Você está na página 1de 15

458

Development of a Vs30 (NEHRP) map for the city of Ottawa, Ontario, Canada
D. Motazedian, J.A. Hunter, A. Pugin, and H. Crow

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Abstract: Four different seismic methods were used extensively to evaluate the shear wave velocity of soils and rock in the city of Ottawa, Canada, from which the travel-time weighted average shear wave velocity (Vs) from surface to 30 m in depth (Vs30) and the fundamental frequency (F0) were computed. Three main geological or geotechnical units were identified with distinct shear wave velocities: these consist of very loose thick post-glacial fine-grained sands, silts, and clays (Vs <150 m/s, thickness up to 110 m), firm glacial sediments (Vs *580 m/s, thickness *3 m), and very firm bedrock (Vs *17503550 m/s). The seismic methods applied were downhole interval Vs measurements at 15 borehole sites, seismic refractionreflection profile measurements for 686 sites, high-resolution shear wave reflection landstreamer profiling for 25 km in total, and horizontal-to-vertical spectral ratio (HVSR) of ambient seismic noise to evaluate the fundamental frequency for *400 sites. Most of these methods are able to distinguish the very high shear wave impedance of and depth to bedrock. Sparse earthquake recordings show that the soil amplification is large for weak motion when the soil behaves linearly. Key words: seismic site classification, shear wave velocity, seismic refractionreflection, downhole. Resume : Quatre methodes sismiques differentes ont ete grandement utilisees afin devaluer la vitesse des ondes de cisail` lement des sols et roches dans la ville dOttawa, Canada, a partir desquelles la vitesse moyenne des ondes de cisaillement ` ponderee selon le temps de parcours (Vs) de la surface jusqua une profondeur de 30 m (Vs30) et la frequence fondamen tale (F0) ont ete calculees. Trois unites geologiques ou geotechniques principales ont ete identifiees selon des vitesses des ` ondes de cisaillement distinctes : des sables, silts et argiles post-glaciaires fins, laches et epais (Vs <150 m/s, jusqua ` 110 m depaisseur), des sediments glaciaires fermes (Vs *580 m/s, *3 m depaisseur) et du substratum rocheux tres ferme (Vs *17503550 m/s). Les methodes sismiques appliquees etaient des mesures de Vs par intervalle en fond de fo rage pour 15 sites de forage, des mesures du profil de refractionreflexion sismique pour 686 sites, du profilage de la re` flexion des ondes de cisaillement a haute resolution landstreamer pour 25 km lineaire au total, et le ratio spectral horizontalvertical (RSHV) du bruit sismique ambiant pour levaluation de la frequence fondamentale sur environ 400 sites. ` La majorite de ces methodes sont capables de distinguer limpedance tres elevee aux ondes de cisaillement et la pro fondeur jusquau substratum rocheux. Quelques mesures de seismes montrent que lamplification du sol est grande pour des mouvements faibles lorsque le sol de comporte de facon lineaire. Mots-cles : classification sismique des sites, vitesse des ondes de cisaillement, refractionreflexion sismique, fond de forage. [Traduit par la Redaction]

Introduction
The National Building Code of Canada (NBCC 2005) has been amended to include the influence of local geology on the prediction of earthquake ground motion based on the growing body of knowledge on the response of soft soils to earthquake shaking. The NBCC (2005) is a step forward in the integration of soft-soil response for routine earthquake
Received 13 August 2009. Accepted 30 August 2010. Published on the NRC Research Press Web site at cgj.nrc.ca on 11 March 2011. D. Motazedian.1 Earth Sciences Department, Carleton University, 1125 Colonel By Drive, Ottawa, ON K1S 5B6, Canada. J.A. Hunter, A. Pugin, and H. Crow. Terrain Geophysics Section, Northern Division, Geological Survey of Canada, 601 Booth Street, Ottawa, ON K1A 0E8, Canada.
1Corresponding

author (e-mail: Dariush_Motazedian@Carleton.ca).

design by utilizing the geotechnical or geophysical description of soils and rocks. The description of sites is primarily based on the measured travel-time weighted average shear wave velocity (Vs) of a site from surface to a depth of 30 m (Vs30). The city of Ottawa is located in a seismically active area called the Ottawa West Quebec Seismic Zone (see map in Adams and Halchuk 2003), which extends from Montreal, Quebec, to Ottawa, Ontario, Canada. Ottawa has been identified as the city with the third highest earthquake risk in Canada. Since 2005, Carleton University and the Geological Survey of Canada (GSC) have applied different geophysical methods to carry out the site classification measurements within the city of Ottawa. The current paper provides Vs30 and fundamental frequency maps for the city of Ottawa along with the description of different geophysical methods applied in the study area. The distribution of the seismic site classes is directly relevant to emergency response planning and seismic mitigation strategies because the densely populated urban areas on soft soils are likely to experience more
Published by NRC Research Press

Can. Geotech. J. 48: 458472 (2011)

doi:10.1139/T10-081

Motazedian et al.

459

damage than firm rock during a significant earthquake, all factors being considered equal. Classification of the amplification characteristics of soils allows a qualitative assessment to be made for the prioritization of buildings for seismic retrofitting, the siting of new critical infrastructure as well as assessing the vulnerability of linear utilities (e.g., gas lines, water mains, and power lines) and linear transportation corridors (e.g., railways, highways) (see Levson et al. 1998). The map is also relevant to the insurance industry for better assessing their exposure to earthquake risk and to aid in calculating fair premiums that better the reflect variability of local seismic hazards (Clark and Khadilkar 1991; Smolka and Berz 1991; Finn et al. 2004). Detailed bedrock and surficial geology maps are available for the Ottawa area (Belanger 1998). In addition, the threedimensional configuration of soil and rock can be derived from a geological compilation from approximately 21 000 water well and geotechnical boreholes (available from the GSC website: gsc.nrcan.gc.ca/urbgeo/natcap/index_e.php). From these data, and prior knowledge of the shear wave velocity characteristics of the rock and soils in the area, a generalized surficial geology map of the city was developed (see Fig. 2 in Motazedian and Hunter 2008), consisting of three basic outcrop units: bedrock, glacial till, and late glacial or post-glacial sediments (hereafter referred to as postglacial). The bedrock outcrop, which covers 20% of the city of Ottawa, consists of either Pre-Cambrian gneisses or lower Paleozoic sedimentary rock (limestone, dolostone, sandstone, or shale): very little mechanical weathering of bedrock is evident because of late-stage Pleistocene glacial scouring, and most outcrops indicate firm hard rock. Glacial till and glacially derived sediments, covering 15% of the area, overlie bedrock and are relatively thin (14 m), but can thicken locally in narrow bedrock topographic lows. Post-glacial sediments, which cover about 65% of the area, consist of fine sand, silt, and clay that were deposited during or immediately after the occupation of the area by an extensive body of seawater called the Champlain Sea in the Ottawa area (11 5009800 BP). Thus, from a generalized geotechnical point of view, 65% of the area in the Ottawa region is covered by deposits of soft post-glacial sediments with thicknesses that can exceed 100 m overlaying very firm bedrock (see Fig. 2 in Motazedian and Hunter 2008). The National Earthquake Reduction Program of the United States (NEHRP) has recommended a classification of soil and rock sites based on one of three geotechnical schemes: average shear strength (Cu), average standard blow count (N) to a depth of 30 m, or the travel-time weighted average shear wave velocity to a depth of 30 m (Vs30). Of these, Vs30 is currently in use worldwide. Because of its correlation with the seismic soil amplification characteristics, it appears to be the most versatile of the three approaches. The National Building Code of Canada (NBCC 2005) has also adopted the NEHRP recommendations for site classes (BSSC 1997). The fundamental frequency (F0) of a site (or site period T0 = 1/F0) is another important factor that is investigated in our research activities. This is important for the evaluation of a possible match between the resonance frequency of the building and the fundamental frequency of the site. The value of F0 is added to our research activities to complement the application of Vs30. Be-

cause of the unique geological setting in eastern Canada (very loose soil underlain by very hard bedrock), we are planning to provide region-specific soil amplification factors for eastern Canada based on both Vs30 and F0 in future. However, the focus of this article is on techniques applied for the measurement of Vs30 and F0. As shown in Table 1, five classes (A through E) are defined based on the ranges of Vs30. A sixth class (F) requires site-specific geotechnical investigations for soils, which are identified as containing organics, liquefiable, highly sensitive, or are otherwise susceptible to failure under seismic loading (NBCC 2005). The value of Vs30 can be sensitive to the thickness and shear wave interval velocity of the layers when the value of this parameter is near to the border line between site classes (e.g., 180, 360, 760, and 1500 m/s). This shortcoming is most severe for cases where the shear wave velocity contrast between geological layers is very high and the boundary between soil and the bedrock lies at depths less than 30 m. For example, for a simple model of a two-layer site in the Ottawa area located on a soft soil with an average shear wave velocity of 140 m/s overlying bedrock with a shear wave velocity of 2300 m/s, a small overestimation or underestimation of the depth of bedrock can lead to differing NEHRP site classes (D versus E). Extreme conditions could exist where the shear wave velocity of the soft sediments of post-glacial sediments are as low as 70 m/s and that of the bedrock as high as 3550 m/s, which could increase the sensitivity of NEHRP site classification to define more accurately the shear wave velocities for any particular site. Thus, it is necessary, and interesting, to investigate the capability of different methods to find more accurate shear wave velocities and thicknesses within each geological unit. To provide reliable and robust shear wave velocity depth profiles unique to the study area, several methods, including seismic refractionreflection (Whiteley and Greenhalgh 1979; Williams et al. 2000, 2003; Motazedian et al. 2006; Hunter et al. 2007b; Motazedian and Hunter 2008), horizontal-to-vertical spectral ratio (HVSR) (see Nakamura 1989; SESAME 2004), borehole measurements (Galperin 1985; Hunter et al. 1998a, 1998b, 1998c, 2002, 2007a), and high-resolution seismic reflection profiling (Pugin et al. 2007, 2008), have been applied to develop a microzonation map based on Vs30 and F0 measurements (see Fig. 1 for the location of the sites). It is of interest to apply different methods, as they may contribute complementary information. The current paper provides an overview of different geophysical methods being conducted in the study area, along with results to date. It should be mentioned that a further method, the multichannel analysis of surface waves (Park et al. 2005), was applied at 36 sites in the study area and has been published as a separate article (Khaheshi Banab and Motazedian 2010).

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Downhole shear wave seismic method


Downhole seismic surveying techniques for seismic prospecting for oil in deep boreholes have been used worldwide for many years (Galperin 1985). More recently, these have been used for engineering purposes in shallow boreholes (Hunter et al. 1998a). In the last 40 years, shear wave techniques have been applied to measure soil and rock
Published by NRC Research Press

460

Can. Geotech. J. Vol. 48, 2011 Table 1. Vs30 site classification for seismic site response as defined by NEHRP (1994) and adapted by the 2005 National Building Code of Canada (see Finn and Wightman 2003). Site class A B C D E F Generic description Hard rock Rock Very dense soil and soft rock (firm horizon) Stiff soil Soil profile with soft clay Site-specific geotechnical investigation required Range of Vs30 (m/s) >1500 7601500 360760 180360 <180

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Fig. 1. Locations of shear wave velocity measurement sites within city of Ottawa.

properties to examine soft-soil earthquake response (Warrick 1974). In Canada, previous surveys for downhole shear wave velocity measurements in soils related to earthquake response were conducted by Hunter et al. (1998a, 1998b, 1998c). For this present work, we followed the procedure developed by Hunter et al. (1998a, 2002). We used a surface polarized shear wave source positioned between 3 and 5 m offset from the borehole, a single three-component 10 Hz Geostuff wall-lock geophone, and a Geometrics Geode seismograph as a recorder. The seismic source used was either a 7.5 kg hammer striking the side of an imbedded steel I-beam

or an IVI Minivib Mark I swept-frequency vibrator source (Pugin et al. 2007) positioned in horizontal mode at a known azimuth from the horizontal geophone components. The three-component receiver was moved up from the bottom of the cased borehole in 0.5 m increments. A composite suite of borehole records from one of the radial horizontal components is shown in Fig. 2. At most sites, the dominant frequencies recorded downhole were in the 2090 Hz range. Interactive shear wave event picking utilized IXSeg2Segy software (Interpex 2008). The IXSeg2Segy software is a program designed to reach multichannel seismic data and allows the user to process the data using
Published by NRC Research Press

Motazedian et al. Fig. 2. Sample downhole seismic record in post-glacial soft sediments showing onset of shear wave energy. Note interfering tubewave noise at locations where casing is poorly bonded to the formation. Note also reflection from bedrock surface, which is at 28 m in depth.

461

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

band-pass filtering, linear refraction, and hyperbolic reflection arrival time velocity fitting. The software allows the user to model a layered earth model based on reflected and refracted seismic waves. Independent measurements were made of shear wave arrival times with each horizontal component, and the average shear wave velocity values were computed from all results. Where the coupling between borehole casing and formation was poor, because of bridging during grouting, or where a substantial unfilled void occurred behind the casing, signal-generated tube-wave noise was generated; in such cases, three-component particle motion hodographs were examined to estimate shear wave onset time using interactive V-Shear software (Kassenaar 2008). Shear wave arrival times could commonly be estimated to the nearest sample interval using hodograph assistance; hence, seismic traces were routinely over-sampled (e.g., 0.125 ms sample rate) for all downhole surveys. Exceptions occurred (0.5 ms sample rate) when the Minivib was employed, because of limitations on recording capabilities, since all data was recorded in an uncorrelated format (7.5 s of data) for later processing. The travel time was corrected for the offset distance (the distance between the seismic source and borehole), and shear wave arrival times were processed using running least-squares fits. Interval velocities were obtained over vertical distances of 1 m (three-point fit) or 2 m (five-point fit) and plotted at 0.5 m spacings. When the hammer source was used, it was possible to compute average velocities directly from the surface to the detector at depth, so that Vs30 could be estimated directly from a single downhole measurement (although this approach is not recommended). Direct measurement of the average shear wave velocity (Vs) is not always possible when using a Minivib source, since the first arrival is commonly emergent and poorly constrained. However, Minivib zero-phase correlated traces usually give superior signal-to-noise results, and interval velocity deter-

minations are usually well constrained. It is preferable that all downhole surveys for Vs30 estimations be conducted using short increments (0.5 m) between sonde locations, so that correlations of downhole shear wave arrival times and interval velocities can be used to guide the estimate of travel time at the depth of 30 m. Figure 3 shows first arrival travel-time picks for the same borehole, as given by the composite record shown in Fig. 2. The surface hammer source was at 3 m offset from the borehole, and the travel times shown are the average of towards and away picks from both horizontal components of motion. A three-point running least-squares fit was applied to the depth-corrected data to obtain interval velocities over a depth interval of 1 m. Since the error associated with the estimation of 95% confidence limits (conf. limits) for a three-point is itself very large, the 2s error limits shown can only be used as a qualitative estimate of the scatter in the travel-time data. Downhole shear wave velocity surveys were performed in 15 boreholes throughout the city of Ottawa. Some boreholes were drilled by GSC as part of the current research, and others were offered by geotechnical companies as part of other site surveys. All surveys were done using the methodology given above. This invasive form of seismic measurements is relatively expensive compared to other seismic techniques because of the costs of drilling and casing grouting; however, such measurements are commonly well constrained and (in the experience of one of us, J.A. Hunter) probably with superior vertical shear wave velocity accuracy compared to most other techniques (the exception being the seismic cone penetrometer test, SCPT, which was not used in this survey; see Hunter et al. 1991).

Seismic refraction and reflection methods


Details of seismic refraction and reflection methods for the microzonation studies for the city of Ottawa are given in Motazedian and Hunter (2008). Open green spaces, which occur throughout the city of Ottawa, gave an opportunity to design surface geophone arrays so as to obtain estimates of shear wave velocity versus depth with refraction and reflection methods (Whiteley and Greenhalgh 1979; Hunter et al. 1998b; Williams et al. 2000, 2003). Because of the large shear wave velocity contrasts between post-glacial sediments and bedrock, a basic linear horizontal geophone array of 24 geophones (8 Hz) at 3 m spacings, and various source offsets (array midpoint and reversed off-end locations up to 30 m) were used. In many cases, this configuration afforded the opportunity to observe SH refraction events from the bedrock interface to a depth of about 30 m, since a location for a linear configuration of *120 m length could be found in most parks and roadside zones. Where the bedrock interface was found to be deeper (up to 100+ m below surface), the same array configuration could be used to observe SH reflections from the interface, as well as SH direct arrivals from the overburden. Also, where more surface space was available, the array spacing was occasionally changed to 5 m geophone intervals to capture bedrock refraction arrivals from interfaces >30 m or to obtain improved move-out for bedrock
Published by NRC Research Press

462

Can. Geotech. J. Vol. 48, 2011

Fig. 3. Downhole shear wave survey results: (a) first arrival travel time and (b) interpreted interval Vs and average Vs. s, standard deviation; conf. limits, confidence limits.

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

reflections. The seismic source used was a 7.5 kg hammer striking the side of an imbedded steel I-beam. An example of a field record showing refraction first arrivals and the main reflection event, as well as a refraction and reflection velocity model developed from the refraction reflection method, is given in Fig. 4 in Motazedian and Hunter (2008). The refraction records of many site locations show a thin high-velocity surface layer. In undisturbed terrain, this layer has been identified as an overconsolidated zone associated with the post-glacial sediments 15 m thick (Eden and Crawford 1957); in other areas, this surface layer is interpreted to be associated with consolidated fill materials in parks and green spaces or with subgrade compaction along roadbeds. Commonly, this event is registered as signal-generated noise only on the first few near-source traces. The Interpex Seg2SegY software (Interpex 2008) was used to display, filter, and pick the first arrival events on all records. Standard refraction interpretations were employed using the slopeintercept refraction method, after first arithmetically averaging forward and reverse velocity segments to remove the apparent velocity effect of dipping layers. Interval velocitydepth layering was then converted to average shear wave velocity (Vsav) versus depth data to compute a travel-time weighted mean curve through all data points. Having obtained a plot of Vsav versus depth from the refraction data, the value for Vs30 was then obtained from the curve for the first 30 m. At sites where bedrock occurred at depths less than 30 m, the interpreted bedrock shear wave velocity was assumed down to a depth of 30 m. Figure 4 shows a typical field record suite for a case where the bedrock surface is very deep. In this situation, hyperbolae were fit to the reflection data, as shown, using the Interpex IXSEG2SEGY software. This was done for all visible wide-angle reflections; depth versus average velocities

were plotted for all forward and reverse records collected at each site. Dipping layers were identified by nonhyperbolic travel times of wide-angle reflections (either flattened or steep curves) and the differing forward and reverse intercept times. The Vsav versus depth curves were fit through all forward and reverse data including dipping layers to obtain the average value for shear wave velocity. In addition, surface refraction velocities were commonly assigned to both the overconsolidated zone (rapidly attenuating first arrival visible on the near traces only) and the underlying near-surface post-glacial sediments. If glacial materials were present at a site, two prominent reflections were generally observed in the sections at depth, which were interpreted as the top of the glacial sequence as well as the bedrock surface. The sharp impedance contrast between the post-glacial soils and the glacial tills causes this reflection to become more prominent than the underlying reflection from the tillbedrock interface (because of strong energy partitioning at the post-glacial and glacial surface as well as masking defractions from cobbles within the glacial materials). Because of the compact nature of the glacial till and its high velocities (500800 m/s), the glacial material is considered a firm horizon and is not distinguished from the bedrock for the purposes of estimating depth to firm ground for site period calculation (Crow et al. 2010). Both refraction and reflection interpretations contribute to Vsav estimates, depending on the particular site. Within the areas of the city underlain by thick post-glacial sediments (65% of the area), wide-angle reflection techniques are preferred; however, such techniques do not offer any shear wave velocity information about the underlying bedrock. Refraction methods, on the other hand, are best applied in the other 35% of the Ottawa area underlain by glacial materials and (or) bedrock, where accurate estimates of shear
Published by NRC Research Press

Motazedian et al. Fig. 4. Time-averaged shear wave velocities were measured using hyperbolic curves fit to the reflections. Average shear wave velocity profiles were derived from reflection velocities fit to all forward and reverse shots collected at each site.

463

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

wave velocities can be obtained for glacial and bedrock materials. It should be cautioned that, in some cases, where substantial thicknesses of glacial material underlie significant amounts of post-glacial sediments, the velocity structure of these and the underlying bedrock, could possibly result in significant errors in layer thickness estimation using refraction techniques, since a hidden layer situation could result (Kaila and Narain 1970). Errors as large as 36% in depths to interfaces could occur in such isolated cases. Fortunately, glacial material beneath the post-glacial sediments is usually quite thin (an analysis of 31 000 area boreholes yields a median thickness value of 1.52 m); hence, the error of the depth estimate to bedrock is normally low (<10%) using refraction techniques. Even where the hidden layer situation exists, the estimated shear wave velocity of the bedrock refractor is not affected, and the arithmetically averaged forward and reverse shear wave velocities remain the best estimates using surface noninvasive techniques. The refractionreflection site method was applied as the mainstay of our work within the city of Ottawa. Six hundred and eighty-six sites were studied throughout the area to provide variations in the regional shear wave velocity depth structure.

Landstreamer shear wave reflection profiling


High-resolution P-wave seismic methods have been applied in the areas of Champlain Sea sediments near Ottawa for many years (Hunter et al. 1984). Such studies have shown the unusually transmissive characteristics of these soft high-water-content soils. Preliminary tests (Neave and

Pullan 1989; Pullan et al. 1991) also indicated that useable high-resolution shear wave reflections may also be obtained. A technique to obtain continuous common midpoint (CMP) shear wave reflection survey sections in an urban environment has been developed by Pugin et al. (2007). This method uses a 2448 channel streamer of horizontal geophones mounted on sleds behind a horizontally polarized swept-frequency vibrator source (Minivib). This technique was used at sites where there were considerable thickness variations of the post-glacial materials and where buried bedrock valleys were suspected from limited borings. Where the streamer length parameters were adequate, it was possible to obtain both average shear wave velocities down to infra-overburden layers as well as to bedrock. With the design of landstreamer technologies, Pugin et al. (2008) has shown the cost effectiveness of towed P and S arrays in a noisy urban environment. With current array designs consisting of source spacings of 1.5 m and 48 channels at 0.75 m geophone spacing, data acquisition rates of more than 3 km/day can be achieved. The velocitydepth curves were accurate enough to yield estimates of Vs30, and velocity analyses at horizontal intervals as close as 15 m spacing were obtained. Over 25 km of continuous seismic sections with Vs30 measurements were obtained in key areas of the city. Figure 5 shows a typical field array using a Minivib horizontal vibrator. (For most surveys, we used a sweptfrequency signal of 5120 Hz over a transmit time of 7.5 s). An example of landstreamer velocity determination is shown in Fig. 6 from the Kinburn area in the northwestern portion of Ottawa. During the development of the landstreamer, the length of streamer was adjusted to obtain
Published by NRC Research Press

464 Fig. 5. A 48-channel three-component landstreamer with IVI MiniVibTM source.

Can. Geotech. J. Vol. 48, 2011

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

higher accuracy in shear wave velocity determination (Pugin et al. 2008). The estimated error of average velocities to the base of the unconsolidated overburden area is approximately 5%. The associated interval velocity between successive reflections is on the order of 10%. Hence, the shear wave velocity structure of the overburden can be contoured with a considerable degree of accuracy as an aid to determination of the primary stratigraphic units in the Ottawa region. This example is characteristic of the Ottawa area in that the upper portion of the Champlain Sea sediments contains low shear wave velocity material, whereas the lower portion of the post-glacial unit has somewhat higher velocities. Both these units overlie glacially derived material, which exhibits a considerably higher interval velocity. If the depth of the overburden exceeds 30 m, then it is possible to estimate Vs30 by interpolation between the reflecting horizons using the average velocities. Where the overburden thickness is less than 30 m, the average velocities to the bedrock surface along with estimates of the bedrock shear wave velocity can be combined to compute Vs30.

Spectral ratio method from ambient background noise


Measurements of site fundamental frequency, bedrock depth, and site amplification using spectral ratio methods are commonly used to characterize soft-soil response. Three spectral ratio methods have been considered in our site effects studies: standard spectral ratio (SSR) between earthquakes recorded on soft soil and adjacent rock outcrop sites, HVSR from earthquake ground motions, and HVSR from background noise. The SSR method will be discussed later.

For HVSR applications using ambient background noise, the ratio of the horizontal spectrum to vertical spectrum of background noise recorded at the surface of a soil layer site was considered to provide two important parameters: (i) the fundamental site frequency (or period) corresponding to a peak value in the horizontal to vertical ratio (H/V) curve versus frequency (Nakamura 1989; Bard 1999) and (ii) a possible indication of seismic site amplification, assuming that the vertical component is not significantly amplified relative to the horizontal; however, we suspect there will be at least some amplification of the vertical, but it may be compensated by the turning of rays towards the vertical (Sv). In general, this method is thought to provide accurate estimates of fundamental site resonance; the resonance peak amplitude is also believed to yield an estimate of soil amplification. In our studies, we used SESAME (2004) guidelines and software, which were developed by 14 European research institutes involving 85 researchers (Site Effects Assessment Using Ambient Excitations (SESAME), European Commission). These researchers have shown that fundamental site periods computed using HVSR correlate closely to those computed from the shear wave transfer function for numerous soilrock structures. They have also suggested that the estimated resonant peak amplitudes may represent a lower bound to broadband amplification for small-strain (linear soil response) seismic events. Our survey procedure for measurement of resonant frequencies using HVSR ambient noise included the use of a Tromino seismograph especially designed for such measurements (available from www.tromino.It). At each site, we recorded three components of ambient noise for 30 min at a digital rate of 128 samples/s. The accompanying Tromino software was fashioned after the recommendations of the
Published by NRC Research Press

Motazedian et al.

465

Fig. 6. Shear wave landstreamer seismic section within Ottawa showing average and interval shear wave velocity variations within postglacial Champlain Sea sediments. Structure within overburden is also distinguishable.

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

SESAME group; we used 60 s time series windows to obtain Fourier spectra and summed all stable spectral windows with KonnoOmachi smoothing. We occupied approximately 400 sites where soft-soil thicknesses were >10 m. Of these, 185 sites were colocated where we had obtained refractionreflection information or downhole shear wave velocities (as given in Fig. 1) and where the shear wave velocity depth structure was well determined. Figure 7 shows an example of HVSR calculated for broadband seismic station ORHO, Orleans subdivision, city of Ottawa, using the Tromino equipment. A prominent peak at 0.77 Hz is interpreted to be the fundamental resonance frequency, F0. The zone of low spectral ratio (below a value of 1) between about 1.5 and 40 Hz results from the effect of a high shear wave velocity screening layer, i.e., the near-surface overconsolidated zone (Castellaro and Mulargia 2009) and has no bearing on the amplification or deamplification of ground motion in that area of the frequency spectrum. The fundamental site periods obtained from the HVSR method for the 185 sites were compared with the fundamen-

tal period given by T0 = 4H/Vsav, where H is the soil thickness and Vs is the soil shear wave velocity using a single layer overburden model (as discussed in NBCC (2005)) and the measured shear wave velocity site data. This equation was derived by Okamoto (1973) for the case of a shear wave propagating vertically through the bedrock. When the shear wave velocity is not constant within a layer, as a widely used approximation, the fundamental period is estimated by T0 = 4H/Vsav, where Vsav is the travel-time weighted average shear wave velocity from ground surface to the bedrock (Madera 1970; Dobry et al. 1976; Hadjian 2002; NBCC 2005). The fundamental site frequencies obtained from the HVSR method for 185 of these sites were compared with the fundamental period given by T0 = 4H/ Vsav. A single layer overburden model is discussed in NBCC (2005); hence, in consideration of the large velocity contrast between soil and rock, a two-layer model (soil over bedrock) was assumed throughout. The depth to bedrock and the average shear wave velocity of soft soils were determined at all seismic reflectionrefraction sites and borehole locations.
Published by NRC Research Press

466 Fig. 7. Horizontal/vertical spectral ratios for a site in Ottawa (adjacent to broadband seismic station ORHO, Orleans subdivision, city of Ottawa).

Can. Geotech. J. Vol. 48, 2011

Earthquake recordings
In addition to a broadband seismic station of the Canadian National Seismic Network (CNSN) and GSC strong motion stations in the city of Ottawa, two pilot sites have been installed in the eastern part of city, one on thick soil and one on bedrock, to compare the level of ground shaking from small local and large teleseismic earthquakes. Identical weak motion broadband seismometers and digitizers (Nanomertics Trillium P120 and Taurus) have been installed at both sites. The soil site (Heritage Park area, station ORHO) is located on 80.8 m of thick soft soil with low shear wave velocity (Hunter and Motazedian 2006), and the rock site (station ORIO) is located on a bedrock outcrop 1.5 km away from the soil site. For applications of the SSR method using earthquake data, both the amplitude spectra of horizontal and vertical components of motion shear wave from an earthquake recorded at a soil site was divided by the respective horizontal and vertical motion recorded on a nearby bedrock site (Borcherdt 1970). In the course of data processing, the shear-wave portion of signal (including direct, reflected, and refracted phases) was windowed. For each record the general procedures are as follows: (i) tapering the windowed time series using a 5% cosine taper on each end of the signal; (ii) zero-padding the time series to the next greatest power of 2; (iii) transforming to frequency domain by Fast Fourier Transform; (iv) removing instrument response; (v) transferring to time domain by applying the inverse Fourier transform; (vi) calculating response spectra for 5% damping from corrected acceleration time series; and (vii) discarding the frequencies with signal to noise ratio less than 2 in the frequency domain, where the pre-event noise is available (the record is discarded where the preevent noise is not available and the quality of signal is poor). Both vertical and horizontal components were compiled. In this method, the epicentral distance of the earthquake should be much larger than the distance between two nearby stations, which in our case is less than 1.5 km. The bedrock site was assumed to be free from amplification, and the SSR ratios are the response characteristics of the soil. The assumption was made that such effects as source, travel path, and recording instruments have been removed from the spectral ratios. Sixteen local and small earthquakes have been recorded at both soil and bedrock sites in Ottawa, including an earthquake with M2.1 (M, magnitude), depth of 10 km, and an epicentral distance of 15 km from both stations. Figure 9 shows the soil-to-rock spectral ratios for both root mean square (rms) combined horizontal components and vertical component. The soil amplification for each frequency at the soil station was obtained by dividing the spectrum of the soil station by the spectrum of the rock station. It is interesting to note that, for weak motion, the spectral seismic soil amplification factor could be as large as 100:1 at the fundamental resonance frequency (for the rms-combined horizontal components of motion). A large ratio was also found at higher frequencies for both horizontal and vertical components (the fundamental resonance frequency is 0.77 Hz for the soil site, as given by the ambient noise raPublished by NRC Research Press

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Fig. 8. Comparison between fundamental site period measured by passive noise monitoring, Tpassive (Tromino), and T0 = 4H/Vsav calculated from average shear wave velocity, Vsav, soil thickness, H, and travel-time measurements at seismic site locations, Tsite. Data was obtained from 185 independent measurement sites. Rcoef, coefficient of determination; Std. Err, standard error.

Figure 8 shows a comparison between the observed HVSR fundamental period and the T0 = 4H/Vsav calculations. As shown, the HVSR method systematically deviated from the calculated fundamental periods, even at relatively short periods (or shallow resonators). The variance is >30% for periods larger than 2 s, corresponding to impedance boundary depths >75 m.

Motazedian et al. Fig. 9. Comparison of vertical and horizontal component amplitude ratios of soil to rock from a magnitude M2.1 earthquake, which occurred 15 km from Ottawa. The soil station sits atop 93 m of soft post-glacial soil. The horizontal ratio indicates that significant amplification is occurring at the soil site at its fundamental frequency (0.77 Hz).

467

made here, using near-surface geophysics or passive monitoring approaches are all in the small-strain domain, and that strong motion large-strain measurements have not been made in post-glacial sediments in the study area.

Results and conclusions


To evaluate various geophysical techniques for seismic soil classification methodologies for the Ottawa region, measurements were made at 686 shear wave seismic reflectionrefraction sites, 15 borehole sites for shear wave downhole studies, and 400 HVSR sites. Approximately 25 km of high-resolution shear and P-wave landstreamer profiling was done to examine detailed subsurface structure and shear wave velocities. Lastly, two temporary seismic stations were deployed, one on bedrock and the other adjacent one on thick soft soil, to examine soil amplification effects using local earthquakes. The summaries and discussions of our results are as follows:  Downhole seismic method Downhole measurements were done in 15 boreholes in the study area. This is one of the most direct measurements of travel time from the surface to a specific depth; this technique provides accurate shear wave velocity when the casing in the borehole is properly grouted to the soft-soil formation. The downhole technique may not be cost effective for small projects because of the cost of drilling.  Seismic refractionreflection method The high impedance contrast between sedimentary deposits and the underlying bedrock in the Ottawa area provides clear refracted and reflected waves, which in the absence of strong background noise leads to reliable shear wave velocity depth profiles for soil site classification. This method is fast, practical, reliable, and inexpensive. We have customized the geophone array and field setup for this region and have covered more than 686 sites in the study area. This method has become the prominent method for soil classification in the Ottawa region.  High-resolution seismic reflection profiling method High-resolution P and S reflection profiling using multichannel CMP techniques and landstreamer technology is an important method where thick soft-soil conditions occur. In its present form, this technology can provide detailed lateral variations in subsurface stratigraphy in a timely and cost-effective manner. In addition, where array lengths are greater than twice the depth to subsurface targets, it is possible to obtain very detailed estimates of Vs30 in such stratified soils using average velocities derived from velocity analyses routinely done as part of the CMP processing. Where conditions allow, acquisition of accurate interval velocities, from several intra-overburden reflectors with sufficient move-out, are possible.  Spectral ratio methods HVSR is a rapid and inexpensive method that can be used to locate significant seismic impedance boundaries in the subsurface. In the Ottawa area, this boundary is on the top of glacial materials or the bedrock surface. It is well known that the amplitude of the fundamental spectral peak can commonly be viewed as the amplification at a soft-soil site and that the frequency or period of the peak is associated with the fundamental resonance of the site. We have shown
Published by NRC Research Press

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

tio shown in Fig. 7). The above-mentioned large soil amplification factors and the fundamental resonance frequency of about 0.77 Hz were observed for all 16 recorded earthquakes. Pugin et al. (2007) also obtained similar high amplification factors for the same soil site. Using the value of F0 = Vsav/4H, where H is the soil thickness. As noted above in Fig. 9, a sharp fundamental resonance peak occurs at 0.77 Hz, whereas this peak is not evident on the vertical component ratio. Similar amplifications have been found for other temporary seismic stations that we have installed on thick soil sites in the area. This unusually large seismic soil amplification for the weak motions cannot be adequately explained from impedance contrasts alone, since the largest impedance ratio is approximately 45:1; however, the site is sufficiently remote from the edge of the buried bedrock valley that a one-dimensional (1-D) response should occur (Bard and Bouchon 1985). It is possible that more detailed interval velocitydepth data may be required to develop more accurate 1-D models. It should be noted that the above-mentioned observations are for a linear elastic soil response. Soil nonlinearity is an important issue when the level of shaking exceeds a threshold level; this behaviour has been studied for decades (Chang et al. 1989; Chin and Aki 1991; Beresnev and Wen 1996; Frankel et al. 2002; Tsuda et al. 2005). Nonlinear effects of soft soils are expected to decrease the amplification effects as well as shift the energy to longer periods, relative to the weak-motion case. Finn and Wightman (2003) describe the soil amplification factors (foundation factors), which were implemented in the NBCC (2005) provisions, as functions of NEHRP site class and input spectral acceleration. For the Ottawa area, amplification factor differences between class A and class E are approximately 4:1 for the design ground motion (1:2475 year return period). It must be noted that the measurements

468

Can. Geotech. J. Vol. 48, 2011 Fig. 10. All average shear wave velocity data from surface reflection sites, landstreamer profiles, and downhole surveys. Note the elevated velocities in the top 010 m of the ground surface where freezethaw cycles have led to compaction or densification of the soil.

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

(Fig. 8) a systematic variance between the HVSR measured from ambient noise and the resonant peak measured from estimates of soft-soil thickness and average shear wave velocity. Although the nature of the variance is not well studied to date, we have gathered sufficient data so that, within the city of Ottawa, one parameter can be predicted from the other across a broad range of fundamental site periods. Preliminary modeling of the H/V ratio using Rayleigh waves suggests that the systematic difference may result from a significant shear wave velocity gradient within the post-glacial sediments. A thorough study of this postulate is beyond the scope of this paper; however, the practitioner should note that a model consisting of a single soil layer overlying bedrock may yield only approximate estimates of the fundamental site period, despite the very large impedance contrast at the bedrock surface.  Earthquake recordings Installation of two broadband seismic stations (one on a thick soil site and one on a nearby bedrock outcrop) provided valuable information on an unusual seismic soil amplification for weak motion (up to 100 times) from 16 earthquakes. Weak-motion recordings from local and remote earthquakes at both soil and rock sites indicate that site amplification is an important concern in geotechnical practice in the Ottawa area, since 65% of area within the city of Ottawa is located on loose late or post-glacial sediments with very low shear wave velocities. In those parts of Ottawa where thick overburden exists, there is a possibility of a site resonance amplification phenomenon at the fundamental frequency. This could be of concern during the design and construction of tall structures in high-earthquake-hazard areas. The phenomenon of seismic soil amplification by soft sediments has been observed in several locations in the Ottawa region and studied by Hunter et al. (2002, 2008), Hunter and Motazedian (2006), Motazedian et al. (2006), Adams (2007), and Pugin et al. (2007). Shear wave velocities of soil and rock Velocitydepth relationships for post-glacial soils were developed both at citywide and site-specific levels. To establish an average shear wave velocity versus depth relationship throughout the city, reflection and refraction information from surface sites, landstreamer velocity analyses, and downhole velocity measurements were compiled as shown in Fig. 10. Because of the influence of the variable thickness of the high-velocity surface layer, and the lack of data at extreme depth, data points between depths of 10 and 100 m were used in the analysis. The citywide average shear wave velocity was determined to be Vsav 123:86 0:880z 20:3 m=s 1 standard deviation for 10 < z > 100 m where z is the depth of the surface layer. Details of the data editing and analyses are given by Hunter et al. (2010). Close inspection of average shear wave velocities between 10 and 100 m in depth at individual sites indicated that all Vsav versus depth curves for specific

sites exhibited linear functions similar to the citywide expression given above. Approximately 65 shear wave measurement sites within the city yielded reliable measurements of glacial deposits; no evidence of velocity variation with depth within these deposits or with total thickness was found. Hence, an arithmetic average velocity of 580 m/s was established for glacial sediments, with a standard deviation of 174 m/s. Approximately 402 determinations of bedrock velocity were obtained within the project area, and the data were subdivided by rock type to produce average velocities as given in Table 2. Combined data from all rocks types gave a mean of 2700 m/s with a standard deviation of 680 m/s. We have merged all shear wave velocity information with the three surficial geological or geotechnical units identified in each of the 21 000 boreholes database. To obtain unique average shear wave velocity depth functions (with statistical error limits) from ground surface to bedrock for each borehole, we utilized the nearest ground-truth velocity data and the inverse distance weighting (IDW) technique. The details of the geostatistical analyses are given in Hunter et al. (2010). Figure 11 shows a typical example of shear wave velocity depth variation assigned to geological units for a particular borehole site.
Published by NRC Research Press

Motazedian et al. Table 2. Tabulation of bedrock shear wave velocities of Paleozoic sedimentary and Pre-Cambrian metamorphic rocks. Bedrock type Dolomite and limestone Limestone and shale interbed Sandstone and dolomite interbed Pre-Cambrian Nepean sandstone Shale Subtypes merged Dolomite Limestone Interval velocity (m/s) 2890 2815 2808 2783 2328 2166 Standard deviation (m/s) 675 580 682 504 282 401 Number of shear wave refraction sites 182 70 20 31 8 111

469

Migmatic Metasedimentary

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Fig. 11. Typical average shear wave velocity profile (Vsav) for Ottawa area, calculated using stratigraphic thicknesses provided by borehole database. Average velocities in post-glacial materials are combined with interval velocities from glacial and bedrock materials to calculate travel-time weighted Vs30 value. Calculations are repeated three times using mean, upper, and lower velocity limits of materials to provide measure of error on Vs30 value. As shown, range can straddle two seismic site categories. z, depth of surface layer.

From the 22 000 point data set we have extracted Vs30, as defined by NBCC (2005). The results are contoured in terms of NBCC (2005) site classes A through E as given in Fig. 12. Areas of the map that are shown as classes D and E are commonly associated with thick post-glacial Champlain Sea sediments; in contrast, classes A and B are associated with rock outcrop or areas of thin soil over rock. Where data points were widely spaced, additional editing of class boundaries were done by inspection to respect known surficial geological boundaries. This map along with a funda-

mental site period map and other ancillary geological maps can be found in Hunter et al. (2010). It is important to continue investigating ground motion amplification in the Ottawa region as well as other locations, ` such as Montreal, Trois-Rivieres, and the city of Quebec where Champlain Sea sediments are widespread. The unusually high amplification of small-strain earthquakes on these soils suggests the importance of damping and nonlinear soil response for large-strain events. It may also be possible that some of the factors that affect small-strain amplification
Published by NRC Research Press

470

Can. Geotech. J. Vol. 48, 2011

Fig. 12. Vs30 map depicting seismic site classes within the city of Ottawa based on Vs30, as specified in the 2005 NBCC, www.carleton.ca/ ~dariush/Microzonation/main.html (modified from Hunter et al. 2010).

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

(e.g., velocity gradients and resonance) may also be present for large-strain events. It may be possible that the proper amplification factors for the Ottawa area may exceed those factors given in NBCC (2005), since the background studies given by Finn and Wightman (2003) were based on data observed for soilbedrock models that may not be applicable in some regions of eastern Canada.

of Carleton University and Robert Burns, Ron Good, and Tim Cartwright of the Geological Survey of Canada for field data acquisition and data interpretation support.

References
Adams, J. 2007. Soil amplification in Ottawa from urban strong ground motion records. In Proceedings of the Ninth Canadian Conference on Earthquake Engineering, Ottawa, Ont., 2629 June 2007. Canadian Association for Earthquake Engineering. Paper No. 1162. Adams, J., and Halchuk, S. 2003. Fourth generation seismic hazard maps of Canada: values for over 650 Canadian localities for the 2005 National Building Code of Canada. Earthquakes Canada, Geological Survey of Canada, Ottawa, Ont. Open File 4459. Bard, P.Y. 1999. Microtremor measurements: a tool for site effect estimation? In The effects of surface geology on seismic motion. Edited by K. Irikura, K. Kudo, H. Okada, and T. Sasatani. Balkema, Rotterdam, the Netherlands. Bard, P.Y., and Bouchon, M. 1985. The two-dimensional resonance of sediment-filled valleys. Bulletin of the Seismological Society of America, 75: 519541. Belanger, J.R. 1998. Urban geology of Canadas National Capital Area [online]. In Urban geology of Canadian cities. Edited by P.F. Karrow and O.L. White. Geological Association of Canada
Published by NRC Research Press

Acknowledgements
The researchers of the Canadian Seismic Research Network gratefully acknowledge the financial support of the Natural Sciences and Engineering Research Council of Canada (NSERC) under the Strategic Research Networks program. This work has also been made possible through funding from the Natural Hazards Program of the Geological Survey of Canada, ORDCF (The Ontario Research and Development Challenge Fund), and POLARIS (Portable Observatories for Lithospheric Analysis and Research Investigating Seismicity). The authors would like to thank the anonymous reviewers for their constructive comments. The authors would like to thank Danika Muir, Adam Jones, Amanda Landriault, Viktor Ter-Emmanuilyan, Alexander Duxbury, Raymond Caron, Laura Dixon, and Michal Kolaj

Motazedian et al. Special Paper 42. Geological Association of Canada, St. Johns, N.L. pp. 365384. Available from gsc.nrcan.gc.ca/urbgeo/ natcap/index_e.php [accessed 2010]. Beresnev, I.A., and Wen, K.L. 1996. Nonlinear soil response a reality? Bulletin of the Seismological Society of America, 86: 19641978. Borcherdt, R.D. 1970. Effects of local geology on ground motion near San Francisco Bay. Bulletin of the Seismological Society of America, 60: 2961. BSSC. 1997. NEHRP recommended provisions for seismic regulations for new buildings and other structures. Part 1: provisions (FEMA 302). Building Seismic Safety Council (BSSC), Federal Emergency Management Agency, Washington, D.C. 290 pp. Castellaro, S., and Mulargia, F. 2009. The effect of velocity inversions on H/V. Pure and Applied Geophysics, 166(4): 567592. doi:10.1007/s00024-009-0474-5. Chang, C.Y., Power, M.S., Tang, Y.K., and Mok, C.M. 1989. Evidence of nonlinear soil response during a moderate earthquake. In Proceedings of the 12th International Conference on Soil Mechanics and Foundation Engineering, Rio de Janeiro, Brazil, 13 18 August 1989. Vol. 3, pp. 14. Chin, B.H., and Aki, K. 1991. Simultaneous study of the source, path, and site effects on strong ground motion during the 1989 Loma Prieta earthquake: a preliminary result on pervasive nonlinear site effects. Bulletin of the Seismological Society of America, 81: 18591884. Clark, K.M., and Khadilkar, J. 1991. Characterization and mapping of earthquake shaking for seismic zonation. In Proceedings of the Fourth International Conference on Seismic Microzonation, Stanford University, Stanford, Calif., 2529 August 1991. Edited by C. Wikander. Earthquake Engineering Research Institute, Oakland, Calif. Vol. 3, pp. 261268. Crow, H.L., Hunter, J.A., Pugin, A.J.-M., and Motazedian, D. 2010. Application of near-surface geophysical techniques for earthquake microzonation mapping in the Ottawa, Ontario Region. In Proceedings of the 2009 Joint Assembly, American Geophysical Union, Toronto, Ont., 2427 May 2009. American Geophysical Union (AGU), Washington, D.C. Abstract NS32A02. Dobry, R., Oweis, I., and Urzua, A. 1976. Simplified procedures for estimating the fundamental period of a soil profile. Bulletin of the Seismological Society of America, 66: 12931321. Eden, W.J., and Crawford, C.B. 1957. Geotechnical properties of Leda clay in the Ottawa area. In Proceedings of the 4th International Conference of the International Society of Soil Mechanics and Foundation Engineering, London, England. Butterworths Scientific Publications, London. Vol. I, pp. 2227. Finn, W.D.L., and Wightman, A. 2003. Ground motion amplification factors for the proposed 2005 edition of the National Building Code of Canada. Canadian Journal of Civil Engineering, 30(2): 272278. doi:10.1139/l02-081. Finn, W.D.L., Onur, T., and Ventura, C.E. 2004. Microzonation: developments and applications. In Recent advances in earthquake geotechnical engineering and microzonation. Edited by A. Ansal. Kluwer Academic Publishers, Dordrecht, the Netherlands. pp. 326. Frankel, A.D., Carver, D.L., and Williams, R.A. 2002. Nonlinear and linear site response and basin effects in Seattle for the M 6.8 Nisqually, Washington, earthquake. Bulletin of the Seismological Society of America, 92(6): 20902109. doi:10.1785/0120010254. Galperin, E.I. 1985. Vertical seismic profiling and its exploration potential. In Modern approaches in geophysics. Springer, Dordrecht, the Netherlands. Vol. 1, 468 pp. Hadjian, A.H. 2002. Fundamental period and mode shape of

471 layered soil profiles. Soil Dynamics and Earthquake Engineering, 22(912): 885891. doi:10.1016/S0267-7261(02)00111-2. Hunter, J.A., and Motazedian, D. 2006. Shear wave velocity measurements for soft soil earthquake response evaluation in the Eastern Ottawa Region, Ontario, Canada. SAGEEP Conference Bulletin. Environmental and Engineering Geophysical Society (EEGS), Denver, Colo. Hunter, J.A., Pullan, S.E., Burns, R.A., Gagne, R.M., and Good, R.L. 1984. Shallow seismic reflection mapping of the overburden-bedrock interface with the engineering seismograph some simple techniques. Geophysics, 49(8): 13811385. doi:10. 1190/1.1441766. Hunter, J.A., Woeller, D.J., and Lunternauer, J.L. 1991. Comparison of surface, borehole and seismic cone penetrometer methods of determining the shallow shear wave velocity structure in the Fraser River Delta, British Columbia. Current Research, Part A, Geological Survey of Canada, Ottawa, Ont. Paper 91-1A, pp. 2326. Hunter, J.A., Pullan, S.E., Burns, R.A., Good, R.L., Harris, J.B., Pugin, A., Skvortsov, A., and Goriainov, N.N. 1998a. Downhole seismic logging for high-resolution reflection surveying in unconsolidated overburden. Geophysics, 63: 13711384. doi:10. 1190/1.1444439. Hunter, J.A., Burns, R.A., Good, R.L., and Pelletier, C.F. 1998b. A compilation of shear wave velocities and borehole geophysical logs in unconsolidated sediments of the Fraser River Delta, British Columbia. Geological Survey of Canada, Ottawa, Ont. Open File D3622. Hunter, J.A., Douma, M., Burns, R.A., Good, R.L., Pullan, S.E., Harris, J.B., Luternauer, J.L., and Best, M.E.. 1998c. Testing and application of near-surface geophysical techniques for earthquake hazards studies, Fraser River Delta, British Columbia. In Geology and natural hazards of the Fraser River Delta, British Columbia. Edited by J.J. Clague, J.L. Luternauer, and D.C. Mosher. Geological Survey of Canada Bulletin, 525: 123145. Hunter, J.A., Benjumea, B., Harris, J.B., Miller, R.D., Pullan, S.E., Burns, R.A., and Good, R.L. 2002. Surface and downhole shear wave seismic methods for thick soil site investigations. Soil Dynamics and Earthquake Engineering, 22(912): 931941. doi:10. 1016/S0267-7261(02)00117-3. Hunter, J.A., Burns, R.A., Good, R.L., Aylsworth, J.M., Pullan, S.E., Perret, D., and Douma, M. 2007a. Borehole shear wave velocity measurements of Champlain Sea sediments in the Ottawa Montreal region. Geological Survey of Canada, Ottawa, Ont. Open File 5345. Hunter, J.A., Motazedian, D., and Aylsworth, J. 2007b. Near surface geophysical mapping techniques for soft soil earthquake ground motion response in the Ottawa region some recent results. POLO Workshop, London, Ont. Hunter, J.A., Motazedian, D., Brooks, G., Lamontagne, M., Crow, H., Cartwright, T., Pugin, A., and Pyne, M. 2008. Earthquake hazard mapping in the Ottawa area, 2008. In 4th Canadian Con ference on Geohazards, Universite Laval, Quebec, Que., 2024 May 2008. Hunter, J.A., Crow, H.L., Brooks, G., Motazedian, D., et al. 2010. Seismic site classification and site period mapping in the Ottawa area using geophysical methods. Geological Survey of Canada. Open File 6273. Interpex. 2008. IXSeg2Segy v3.30 [computer program]. Available from www.interpex.com [accessed 2010]. Kaila, K.L., and Narain, H. 1970. Interpretation of seismic refraction data and the solution of the hidden layer problem. Geophysics, 35(4): 613623. doi:10.1190/1.1440119.
Published by NRC Research Press

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

472 Kassenaar, D. 2008. VIEWLOG Systems [online]. Available from www.interlog.com/~dirk/viewlog.html [accessed 2010]. Khaheshi Banab, K., and Motazedian, D. 2010. On the efficiency of the multi-channel analysis of surface wave method for shallow and semi-deep loose soil layers. International Journal of Geophysics, Vol. 2010, Article ID 403016, 13 pp.: doi:10.1155/ 2010/403016. Levson, V.M., Monahan, P.A., Meldrum, D.G., Watts, B.D., Sy, A., and Yan, L. 1998. Seismic microzonation in the Pacific Northwest, with an example of earthquake hazard mapping in Southwest British Columbia. In A paradox of power; voices of warning and reason in the geosciences. Edited by C.W. Welby and M.E. Gowan. Reviews in Engineering Geology, 12: 7588. Madera, G.A. 1970. Fundamental period and amplification of peak acceleration in layered systems. MIT Press, Cambridge, Mass. Research Report R 70-37 [June]. Motazedian, D., and Hunter, J.A. 2008. Development of an NEHRP map for the Orleans suburb of Ottawa, Ontario. Canadian Geotechnical Journal, 45(8): 11801188. doi:10.1139/T08051. Motazedian, D., Hunter, J.A., Khaheshi Banab, K., Plastow, G., and Jones, A. 2006. Microzonation measurements for Ottawa city. In Proceedings of the 2006 SSA Eastern Section Annual Meeting, Ottawa, Ont., 24 October 2006. Seismological Society of America (SSA), El Cerrito, Calif. Nakamura, Y. 1989. A method for dynamic characteristics estimation of subsurface using microtremor on the ground surface. Railway Technical Research Institute (Quarterly Report), 30(1): 2530. NBCC. 2005. Canadian commission on building and fire codes. National Building Code of Canada (NBCC), National Research Council Canada, Ottawa, Ont. Neave, K.G., and Pullan, S.E. 1989. Shallow seismic reflections using SV Waves. Geological Survey of Canada, Ottawa, Ont. Paper 89-1F, pp. 6164. NEHRP. 1994. Recommended provisions for seismic regulations of new buildings: Part 1, provisions. FEMA 222A. National Earthquake Hazards Reduction Program (NEHRP), Federal Emergency Management Agency, Washington, D.C. Okamoto, S. 1973. Introduction to earthquake engineering, University of Tokyo Press, Tokyo, Japan. Park, C.B., Miller, R.D., Ryden, N., Xia, J., and Ivanov, J. 2005. Combined use of active and passive surface waves. Journal of Environmental & Engineering Geophysics, 10(3): 323334. doi:10.2113/JEEG10.3.323. Pugin, A., Hunter, J.A., Motazedian, D., and Khaheshi Banab, K. 2007. An application of shear wave reflection landstreamer technology to soil response evaluation of earthquake shaking in an

Can. Geotech. J. Vol. 48, 2011 urban area, Ottawa, Ontario. SAGEEP Conference Bulletin. Environmental and Engineering Geophysical Society (EEGS), Denver, Colo. Pugin, A.J.-M., Pullan, S.E., and Hunter, J.A. 2008. SV-wave and P-wave high resolution seismic reflection using vertical impacting and vibrating sources. In Proceedings of SAGEEP08 (Symposium on the Application of Geophysics to Engineering and Environmental Problems), Philadelphia, Pa., 610 April 2008. Environmental and Engineering Geophysical Society (EEGS), Denver, Colo. Pullan, S.E., Hunter, J.A., and Neave, K.G. 1991. Shallow shearwave reflection tests. In Proceedings of the Annual Meeting of the Society of Exploration Geophysicists, Extended Abstracts, pp. 380382. SESAME. 2004. Guidelines for the implementation of the H/V spectral ratio technique on ambient vibrations measurements, processing and interpretation [online]. Site effects assessment using ambient excitations (SESAME), European Commission, European Research Project WP12, Project No. EVG1-CT-200000026 SESAME, Deliverable D23.12, March 2006. Available from sesame-fp5.obs.ujf-grenoble.fr [accessed 2010]. Smolka, A., and Berz, G. 1991. Seismic zoning in earthquake insurance: past and future. In Proceedings of the Fourth International Conference on Seismic Microzonation, Stanford University, Stanford, Calif., 2529 August 1991. Edited by C. Wikander. Earthquake Engineering Research Institute, Oakland, Calif. Vol. 3, pp. 245252. Tsuda, K., Archuleta, R., and Steidl, J. 2005. Confirmation of nonlinear site response: case study from 2003 and 2005 Miyagi-Oki earthquakes. In International workshop on strong ground motion prediction and earthquake tectonics in urban areas. pp. 115118. Warrick, R.E. 1974. Seismic investigation of a San Francisco bay mud site. Bulletin of the Seismological Society of America, 64(2): 375385. Whiteley, R.J., and Greenhalgh, S.A. 1979. Velocity inversion and the shallow seismic refraction method. Geoexploration, 17(2): 125141. doi:10.1016/0016-7142(79)90036-X. Williams, R.A., Odum, J.K., Stephenson, W.J., and Worley, D.M. 2000. Surface seismic reflection/refraction measurements of Pand S-wave velocities in the Memphis, TN, region. Seismological Research Letters, 2: 71113. Williams, R.A., Wood, S., Stephenson, W.J., Odum, J.K., Meremonte, M.E., Street, R., and Worley, D.M. 2003. Surface seismic refraction/reflection measurement determinations of potential site resonances and the areal uniformity of NEHRP site class D in Memphis, Tennessee. Earthquake Spectra, 19(1): 159189. doi:10.1193/1.1543161.

Can. Geotech. J. Downloaded from www.nrcresearchpress.com by CARLETON UNIV on 06/06/11 For personal use only.

Published by NRC Research Press

Você também pode gostar