in contradiction with the first conclusion.

3. Empirical correlations that contain no function taking into account the particle characteristics

and/or that didn’t impose particle-size limits for the validity of the proposed relationship should

be used with cautions.

4. The prime effect of grain size and gradation was to change the range of possible void ratios

which in turn had a substantial impact on Vs values.

5. New Vs1-(N1)60 correlations accounts for mean grain sizes are proposed by combining the recent

published Vs1-qc1 formulation by Karray et al. (2011) and available (N1)60-qc1 relationships. The

validity of the proposed correlations has been confirmed by comparing the calculated Vs1 based

on the proposed correlations with reliable laboratory and field measurements. The proposed for-

mulations are utilized to examine the applicability of some existing widely-used Vs-N correla-

tions.

6. A new Vs1-(N1)60 chart has been proposed allowing the practitioner to estimate Vs1 based on a

combination of data including N-SPT, e, D50, and Id. The new chart has the advantage that it ap-

plies over a wide range of uncemented, Holocene-age soils with the separation of Vs1-(N1)60 cor-

relation trend according to mean grain sizes and compared favorably against field as well as reli-

able laboratory measurements.

24
Page 25 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

ACKNOWLEDGEMENTS

The authors would like to thank NSERC for their financial support throughout this research project.

References
Anagnostopoulous, A., Koukis, G., Sabatakis, N., and Tsiambaos, G. 2003. Empirical correlations of soil parame-

ters based on cone penetration test for Greek soils. Geotechnical and Geological Engineering, 21: 377–387.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Anastasiadis, A., Senetakis, K., and Pitilakis, K. 2012. Small-strain shear modulus and damping ratio of sand-

rubber and gravel-rubber mixtures. Geotechnical and Geological Engineering, 30(2): 363–382.

Anbazhagan, P., Parihar, A., and Rashmi, H. N. 2012. Review of correlations between SPT-N and shear modulus:

A new correlation applicable to any region. Soil Dynamic and Earthquake Engineering, 36: 52–69.

Andrus, R.D., and Youd, T.L. 1989. Penetration tests in liquefiable gravels. In Proceedings of the 12th Interna-

tional Conference on Soil Mechanics and Foundation Engineering, A. A. Balkema, Rotterdam, The Nether-

lands, 679-682.

Andrus, R.D., Mohanan, N.P., Piratheepan, P., Ellis, B.S. and Holzer, T.L. 2007. Predicting shear-wave velocity

from cone penetration resistance. In Proceedings of the 4th International Conference on Earthquake Geotech-

nical Engineering, Thessaloniki, Greece, K.D. Pitilakis, ed., Springer, The Netherlands, paper No. 1454.

ASTM 2011a. Standard D4253‒Maximum index density and unit weight of soils using a vibratory table. In:

ASTM annual CDs of standards, vol 04.08, West Conshohocken, PA.

ASTM 2011b. Standard D4254‒Minimum index Density and unit weight of soils and calculation of relative den-

sity. In: ASTM annual CDs of standards, vol 04.08, West Conshohocken, PA.

Baldi, G., Bellotti, R., Ghionna, V., Jamiolkwski, M., and Pasqualini, E. 1986. Interpretation of CPTs and CPTUs,

2nd Part: Drained penetration of sands. In Proceedings of the 4th International Geotechnical Seminar, Field In-

strumentation and In-situ Measurements, Nanyang Technological Institute, Singapore, pp. 143–156.

Bartake, P.P., and Singh, D.N. 2007. Studies on the determination of shear wave velocity in sands. Geomechanics

and Geoengineering, 2(1): 41–49.

25
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 26 of 54

Ben Romdhan, M., Hussien, M.N., Karray, M., Chekirad, M., and Varvara, R. 2014. The use of piezoelectric ring-

actuator technique (P-RAT) in shear wave velocity measurement in granular media. In Proceedings of the 67th

Canadian geotechnical conference, Regina, Saskatchewan, paper no. 307.

Brandenberg, S.J., Bellana, N., and Shantz, T. 2010. Shear wave velocity as function of standard penetration test

resistance and vertical effective stress at California bridge sites. Soil Dynamics and Earthquake Engineering,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

30(10): 1026–1035.

Brignoli, E.G.M., Gotti, M., and Stokoe, K.H.II. 1996. Measurement of shear waves in laboratory specimens by

means of piezoelectric transducers. Geotechnical Testing Journal, 19(4): 384–397.

Bui, M.T. 2009. Influence of some particle characteristics on the small strain response of granular materials. Ph.D.

thesis, University of Southampton.

Cai, G.J., Puppala, A.J., and Liu S.Y. 2014. Characterization on the correlation between shear wave velocity and

piezocone tip resistance of Jiangsu soft clays. Engineering Geology, 171: 96–103.

Campanella, R. G., Robertson, P. K., and Gillespie, D. G. 1986. Seismic cone penetration test: Proceed-
ings of In Situ ’86, 116–130, ASCE.
Chang, N.-Y., and Ko, H.-Y. 1982. Effect of grain size distribution on dynamic properties and liquefaction poten-

tial of granular soils. Research Repot R82-103, University of Colorado at Denver, 145 p.

Chien, L.-K., and Oh Y.-N. 2000. Laboratory and field shear wave measurement at a reclaimed site in west Tai-

wan. Geotechnical Testing Journal, 23(1): 021–035.

Choi, Y., and Stewart, J.P. 2005. Nonlinear site amplification as function of 30 m shear wave velocity. Earth-

quake Spectra, 21(1): 1–30.

Clayton, C.R.I. 2011. Stiffness at small strain: research and practice. Géotechnique, 61(1): 5–38.

Cubrinovski, M., and Ishihara, K. 1999. Empirical correlation between SPT N value and relative density for sandy

soils. Soils and Foundations, 39(5): 61–71.

Cubrinovski, M., and Ishihara, K. 2002. Maximum and minimum void ratio characteristics of sands. Soils and

Foundations, 42(6): 65–78.

26
Page 27 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Dyvik, R., and Madshus, C. 1985. Lab measurements of Gmax using bender elements. In Proceedings of the con-

ference on the Advances in the Art of Testing Soil under Cyclic Conditions. ASCE Geotechnical Engineering

Division, New York., pp. 186-196.

Fam, M.A., Cascante, G., and Dusseault, M.B. 2002. Large and small strain properties of sands subjected to local

void increase. Journal of Geotechnical and Geoenvironmental Engineering ASCE, 128(12): 1018‒1025.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Gazetas, G. 1991. Chapter 15: Foundation vibrations. Foundation engineering handbook, 2nd Ed., H.-Y. Fang, ed.,

Chapman & Hall, New York/London, 553–593.

Gibbs, K.J., and Holtz, W.G. 1957. Research on determining the density of sands by spoon penetration testing. In

Proceedings of 4th ICSMFE, 1, 35–39.

Hardin, B.O. and Drnevich, V.P. 1972. Shear modulus and damping in soil: Measurement and parameter effects.

Journal of the Soil Mechanics and Foundation Division ASCE, 98(7): 603–624.

Hardin, B.O., and Black, W.L. 1966. Sand stiffness under various triaxial stresses. Journal of the Soil Mechanics

and Foundations Division ASCE, 92(2): 27–42.

Hardin, B.O., and Richart, F.E. 1963. Elastic wave velocities in granular soils. Journal of Soil Mechanics and

Foundation Engineering Division ASCE, 89(1): 39–56.

Hasancebi, N., and Ulusay, R. 2007. Empirical correlations between shear wave velocity and penetration re-

sistance for ground shaking assessments. Bulletin of Engineering Geology and the Environment, 66(2): 203–

13.

Huang, Y.-T., Huang, A.-B., Kuo, Y.-C., and Tsai, M.-D. 2004. A laboratory study on the undrained strength of

silty sand from Central Western Taiwan. Soil Dynamics and Earthquake Engineering, 24: 733–743.

Imai, T., and Tonouchi, K. 1982. Correlation of N-value with S-wave velocity and shear modulus. In Proceedings

of the 2nd European symposium on penetration testing, pp. 57–72.

Imai, T., and Yoshimura, Y. 1970. Elastic wave velocity and soil properties in soft soil. Tsuchito-Kiso 18(1): 17–

22. (in Japanese).

Ishihara, K. 1996. Soil behaviour in earthquake geotechnics. Oxford University Press, New York. ISBN 0-19-

856224-1.

27
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 28 of 54

Iwasaki, T., and Tatsuoka, F. 1977. Effect of grain size and grading on dynamic shear moduli of sand. Soils and

Foundations, 38(1): 19–35.

Kanai, K. 1966. Observation of microtremors, XI: Matsushiro earthquake swarm areas. Bulletin of Earthquake

Research Institute, vol. XLIV, part 3, University of Tokyo, Tokyo, Japan.

Karray, M. 2010. Shear wave velocity in geotechnical engineering. In Proceedings of the International Conference
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

on Geotechnical Engineering, France.

Karray, M., and Hussien, M.N. 2015. On the influence of grain size distribution on Vs˗qc correlation. Submitted to

Canadian Geotechnical Journal for possible publication.

Karray, M., and Lefebvre, G. 2008. Significance and evaluation of Poisson's ratio in Rayleigh wave testing. Ca-

nadian Geotechnical Journal, 45(5): 624-635, 10.1139/T08-016.

Karray, M., Ben Romdhan, M., Hussien, M.N., and Éthier, Y. 2015. Measuring shear wave velocity of granular

material using the Piezoelectric Ring Actuator Technique (P-RAT). Canadian Geotechnical Journal, Published

on the web 28 January 2015, 10.1139/cgj-2014-0306

Karray, M., Lefebvre, G., Ethier, Y., and Bigras, A. 2010. Assessment of deep compaction of the Péribonka dam

foundation using “modal analysis of surface waves” (MASW). Canadian Geotechnical Journal, 47(3): 312–

326. doi:10.1139/T09-108.

Karray, M., Lefebvre, G., Ethier, Y., and Bigras, A. 2011. Influence of particle size on the correlation between

shear wave velocity and cone tip resistance. Canadian Geotechnical Journal, 48(4): 599–615. doi:10.1139/T10-

092.

Kokusho, T., and Yoshida, Y. 1997. SPT N-value and S-wave velocity for gravelly soils with different grain size

distribution. Soils and Foundations, 37(4): 105–113.

Konno, T., Suzuki, Y., Tateishi, A., Ishihara, K., Akino, K., Iizuka, S. 1993. Gravely soil properties by field and

laboratory tests. In Proceedings of the 3th International Conference on Case Histories in Geotechnical Engi-

neering, St. Louis, Missouri, 575‒594.

Kulhawy, F.H., and Mayne, P.W. 1990. Manual on estimating soil properties for foundation design. Final Rep.

1493-6, EL-6800, Electric Power Research Institute, Palo Alto, Calif.

28
Page 29 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Lee, C-J., Hung, W-Y., Tsai, C-H., Chen T., Tu, Y., and Huang, C-C. 2014. Shear wave velocity measurements

and soil–pile system identifications in dynamic centrifuge tests. Bulletin of Earthquake Engineering, 12: 717–

734.

Lee, S.H. 1990. Regression models of shear wave velocities. Journal of the Chinese Institute of Engineers, 13:

519–532.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Lee, S.H. 1992. Analysis of the multicollinearity of regression equations of shear wave velocities. Soils and

Foundations, 32 (1): 205–214.

Liao, T., Davie, J.R., and Chey, M. 2011. Characterization of fluvial sand deposits on floodplain of Ohaio River.

Geo-Frontiers, 2345-2355.

Lo Presti, D.C.F. 1987. Mechanical behavior of Ticino sand from resonant column tests. Ph.D thesis, Politecnico

di Torino, Turin, Italy.

Lo Presti, D.C.F., Pedroni, S., and Crippa, V. 1992. Maximum dry density of cohesionless soil by pluviation and

by ASTM D 4253-83: a comparative study. Geotechnical Testing Journal, 15(2): 180–189.

Marcuson III, W.F., and Bieganousky, W.A. 1977. Laboratory Standard Penetration Tests on Fine Sands. Journal

of the Geotechnical Engineering Division ASCE, 103(6): 565–588.

Mayne, P.W., and Rix, G.J. 1995. Correlations between shear wave velocity and cone tip resistance in natural

clays. Soils and Foundations, 35(2): 107–110.

Mayne, P.W., Mitchell, J.K., Auxt, J.A., and Yilmaz, R. 1995. U.S. national report on CPT. In Proceedings of

Symposium on Cone Penetration Testing, Vol. 1, Swedish Geotechnical Society, 263–276.

McGillivray, A., and Mayne, P.W. 2004. Seismic piezocone and seismic flat dilatometer tests at Treporti. Ge-

otechnical and Geophysical Site Characterization, ISC-2, Porto, Portugal, Millpress, Rotterdam, 2, 1623–1628.

Menq, F.Y. 2003. Dynamic Properties of Sandy and Gravelly Soils. Ph.D. Dissertation, The University of Texas

at Austin.

Menq, F.Y. and Stokoe, K.H.II. 2003. Linear dynamic properties of sandy and gravelly soils from large-scale

resonant tests. In: Di, B.H., Doanh, T., Geoffroy, H. & Sauzeat, C. (eds.), The 3rd International Symposium on

Deformation Characteristics of Geomaterial, pp. 63–71. Rotterdam A. A. Balkema Publishers, Lyon, France.

29
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 30 of 54

Muromachi, T. 1981. Cone penetration testing in Japan. In Proceedings of Symposium on Cone Penetration Test-

ing and Experience, Geotechnical Engineering Division, ASCE, St. Louis, Missouri, pp. 49–75.

Ohba, S., and Toriumi, I. 1970. Research on vibration characteristics of soil deposits in Osaka, part 2, on veloci-

ties of wave propagation and predominant periods of soil deposits. Technical meeting of Architectural Institute

of Japan. (in Japanese).


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Ohsaki, Y., and Iwasaki, R. 1973. On dynamic shear moduli and Poisson’s ratio of soil deposits. Soils and Foun-

dations, 13(4): 61–73.

Ohta, T., Hara, A., Niwa, M., and Sakano, T. 1972. Elastic moduli of soil deposits estimated by N-values. In Pro-

ceedings of the 7th annual conference. The Japanese Society of Soil Mechanics and Foundation Engineering,

pp. 265–268.

Ohta, Y., and Goto, N. 1978. Empirical shear wave velocity equations in terms of characteristic soil indices.

Earthquake and Engineering Structure Dynamics, 6: 167–187.

Okamoto, T., Kokusho, T., Yoshida, Y., and Kusuonoki, K. 1989. Comparison of surface versus subsurface wave

source for P-S logging in sand layer. In Proceedings of 44th Annual Conference, JSCE, 3, pp. 996–997 (in

Japanese)

Patel, A., Bartake, P.P., and Singh, D.N. 2008. An empirical relationship for determining shear wave velocity in

granular materials accounting for grain morphology. Geotechnical Testing Journal, 32(1): 1–10.

Pitilakis, K., Raptakis, D., Lontzetidis, K., Tika-Vassilikou, T., and Jongmans, D. 1999. Geotechnical and geo-

physical description of Euro-Seistests, using field, and laboratory tests and moderate strong ground motions,

Journal of Earthquake Engineering, 3(3): 381–409.

Pitilakis, K.D., Anastasiadis, A., and Raptakis, D. 1992. Field and laboratory determination of dynamic properties

of natural soil deposits. In Proceedings of 10th World Conf. Earthquake Engineering, Rotterdam, pp. 1275–

1280.

Raptakis, D.G., Anastasiadis, S.A.J., Pitilakis, K.D., and Lontzetidis, K.S. 1995. Shear wave velocities and damp-

ing of Greek natural soils. In Proceedings of 10th European Conference of Earthquake Engineering, Vienna,

pp. 477–482.

30
Page 31 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Richart, F.E., Hall, J.R., and Woods, R.D. 1970. Vibrations of soils and Foundations. Prentice Hall, Englewood

Cliffs.

Rix, G.J., and Stokoe, K.H. 1991. Correlation of initial tangent moduli and cone penetration resistance. In Cali-

bration chamber testing. Edited by A.B. Huang, Elsevier, pp. 351–362.

Robertson, P.K. 2009. Interpretation of cone penetration tests‒a unified approach. Canadian Geotechnical Journal,
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

46: 1337–1355.

Robertson, P.K. 2012. Discussion of “Influence of particle size on the correlation between shear wave velocity

and cone tip resistanceˮ. Canadian Geotechnical Journal, 48(4): 599–615.

Robertson, P.K., and Wride, C.E. 1998. Evaluating cyclic liquefaction potential using the cone penetration test.

Canadian Geotechnical Journal, 35 (3): 442–459.

Robertson, P.K., Campanella, R.G., and Wightman, A. 1983. SPT-CPT correlations, Journal of Geotechnical En-

gineering, 109(11): 1449–1459.

Robertson, P.K., Sasitharan, S., Cunning, J.C., and Sego, D.C. 1995. Shear-wave velocity to evaluate in-situ state

of Ottawa sand. Journal of Geotechnical Engineering, 121(3): 262–273.

Rollins, K.M., Evans, M.D., Diehl, N.B., and Daily, W.D. 1998. Shear modulus and damping relationships for

gravels. . Journal of Geotechnical and Geoenvironmental Engineering, 124(5): 396–405.

Salgado, R., Bandini, P., and Karim, A. 2000. Shear strength and stiffness of silty sand. Journal of Geotechnical

and Geoenvironmental Engineering, 126(5): 451–462.

Santamarina, J.C., Klein, K.A., and Fam, M.A. 2001. Soils and Waves – particular materials behavior, characteri-

zation and process monitoring, Wiley, New York.

Seed, H.B., and de Alba, P. 1986. Use of SPT and CPT tests for evaluating the liquefaction resistance of sands.

Use of in situ tests in geotechnical engineering, Geotechnical Special Publication No. 6, ASCE, New York,

1249–1273.

Seed, H.B., and Idriss, I.M. 1970. Soil moduli and damping factors for dynamic response analyses. Research Re-

port EERC 7010, Earthquake Engineering Research Center, University of California, Berkeley, CA.

31
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 32 of 54

Seed, H.B., and Idriss, I.M. 1981. Evaluation of liquefaction potential in previous earthquakes. In Proceedings of

the Conference on In Situ Testing to Evaluate Liquefaction Susceptibility, American Society of Civil Engi-

neers, St. Louis, Mo.

Seed, H.B., Idriss, I.M. and Arango, I. 1983. Evaluation of liquefaction potential using field performance data.

Journal of Geotechnical Engineering Division, 109(3): 458–482.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Seed, H.B., Wong, R.T., Idriss, I.M., and Tokimatsu, K. 1986. Moduli and damping factors for dynamic analysis

of cohesionless soils. Journal of Geotechnical Engineering, 112 (11): 1016–1032.

Shibata, T., 1970. Analysis of liquefaction of saturated sand during cyclic loading. Disaster Prevention Research

Institute Bulletin 13, 563–570.

Skempton, A.W. 1986. Standard penetration test procedures and the effects in sands of overburden pressure, rela-

tive density, particle size, aging and overconsolidation. Géotechnique, 36(3): 425–557.

Sykora, D.E., and Stokoe, K.H. 1983. Correlations of in-situ measurements in sands of shear wave velocity. Soil

Dynamic and Earthquake Engineering, 20: 125–136.

Sykora, D.W. 1987. Examination of existing shear wave velocity and shear modulus correlations in soils, De-

partment of the Army, Waterways Experiment Station, Corps of Engineers, Miscellaneous Paper GL-87-22.

Taya, T., Hatanaka, M., and Uchida, A. 1999. Effect of principle stress ratio and stress history on Vs for undis-

turbed sand and gravel samples. In Proceedings of the 2nd International Symposium on Pre-failure Defor-

mation Characteristics of Geomaterials, Torino, V.1, p. 331‒340.

Viggiani, G. and Atkinson, J.H. 1995. Interpretation of bender element tests. Géotechnique, 45(1): 149–154.

Wei, B.Z., Pezeshk, S., Chang, T.S., Hall, K.H., and Liu, H.P. 1996. An empirical method to estimate shear wave

velocity of soils in the New Madrid seismic zone. Soil Dynamics and Earthquake Engineering, 15(6): 399–408.

Wichtmann, T., and Triantafyllidis, T.h. 2009. Influence of the grain size distribution curve of quartz sand on the

small strain shear modulus Gmax. Journal of Geotechnical and Geoenvironmental Engineering ASCE, 135(10):

1404–1418.

32
Page 33 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Wride, C.E., Robertson, P.K., Biggar, R.G., Campanella, R.G., Hofmann, B.A., Hughes, J.M.O., Kupper, A., and

Woeller, D.J. 2000. Interpretation of in situ test results from the CANLEX sites. Canadian Geotechnical Jour-

nal, 37(3): 505–529.

Yamashita, S., Kawaguchi, T., Nakata, Y., Mikami, T., Fujiwara, T. and Shibuya, S. 2009. Interpretation of inter-

national parallel test on the measurement of Gmax using bender elements. Soils and Foundations, 49(4): 631–
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

650.

Yang, J., and Gu, X.Q. 2013. Shear stiffness of granular material at small strains: does it depend on grain size?.

Géotechnique 63(2): 165–179.

Yimsiri, S., and Soga, K. 2000. Micromechanics-based stress-strain behaviour of soils at small strains. Géotech-

nique, 50(5): 559–571

Youd, T.L., Indriss, L.M., Andrus, R.D., Arango, I., Castro, G., Christian, J.T., Dobry, R., Liam Finn, W.D.,

Harder, L.F., et al. 2001. Liquefaction resistance of soils: summary report from the 1996 NCEER and 1998

NCEER/NSF workshops on evaluation of liquefaction resistance of soils. Journal of Geotechnical and Geoen-

vironmental Engineering, 127(10): 817–33.

33
al use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version Page 34 of 54

Table 1. Some of the existing correlations between Vs and geotechnical parameters for granular soils and calculation of Vs1 for different stress state
and void ratios.
Author (s) Type of soil Packing A F(e) B K0 = 1.0 K0 = 0.5
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

emin emax Vs1max Vs1min ∆Vs1 Vs1max Vs1min ∆Vs1


Id ≈ 100% Id ≈ 0% (m/s) (m/s) (m/s) (m/s) (m/s) (m/s)

Hardin and Richard (1963) Round-grain Ottawa sand 0.35 0.85 51 (2.17 − e) 0.25 294 214 80 266 193 73
Hardin and Richard (1963) Angular-grain crushed quartz 0.60 1.30 35 (2.97 − e) 0.25 263 185 78 238 168 70
Lo Presti (1987) Hokksund sand 0.55 0.87 60 (1.96 − e) 0.23 244 189 55 222 172 50
Lo Presti et al. (1992) Ticino sand 0.58 0.93 52 (2.27 − e) 0.22 242 192 50 221 176 45
Robertson et al. (1995) Ottawa sand 0.56 0.90 78 (1.47 − e) 0.26 236 148 88 212 133 79
Wei et al. (1996) Poorly graded fine sand 0.47 0.88 87 (1.56 − e) 0.25 300 187 113 272 169 103
Wei et al. (1996) Fine to coarse sand 0.38 0.79 72 (1.66 − e) 0.25 291 198 93 263 179 84
Salgado et al. (2000) Silty Ottawa sand 0.48 0.78 48 (2.17 − e) 0.22 223 184 39 204 168 36
Huang et al. (2004) Fine sand 0.65 1.13 44 (2.04 − e) 0.25 193 127 66 175 114 61
Witchmann and
Triantafyllidis 2009
Poor-graded quartz sand 0.63 1.12 77 (1.76 − e) 0.22 240 136 104 219 124 95
Witchmann and
Triantafyllidis 2009
Well-graded quartz sand 0.35 0.67 109 (1.14 − e) 0.29 327 195 132 291 173 118

34
al use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version
Page 35 of 54

Table 2. Existing correlations between Vs and N-SPT.


Author (s) Type of soil N value used No. of data Correlation Correlation Method of Vs measurement Country
Min Max set coefficient
Kanai (1966) Sandy soil 1 50 70 V s = 1 8 .9 N 0 .6 0 Microtremor Japan
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Shibata (1970) Sandy soil V s = 3 1 .7 N 0 .5 4 Japan


Imai and Yoshimura (1970) Mixed soil type V s = 7 6 .0 N 0 .3 9 Down-hole Japan
Ohba and Toriumi (1970) Alluvial sand and clay V s = 8 4 .0 N 0 .3 1 Rayleigh wave velocity Japan
Ohta et al. (1972) Tertiary soil, diluvial 0.5 150 100 V s = 8 7 .2 N 0 .3 6 Japan
sandy and cohesive soil
Ohsaki and Iwasaki (1973) Sandy soil 1 100 220 V s = 5 9 .0 N 0 .4 7 0.886 Down-hole Japan
Ohta and Goto (1978) Granular soils 2 200 289 Vs = XN 0 .2 2
D 0 .2 3 X=62.14 (clay) Japan
X=67.79 (fine sand)
X=63.94 (medium sand)
X=66.68 (coarse sand)
X=71.52 (gravelly sand)
X=92.28 (gravel)
Imai and Tonouchi (1982) Alluvial sand 1 70 294 V s = 9 7 .0 N 0 .3 1 0.690 OYO suspension Japan
Seed and Idriss (1981) All soils V s = 6 1 .0 N 0 .5 0 USA
Seed et al. (1983) Sands V s = 5 6 .4 N 0 .5 0 USA
Sykora and Stokoe (1983) Granular soils 1 350 229 V s = 1 0 6 .7 N 0 .2 7 Japan
Okamoto et al. (1989) Sands V s = 1 2 5 .0 N 0 .3 0 Japan
Lee (1990) Sands V s = 5 7 .0 N 0 .4 9 0.620 Taiwan
Lee (1992) Silty sand 126 V s = 1 0 4 .7 N 0 .3 0 0.670 Down-hole Taiwan
Pitilakis et al. (1992) Sands 8 60 V s = 1 6 2 .0 N 0 .1 7 Cross-hole and Greece
resonant column
Raptakis et al. (1995) Sands V s = 1 0 0 .0 N 0 .2 4 Greece
Rollins et al. (1998) Holocene gravels 1 60 186 V s = 6 3 .0 N 6 0 0 .4 3 0.770 Becker penetration USA
Rollins et al. (1998) Pleistocene gravels 1 50 105 V s = 1 3 2 .0 N 6 0 0 .3 2 0.690 Becker penetration USA
Pitilakis et al. (1999) Clays, silts, and sands 1 110 300 V s = 1 4 5 .0 N 6 0 0 .17 8 0.700 Cross-hole, down-hole, seismic refraction Greece
Hasancebi and Ulusay (2007) Clays, silts, and sands 5 45 V s = 1 0 4 .7 9 N 6 0 0 .2 6 0.710 Seismic refraction Turkey
Wride et al. (2000) Fine sand 0.16<D50<0.25 V s 1 = Y ( N 1 ) 60 0.2 5 74.8≤Y≤102.1 CANLEX
project
Note: Vs and Vs1 are presented in m/s; depth, D is presented in meters; and effective mean diameter, D50 is presented in mm.

35
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 36 of 54

Figure Captions

Fig. 1. Illustration of seismic wave propagation and relation with elastic parameters: (a) P-wave, and (b)
S-wave.
Fig. 2. Distribution of normalized shear wave velocities collected from MMASW and SCPT at different
sites in Québec, Canada, and at Yarbro-Excavation, USA.
Fig. 3. Relation between N-SPT, relative density of soil, and particle size: (a) N-SPT versus vertical ef-
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

fective stress for different relative densities, and (b) N1-SPT versus relative density for different particle
sizes.
Fig. 4. SPT-N and Vs empirical relations presented in Table 2.
Fig. 5. Variation of qc1/(N1)60 ratio with mean grain size, D50.
Fig. 6. Variations of the upper and lower bounds of Vs1 with e.
Fig. 7. Variations of the upper and lower bounds of Vs1 with e from Menq (2003) experimental data.
Fig. 8. Comparison between experimental Vs1‒e data on TS sand with D50 = 0.34 mm and Cu = 1.95
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970) and
Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 0.82.
Fig. 9. Comparison between experimental Vs1‒e data on G25 gravel with D50 = 1.13 mm and Cu = 5.65
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970) and
Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.10.
Fig. 10. Comparison between experimental Vs1‒e data on G50 gravel with D50 = 2.28 mm and Cu = 11.3
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970) and
Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.25.
Fig. 11. Comparison between experimental Vs1‒e data on G75 gravel with D50 = 7.30 mm and Cu = 31.1
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970) and
Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.50.
Fig. 12. Correlation between mean grain size (D50) and the correction factor to be applied to Seed and
Idriss (1970) correlation.
Fig. 13. Comparison between Péribonka site results with those predicted by Eqs. 23 and 24: (a) Vs1 and
(N1)60 profiles; and (b) D50 profile.
Fig. 14. Comparison between CANLEX project (Kidd site) results with those predicted by Eqs. 23 and
24: (a) Vs1 and (N1)60 profiles; and (b) D50 profile.

36
Page 37 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 15. Comparison between Ohio River (site C) results with those predicted by Eqs. 23 and 24: (a) Vs1
and (N1)60 profiles; and (b) D50 profile.
Fig. 16. Comparison between Vs1‒(N1)60 ranges estimated by Eqs. 23 and 24 at different grain sizes and
those predicted by Seed et al. (1983) and Seed and Idriss (1981) correlations.
Fig. 17. Normalized shear wave velocity, Vs1 as function of (N1)60, e, Id, and D50 for uncemented, Holo-
cene-age soils.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

37
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

(b) S-wave.

Primary wave (P)

Shear wave (S)


σ

1
1
Emax

Gmax
ε

γ
Fig. 1. Illustration of seismic wave propagation and relation with elastic parameters: (a) P-wave, and
Page 38 of 54
Page 39 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 2. Distribution of normalized shear wave velocities collected from MMASW and SCPT at
different sites in Québec, Canada, and at Yarbro-Excavation, USA.

0.2
Péribonka, Québec (Canada)
0.16
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

n=1082 Gravely medium sand


0.12 0.2 mm < D50 < 10mm
15 < N1 < 60
0.08
0.04 Vs, MMASW
0 0.2
Mékinac, Québec (Canada)
n=377 0.16
Fine to medium sand 0.12
0.1 mm < D50 < 0.7 mm
7 < N1 < 30 0.08
0.04
Vs, MMASW
0.2
Relative frequency

0
Yarbro-Excavation, USA
0.16 n=31
0.12 Clean sand
D50 = 0.41 mm
0.08 5 < qc1 < 28
0.04
Vs, SCPT
0 0.2
Fortierville,Québec
Fortieville, Québec(Canada)
(Canada)
0.16
n=377 Silt and sandy silt
0.02 mm < D50 < 0.09 mm 0.12
30 < N1 < 80 0.08
0.04
Vs, MMASW
0.3 0
0.25 Sorel, Québec (Canada)
n=588 Clayey silt
0.2
0.15 D50 < 0.02 mm
N1(average) = 4.13±1.25
0.1
0.05 Vs, MMASW
0
50 100 150 200 250 300 350 400 450
Normalized shear wave velocity, Vs1 (m/s)
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 40 of 54

Fig. 3. Relation between N-SPT, relative density of soil, and particle size: (a) N-SPT versus vertical
effective stress for different relative densities, and (b) N1-SPT versus relative density for different
particle sizes.

N-SPT
0 10 20 30 40 50 60
0
(a) Gibbs and Holtz (1957)
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

50
Vertical effective stress (kPa)

100

150

200 Id=90%

80%
20% 30% 50% 60% 70%
250
0
(b) Gibbs and Holtz (1957)
Kulhawy and Mayne (1990)
20 Cubrinovski and Ishihara (1999)
Relative density, Id

40

D50=2 mm
60

D50=1 mm
80
D50=0.5 mm
D50=0.15 mm
100
0 10 20 30 40 50 60 70 80
N1-SPT
Page 41 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 4. SPT-N and Vs empirical relations presented in Table 2.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

450 Kanai (1966)


Shibata (1970)
Imai and Yoshimura (1970)
Ohba and Toriumi (1970)
400
Ohta et al. (1972)
Ohsaki and Iwasaki (1973)
Ohta and Goto (1978)
350 Imai and Tonouchi (1982)
Seed et al. (1983)
Sykora and Stokoe (1983)
Shear wave velocity, Vs (m/s)

300 Okamoto et al. (1989)


Lee (1990)
Lee (1992)
Pitilakis et al. (1992)
250 Raptakis et al. (1995)
Rollins et al. (1998)
Pitilakis et al. (1999)
200 Hasancebi and Ulusay (2007)

150

100

50

0
0 10 20 30 40 50 60
N (SPT)
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 42 of 54

Fig. 5. Variation of qc1/(N1)60 ratio with mean grain size, D50.


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

1.6
Mayne et al. (1995) - Eq. 15
Anagnostopoulos et al. (2003) - Eq. 17
Robertson et al. (1983)
Andrus and Youd (1989)
1.2 Seed and De Alba (1986)

[qc1/(N1)60]max=0.764D500.26
qc1/(N1)60

0.8
CANLEX

PÉRIBONKA
0.4

[qc1/(N1)60]min=0.483D500.19

0
0.1 1 10
D50 (mm)
Page 43 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 6. Variations of the upper and lower bounds of Vs1 with e.

500
Vs1 interval
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Seed et al. (1986) - Pyramid material (D50 ≈ 15mm)

450 Seed et al. (1986) - Orivil


Normalised shear wave velocity, Vs1 (m/s)

Seed et al. (1986) - Sandstone


Kokusho and Yoshida (1997) - D50 = 0.34 mm
Kokusho and Yoshida (1997) - D50 = 1.13 mm
400 Kokusho and Yoshida (1997) - D50 =7.3 mm
Kokusho and Yoshida (1997) - D50 = 2.28 mm
Huang et al. (2004)
350 Wichtmann and Triantafyllidis (2009)
Wichtmann and Triantafyllidis (2009)
Hardin and Richart (1963)
Wei et al. (1996)
300 Wei et al. (1996)
Robertson et al. (1995)
Menq (2003) - D50 = 5 mm

250 Lo Presti et al. (1992)


Hardin and Richart (1963)
Lo Presti et al. (1987)
Salgado et al. (2000)
200

150

100 Fine to coarse sand

Gravel content 0 % Gravel content ≥ 50 %


50
FC ≥ 50 %

0
1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1
Void ratio, e
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 44 of 54

Fig. 7. Variations of the upper and lower bounds of Vs1 with e from Menq (2003) experimental data.

500

Increasing Cu
Vs1 Interval according to relations of table 1
Menq (2003) - D50 =5 mm (Cu=2, 10 and 30)
450
Normalised shear wave velocity, Vs1 (m/s)

Menq (2003) - D50 = 2 mm (Cu=2, 10 and 30)


Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Menq (2003) - D50 =0.5 mm (Cu=2, 10 and 30)


400 Menq (2003) - D50 =0.2 mm (Cu=2, 10 and 30)

350

300

250

200

150

100 Fine to coarse Sand

Gravel content 0 % Gravel content ≥ 50 %


50 FC ≥ 50 %

0
1.2 1.1 1 0.9 0.8 0.7 0.6 0.5 0.4 0.3 0.2 0.1
Void ratio, e
Page 45 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 8. Comparison between experimental Vs1‒e data on TS sand with D50 = 0.34 mm and Cu = 1.95
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970)
and Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 0.82.

450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1) 0.25(D50)0.1625 with ((N1)60-1.6)/Id2=60+25Log(D50)
60
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


TS Sand (D50=0.34 mm, Cu=1.95)
Seed and Idriss (1970) - D50 unkown
350
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Seed et al. (1983) - Vs=56.4N0.5


with ((N1)60-1.6)/Id2=60+25Log(D50)
300 See Eq. 13 and section 3.3

Id ≈ 0 %
250 (N1)60=1.6

∆Vs1 ≤ 150 m/s


200

150 Id ≈ 100%
(N1)60= 49.9

100

50
(a)
0
450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1) 0.25(D50)0.1625 with ((N1)60-1.6)/Id2=60+25Log(D50)
60
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


TS Sand (D50=0.34 mm, Cu=1.95)
Data from Robertson et al. (1995)
350 Ottawa sand (D50 = 0.35 mm)
Seed and Idriss (1970) - ajusted by a factor of 0.82
Data from Menq (2003) - D50=0.5 and Cu=1.5
300 Data from Taya et al. (1999) - D50=0.58 and Cu=3.8
Data from Anastasiadis et al. (2012)
D50 = 0.27 mm (Cu=1.58) and D50 = 0.56 mm (Cu=2.76)
250
Id ≈ 0 %
∆Vs1 ≤ 150 m/s

200 (N1)60= 1.6

150

100 Id ≈ 100%
(N1)60= 49.9

50
(b)
0
1 0.9 0.8 0.7 0.6 0.5
Void ratio, e
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 46 of 54

Fig. 9. Comparison between experimental Vs1‒e data on G25 gravel with D50 = 1.13 mm and Cu = 5.65
(Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss (1970)
and Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.10.

450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; with ((N1)60-2.3)/Id2=60+25Log(D50)
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=1.13 mm, Cu=5.65)
Seed and Idriss (1970) - D50 unkown
350 Seed et al. (1983) - Vs=56.4N0.5
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

with ((N1)60-2.3)/Id2=60+25Log(D50)
See Eq. 13 and section 3.3
300

∆Vs1 ≈ 150 m/s


Id ≈ 0 %
250 (N1)60=2.3

200

150 Id ≈ 100%
(N1)60≈ 63.6

100

50
(a)
0
450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; with ((N1)60-2.3)/Id2=60+25Log(D50)
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=1.13 mm, Cu=5.65)
Bui (2009)
350 Seed and Idriss (1970) - Ajusted by a factor of 1.1

300
∆Vs1 ≈ 150 m/s

Id ≈ 0 %
250 (N1)60=2.3

200
Id ≈ 100%
150
(N1)60≈ 63.6

100

50
(b)
0
0.6 0.55 0.5 0.45 0.4 0.35 0.3 0.25
Void ratio, e
Page 47 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 10. Comparison between experimental Vs1‒e data on G50 gravel with D50 = 2.28 mm and Cu =
11.3 (Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss
(1970) and Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.25.

450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; With ((N1)60-2.3)/Id2=60+25Log(D50)
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=2.28 mm, Cu=11.3)
Seed and Idriss (1970) - D50 unkown
350 Seed et al. (1983) - Vs=56.4N0.5 with
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

((N1)60-2.3)/Id2=60+25Log(D50)
See Eq. 13 and section 3.3

∆Vs1 ≈ 200 m/s


300
Id ≈ 0 %
250 (N1)60=2.3

200

150 Id ≈ 100%
(N1)60≈ 71.3

100

50
(a)
0
450
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; With ((N1)60-2.3)/Id2=60+25Log(D50)
400
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=2.28 mm, Cu=11.3)
Seed and Idriss (1970) - Ajusted by a factor of 1.25
350 Konno et al. (1993)
∆Vs1 ≈ 200 m/s

300
Id ≈ 0 %
250 (N1)60=2.3

200

150 Id ≈ 100%
(N1)60≈ 71.3

100

50
(b)
0
0.5 0.45 0.4 0.35 0.3 0.25 0.2
Void ratio, e
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 48 of 54

Fig. 11. Comparison between experimental Vs1‒e data on G75 gravel with D50 = 7.30 mm and Cu =
31.1 (Kokusho and Yoshida 1997) and those predicted by Eqs. 23 and 24 and: (a) Seed and Idriss
(1970) and Seed et al. (1983) correlations; (b) Seed and Idriss (1970) adjusted by a factor of 1.50.

550
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; With ((N1)60-2.3)/Id2=60+25Log(D50)
500
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=7.3 mm, Cu=31.1)
450 Seed et al. (1983) - Vs=56.4N0.5 with
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

((N1)60-2.3)/Id2=60+25Log(D50)
400 See Eq. 13 and section 3.3

∆Vs1 ≈ 200 m/s


Seed and Idriss (1970) - D50 unkown

350
Id ≈ 0 %
300 (N1)60=2.3

250
200
150 Id ≈ 100%
(N1)60≈ 83.9
100
50
(a)
0
550
Vs1max=118(N1)600.25(D50)0.18
Vs1min=104.6(N1)600.25(D50)0.1625; With ((N1)60-2.3)/Id2=60+25Log(D50)
500
Normalised shear wave velocity, Vs1 (m/s)

Kokusho and Yoshida (1997)


Sand (D50=7.3 mm, Cu=31.1)
Seed and Idriss (1970) - Ajusted by a factor of 1.5
450 Seed et al. (1986) - D50 = 15 mm - Oriville
Menq (2003) - D50= 8 mm, Cu=20
400
∆Vs1 ≈ 200 m/s

350
Id ≈ 0 %
300 (N1)60=2.3

250

200

150 Id ≈ 100%
(b) (N1)60 ≈ 83.9

100
0.4 0.35 0.3 0.25 0.2 0.15 0.1
Void ratio, e
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 49 of 54

Factor

Idriss (1970) correlation.

0.4
0.6
0.8
1.2
1.4
1.6
1.8

0.1
Fact= 0.22* ln(D50) + 1.065

1
Mean grain size, D50 (mm)
10
Fig. 12. Correlation between mean grain size (D50) and the correction factor to be applied to Seed and
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 50 of 54

Fig. 13. Comparison between Péribonka site results with those predicted by Eqs. 23 and 24: (a) Vs1 and
(N1)60 profiles; and (b) D50 profile.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Vs1 (m/s) D50 (mm)


0 50 100 150 200 250 300 350 400 450 500 1 10
180 180
a) Péribonka b)

175
175

170
170

165
Elevation (m)

165

160

160
155

155
150

150
145 (N1)60 - TF-88-05
Vs1 (MMASW) - PM 512 - line 15 upstream D50 from borings
Vs1 calculated D50 calculated
140 145
0 20 40 60 80 100 120 140 160 1 10

(N1)60
Page 51 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 14. Comparison between CANLEX project (Kidd site) results with those predicted by Eqs. 23 and
24: (a) Vs1 and (N1)60 profiles; and (b) D50 profile.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Vs1 (m/s) D50 (mm)


0 50 100 150 200 250 0.1 1
8 8
a) CANLEX - Kidd site b)

10 10

D50 av. = 0.2 mm


12 12
Elevation (m)

14 14

16 16

18 18

KD9305 KD9404
20 20
SPT/E KD9403
D50 estimated from KD9305
KD9402 Estimated velocity from KD9305
Estimated velocity from SPT/E D50 estimated from SPT/E
KD9405
22 22
0 20 40 60 80 0.1 1

(N1)60
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 52 of 54

Fig. 15. Comparison between Ohio River (site C) results with those predicted by Eqs. 23 and 24: (a)
Vs1 and (N1)60 profiles; and (b) D50 profile.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Vs1 (m/s) D50 (mm)


0 50 100 150 200 250 300 350 0.1 1 10
4 4
a) Ohio River b)
6 6

8 8

10 10

12 12
Elevation (m)

14 14

16 16

18 18

20 20

22 22
SPT - site C average
24 Vs1 average measured 24 D50 meseared
Vs1 calculted from SPT D50 estimated from SPT
26 26
0 20 40 60 80 0.1 1 10

(N1)60
Page 53 of 54
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record.

Fig. 16. Comparison between Vs1‒(N1)60 ranges estimated by Eqs. 23 and 24 at different grain sizes and
those predicted by Seed et al. (1983) and Seed and Idriss (1981) correlations.
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

500
Vs1min = 104.6(N1)600.25.(D50)0.1625
Vs1max = 118(N1)600.25.(D50)0.18
(N1)60/Id2=60+Log(D50)
Id=100%
450 Seed et al. (1983) - Sand
Seed and Idriss (1981) - All
CANLEX - D50av =0.2 mm m
5m
Péribonka - D50av = 1.8 mm D 50=
400 Ohio River - D50=2.15±0.39
Ohio River - D50=1.12±0.24
Ohio River - D50=0.46±0.04
Ohio River - D50=2.24±0.35

350
2 mm
D 50=

∆Vs1 > 325 m/s


Id=30%

∆Vs1 ∼ 150 m/s


300
Vs1 (m/s)

m
0.8 m
D 50=

250

200 0.2 mm
D 50=
∆Vs1 ∼ 100 m/s

150

100

50
0 10 20 30 40 50 60 70 80
(N1)60
For personal use only. This Just-IN manuscript is the accepted manuscript prior to copy editing and page composition. It may differ from the final official version of record. Page 54 of 54

Fig. 17. Normalized shear wave velocity, Vs1 as function of (N1)60, e, Id, and D50 for uncemented,
Holocene-age soils.

500
Can. Geotech. J. Downloaded from www.nrcresearchpress.com by San Diego (UCSD) on 09/10/15

Id=85% Id=100%
450

Id=65%
400 e=0.2

5 mm
D 50=
Id=50% e=0.25
350
e=0.3
Id=30% m
2m
300 D 50=
Vs1 (m/s)

e=0.4

1 mm
D 50= e=0.5
250
m e=0.6
0.5 m
D 50=
e=0.7
200
m
0.2 m
D 50= e=1

150
Vs1min = 104.6(N1)600.25(D50)0.1625
Vs1max = 118(N1)600.25(D50)0.18
emin=0.3346 D50-0.491
100
emax-emin=0.23+0.06/D50
(N1)60/Id2=60+Log(D50)
50
0 10 20 30 40 50 60 70 80
(N1)60

You might also like